Anda di halaman 1dari 129

Lecture Notes

Advanced Dynamics
Alejandro L. Garcia
San Jose State University

December 5, 2012

PHYSICS 205 / Fall 2012

c
2012
by Alejandro L. Garcia
Department of Physics, San Jose State University, San Jose CA 95192-0106
Creative Commons Attribution- Noncommercial-Share Alike 3.0 United States License

Preface
These lecture notes should supplement your notes, especially for those of you (like myself) who are poor
note takers. But I dont recommend reading these notes while I lecture; when you go to see a play you
dont take the script to read along. Continue taking notes in class but relax and dont worry about
catching every little detail. WARNING: THESE NOTES ARE IN DRAFT FORM AND
PROBABLY HAVE NUMEROUS TYPOS; USE AT YOUR OWN RISK.
Alejandro L. Garcia

Chapter 1

Equations of Motion
Generalized Coordinates (1)
Lecture 1
We begin with mechanical systems composed of one or more particles; rigid body mechanics will be
considered in Chapter VI.
. Sometimes well use
A particle has position r, velocity v = dr/dt = r , and acceleration, a = v = v
a Cartesian coordinate system so

r = xi + yj + z k
but well often also use other systems (polar, spherical, etc.) or not specify any specic coordinates.
The interactions on the particles are:
Explicit forces (e.g., gravity)
Implicit constraints (e.g., rigid supports)
The simple pendulum is a good example with both types of interactions. We will only consider constraints that are holonomic, that is, they may be expressed as
f (r1 , . . . , rN , t) = 0
For example, for a simple pendulum of length the constraint is |r| = 0. An example of a nonholonomic constraint would be for a rigid ball bouncing o a rigid surface since the constraint is an
inequality (distance between the center of the ball and the surface is greater than or equal to the
radius of the ball). Constraints may be independent of time (scleronomous) or dependent on time
(rheonomous).
A system of N particles has 3N Cartesian coordinates but due to implicit constraints these may
not all be independent. Suppose that, given the constraints, the positions of the particles are uniquely
specied by the generalized coordinates, q1 , q2 , . . . , qs , where s 3N . The system then has s degrees
of freedom.
Example For a system of two particles (N = 2) the cartesian coordinates are (x1 , y1 , z1 , x2 , y2 , z2 ).
Suppose that the particles are constrained to move in the xy-plane (see Figure 1.1); then the number
of degrees of freedom is reduced from 6 to 4.
I

encourage you to introduce these terms into your daily conversation (e.g., My love for you is scleronomous.)

CHAPTER 1. EQUATIONS OF MOTION

Figure 1.1: Two masses in the xy-plane with a constraint of xed separation .
Lets introduce a further constraint that the separation between the particles is xed to be a distance
, that is,

(x2 x1 )2 + (y2 y1 )2 =
We can now take the s = 3 generalized coordinates to be,
q 1 = x1 ,

q2 = y1 ,

q3 = = arctan

x2 x1
y2 y1

where is simply the angle between the +y-axis and the line connecting the particles.
We can also express the original coordinates in terms of the generalized coordinates,
x1 = q 1 ,

y1 = q2 ,

x2 = q1 + sin q3 ,

y2 = q2 + cos q3

This allows us to relate the Cartesian velocities to the generalized velocities, for example,
x 2 = q1 + q3 cos q3
Expressions like this are convenient for transforming quantities like kinetic and potential energy, which
are easier to express in Cartesian coordinates, into functions of the generalized coordinates.
Note that the qs can be a mixture of position, angle, or even more complicated forms. They do not
have to have units of length; thats what we mean by generalized.
One of the tasks (though not the only one) of classical mechanics is to determine the motion of a
system from a given initial condition. That is, wed like to be able to nd q1 (t), . . . , qs (t) given the
initial positions, q1 (0), . . . , qs (0), and velocities, q1 (0), . . . , q(0).

The general approach for this task will


be,
Step A Specify the implicit interactions (i.e., constraints) and establish the convenient generalized
coordinates q1 , . . . , qs .
Step B Specify the explicit interactions (i.e., forces) and establish the mechanical properties by formulating the Lagrangian for the system, L(q1 , . . . , qs , q1 , . . . , qs , t).

CHAPTER 1. EQUATIONS OF MOTION

Step C Using the Lagrangian, obtain the equations of motion, which will be a set of ordinary dierential equations (ODEs) of the form, fj (qi , qi , qi , t) = 0 for i, j = 1, . . . , s.
Step D Solve this set of second order ODEs, xing the 2s constants in the solution using the initial
conditions. For complicated systems the solution is obtained numerically.

Example Lets take our previous example and simplify it further to formulate the simple pendulum.
Our mechanical system is a particle of mass m constrained to move in the xy-plane at a xed distance
from the origin; there is a constant force on the particle equal to mgj. Following the recipe outlined
above,
Step A: To satisfy the constraint well use as our generalized coordinate q = where the angle is
from the +y-axis. Notice that q = 0 is highest position and q = is the lowest position. Comparing
with our previous example were taking q1 = 0, q2 = 0, q3 = q.
Step B: The Lagrangian for this system happens to be,
L(q, q,
t) =

1 2 2
m q mg cos q
2

Later youll see how this is formulated; note that the rst term is given by the rotational kinetic energy
and the second is from the gravitational potential energy. For this system the Lagrangian does not
depend explicitly on time.
Step C: In the next section well show that the equations of motion are obtained from Lagranges
equation,
(
)
d L
L

=0
dt q
q
For the Lagrangian in our example the equation of motion turns out to be,
m2 q mg sin q = 0
Later in this chapter youll see how we obtain this from the Lagrangian.
Step D: This ODE is easy to solve numerically but not so simple to solve analytically. However,
suppose were interested in the motion of small oscillations when the pendulum swings near the lowest
position, q0 = . We can introduce a new coordinate, x = q q0 and make use of the small angle
approximation,
sin q = sin(x + ) = sin x x
since we assume that x is a small displacement. The equation of motion simplies to,
g
x
= x

so the solution is,


x(t) = C1 sin(t) + C2 cos(t)

where = g/. The constants C1 and C2 are xed by the initial conditions. For example, if x(0) = 0 ,
x(0)

= 0 then C1 = 0, C2 = 0 . Well use the small oscillation approximation extensively in Chapter V


for a variety of systems.
This

x is not the Cartesian coordinate.

CHAPTER 1. EQUATIONS OF MOTION

We wont always do all of these steps when studying a particular system. In Chapter I we focus on
steps A and B but will often skip the explicit calculation of step C since its just turning the crank,
evaluating derivatives of the Lagrangian. Step D is the hardest and in Chapters II through V well
explore a variety of tools for either solving the equations of motion approximately (as in the example
above) or for predicting certain properties, such as using conservation of energy.
Finally, we will discover some unexpected properties in mechanical systems. Lets return to the
example weve been looking at and modify it so that the pendulums pivot point oscillates up and
down. That is, our original generalized coordinates are rewritten as, q1 = 0, q2 = f (t), q3 = q, where
f (t) is a given periodic function (e.g., f (t) = a sin a t). The Lagrangian is easy to formulate and
well study this system in sections 27 and 30. The results will surprise you; for a preview, see:
http://www.youtube.com/watch?v=rwGAzy0noU0 .

Principle of Least Action (2)


Lecture 2
Consider a mechanical system with generalized coordinates q1 , . . . , qs and velocities q1 , . . . , qs (or briey,
q and q).
The Principle of Least Action states that for every mechanical system there exists a function,
L(q, q,
t) = L(q1 , . . . , qs , q1 , . . . , qs , t)
called the Lagrangian from which we dene the action, S, as
t2
S=
L(q, q,
t)dt
t1

This action is a minimum when the integral is evaluated taking the path from q (1) = q(t1 ) to q (2) = q(t2 )
that the mechanical system follows by its physical motion.
For example, consider free-fall motion in one-dimensional motion (e.g., a particle falling straight
downward). Lets look at two dierent paths, q A and q B , which have the same starting and ending
positions. Path A happens to be the correct physical motion while Path B is the kind of motion you
might see in a cartoon (the particle remains at rest for a time and then falls with constant speed).
These two paths are illustrated in Fig. 1.2.
The Lagrangian for this system (as well later derive) happens to be,
L(q, q,
t) = 21 mq2 mgq
Notice that the rst term is the kinetic energy and the second term is the gravitational potential energy.
By the Principle of Least Action,
t2
t2
A
A
L(q (t), q (t), t)dt
L(q B (t), qB (t), t)dt
t1

t1

While this integral form of Hamiltons principle is beautiful and fundamental, its not very practical
if we want to nd q A (t), the physical trajectory of the motion. Landau and Lifshitz show that by the
calculus of variations the Principle of Least Action leads to the result that q A (t) is the solution to
Lagranges equations,
d L
L

=0
i = 1, . . . , s
dt qi
qi
Also

known as Hamiltons Principle.

CHAPTER 1. EQUATIONS OF MOTION

Figure 1.2: Illustration of two possible paths, A and B, for falling motion.
Ill sketch out the proof at the end of this lecture but rst lets see some examples of how Lagranges
equations give us the ordinary dierential equations (ODEs) that are the equations of motion for the
dynamics.
Example Lets return to the example of free-fall motion. Well evaluate each term in Lagranges
equations step-by-step,
L
q
(
)
d L
dt q
L
q
so

)
(1 2
mq mgq = mq
2
q

= m
q
= mg
d L L

=0
dt q
q

gives the ODE,


m
q + mg = 0
or simply, q = g. The general solution of this ODE is
q(t) = 12 gt2 + C1 t + C2
where the constants C1 and C2 may be xed by the initial conditions, q(0) and q(0).

Example Lets consider a more complicated example: the co-planar double pendulum (see pg. 11 in
Landau and Figure 1.3). The generalized coordinates are the angles from the vertical, that is, q1 = 1

CHAPTER 1. EQUATIONS OF MOTION

Figure 1.3: Double pendulum illustration.


and q2 = 2 . Landau writes the Lagrangian for this system as,
L =

+ m2 )21 21 + 21 m2 22 22
+m1 1 2 1 2 cos(1 2 )
1
2 (m1

+(m1 + m2 )g1 cos 1 + m2 g2 cos 2


Later well see how to construct this Lagrangian ourselves but for now well take it as given.
There are two equations of motion; lets work out the rst one by evaluating
d L
L

=0
dt 1
1
The expressions are a bit messy so lets do this one step at a time. First,
L
= (m1 + m2 )21 21 + m2 1 2 2 cos(1 2 )
1
so

d L
= (m1 + m2 )21 21 + m2 1 2 2 cos(1 2 ) m2 1 2 2 sin(1 2 )( 1 2 )
dt 1

Finally, we need to evaluate,


L
= m2 1 2 1 2 sin(1 2 ) (m1 + m2 )g1 sin 1
1
Combining all the above into Lagranges equation gives,
(m1 + m2 )21 21 + m2 1 2 2 cos(1 2 ) m2 1 2 2 sin(1 2 )( 1 2 )
+m2 1 2 1 2 sin(1 2 ) + (m1 + m2 )g1 sin 1 = 0

CHAPTER 1. EQUATIONS OF MOTION

which is a rather complicated ODE containing 1 , 2 , 1 , 2 , 1 , and 2 . But thats not all because we
need to evaluate the second Lagrange equation, which yields a similarly complicated ODE. Together this
pair of ODEs describes the motion; although theyre too complicated to yield an analytic solution (in
fact, the motion can be chaotic) they can easily be solved numerically to compute 1 (t) and 2 (t). See
the Wikipedia entry for double pendulum for numerical animations and photographs from laboratory
experiments.
Although the general solution is rather complicated its interesting to check certain limiting cases.
For example, in the limit that 2 0 you can verify that the equation of motion becomes,
(m1 + m2 )21 1 + (m1 + m2 )g1 sin 1 = 0
or

g
1 = sin 1
1

which, as expected, is the equation for a simple pendulum. What do you get if m2 0?

General Properties of Lagrangians


The next few sections in Landau discuss how to formulate the Lagrangian for various mechanical
systems. Before doing so, lets outline some of the general properties of Lagrangians,
* If systems A and B are independent then the two systems, taken together as a single system, have
a Lagrangian, LA+B = LA + LB . This can be useful for checking limiting cases, such as neglecting the
coupling between systems.
* The Lagrangian for a system is dened up to a multiplicative constant, c. This is easily seen by
inspection from Lagranges equation,
d L
L

=0
dt qi
qi

i = 1, . . . , s

which gives the same equations of motion if we replace L cL. This constant may be viewed as xing
the units of mass.
* Similarly, the Lagrangian is dened up to an additive constant; the equations of motion are
unchanged if we replace L L + c. This is the same arbitrary additive constant that appears in
potential energy.
* Landau establishes the more general result that the Lagrangian is dened up to an additive function
of the form,
d
f (q1 , . . . , qs , t)
dt
dq1 f
dqs f
f
L+
+ ... +
+
dt q1
dt qs
t
f
f
f
+ . . . + qs
+
L + q1
q1
qs
t

L L+
=
=

This result is useful for simplifying Lagrangians by allowing us to drop any terms of this form. An
important special case is that the Lagrangian is dened up to an additive function of time, so the
equations of motion are unchanged if L L + g(t).

CHAPTER 1. EQUATIONS OF MOTION

Optional Proof Consider two Lagrangians that dier by the total time derivative of an arbitrary
function of the form f (q, t). That is, we may write them as,
L (q, q,
t) = L(q, q,
t) +

d
f (q, t)
dt

(see eqn. (2.8)). Using L we obtain the equation of motion from Lagranges equation,
d L
L

=0
dt q
q
But since L = L + df /dt,
(

d L L
d

+
dt q
q
dt
But

so

df

=
q dt
q
d L L

+
dt q
q

df
q dt

f
f
q +
q
t

d f
dt q

df
q dt

)
=0

f
q

df
q dt

)
=0

The last two terms cancel since


d f
dt q

=
=

and

df
dt

d
g
g
g(q, t) =
q +
dt
q
t
2
2
f
f
q +
q 2
qt

)
=
=

(
)
f
f
q +
q q
t
2
2
f
f
q +
q 2
qt

so the equation of motion obtained from L is the same as that obtained from L.

Derivation of Lagranges Equations


Now that were a bit familiar with Lagrangians and Lagranges equations, lets go back a few steps
and see how Lagranges equations are a consequence of the Principle of Least Action. This is only a
sketch of the derivation since details are found in every mechanics book; for variety well formulate the
derivation a bit dierently than Landau.
Take a system with a single generalized coordinate, q. Call q A (t) the physical trajectory and q B (t)
some other (unphysical) trajectory for the motion. The dierence between these two paths is
q(t) = q B (t) q A (t)
which is called the variation. Both paths start and end at the same points, that is, q A (t1 ) = q B (t1 ) and
q A (t2 ) = q B (t2 ), which means q(t1 ) = q(t2 ) = 0.

CHAPTER 1. EQUATIONS OF MOTION

The variation in the action is

S = S B S A =

t2

t1
t2

=
t1
t2

10

L(q B (t), qB (t), t)dt

t2

L(q A (t), qA (t), t)dt


t1

[L(q B (t), qB (t), t) L(q A (t), qA (t), t)]dt


[L(q A (t) + q(t), qA (t) + q(t),

t) L(q A (t), qA (t), t)]dt

t1

This expression is just begging to be Taylor expanded so lets do so,


L(q A (t) + q(t), qA (t) + q(t),

t) = L(q A (t), qA (t), t) + q

L
L
+ q
+ ...
q
q

The Principle of Least Action tells us that the action, S, is a minimum when q A is truly the physical
path so we can consider small variations and neglect higher order terms.
Combining the above gives,
]
t2 [
L
L
S =
q
dt
+ q
q
q
t1
(
)
]
t2 [
L
d
L
=
q
+
q
dt
q
dt
q
t1
To pull the q out of the square brace we perform integration by parts on the second term and get,
[
]t
]
t2 [
L 2
L
d L
S = q
+
q
dt
+
q t1
q
dt q
t1
The rst term on the right hand side is zero since q(t1 ) = q(t2 ) = 0. Furthermore, by the Principle
of Least Action S = 0; this is the same as the slope being zero at an extremum since,

df
f = f (x0 + x) f (x0 ) =
x
dx x=x0
so f = 0 if x0 is an extremum. Since S = 0 for arbitrary q(t) this implies that the expression in
the square brace is zero. By now that term in the square brace should be familiar since its Lagranges
equation.

Inertial Frames and Free Particle Lagrangian (3,4)


Lecture 3
In these two sections Landau establishes the Lagrangian for a free particle, that is, a particle with no
forces acting on it. Along the way he also obtains some fundamental results in mechanics such as the
law of inertia and galilean relativity.
The general form for the Lagrangian for a single particle is L(r, v, t). We will consider the motion
of a free particle in an inertial frame of reference and use the following three properties:
Space is homogenous; the equations of motion are unchanged if we shift the origin.
Space is isotropic; the equations of motion are unchanged if we rotate the coordinate axes.
Time is homogeneous; the equations of motion are unchanged if we shift the time origin (t = 0).

CHAPTER 1. EQUATIONS OF MOTION

11

Figure 1.4: Free particle motion as measured by two dierent coordinate systems in single frame of
reference.
Consider two coordinate systems in our inertial frame (see Figure 1.4). Well write the velocity of
the particle in spherical coordinates but the position in Cartesian coordinates. Since the equations of
motion are the same in the two coordinate systems then we have,
L(r, v, t) = L(r , v , t )
or
L(x, y, z, v, , , t) = L(x , y , z , v , , , t )
Due to the spatial shift x = x , y = y , and z = z ; due to the rotation of the coordinates =
and = ; nally, due to the time shift t = t . Yet the magnitude of the velocity is the same for the
two coordinate systems so v = v , which means that for a free particle the Lagrangian can only be a
function of v.

Law of Inertia
Recall Lagranges equations,
d
dt

L
qi

L
=0
qi

L
=0
x

Taking q1 = x, q2 = y, and q3 = z, we have,


d
dt

L
x

The two coordinate systems are in the same inertial frame; later well look at transforming from one inertial frame
to another.

CHAPTER 1. EQUATIONS OF MOTION

12

with similar expressions for y and z. We may write these component equations in a compact form as,
(
)
L
d L

=0
dt v
r
where the vector derivative notation means,
( )
( )
( )
f
f
f
f
=
i+
j+
k
r
x
y
z
which is the same as the gradient operator, f . We just established that for a free particle L(v) so
L/r = 0; from Lagranges equation this implies that,
(
)
L
d L
=0
so
= constant
dt v
v
This presents three possibilities for the free particle Lagrangian:
L = constant independent of v. But this is equivalent to L = 0 (why?), which is the null result.
L = (constant)v. But this contradicts our result that L depends only on v = |v|.
L(v) is an arbitrary function of v but v is constant in time.
The last option is the one were forced to take. Furthermore, since the direction of the velocity is absent
from the equation of motion then it must be constant so not only is v constant but so it v. This is
Newtons First Law of Motion, that is, the law of inertia.

Galilean Relativity
Landau reminds us that there isnt a single holy inertial frame of reference. Consider our original
frame of reference F and another frame of reference F such that F moves with constant velocity
V relative to F. Both are inertial frames and the coordinates in the two are related by the galilean
transformations,
r = r + Vt
v = v + V
a = a
Well make use of this result to nish the derivation of the Lagrangian for a free particle.

Free-particle Lagrangian
Consider a galilean transformation V = where is innitesimally small. The equations of motion
of a free particle are the same in the two frames of reference so L(v) and L (v ) must be equivalent. As
we know, that doesnt mean that they are identical since the Lagrangian is specied up to an additive
function of the form,
L (v )

where f (r, t) is arbitrary.

d
f dr f
f (r, t) = L(v) +

+
dt
r dt
t
f
f
= L(v) +
v+
()
r
t

= L(v) +

CHAPTER 1. EQUATIONS OF MOTION

13

The next step is to express L (v ) as L (v + g(v) ) in order to Taylor expand this and relate and
L (v ) to L(v). To determine the form of the function g from our galilean transformation we note that

(v )2 = v v

(v + ) (v + )

= v v + 2v +
= v 2 + 2v
since || is innitesimally small. Thus,

v = v 2 + 2v = v 1 + 2v /v 2 = v + v /v

where we used the fact that 1 + x = 1 + 12 x in the limit that |x| 0. We now see that for this galilean
transformation g(v) = v/v so
L (v ) = L (v + v /v) = L(v) +

v dL
v dv

()

where we performed a Taylor expansion and discarded the terms that are quadratic in . Note that
L (v) = L(v) thought L (v ) = L(v) (why?).
Comparing equations (*) and (**) we see that,
d

f (r, t) =
L(v)
r
v dv
The left hand side is only a function of r and t while the right hand side is only a function of v. The
only (non-null) solution for L(v) occurs if the left hand side is a constant so
1 d
L(v) = constant
v dv
or
L(v) = Cv 2
where C is a constant. In order to preserve our traditional denition of mass (given by F = ma) it
turns out that C = 12 m.
L(v) =

1
mv 2
2

(Free particle Lagrangian)

Finally, for N non-interacting particles the Lagrangian is


L=

1
1
2
m1 v12 + . . . + mN vN
2
2

by the additivity of Lagrangians of independent systems.


The Lagrangian for a free particle in cartesian coordinates is,
L = 21 mv 2 = 12 m(x 2 + y 2 + z 2 )
Writing Lagranges equations in these coordinates,
(
)
d L
L
=0

dt x
x
we get x
= y = z = 0, as expected from the law of inertia.

CHAPTER 1. EQUATIONS OF MOTION

14

In cylindrical coordinates we replace x and y with r and using the transformation,


x = r cos

y = r sin

To nd v 2 in these coordinates we evaluate


x = r cos r sin
y = r sin + r cos
so
v 2 = x 2 + y 2 = r 2 + r2 2
Finally, this gives the Lagrangian of a free particle in cylindrical coordinates as
L=

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

A similar calculation in spherical coordinates gives,


L=

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

There are more complicated coordinate systems but the three above are the most common.
In some cases we impose a constraint on the coordinates, for example, if a free particle is constrained
to move along the x-axis then the Lagrangian is simply,
L=

1
mx 2
2

since the constraint implies y = z = 0. Another example would be a free particle thats constrained to
move in a circle of radius R; from the result above for cylindrical coordinates,
L=

1
mR2 2
2

since the constraint implies r = R and r = 0. From Lagranges equation we nd that = 0 so


the particle moves with constant angular speed ( = constant). Lets nish the lecture with a more
complicated example of motion with a constraint.
Example A particle (mass m) is constrained such that it is a constant distance , from a pivot
point. That pivot point rotates in a circle (radius a) with constant angular velocity . The motion is
further constrained to be in the x y plane. Find the Lagragian (note that this is similar to Problem
3a, pg. 11, in Landau but without gravity).
Solution The geometry is illustrated in Figure 1.5 The Lagrangian for a particle in the x y plane
is
1
L = m(x 2 + y 2 )
2
By some basic trig we can convert to the generalized coordinate by the transformation
x = x1 + x2 = a cos t + sin
y

= y1 + y2 = a sin t cos

CHAPTER 1. EQUATIONS OF MOTION

15

Figure 1.5: Particle constrained to move a xed distance from a pivot that uniformly rotates in a circle.

CHAPTER 1. EQUATIONS OF MOTION

16

so
x = a sin t + cos
y = a cos t + sin
Thus
t)
L(, ,

1 ( 2 2 2
m a sin t 2a sin t + 2 2 cos2
2
)
+a2 2 cos2 t + a cos t + 2 2 sin2

)
1 ( 2 2
m a + 2a sin( t) + 2 2
2

This expression may be simplied further; rst, the term a2 2 may be dropped since its a constant.
Second, we may rewrite the middle term using,
d
sin( t) = sin( t)
cos( t)
dt
then we may drop the cosine term since it is of the form d/dtf (q, t). Finally, we nd that the Lagrangian
simplies to the form,
t) = am 2 sin( t) + 1 m2 2
L(, ,
2
which matches Landaus expression.

Lagrangian for a System of Particles


We already derived the Lagrangian for a single free particle, L = 12 mv 2 and by the additivity of
independent Lagrangians, for a system of N independent (i.e., non-interacting) free particles,
L=

1
1
2
m1 v12 + . . . mN vN
2
2

We now consider closed systems in which particles interact among themselves but with no external
interactions. We take the form of this interaction to be such that the Lagrangian for the closed system
of particles may be expressed as,
L =
=

1
1
2
m1 v12 + . . . mN vN
U (r1 , . . . , rN )
2
2
T (v1 , . . . , vN ) U (r1 , . . . , rN )

We call T the kinetic energy and U the potential energy. In classical, non-relativistic mechanics all
physical interactions of interest for a closed system may be represented by a potential energy of this
form.
Example The Lagrangian for two similar particles interacting by an elastic (spring) attraction is,
L=

1
1
1
mv 2 + mv 2 k(|r1 r2 | )2
2 1 2 2 2

Lecture 4

CHAPTER 1. EQUATIONS OF MOTION

17

where k is the spring constant and is the rest length (recall that the spring potential energy is U = 21 kx2
where x is the extension from the rest position).
From Lagranges equations,
d
dt

L
va

L
=0
ra

a = 1, . . . , N

and the form of the Lagrangian as L = T (v) U (r),


d T
U
=
dt va
ra
Since

va

1
mv 2
2 a

)
=

va

1
mva va
2

we have
mv a =

)
= mva

U
ra

so we identify the net force on particle a as


Fa =

U
ra

in agreement with Newtons Second Law of Motion.


Example From the earlier example of two particles with an elastic interaction, the force on particle
1 is,
(
)
U

1
2
F1 =
=
k(|r1 r2 | )
r1
r1 2
)2
1 (
= k
(x1 x2 )2 + (y1 y2 )2 + (z1 z2 )2 i + (Similar terms for y and z)
2 x1
r1 r2
= k (|r1 r2 | )
|r1 r2 |
= k(|r12 | )r12
where r12 = r1 r2 is the separation vector. Note that evaluating the force on the other particle gives
F2 = F1 , as expected from Newtons Third Law of Motion.

Generalized Coordinates
The simple form of the Lagrangian in vector form does not necessarily carry over to generalized coordinates. Given that the constraints may be written in the form,
xa = fa (q1 , . . . , qs )
with similar expressions for ya and za we have that
x a =

fa dqk
k=1

qk dt

fa
k=1

qk

qk

CHAPTER 1. EQUATIONS OF MOTION

18

To nd the expression for the kinetic energy in terms of the generalized coordinates we use,
)(
)
s
s (
s
s

fa
fa
x 2a =
qi qk =
ga,i,k (q1 , . . . , qs )qi qk
qi
qk
i=1
i=1
k=1

k=1

and similar expressions for the y and z components in


1
ma (x 2a + y a2 + za2 ) U (x1 , . . . , zN )
2 a=1
N

L=
From the above we have,

1
aik (q1 , . . . , qs )qi qk U (q1 , . . . , qs )
2 i=1
s

L=

k=1

A simple example is the Lagrangian for a free particle in cylindrical coordinates,


)
1 (
L = m r 2 + r2 2 + z 2
2
The generalized coordinates are q1 = r, q2 = , and q3 = z; by inspection
a11 = m,

a22 = mq12 ,

a33 = m

and aik = 0 if i = k. For more complicated examples, see the worked problems on pages 11 and 12 of
Landau.

Open Systems
We can extend the formulation above to open systems by the decomposition:
System A The system of interest
System B A large external system coupled to system A
System A+B A closed system that is the combination of systems A and B
Systems A and B interact but the motion of system B is assumed to be unaected (in any signicant
way) by system A (e.g., A is an apple and B is the Earth).
The Lagrangian for system A+B is
LA+B = TA (qA , qA ) + TB (qB , qB ) U (qA , qB )
Since system B is unaected by system A we may take qB (t) and qB (t) as given functions of time and
write,
LA+B

TA (qA , qA ) + TB (qB (t), qB (t)) U (qA , qB (t))

TA (qA , qA ) + TB (t) U (qA , t)

TA (qA , qA ) U (qA , t)

where we may drop the term TB (t) (why?). So the Lagrangian for an open system is like that of a
closed system except that U may be an explicit function of time (Problem 3 on page 11 of Landau is a
good example).

CHAPTER 1. EQUATIONS OF MOTION

19

A single particle moving in an external eld has a Lagrangian,


L=

1
mv 2 U (r, t)
2

where U is the potential for the external eld. The equation of motion is
mr =

U
= F(r, t)
r

where F is the force on the particle due to the external eld. If the force is constant then
U (r) = F r

(Constant force)

For example, near Earths surface the force of gravity is F = mg k so U = mgz.

Worked Examples
Landau closes the chapter with some worked examples; he leaves out many of the steps so lets do one
of them in detail.
Example (Problem 4, pg. 12) This example is a mechanical system similar to the centrifugal
governor used in steam engines (see http://en.wikipedia.org/wiki/Centrifugal_governor). Two
similar particles (mass m1 ) on the sides are connected by four rigid rods of equal length a (see Fig. 1.6)
The upper rods are linked to a xed hinge; the lower rods are linked by a hinge to a particle (mass
m2 ) that can slide up and down. The central axis spins with constant angular speed = ; by the
constraint the masses on the sides rotate at the same angular speed. Our task is to nd the Lagrangian,
t) where the generalized coordinate is the angle between the side masses and the (vertical)
L(, ,
z-axis.
The kinetic energy for a free particle in spherical coordinates is
T =

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

For the particles on the sides the constraints make r = a, r = 0, and = so the kinetic energy for
each of them is,
1
T1 = m(a2 2 + a2 2 sin2 )
2
For the particle constrained to move along the vertical axis the kinetic energy is T2 =
transform this into our generalized coordinate we use
z2 = 2a cos

so

z2 = 2a sin

which gives T2 = 2m2 a2 2 sin2 . Note that were taking the top hinge as the origin.
The potential energy for each of the particles on the sides is
U1 = m1 gz1 = m1 ga cos
For the particle moving along the vertical axis,
U2 = m2 gz2 = 2m2 ga cos

1
2
2 m2 z2 .

To

CHAPTER 1. EQUATIONS OF MOTION

Figure 1.6: Simple centrifugal governor (Problem 4, pg. 12)

20

CHAPTER 1. EQUATIONS OF MOTION

21

Putting it all together (and remembering that there are two particles on the sides) we have,
L = 2T1 + T2 (2U1 + U2 )
= m1 a2 (2 + 2 sin2 ) + 2m2 a2 2 sin2 + 2(m1 + m2 )ga cos
which matches Landaus result (but his approach is much more elegant and compact). Finally, notice
that we may rearrange the terms and write this as,
L = a2 (m1 + 2m2 sin2 )2 + 2(m1 + m2 )ga cos + m1 a2 sin2 2
The last term may be viewed as an eective potential energy for the centrifugal force
Fc =

Uc
= 2m1 2 R
R

where R is the radius vector in cylindrical coordinates; recall that the masses on the sides move in a
circle of radius R = a sin .

Chapter 2

Conservation Laws
Energy (6)
Lecture 5
In this chapter we consider various integrals of motion, which are various functions of co-ordinates and
velocities that, under certain conditions, are constants of the motion. We focus on the seven additive
integrals of motion:
Energy (scalar)
Linear momentum (3 components)
Angular momentum (3 components)
These quantities are called conserved variables and their constancy arises from homogeneity and isotropy
of space and time. Other integrals of motion are specic to particular systems (e.g., three body gravitational system) so theyre not as important.
The rst conserved variable well study is the energy. Consider a system whose Lagrangian does not
depend explicitly on time, that is, L(q, q).
Important examples are closed systems but in some cases
open systems (e.g., Problem 4, pg. 12) have Lagrangians of this form. By the chain rule,
d
L(q1 , . . . , qs , q1 , . . . , qs )
dt

L dq1
L dqs
L dq1
L dqs
+ ...
+
+ ...
q1 dt
qs dt
q1 dt
qs dt
s
s

L
L
=
qi +
qi
q

qi
i
i=1
i=1

From Lagranges equation,


d
L
=
qi
dt

L
qi

so using this in the rst summation gives,


dL
dt

(
)]
s [
s

d L
L
=
qi
qi +
dt

qi
i
i=1
i=1
{(
) }
s

d
L
=
qi
dt
qi
i=1

22

CHAPTER 2. CONSERVATION LAWS

23

Finally, collect everything to the left hand side and we may write,
[{ s (
}
]
L )
d
qi L
dt
qi
i=1
This means that dE/dt = 0 where
E=

)
s (

L
i=1

qi

qi L

is the denition of the energy.


Example For projectile motion in 2D,
L=
so

)
L
x +
y L
=
y
(
)
1
2
2
2
2
= mx + my
m(x + y ) mgy
2
1
=
m(x 2 + y 2 ) + mgy
2
(

1
m(x 2 + y 2 ) mgy
2

L
x

This example illustrates the general result that for a closed system (or one in a constant eld) the
energy is simply E = T + U . However, note that the open system in Problem 4, pg. 12, is an example of
a system in which E = T + U however E is conserved since the Lagrangian does not depend explicitly
on time.
Conservation of energy is a consequence of the homogeneity of time, that is, if the Lagrangian does
not depend explicitly on time then the equations of motion are unchanged if t t + t0 .

Momentum (7)
Consider now the case where the Lagrangian of a system is unchanged by a spatial shift of positions,
r r + where is an innitesimal vector displacement. For a Lagrangian with this property we have,
L(r1 + , . . . , rN + , v1 , . . . , vN , t) = L(r1 , . . . , rN , v1 , . . . , vN , t)
By Taylor expansion of the left hand side we have,
L(r1 + , . . . , rN + , v1 , . . . , vN , t) = L(r1 , . . . , rN , v1 , . . . , vN , t) +
Since is innitesimal we drop the higher order terms, which gives,

L
=0
r
a
a=1

L
L
+
+ (h.o.t.)
r1
rN

CHAPTER 2. CONSERVATION LAWS

24

If the Lagrangian is unchanged by spatial shift in any direction (i.e., if the direction of is arbitrary)
then
N

L
=0
()
ra
a=1
Two important results follow from this:
* For a closed system of particles the Lagrangian has the general form,
L=

2
a=1

ma va2 U (r1 , . . . , rN )

so the net force on particle a is


Fa =

U
L
=
ra
ra

From our result (*) above,


N

Fa = 0

a=1

For two particles this is simply F1 + F2 = 0 so F1 = F2 , which is Newtons Third Law of Motion.
* Lagranges equations may be written as,
(
)
d
L
L

=0
dt va
ra
so from (*)
d L
=0
dt a=1 va
N

Dene the momentum of a particle and the total momentum as,


pa =

L
va

and

P=

pa

a=1

so we have that the total momentum is constant (i.e., dP/dt = 0).


We may also dene a generalized momentum and generalized force as,
pi =

L
qi

and

Fi =

L
qi

so by Lagranges equations for generalized coordinates,


(
)
d L
L
=0

dt qi
qi
which gives us the generalized form of Newtons Second Law of Motion,
pi = Fi
Note that this result is just a way to re-write Lagranges equations in terms of pi and Fi and it applies
to arbitrary Lagrangian. If the Lagrangian does not contain a given coordinate qi then L/qi = 0,
implying that the corresponding momentum pi is constant (i.e., pi = 0). In that case we say that qi is
a cyclic coordinate.

CHAPTER 2. CONSERVATION LAWS

25

Figure 2.1: Pendulum (mass m2 ) with a pivot (mass m1 ) that slides horizontally.

Example The Lagrangian for a particle in a constant gravity eld is


L=

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

Since this Lagrangian does not explicitly depend on x or y these are cyclic coordinates which means
that
L
L
px =
= mx
and
py =
= my
x
y
are constant. Note that pz = L/ z = mz is not constant.
Example Consider the system in Problem 2, pg. 11, which is a pendulum whose pivot slides freely
on a horizontal wire (see Figure 2.1). The Lagrangian is
L=

1
1
(m1 + m2 )x 2 + m2 (2 2 + 2x cos ) + m2 g cos
2
2

By inspection x is a cyclic coordinate so its corresponding momentum,


px =

L
= (m1 + m2 )x + m2 cos
x

is constant. To understand this result, consider that the horizontal position and velocity of the pendulum
mass are,
x2 = x + sin
and
x 2 = x + cos
so the generalized momentum px may be written as,
px

(m1 + m2 )x + m2 ( cos )

(m1 + m2 )x + m2 (x 2 x)

= m1 x 1 + m2 x 2
which is the total horizontal momentum of the particles.

CHAPTER 2. CONSERVATION LAWS

26

Note that we may write this momentum as


px = X
where = m1 + m2 is the total mass and
X=

m1 x1 + m2 x2
m1 + m2

is the horizontal coordinate of the center of mass. Since px is constant the velocity of the center of
is also constant. In the next section well see more general results for the motion of the center
mass, X,
of mass.
Example Landau solves the following problem on page 16: A particle moving freely with velocity
v1 on the xy-plane encounters a steep ramp parallel to the x-axis. The potential energy of the particle
is U1 and U2 before and after the encountering the ramp, respectively. The direction of motion changes,
as shown in Figure 2.2; nd the ratio sin 1 / sin 2 .
Solution The particles Lagrangian is
{
1
U1 y 0
whereU (y) =
L = m(x 2 + y 2 ) + U (y)
U2 y < 0
2
By inspection the x component is a cyclic coordinate so the component of momentum in the x direction
is conserved. Equating this component of momentum before and after the particle passes the ramp
gives,
v1 sin 1 = v2 sin 2
By inspection the Lagrangian does not depend explicitly on time so energy is conserved so
1
1
mv 2 + U1 = mv22 + U2
2 1
2
With a little algebra this gives,

v2 = v1

1+

2
(U1 U2 ) = v1
mv12

U
T1

where T1 is the initial kinetic energy. Finally, collecting the above gives,

sin 1
U
= 1
sin 2
T1
Notice that this result holds even if the ramp is not a step function as long as the gradient is perpendicular to the x-axis (i.e., as long as x is a cyclic coordinate). Finally, this result is the Newtonian model
for the refraction of light.

CHAPTER 2. CONSERVATION LAWS

27

Figure 2.2: Particle deected by crossing a ramp.

Center of Mass (8)


Lecture 6
The total momentum for a system of N particles is,
P=

ma va

a=1

then P n
= 0;
If the Lagrangian is unchanged by a shift in the coordinates in a specied direction n
if its unchanged for all coordinate shifts then P = 0.
Consider a Galilean transformation from the current frame of reference (call it K) to another inertial
frame of reference (called K ) moving with velocity V relative to K (see Figure 2.3).
By galilean relativity the particle velocities in the two frames are related by

va = va + V
The total momentum in frame K is
P =

ma va =

a=1

ma (va + V)

a=1

ma va + V

a=1

ma

a=1

P + V

where is the total mass. Note that P = 0 if V = P/; in that case we call K the center of
mass frame of reference. In a sense, the global motion of the system is zero in this frame, though the
What

happens if U > T1 ?

CHAPTER 2. CONSERVATION LAWS

28

Figure 2.3: Galilean transformation of particle velocities in two frames of reference K and K .
particles move relative to each other. Call V the velocity of the center of mass in frame K; we may
write it as,

d a ma ra
a ma va

V=
=
=R
dt
a ma
a ma
where

ma ra
R = a
a ma

is the position of the center of mass.


Example Consider two particles moving on the x-axis with an elastic attraction; well take the
Lagrangian to be,
1
1
1
L = m1 x 21 + m2 x 22 k(x1 x2 )2
2
2
2
We expect that the total momentum is conserved since the Lagrangian is unchanged if x1 x1 + ,
x2 x2 + .
The equations of motion turn out to be,
m1 x
1 + k(x1 x2 )

m2 x
2 k(x1 x2 )

Adding them gives,


m1 x
21 + m2 x
22 = 0
so

d m1 x 1 + m2 x 2
d
= V=0
dt m1 + m2
dt

which means that the center of mass velocity is constant, as we expect since the total momentum is
conserved for this system.

CHAPTER 2. CONSERVATION LAWS

29

Figure 2.4: Innitesimal rotation of the position of a particle (not to scale).

Angular Momentum (9)


The derivation for conservation of angular momentum is similar to that of linear momentum except that
we consider Lagrangians that are invariant under an innitesimal rotation instead of a displacement
(See Figure 2.4). After a few manipulations, Landau arrives at the result,


ra pa = 0
a

Dene the angular momentum as


M=

ra pa

and our result may be expressed as,


d
( M) = 0
dt
If the Lagrangian is invariant under any rotation then M is constant. If the Lagrangian is invariant only
for a rotation around a specic direction, say the z-axis, then only that component of M is constant.
In frame K is the center of mass frame then

M = M + R V = M + R P

The term M is called the intrinsic angular momentum and the term RP gives the angular momentum
of the motion as a whole.

CHAPTER 2. CONSERVATION LAWS

30

Figure 2.5: Two particles moving together with equal velocities.

Example Consider two particles moving freely in the xy-plane. The z-component of angular momentum is

Mz = k M = k
ra pa
a

= m1 (x1 y 1 y1 x 1 ) + m2 (x2 y 2 y2 x 2 )
Now suppose that we take the special case that the particles move in a straight line with the same
velocity, specically, lets take x 1 = x 2 and y1 = y2 (see Figure 2.5). In that case,
Mz = y1 (m1 x 1 + m2 x 2 ) = Y X = (R V) k
This simple example demonstrates our
where R = Xi + Y j is the center of mass position and V = R.
result that

Mz = Mz + (R V) k

since obviously Mz = 0 for this system (particles are at rest in the center of mass frame). By the way,
its easy to show that Mx = My = 0 since the motion is in the xy-plane so z1 = z2 = 0.
For a given coordinate axis, if is the angle about that axis (e.g., the z-axis in cylindrical and
spherical coordinates) then
L
Mz =

a a
For example, for a free particle in cylindrical coordinates Mz = mr2 since the Lagrangian may be
written as L = 21 m(r 2 + r2 2 + z 2 ).
Example Consider a particle that moves freely in the interior of a 45 degree cone (so z = r); there
is a constant downward gravitational force acting on the particle. See Figure 2.6 Using conservation of
energy and angular momentum, express z and as a functions of z.

CHAPTER 2. CONSERVATION LAWS

31

Figure 2.6: Particle moving freely inside a 45 degree cone.


Solution We have the general result that for a closed system (or one in a constant eld) the energy
is simply E = T + U . In cylindrical coordinates the kinetic energy of a free particle,
T =

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

The potential energy in a constant gravitational eld with acceleration g is U = mgz. If you have any
doubts you can use the general denition of energy and verify that this result is correct. Using the
constraint (r = z, r = z)
the energy is,
E=

1
mz 2 2 + mz 2 + mgz
2

Since the Lagrangian does not depend explicitly on time the energy is constant.
For the angular momentum only the z component is conserved since the Lagrangian is invariant
under rotation about the z axis but not about any other axis. In cylindrical coordinates Mz = mr2

and with our constraint this is Mz = mz 2 .


By inspection we may write
Mz2
E=
+ mz 2 + mgz
2mz 2
or

E
Mz2
z =

gz
m 2m2 z 2
From our expression for Mz ,
Mz
=
mz 2

CHAPTER 2. CONSERVATION LAWS

32

Figure 2.7: Graph indicating maximum and minimum heights for particle rolling inside a cone for a
given value of angular momentum, Mz , and energy, E.
This pair of rst-order ODEs for z and may (in theory) be solved to nd z(t) and (t). The initial

conditions z(0), z(0),

(0), and (0)


x the two constants of integration of these ODEs as well as the
values of E and Mz .
Before leaving this example, lets examine some of the qualitative features about the motion that
we can extract from the result above. From the expression for z we see that as the particles height
decreases this vertical velocity may increase (due to the gravity term) or decrease (due to the angular
momentum term). Physically, the particle spins in the cone rising to a maximum height, constrained
by the gravity term, and diving to a minimum height, constrained by the angular momentum term.
The turning points occur when z = 0, which means
E=

Mz2
+ mgz = f (z)
2mz 2

This cubic equation has two real roots, zmin and zmax ; we wont bother with their explicit form but
instead just consider the graphical solution shown in Figure 2.7. For a given value of Mz the turning
points depend on the energy E, as shown in the gure. Notice that the curve of f (z) has a minimum
value, which is the case when the particle orbits in a circle of constant height.

Finally, in the absence of gravity the particle will move out towards z (with z E/m and
0) unless the initial condition is z(0)

= 0, in which case the orbit is circular; Landau considers


this system in Problem 2 on page 34. The next chapter extends the ideas in this example to obtain
properties of the motion using the constants of the motion.

CHAPTER 2. CONSERVATION LAWS

33

Mechanical Similarity (10)


Landau points out the a number of interesting results may be inferred about the motion of a system
when the potential energy is a homogeneous function of the coordinates, that is, if,
U (r1 , . . . , rn ) = k U (r1 , . . . , rn )
For example, for Newtonian gravitational attraction among n particles of mass m,
1 Gm2
2 i=1 j=1 |ri rj |
n

U (r1 , . . . , rn ) =
In this case,

1
1
Gm2
= U (r1 , . . . , rn )
2 i=1 j=1 |ri rj |

U (r1 , . . . , rn ) =

so = 1. Similarly, for a potential of the form U |ri rj |c we have k = c (e.g., for harmonic
oscillator potential k = 2).
Starting from the Lagrangian for a system of particles,
L(r, v, t) =

2
a=1

ma |va |2 U (r1 , . . . , rn )

Now consider the transformation of the system such that coordinates and the time are changed as,
ra ra

t t

This means that all of the velocities va = dra /dt are changed by a factor of / so the kinetic energy
is changed by a factor of 2 / 2 . The potential energy is multiplied by a factor of k . In other words,
L (r, (/)v, t) =

n
2 1
ma |va |2 k U (r1 , . . . , rn )
2 a=1 2

Now if we choose = 2k then L and L dier by a multiplicative constant and thus they have the
same equation of motion so the two have mechanical similarity. As Landau says, if the potential energy
of the system is a homogeneous function of degree k in the (Cartesian) coordinates, the equations of
motion permit a series of geometrically similar paths, and the times of the motion between corresponding
points are in the ratio,
1
t /t = ( /)1 2 k
where / is the ratio of linear dimensions of the two paths.
Example For the motion of free particles in a constant gravitational eld U (r) = gz
z so k = 1
and
1
t /t = ( /) 2
For example, a particles parabolic trajectory is mechanically similar for a length scale (height and
range) magnied by 16 if the time scale is magnied by a factor of 4.
Mechanical similarity is used by Hollywood to create special eects using high-speed cameras. Normal lm speed is 24 frames per second so when a scene is lmed at 96 frames per second then played

CHAPTER 2. CONSERVATION LAWS

34

Figure 2.8: Scene from Jason and the Argonauts (1963).


back at normal speed what appears to be a boulders falling o a 64 foot cli is really small rocks falling
from a 4 foot height (see Fig. 2.8). In modern lms these special eects are usually done using computer
graphics some are still produced using high-speed cameras. Furthermore, the visual eects artists who
use computer animation need to duplicate this relationship between timing and scale.

Chapter 3

Integration of the Equations of


Motion
Lecture 7
In Chapter 1 we formulated Lagrangian dynamics; in Chapter 2 we derived various important conserved
quantities, such as energy and momentum. We now consider some important cases where its possible
to solve, at least in part, the equations of motion.

Motion in One-Dimension (11)


Consider a Lagrangian of the form,
1
mx 2 U (x)
2
where x is a single coordinate. Many problems reduce to this form (e.g., motion in a central eld, 14).
The physical picture for the motion for such a Lagrangian is that of a particle sliding on a roller coaster
whose shape is the graph of U (x) versus x (see Figure 3.1).
The total energy is
1
E = mx 2 + U (x)
2
L=

Figure 3.1: Analogy between the shape of a potential and the motion of a particle sliding in a similarly
shaped landscape.

35

CHAPTER 3. INTEGRATION OF THE EQUATIONS OF MOTION

36

Figure 3.2: Motion in a potential between endpoints x1 and x2 .


Since the total energy is constant we may obtain an integral expression for the motion without making use of the equations of motion (i.e., without applying Lagranges equations to the Lagrangian).
Specically, the approach is to write
1
mx 2 = E U (x)
2
so

dx
2
=
(E U (x))
Note : E U (x)
dt
m
or
dx
dt =
2
m (E U (x))
Integrating both sides,

t=

m
2

dx
E U (x)

The initial conditions (initial position and velocity) specify the constant of integration and the value of
E.
This result has various useful corollaries, for example we may use it to nd the period of oscillation
for periodic motion. Consider the graph of a potential U (x) such as in Figure 3.2. For a given value of
energy, E and an initial position between points x1 and x2 we know that the motion is periodic between
these turning points (why?). By denition, the time it takes the particle to go from x1 to x2 is half the
period so the period of oscillation, T , is,
x2

dx

T = 2m
E U (x)
x1

Example Consider an innite square-well potential,


{
0 if 0 x L
U (x) =

otherwise
Landau

writes the period as T but Ill write T to distinguish it from the kinetic energy.

CHAPTER 3. INTEGRATION OF THE EQUATIONS OF MOTION

37

The turning points are x1 = 0 and x2 = L for all values of E; since the potential is zero in this interval,

dx
2m
=
L
T = 2m
E
E
0
Since U = 0 in the interval E = 12 mv 2 so T = 2L/v, that is, the period is the time it takes a particle
with speed v to travel a distance of 2L.
Landau works through some more complicated examples on pages 26 and 27, including the simple
pendulum without the small angle approximation. In general the resulting integrals can be dicult to
solve in closed form but are easy to evaluate numerically. One important result that has a complicated
but closed form solution is for the potential U (x) = A|x|n (see Problem 2(a) on page 27); the period of
oscillation is

( )1/n
1
1
2 2m E
(1/n)
T =
E n2
1
n
E
A
( 2 + 1/n)
where is the gamma function. An important special case is n = 2, which is the potential for simple
harmonic motion; by inspection it is the only exponent n that gives a period thats independent of E,
which means its independent of the initial amplitude.

Finding U (x) from T (E) (12)


We just saw how to get the period, T for a given potential; is it possible to do the reverse? Namely,
from knowing T (E) is it possible to obtain U (x)? The answer is yes but the potential is not unique
unless we also impose the condition that it is symmetric, say about x = 0.
After a few manipulations (see page 28) Landau arrives at the result,
x(U ) =

2 2m

T (E)dE

U E

from which we may get U (x) with U (x) = U (x). Lets see how this works in an example,
Example Consider the case for which T (E) = T0 = constant. That is, the period is independent
of the energy. In this case,
x(U )

=
=

U
T
dE
0

U E
2 2m 0
]U
T0 [
T0

2 U E =
U
0
2 2m
2m

so
U (x) =

2 2 m 2
x
T02

You should recognize this as the potential for simple harmonic motion (SHM), which is usually written
as U (x) = 12 kx2 so in our case k = 4 2 m/T02 . Recall that the period for SHM is

TSHM = 2

m
= 2
k

m
= T0
4 2 m/T02

CHAPTER 3. INTEGRATION OF THE EQUATIONS OF MOTION

38

Figure 3.3: Symmetric potential, U (x), and asymmetric potential, U (x), with the same period, T (E).
which conrms the expected result.
Landau shows that an asymmetric potential, U (x) will have the same period as a symmetric
potential, U (x) as long as the distance between points at the same energy is equal for the two potentials.
This is illustrated graphically in Figure 3.3; mathematically the condition is that
x2 (U ) x1 (U ) = 2x(U )

when

U = U.

Example A simple way to build an asymmetric potential is to take x1 (U ) = 0 for all values of U
and x2 (U ) = 2x(U ). The condition on x1 implies that the asymmetric potential U (x) = for x < 0.
The condition on x2 simply means that U (x) is twice as wide as U (x) for all x > 0 (see Figure 3.4).
As an example, consider a simple pendulum of length ; with the pivot at the origin the (symmetric)
potential may be written as,

U = mg( + y) = mg mg 2 x2
(
)
1 x2
mg mg 1
2 2
1 mg 2
=
x
2

using the small angle approximation and the Taylor expansion 1 + x = 1+ 12 x+. . .. The corresponding
asymmetric potential would be U (x) = for x < 0 and
U (x) =

1 mg
1 mg 2
(x/2)2 =
x
2
2 4

Lets interpret this result: The asymmetric potential corresponds to a pendulum that is 4 times longer
and that swings against a rigid, elastic wall located
at x = 0. We know that in the small angle
approximation the period of a pendulum is T = 2 g/ so a pendulum thats 4 times longer would
have twice the period but due to the rigid wall it only travels half the distance so U (x) and U (x) have

CHAPTER 3. INTEGRATION OF THE EQUATIONS OF MOTION

39

Figure 3.4: Simple pendulum example illustrating a symmetric potential, U (x), and an asymmetric
potential, U (x), with the same period, T (E).
the same period for all values of E (i.e., for all initial conditions). Finally, note that we may construct a
similar expression for U without using the small angle approximation but then the resulting potential
is not equivalent to an interrupted pendulum.

Reduced Mass (13)


Lecture 8
In this section and the next two we analyze in detail the two-body problem, that is, a closed system
with two interacting particles. The general form of the Lagrangian is,
L=

1
1
m1 r 21 + m2 r 22 U (|r1 r2 |)
2
2

Notice that the potential can only depend on the magnitude of the separation due to the homogeneity
and isotropy of space (both total linear momentum and angular momentum are constant).
Dene
r = r1 r2
(relative position)
m1 m2
m=
(reduced mass)
m1 + m2
In the center of mass frame of reference m1 r 1 + m2 r 2 = 0 so
m21 r 21 + m22 r 22 + 2m1 m2 (r1 r 2 ) = 0
which means that
1 2
mr
2

=
=
=
=

1
m(r1 r 2 )2
2
1 m1 m2
(r2 + r 22 2(r1 r 2 ))
2 m1 + m2 1
1 m1 m2
m21 2
m22 2
(r21 + r 22 +
r 1 +
r )
2 m1 + m2
m1 m2
m1 m2 2
1
1
m1 r 21 + m2 r 22
2
2

CHAPTER 3. INTEGRATION OF THE EQUATIONS OF MOTION

40

This simplies the Lagrangian to be


1
mr U (|r|)
2
so the two body problem is equivalent to the problem of a single particle with reduced mass m moving
in a central potential.
One last thing: Be careful to remember that
L=

r 2 = r r = r 2
For example, in cylindrical coordinates,
r 2 = r 2 + r2 2 + z 2

Motion in a Central Field (14)


Now we turn to the general problem of a single particle moving in a eld with potential U (|r|) = U (r).
The force on the particle is,
)
(
U
dU
r
F=
= U =
r
dr
where r = r/|r| is the radial unit vector. Notice that if dU/dr > 0 then the direction of F is r so the
potential is attractive; if dU/dr > 0 then the potential is repulsive.
The angular momentum M is constant (why?) and M is perpendicular to r since M = r p. This
means that the motion of the particle, r(t) is in a plane; well take it to be the xy-plane by having the
z-axis parallel to M. Using polar coordinates (i.e., cylindrical coordinates with z = z = 0) the energy
is,
1
E = m(r 2 + r2 2 ) + U (r)
2
2
Using M = Mz = mr = constant we have
E

=
=

1 2 1 M2
mr +
+ U (r)
2
2 mr2
1 2
mr + Ue (r)
2

where
Ue (r) =

1 M2
+ U (r)
2 mr2

is the eective potential.


Note that we have now reduced the problem to one-dimensional motion in a potential Ue (r), which
allows us to use the results from Section 11. For example, the motion r(t) may be found by evaluating

dr

()
t=
2
m (E Ue (r))
with the initial conditions setting the constant of integration as well as xing the values of E and M.
The eective potential leads to an eective force,
Fe =
Recall

Ue
U
M2
r = F + Fc
=
+
r
r
mr2

that Mx = My = 0 since the motion is in the xy-plane.

CHAPTER 3. INTEGRATION OF THE EQUATIONS OF MOTION

41

Figure 3.5: Sketch of eective potential Ue (r) versus r. Notice that for small r the centrifugal repulsion,
M2
2mr 2 , dominates but for large r we have Ue (r) U (r).
where Fc is the centrifugal force, which is an eective force of repulsion.
If U (r) is an attractive potential (i.e., dU/dr > 0) then graphically the eective potential is illustrated
in Figure 3.5. If the potential U (r) has an upper bound (e.g., if U 0 as r ) then the motion
is bounded if the energy is less than this upper bound. If the potential is unbounded (e.g., the space
oscillator potential, U (r) = 21 kr2 ) then the motion is bounded for all values of E. Bounded motion is
conned between r = rmin and rmax where U (rmin ) = U (rmax ) = E.
Using M = mr2 we may write d = (M/mr(t)2 )dt and then use (*) to get

M dr

=
2
r 2m(E Ue (r))
The change in angle during one period of the motion (r = rmin then r = rmax then back to r = rmin ) is
rmax
M dr

=
2
r
2m(E
Ue (r))
rmin
In general = 2 so orbits are typically not closed (see Figure 9 on page 33 in Landau). The
only power-law potentials with closed orbits are U (r) r1 (Kepler problem) and U (r) r2 (space
oscillator). We consider the former in detail in the next section (and next lecture) while Landau discusses
the latter in Problem 3 on page 70.
Example Consider the attractive central potential,
U (r) =

r2

Note that this is not the Kepler problem. For this example well take the initial condition to be such
that M 2 = 2m; for this unique case
Ue (r) =

1 M2
+ U (r) = 0
2 mr2

which greatly simplies all the calculations. Notice that since E = 12 mr 2 + Ue (r) we know that E 0.

CHAPTER 3. INTEGRATION OF THE EQUATIONS OF MOTION

42

Figure3.6: Trajectory for particle in potential U (r) = r2 for the initial conditions r(0) = 0 and
M = 2m. Plots are for short (left) and long (right) times; parameter values are E = 103 and
= 1.
First well consider the case where E > 0 and calculate the motion using (*), which simplies to,

dr
1

t=
=
(r(t) r(0))
2
2
m E
m E

so
r(t) =

2
E t + r(0)
m

We obtain the same result fromthe observation that since in this example Ue (r) = 0 then E = 12 mr 2 .
But energy is conserved so r = 2E/m is constant. Physically, the particle moves away from the origin
with constant radial velocity.
The relation between the angle and radius is found using,

M dr
M 1

=
=
+C
2
r 2mE
2mE r
where C is the constant of integration. Well take r(0) = 0 to see how the motion looks as the particle
ies away from the origin. For this initial condition we may arbitrarily set C = 0 so
=

mM 1
M 1
M
1

=
=
4E 2 t
2
2mE r
2mE
Et
m

The resulting motion is a spiral, as shown in Figure 3.6.


To nish
the example, what about the case E = 0? From the above we know that since Ue (r) = 0
that r = 2E/m so r is constant. Recall that M = mr2 so

2 1
M
=
=
mr2
m r2

CHAPTER 3. INTEGRATION OF THE EQUATIONS OF MOTION

43

Physically, the orbits are circular with a constant angular velocity that goes as r2 so their period,
goes as r2 . Compare this with the period of circular orbits in the Kepler problem for which
T = 2/,
T r3/2 by Keplers Third Law.
One nal aside: The result for the period of orbits may be obtained from the principle of mechanical
similarity (see Section 10). Specically, if we take r = ar, where a is a constant, then

= a2 U (r)
U (r ) = 2 =
(r )
(ar)2
Landau shows that when a potential scales as
U (ar) = ak U (r)
then geometrically similar paths (e.g., circular orbits) scale as
( )

t
=
t

For the potential in this example k = 2 so the period of orbits goes as r2 ; for the Kepler problem
k = 1 so the period goes as r3/2 .

Kepler Problem (15)


We now turn to an important mechanics problem, motion in a central potential where U (r) 1/r;
the gravitational and Coulomb potentials for point objects are well-known cases. We will rst consider
attractive potentials so

U (r) =
( > 0)
r
so the eective potential for the radial motion (as formulated in the previous section) is

M2
+
r
2mr2
Recall that M is constant with direction perpendicular to the plane of the motion. A sketch of the
eective potential is shown in Figure 3.7.
If E < 0 then the motion is bounded between the turning points rmin and rmax . If E 0 then the
motion is unbounded so r as t . If E = min{Ue (x)} then the radius is constant, that is,
the orbit is circular.
From conservation of angular momentum, M , and energy, E, we derived in the previous section
that,

M dr

=
2
r 2m[E Ue (r)
Ue (r) =

For our case the integration yields,

]
M r1 m/M
= arccos
+ 0
2mE + m2 2 /M 2

Taking 0 = 0 and dening the quantities,


p

M2
m

1+

(Semi latus rectum)


2EM 2
m2

(Eccentricity)

Lecture 9

CHAPTER 3. INTEGRATION OF THE EQUATIONS OF MOTION

44

Figure 3.7: Eective potential Ue (r) for the attractive potential, U (r) = r , of the Kepler problem.

Figure 3.8: Geometry of an: ellipse (left); parabola or hyperbola (right)

CHAPTER 3. INTEGRATION OF THE EQUATIONS OF MOTION

45

we may rewrite the expression above as,


r=

p
1 + e cos

which is the equation for a conic section. By inspection, for E < 0 the eccentricity e < 1; the orbit is
an ellipse (Keplers First Law) with semi-major and semi-minor axes,
a=

p
1 e2

b=

p
1 e2

The perihelion and aphelion are


rmin =

p
1+e

rmax =

p
1e

Note that e = 0 for a circular orbit.


Given that angular momentum is conserved in a central eld we may use M = mr2 to write
M
1 2
r d =
dt
2
2m

dA =

where dA is the area swept out by an orbital increment d. This result (A line joining a planet and
the Sun sweeps out equal areas during equal intervals of time) is Keplers Second Law (see Figure 3.9).
Integrating over the period of an orbit we have
T =

2m
A
M

Since the area of an ellipse is A = ab from the previous results we nd that T a3/2 , which is Keplers
Third Law. Why did Kepler express this as a relation between T and a instead of T and A? Because
the angular momentum M diers for each planet.
For E > 0 the eccentricity e > 1 and the trajectory is a hyperbola with perihelion rmin = p/(1 + e);
if E = 0 then e = 1 and the trajectory is a parabola with rmin = p/2.
Landau shows that we may formulate expressions for the position versus time but only in parametric
form. For example, for an ellipse,
r
t

= a(1 e cos )
= e sin

where = /ma3 . Note that (the eccentric anomaly) and


t (the mean anomaly) go from 0 to
2 (see Figure 3.10). Note that the period is T = 2/ = 2 m/ a3/2 ; for gravity m1 m2 and
the sun is massive so m/ (m1 + m2 )1 1/msun . The calculation of for a given time t is a
transcendental problem that was an early application for numerical root nding algorithms.
For a repulsive potential, U (r) = +/r, the results are similar as for an attractive potential except
that there is no bounded motion, that is, the trajectory is either a parabola or a hyperbola. The main
application is the study of scattering due to Coulombic repulsion.

Laplace-Runge-Lenz Vector
Well nish this chapter by deriving a special constant of the motion thats specic to the 1/r potential.
Consider the general case of a central potential; the Lagrangian is,
L=

1 2
mr U (r)
2

CHAPTER 3. INTEGRATION OF THE EQUATIONS OF MOTION

46

Figure 3.9: Illustration of Keplers three laws with two planetary orbits. (1) The orbits are ellipses,
with focal points 1 and 2 for the rst planet and 1 and 3 for the second planet. The Sun is placed in
focal point 1. (2) The two shaded sectors A1 and A2 have the same surface area and the time for planet
1 to cover segment A1 is equal to the time to cover segment A2. (3) The total orbit times for planet 1
and planet 2 have a ratio a13/2 : a23/2 . (From Wikipedia)

Figure 3.10: Parametric solution for an elliptical orbit. Note that tan(/2) =

(1 + e)(1 e) tan(/2).

CHAPTER 3. INTEGRATION OF THE EQUATIONS OF MOTION

47

so using Lagranges equations (in vector form),


(
)
L
d L

=0
dt r
r
gives,

(
mr +
{

or
p +

dU
dr

1 dU
r dr

)
r = 0
}
r=0

Note that we used the vector calculus identity,


f
f
1 f
1 f
r +
= f (r, , ) =
+

r
r
r
r sin
in spherical coordinates.
Applying M to the previous expression gives,
{
}
1 dU
p M +
rM=0
r dr

()

yet
r M = mr (r r ) = m[r(r r r (r r)]
using the identity A (B C) = B(A C) C(A B). This expression simplies to
r M = m[rrr
r2 r ]
Since r r =

1 d
2 dt (r

r) = rr.
Finally, we rewrite this as
r M = mr3 [r2 rr
r1 r ] = mr3

d [r]
dt r

Putting this result back into (*) and recalling that M is a constant gives,
{
} [
d
d
r]
2 dU
[p M] r
m =0
dt
dr dt
r
Up to now this result holds for any central potential U (r). But notice that if the term in the curly
d
brackets is a constant then we can combine the terms into dt
[. . .] = 0, yielding a new constant of the
motion. This is the case for the U (r) = /r potential since r2 dU/dr = so,
r]
d [
p M m = 0
dt
r
our constant of the motion is

r
r
which is called the Laplace-Runge-Lenz vector. Since A M = 0 the vector A lies in the plane of motion.
At perihelion the vector v M is parallel to r so A points along the semi-major axis; with a bit more
algebra you can show that it has magnitude A = e.
Finally, it seems that we have now established seven constants of integration for the Kepler problem:
E, the three components of M and of A. However we really only have ve constants since A M = 0
A=vM

CHAPTER 3. INTEGRATION OF THE EQUATIONS OF MOTION

48

and |A| may be expressed in terms of E and M (via e). Can take these ve constants of the motion
as E, the three components of M, and A , the direction angle of A (which gives the direction of the
semi-major axis). The direction of M xes the plane of the motion while E and M determine the shape
and size of the trajectory. One more value is needed to uniquely specify the motion and we can take
that as the initial angle (0). These six parameters are uniquely specied by the initial conditions r(0)
and v(0) (and vice versa).

Chapter 4

Collisions
In the rst two sections of this chapter we ask the question: What can be said in general about the
motion of interacting particles in certain before and after situations without explicitly evaluating
the complete trajectories (see Figure 4.1) Particles are assumed to have no interaction in the before
and after situations but have an arbitrary interaction in between (e.g., repulsion during a collision).

Disintegration (16)
First consider the case where in the before situation we have a single particle of mass m1 + m2 and
in the after situation there are two particles, m1 and m2 , that are non-interacting. In the center of
mass frame of reference (C system) the initial particle is at rest. By conservation of momentum the
particles have equal but opposite momenta after disintegration so p1 = p2 or m1 v1 = m2 v2 .
By conservation of energy,
p2
p2
Ei = E1i + 0 + E2i + 0
2m1
2m2
where Ei is the internal energy; note that p0 = |p1 | = |p2 |. Call = Ei (E1i + E2i ) the disintegration
energy (i.e., the internal energy released in the disintegration) then
(
)
(
)
1 2 1
1
1 2 m1 + m2
p2
= p0
+
= p0
= 0
2
m1
m2
2
m1 m2
2m
where m is the reduced mass.
Going to a general frame of reference in which the initial particle moves with velocity V (call this
the lab frame or L system) then for either particle we have v = V + v0 , where v is in the L system and
v0 is in the C system. Since v V = v0 ,
(v V) (v V) = v0 v0
or
v 2 + V 2 2vV cos = v02
where is the angle between v and V.
This result has a simple geometric picture: First, take the case where the two particles have the
same mass so their after velocities are v0 . For V < v0 the direction of v1 in the L system can go

49

Lecture 10

CHAPTER 4. COLLISIONS

50

Figure 4.1: Illustration of before and after situations for: (Top) Disintegration (16); (Bottom)
Elastic collisions (17)

Figure 4.2: Diagrams of the post-disintegration velocities in the laboratory frame of reference for (left)
V > v0 and (right) V < v0 ; note that if V = v0 then V equals the radius of the circle.
from = 0 to depending on 0 , the direction of v0 (see Figure 4.2 and Fig. 14 in Landau). For V > v0
the maximum value of is given by,
v0
sin max =
V
(again, see Figure 4.2 and Fig. 14 in Landau). Notice that for v0 = V that max = /2; furthermore
the relation between 0 and is not unique if V > v0 (see points B and C in the Figures). Finally, in
the case that the particles have dierent masses then the geometric construction is the same but each
particle has its own circle of radius v0,1 = p0 /m1 and v0,2 = p0 /m2 .
Example From observations of numerous similar disintegrations (i.e., particle masses are always
m1 and m2 , center of mass velocity is always V) we nd that that the angle for particle 1 never exceeds
1,max however the angle for particle 2 can be arbitrary. What does this tell us about the ratio m2 /m1 ?
Solution Since m1 v01 = m2 v02 then v01 = (m2 /m1 )v02 . Furthermore,
sin 1,max =

v01
m2 v02
=
V
m1 V

CHAPTER 4. COLLISIONS

51

or

m1
v02
sin 1,max =
m2
V

But since particle 2 doesnt have a maximum angle then v02 < V so
m1
sin 1,max < 1
m2
2
so m
m1 < sin 1,max . Physically, when particle 1 is massive compared with particle 2 the deection of
its motiondue to disintegration is limited by the ratio of masses; for example if 1,max = 45 then
m2 < m1 / 2.

Elastic Collisions (17)


Elastic collisions of two particles is a problem thats similar to disintegration but with a few more
elements. Again, we start the analysis in the C system where before the collision,
m1 v10 + m2 v20 = 0
so again p10 = p20 and p10 = p20 = p0 . By conservation of momentum the total momentum after the
collision is still zero so p10 = p20 and p10 = p20 = p0 . By denition the energy in an elastic collision

is conserved so E = p20 /2m = E = p2


0 /2m so p0 = p0 . This means that in the C system (center of
mass system) the elastic collision changes the direction of the momentum but not the magnitude.
Using relative velocity, v = v10 v20 we may write,
v10
v20

m2
v
m1 + m2
m1
=
v
m1 + m2
=

0 in the direction of the post-collision


From the above v = v but v = v . Landau denes a unit vector, n
0 and
relative velocity so v = v n

v10

v20

m2
m2
0
v =
vn
m1 + m2
m1 + m2
m1
0
=
vn
m1 + m2
=

We now shift to the laboratory frame of reference (L system) by the galilean transformation,
v1 = v10 + V
where the center of mass velocity is
V=

v2 = v20 + V
m1 v1 + m2 v2
m1 + m2

Some minor algebra yields,


p1
p2

m1
(p1 + p2 ) = mv + Q1
m1 + m2
m2
0 +
= mv n
(p1 + p2 ) = mv + Q2
m1 + m2
0 +
= mv n

CHAPTER 4. COLLISIONS

52

Figure 4.3: Diagram of the pre- and post-collision momenta in the laboratory frame of reference for an
elastic collision. Note that the circle has radius mv = mv .
This result has a geometric interpretation, as shown in Figure 4.3 and Fig. 16 in Landau. In the
gure the vectors Q1 and Q2 are oriented on the x-axis; notice that they are parallel and the ratio of
their magnitudes is m1 /m2 . Recall that an elastic collision only changes the direction of the relative
velocity, not its magnitude; in the Figure we see how changing this direction changes the direction and
magnitude of the momenta after the collision.
A special case of the result above is the case where (in the L system) particle 2 is initially at rest
(v2 = p2 = 0, v = v1 ), which gives,
Q2 =

m2
m1 m2
p1 =
v1 = mv
m1 + m2
m1 + m2

In this case point B in the diagram lies on the circle. If m1 < m2 then point A is inside the circle and
1 can be any angle. If m1 > m2 then point A is outside the circle since Q1 > Q2 . The angles 1 and
2 of the post-collision velocities may be expressed in terms of the collision angle of the post-collision
relative velocity (see Figure 4.4).
Since OC = OB then the angle formed by OCB equals 2 and from the diagram its clear that
+ 22 = 180 . With a little trig we arrive at,
tan 1 =

m2 sin
m1 + m2 cos

(Note that these expressions for 1 and 2 remain valid if m1 m2 ). From the diagram we see that
the maximum deection of particle 1 occurs when p1 is perpendicular to the circle (point C ) and this

CHAPTER 4. COLLISIONS

53

Figure 4.4: Diagram of the collision when particle 2 is initially at rest and m1 > m2 .

Figure 4.5: Diagram of the collision when particle 2 is initially at rest and m1 = m2 .
maximum angle is,
sin 1,max =

mv
|OC |
=
=
|OA|
Q1

m1 m2
m1 +m2 |v1 |
m21
m1 +m2 |v1 |

m2
m1

Physically, when the mass of particle 1 is very large then the radius of the circle (mv) is small compared
with Q1 so the maximum deection angle is small.
Finally, if m1 = m2 then both points A and B lie on the circle so 21 + 22 = 180 (see Figure 4.5).
This means that p1 is always perpendicular to p2 (and v1 v2 ). This result is well-known to those of
you who play billiards.
Example A very light particle (mass m1 ) moving in the +x direction collides with a very massive
stationary particle (mass m2 , m2 m1 ). The light particle is deected by the collision such that it
moves in the +y direction. Find v2 , that is, the magnitude and direction of the velocity of the massive
particle after collision, in terms of m1 , m2 , and v1 , the magnitude of the velocity of the light particle
before collision.
Solution The post-collision velocity of the second particle has magnitude

p
2p1
2m1
v2 = 2 =
=
v1
m2
m2
m2
with a direction of 45 below the x-axis. These results may be seen directly from the collision diagram
in Figure 4.6.

CHAPTER 4. COLLISIONS

54

Figure 4.6: Collision of a light particle with a stationary massive particle in which the light particle is
deected by 90 degrees.

Scattering (18)
Lecture 11
In the last two sections we worked out all that can be predicted for the interaction of two particles without
specifying the interaction potential. We now consider what can be determined regarding collisions for a
given potential and, conversely, what can we say about an unknown potential based on measurements
of collisions.
Although were interested in the interaction of two particles undergoing a collision we formulate the
problem as that of a single particle scattering o of a xed cental potential U (r). We can later map
the results into the collision of two particles in the center of mass frame with the center of mass at the
origin.
From Section 14 we have the result illustrated in Figure 4.7 (also see Fig. 18 in Landau). The
trajectory is symmetric about the apse line, which is the line from the origin to the nearest point on
the trajectory. The apse line makes an angle 0 with respect to the incoming velocity (which we may
take as being in the -x direction). The scattering angle, , is the angle of deection of the outgoing
velocity (i.e., = 0 for no deection and = if the direction is reversed by the scattering). From
Figure 4.7 its clear that,
= | 20 |
The angle of the apse line is useful since from the results in Section 14 we have that


(M/r2 ) dr
(M/r2 ) dr

0 =
=
()
2
2
2m[E U (r)] M /r
2m[E Ue (r)]
rmin
rmin
since = 0 when the particle is initially at r = . Note that E = Ue (rmin ) since rmin is a turning
point of the motion.

CHAPTER 4. COLLISIONS

55

Figure 4.7: Geometry of scattering illustrating the impact parameter, , and the scattering angle .
(See also Fig. 18 in Landau)
As Landau points out, for comparison with scattering experiments its convenient and traditional
to reformulate this result, replacing E and M with the initial particle speed, v , and the impact
parameter, , which is the distance from the tangent line of v and the origin (see Figure 4.7 again).
The energy and angular momentum are constant so we may use the initial values of position and
velocity to write,
1
2
E = mv
M = mv
2
The result for M is evident if we consider the case of U (r) = 0, for which v is constant and = rmin .
Using these results we may write (*) as,

/r2 dr

0 =
2 )
1 (/r)2 (2U (r)/mv
rmin
An important example is the case U (r) = +/r (e.g., Coulomb scattering) for which the result is,
[
]
A/

0 = cos
A=
2
mv
1 + (A/rho)2
so
= A tan 0 = A cot(/2)
This example is treated at length in Section 19 (Rutherfords formula).
Note

that the value of M is independent of U (r).

CHAPTER 4. COLLISIONS

56

Figure 4.8: Geometry of scattering from a sphere of radius a (See also Fig. 19 in Landau)
Example (Problem 1, pg. 50) Consider scattering of particles from a perfectly rigid sphere of radius
a, that is, for the potential,
{
0 r>a
U (r) =
ra
See Figure 4.8.
The trajectories of particles bouncing o the sphere are straight lines so using simple trigonometry
we have, sin 0 = /a or
(
)
1
= a sin 0 = a sin
( ) = a cos(/2)
2
or
() = 2 cos1

()

a
for a and = 0 for > a (particle misses the sphere).
Finally, we can use the formal, integral expression and evaluate,


/r2 dr
dr

0 =
=

= sin1 (/a)
2
2
2
1 (/r) )
r r )
a
a
which leads to the same results. Notice that due to the unique form of this potential is only a function
of and does not depend on v .

CHAPTER 4. COLLISIONS

57

Figure 4.9: Geometry of scattering illustrating the relation particles with impact parameter between
and + d scattering into angles between and + d.
In scattering experiments one often cannot easily measure the impact parameter ; whats typically
done is to re a beam of particles (with known velocity v ) at a target (the scattering potential at
the origin) and the distribution is measured with a detector placed at a given scattering angle . Call
dN the number of particles per unit time that are scattered between and + d, where d may be
thought of as the angular aperture of the detector. We normalize this by n, the number of particles per
unit time per unit area and write,
d = dN ()/n
To quote Landau, This ratio has the dimensions of area and is called the eective scattering crosssection. It is entirely determined by the form of the scattering eld and is the most important characteristic of the scattering process.
All the particles with impact parameter between and + d scattering into angles between and
+ d (see Figure 4.9). We will assume (as is typically the case) that this is a one-to-one relation so
not only is () but also (). The number of particles scattering into (, + d) equals the product
of n and the area between the two circles of radii and + d so,
dN = (2d)n

and

d() = 2()d

Because its easier to experimentally control (which depends on the position of the detector) than
(since we shoot the target with a beam of particles) we instead use

d
d() = 2() d
d
As Landau mentions, sometimes this expression is written in terms of solid angle o instead of polar
angle ; since the scattering from a central potential is symmetric in azimuthal angle we have do =
2 sin d.

CHAPTER 4. COLLISIONS

58

Example (Problem 1, pg. 50, cont.) Determine the eective cross-section for scattering of particles
from a perfectly rigid sphere of radius a.
Solution Since we already found that = a cos(/2) so

d
d = 2 d
d



1
= 2[a cos(/2)] a sin(/2) d
2
2
1 2
=
a sin d = a do
2
4
Notice that the scattering is isotropic (since d = (constant) do). Furthermore, d is independent of
v for this rigid potential. The integral over solid angle immediately gives,


1 2 2
1 2
1
a do = a
d
=
sin d = a2 (2)(2) = a2
4
4
4
0
0
which is the expected result, the cross-sectional area of a sphere.
In the rest of this section Landau presents several interesting examples in the PROBLEMS section
and also discusses the transformation between the center of mass frame (C system) to the laboratory
frame (L system). In Section 19 he continues the discussion of scattering by presenting an important
application: scattering from a Coulomb eld. The resulting eective cross-sectional area for that
problem is (see (19.2)),
(
)2

cos(/2)
d =
d
2
mv
sin3 (/2)
which is Rutherfords formula. As with most potentials the integral over does not give a nite value
for . Physically, the Coulomb potential has such a long range that we have small deections even when
the impact parameter is very large. Nevertheless, while is not nite the important physical quantity,
d, is well-dened.

FIRST MIDTERM
Lecture 12

Chapter 5

Small Oscillations
Lecture 13
In general, the equations of motion are very complicated to solve for arbitrary initial conditions. However, we are often interested in small displacements from an equilibrium state for the system. Since the
displacements are small we can use Taylor expansions of the Lagrangian. Interestingly, the resulting
motion is the same for a wide variety of systems.

Free oscillations in one variable (21)


Consider the general Lagrangian for a system with just one variable that has a stable equilibrium
position, q0 . For such a system U (q) is independent of time and the potential energy has a minimum
at q = q0 . The Lagrangian is
L(q, q)
= T (q, q)
U (q)
Were interested in small oscillations about the equilibrium position so well Taylor expand the Lagrangian using q = q0 + x,
(
)
(
)
dU
1 2 d2 U
U (q) = U (q0 + x) = U (q0 ) + x
+ x
+ ...
dq q=q0 2
dq 2 q=q0
The rst term on the right hand side, U (q0 ), is a constant so we may discard it. The next term is zero
since q0 is a minimum of the potential so dU/dq = 0 at that point. In the new variable, x, which is the
displacement from equilibrium, we have
(
)
1 2 d2 U
1
U (x) = x
= kx2
2
2
dq q=q0
2
after dropping terms O(x3 ) or higher. Physically were approximating the potential as a parabola
in the vicinity of q = q0 and the constant k is the same as the spring stiness that appears in the
formulation of simple harmonic motion.
Recall from equation (5.5) that the kinetic energy has the general form,
T (q, q)
=
The

1
a(q)q2
2

case where U (q, t) is considered in the Section 22 (Forced Oscillations).


see the contribution of keeping those higher order terms in Section 28 (Anharmonic Oscillations).

Well

59

CHAPTER 5. SMALL OSCILLATIONS

60

We Taylor expand a(q) but in this case the rst term is the one thats kept, that is,
a(q) = a(q0 + x) = a(q0 ) + . . .
and well write this constant as a(q0 ) = m. Finally, since q = x we have T = 12 mx 2 and
L(x, x)
=

1
1
mx 2 kx2
2
2

The equation of motion is that of simple harmonic motion (SHM),


m
x + kx = 0
or
x
+ 2 x = 0

where = k/m is the (angular) frequency of the small oscillations. Note that frequency is also
dened as f = /2 = 1/T .

Example Find the period of small oscillations for a particle (mass m) moving in the one-dimensional
potential U (q) = U0 tan2 (q); compare with the exact result from Section 11,

2m
Te =
E + U0
see Problem 2 (c).
Solution From a quick sketch of U (q) its clear that the only minimum is q0 = 0. Evaluating,
( 2 )
)
(
d U
k=
= 2U0 2 [2 cos(2q)] sec4 (q) q=q0 = 2U0 2
2
dq q=q0
Since T = 2/, the small oscillations period is,

m
m
2m
T = 2
= 2
=
k
2U0 2
U0
By inspection Te T when E U0 .
Finally, Landau points out that there are various ways of writing the solution for SHM,
x(t)

c1 cos t + c2 sin t

= a cos(t + )
[
]
= re Aeit
where

A = aei

where the (real) constants (c1 , c2 ) or (a, ) or the (complex) constant A are xed by the initial conditions
x(0) and x(0).

CHAPTER 5. SMALL OSCILLATIONS

61

Forced Oscillations (22)


Next we consider open systems with one coordinate by adding a time-dependent external potential to
the Lagrangian,
1
L(q, q,
) = a(q)q2 U (q) Ue (q, t)
2
Using the same Taylor expansion about q0 , which is the stable equilibrium state in the absence of Ue ,
we have
1
1
L = mx 2 kx2 + xF (t)
2
2
where
(
)
Ue
F (t) =
q q=q0
is the external driving force. Notice that in Taylor expanding the external potential we discard the rst
term, Ue (q0 , t), since it is only a function of time.
The equation of motion is
1
x
+ 2 x = F (t)
m
which is an inhomogeneous second order ODE. First, lets consider the case where the external driving
force is periodic and of the form,
F (t) = f cos(t + )
where f , , and are the amplitude, angular frequency, and phase of the external force. A little math
yields,
x(t) = a cos(t + ) + b cos(t + + )
{

with
b=

=

=

f
m( 2 2 )
ft
2m

{
;

0
/2

=

=

with (a, ) xed by the initial conditions. Notice that when = we have resonance and the amplitude
diverges in time (and, presumably, the small oscillations approximation breaks down at some point).
For = we may write the solution above as
[
]
x(t) = re Aeit + Beit
where
A = aei

and

B=

f
ei
m( 2 2 )

Dene the dierence frequency = then


Aeit + Bei(+)t
(
)
= A + Beit eit

x(t) =

= C(t)eit
The physical interpretation of this result is that the amplitude of the motion has a time-dependent
complex amplitude C(t). When we have beats (see Figure 5.1).
In

equations like this, where the left hand side is a real function and the other side is complex, its implicitly understood
that we take the real part, re[Right Hand Side].

CHAPTER 5. SMALL OSCILLATIONS

62

Figure 5.1: Beats in the periodic motion of a forced oscillator with natural frequency and driving
frequency = + with .
Finally, we consider the solution for a general forcing F (t). To formulate this general solution we
write the equation of motion,
1
x
+ 2 x = F (t)
()
m
as
1
(
x + i x)
(i x + (i)2 x) = F (t)
m
or
1
d
(x + ix) i(x + ix) = F (t)
dt
m
Introducing the auxiliary function = x + ix we transform our second order real ODE (*) into a rst
order complex ODE,
d
1
i = F (t)
dt
m
Note that x = re[] and x = 1 im[]. Using the trial solution, = A(t)eit , the solution of this rst
order ODE is,
t
1
= eit
F (t)eit dt + 0 eit
m
0
where 0 is xed by the initial conditions.
Example (Problem 1(a), pg. 64) Consider a particle (mass m) at rest in a potential U (x) = 12 kx2 .
At time t = 0 a constant force F0 is applied. Find x(t) and x.

CHAPTER 5. SMALL OSCILLATIONS

63

Solution Using the general result (note 0 = 0 since x(0) = x(0)

= 0),

=
=
=
=
=

1 it
e
m

F0 eit dt

[
]t
F0 it
1 it
e
e
m
i
0
)
F0 it ( it

e
e
1
im
)
F0 (

1 eit
im
F0
(sin t + i(1 cos t))
m

Since = x + ix,
x(t) =
x(t)

1
F0
im[] = 2 (1 cos t)

m
F0
re[] =
sin t
m

Notice that the oscillations go between xmin = 0 and xmax = 2F0 / 2 m = 2F0 /k so F0 xmax = 21 kx2max ,
which we expect from the work-energy relation.

Oscillations of Systems with More Than One Degree of Freedom


(23)
Lecture 14
We return to the problem of free (i.e., un-forced) oscillations but for more than one generalized coordinate. Again, we dene a stable steady state (q10 , . . . , qs0 ) as a point where the potential has a minimum,
that is,
[
]

U (q1 , . . . , qs )
=0
for all i
qi
q=q0
and

]

U (q1 , . . . , qs )
= kik > 0
qi qk
q=q0

for all i, j

Note that kik = kki . As before, we dene the displacement from the steady state as xi = qi qi0 and
Taylor expand the potential to get,
1
kik xi xk
2 i=1
s

U (q1 , . . . , qs ) = U (q10 + x1 , . . . , qs0 + xs ) =

k=1

after discarding the constant term and terms of O(x3 ) and higher.

Example Recall that for the double pendulum (Problem 1, pg. 11) the potential may be written
as,
U (1 , 2 ) = (m1 + m2 )g1 cos 1 m2 g2 cos 2

CHAPTER 5. SMALL OSCILLATIONS

64

The steady state is 10 = 20 = 0 so x1 = 1 and x2 = 2 . We want to expand to quadratic order in


1 and 2 so you can either grind out the derivatives or make use of the Taylor expansion of the cosine
function,
1
cos = 1 2 + . . .
2
so to quadratic order,
1
1
(m1 + m2 )g1 (1 21 ) m2 g2 (1 22 ) + . . .
2
2
1
1
= (m1 1 + m2 1 + m2 2 )g + (m1 + m2 )g1 21 + m2 g2 22
2
2

U (1 , 2 ) =

so k11 = (m1 + m2 )g1 , k22 = m2 g2 , k12 = k21 = 0.


The general form of the kinetic energy is (see eqn. (5.5)),
1
aik (q1 , . . . , qs )qi qk
2 i=1
s

T (q1 , . . . , qs ) =

k=1

In the small oscillation approximation we keep only quadratic terms and since qi = x i ,
1
aik (q10 , . . . , qs0 )x i x k + . . .
2 i=1
s

T (q1 , . . . , qs ) =

k=1

where the higher order terms are of cubic order, O(xx 2 ). We dene the mass matrix, with coecients,
mik = aik (q10 , . . . , qs0 )
Finally, we may write the Lagrangian for small oscillations as,
1
(mik x i x k kik xi xk )
L=
2 i=1
s

k=1

Example Returning to the double pendulum (Problem 1, pg. 11), the kinetic energy is,
T =

1
1
(m1 + m2 )21 21 + m2 22 22 + m2 1 2 cos(1 2 ) 1 2
2
2

We want to match this to the general form,


T =

1
(a11 21 + a22 22 + (a12 + a21 ) 1 2 )
2

Matching terms (and noting that, in general, a12 = a21 ),


(m1 + m2 )21

a11

a22

= m2 22

a21

= m2 1 2 cos(1 2 )

CHAPTER 5. SMALL OSCILLATIONS

65

Finally, since 10 = 20 = 0,
m11

= (m1 + m2 )21

m22

= m2 22

m21

= m2 1 2

and m21 = m12 .


Lets use Lagranges equations to get the equations of motion,
(
)
L
d L

=0
(j = 1, . . . , s)
dt x j
xj
We need to be a little careful in evaluating derivatives of sums, specically,
L
x j

x i x k
mik
2 i=1
x j

1
mik (x i jk + x k ji )
2 i=1

k=1
s

k=1

where the Kronecker delta function, ab , is dened as,


{
1 a=b
ab =
0 a = b
Notice that Kronecker delta functions collapse sums, that is,
s

f (xa )ab = f (xb )

a=1

This is similar to how Dirac delta functions collapse integrals. Using this property,
L
x j

1
1
mij x i +
mjk x k
2 i=1
2
s

=
=
=

s
1

mij x i +

i=1

1
2

k=1
s

mji x i

i=1

mji x i

i=1

Remember that we can change the name of a summation index and that the mass matrix is symmetric
so mji = mij .
As similar analysis gives
s

L
=
kji xi
xj
i=1
so the equations of motion are,
s

i=1

(mji x
i + kji xi ) = 0

(j = 1, . . . , s)

CHAPTER 5. SMALL OSCILLATIONS

66

For these equations of motion we look for periodic solutions of the form
xk (t) = Ak eit

(k = 1, . . . , s)

Since
x
k = 2 Ak eit
then
= 2 x
x
It turns out theres not just one frequency for the oscillations, theres 1 , . . . , s . We now need to nd
a way to nd these oscillation frequencies in terms of the coecients mji and kji .
To analyze the equations of motion well write them in matrix form,

m11
m21
..
.
ms1

m12
m22
..
.
ms2

. . . m1s
. . . m2s
..
..
.
.
. . . mss

x
1
x
2
..
.
x
s

k11
k21
..
.
ks1

k12
k22
..
.
ks2

. . . k1s
. . . k2s
..
..
.
.
. . . kss

x1
x2
..
.
xs

= 0 or
M
x + Kx = 0
where M and K are matrices and x is a column vector.
Example Returning to the double pendulum (Problem 1, pg. 11), the matrices for the equation of
motion are
)
(
)
(
(m1 + m2 )g1
0
(m1 + m2 )21 m2 1 2
K
=
M=
0
m2 g2
m2 1 2
m2 22
For the special case that m1 = m2 = m0 and 1 = 2 = 0 ,
(
)
(
2 1
2
2
M = m0 0
K = m0 g0
1 1
0

0
1

Well use this in the next example in which well evaluate the equations of motion to nd x(t).
Inserting the trial solution,
x(t) = Aeit
into the equations of motion give
2 Mx + Kx = 0
or
(K 2 M)x = 0
This is a matrix eigenvalue problem and for there to be a non-trivial solution (i.e., for x = 0) it must
be that (K 2 M) is a singular matrix with determinant equal to zero, that is,
|K 2 M| = 0

CHAPTER 5. SMALL OSCILLATIONS

67

where the absolute value of a matrix is its determinant. This condition allows us to nd oscillation
frequencies, specically, evaluating the determinant will lead to a polynomial equation for . This
polynomial has 2s roots, 1 , . . . , s , and Landau shows that these roots are real.
Example One last time, lets return to the double pendulum (Problem 1, pg. 11). For the special
case that m1 = m2 = m0 and 1 = 2 = 0 ,
(
)
(
)
2 0
2 1
K 2 M = A
2 B
0 1
1 1
where A = m0 g0 and B = m0 20 . The determinant we need to evaluate is,


2(A 2 B)
2 B
|K 2 M| =
2 B
A 2 B
This determinant gives the quadratic equation (called the characteristic equation),
B 2 2 4AB + 2A2 = 0
where = 2 . The roots of this quadratic are,
= (2

2)

g
A
= (2 2)
B
0

so the frequencies of oscillation are,

g
(2 + 2)
0

g
=
(2 2)
0

These two frequencies correspond to two modes of oscillations, as shown in Figure 5.2. For the high
frequency, 1 , the arms swing out of phase and for the low frequency, 2 , they swing together. By the
way, Landau presents the more general result (for m1 = m2 , 1 = 2 ) in Problem 2, pg. 70.
Each eigenfrequency i has a corresponding has a corresponding eigenvector, i , where,

i =

i1
i2
..
.
is

11
12
..
.
1s

21
22
..
.
2s

...
...
..
.
...

s1
s2
..
.
ss

where is a matrix assembled from the eigenvectors. This formulation is useful in that we may write
the general solution to the equations of motion as,
x(t) = (t)

or

xk (t) =

k (t)

=1

where

[
]
(t) = re C ei t

()

Lecture 15

CHAPTER 5. SMALL OSCILLATIONS

68

Figure 5.2: Modes of oscillation for the double pendulum.

The values of the (complex) constants, C , are xed by the initial conditions, x(0) and x(0).
By the
way, the eigenvectors are dened up to a multiplicative constant and often this constant is xed by
making each be a unit vector (i.e., | | = 1).
Notice from the denition of that,
=

so for each the co-ordinate performs simple harmonic motion that is independent of the other
co-ordinates. This suggests that the motion is much simpler if we transform from our original coordinates, (x1 , . . . , xs ), into new co-ordinates, (1 , . . . , s ). This transformation is accomplished using
(*), specically,
(t) = ()1 x(t)
where the matrix inverse is dened by the property,
()1 () = ()()1 = I
where I = is the identity matrix. Notice that because matrix multiplication is a linear operation
the new co-ordinates, are just a linear combination of the original ones, x.
Example (Problem 1, pg. 69) Consider a pair of similar pendula coupled by a spring (see Figure 5.3).
The Lagrangian, in the small oscillations approximation, turns out to be,
L=

1 2
1
( + 22 ) 02 (21 + 22 ) + 1 2
2 1
2

where 0 is the frequency of oscillation when the pendula are uncoupled and is proportional to the
spring constant. Note that the Wilberforce pendulum has a similar Lagrangian with (1 , 2 ) (, z)
where is the torsional angle of rotation and z is the vertical displacement of the spring.
By inspection,
(
)
( 2
)
1 0
0
M=
;
K=
0 1
02

CHAPTER 5. SMALL OSCILLATIONS

69

Figure 5.3: Pair of similar pendula (mass m, length ) coupled by a spring.


The determinant gives a quadratic characteristic equation,
2
2

2
|K M| = 0

02 2



=0

or
(02 2 )2 2 = 0
so
2 = 02
Well write the two eigenfrequencies as,

02

02 +

Note that if 02 then the motion exhibits beats (see http://www.youtube.com/watch?v=S42lLTlnfZc).


Now lets nd the corresponding eigenvectors. For = + we have,
( 2
) (
)
2
0 +


2
K +
M=
=
2

02 +

Since the corresponding eigenvector, by denition, obeys the equation,
2
(K +
M)+ = 0
Landau

writes these as 1 and 2 , respectively.

CHAPTER 5. SMALL OSCILLATIONS


then

so +1 = +2 or

70

1
1

1
1

)(

+1
+2

)
=0

)
1
(Eigenvector for frequency + )
1

where the multiplicative factor of 1/ 2 is included to make + a unit vector. A similar analysis gives
the other eigenvector,
(
)
1
1

=
(Eigenvector for frequency )
1
2
1
+ =
2

Now that weve formulated the eigenvectors well nd the corresponding normal modes. We start
from the general expression, x = , which in this example is,
(
)(
)
)
(
1
+
1
1 1
=
1 1

2
2
or
1

1
(+ + )
2
1
(+ + )
2

We can invert these equations to get,


+

1
(1 2 )
2
1
(1 + 2 )
2

which are the two normal modes for the high and low frequency oscillations, respectively.
Lets nish by establishing the physical meaning for these normal modes. Consider the case shown in
Figure 5.4 (left) in which the two pendula swing 180 degrees out of phase (i.e., when the rst pendulum
is swinging towards the right the other one is swinging towards the left). In that case, 1 (t) = 2 (t)
so

+ (t) = 21 (t) = 21 (0) cos(+ t + 0 )


(Pendula out of phase)
and (t) = 0. This is the physical motion that corresponds to the high frequency normal mode. For
the low frequency mode the motion is shown in Figure 5.4 (right) in which the pendula swing in phase.
In that case, 1 (t) = 2 (t) so + (t) = 0 and

(t) = 21 (t) = 21 (0) cos( t + 0 )


(Pendula in phase)
For an arbitrary initial condition the resulting motion may supercially look complicated but its always
a linear combination of these two normal modes.
Remember that were labeling the eigenfrequencies and eigenvectors + and instead of 1 and 2; if you dont like
that then use the numbers since its just a label.
In this expression is the phase that comes from the initial conditions.
0

CHAPTER 5. SMALL OSCILLATIONS

71

Figure 5.4: Normal modes for a pair of coupled pendula for the (left) high frequency, + , and (right)
low frequency, .

A few nal comments:


* Using the normal coordinates its possible to write the Lagrangian as,
L=

1
2 2 2 )
m (
2
=1

where the m are positive constants. Notice that these ms are no longer a matrix.
* There are two kinds of normal coordinates, and Q , but they dier only by a multiplicative

scale factor, specically m = Q so,


L=

1 2
(Q 2 Q2 )
2
=1

In the example above m = 1 so = Q .


* Forced oscillations are simple to treat using normal coordinates. If Fk (t) is the force in the original
coordinates then in the normal coordinates,
F (t) =

k Fk (t)

k=1

and

+ 2 = 1 F (t)

m
Note that the variables (t) are independent so each may be treated as a single oscillator, that is, as
in Section 22. The formulation is similar and slightly simpler using Q .

Vibrations of Molecules (24)


Lecture 16
Formulating the small oscillations approximation is sometimes complicated because the motion of the
system contains elements that are not considered as vibrations. The vibrations in molecules is a common,
important example.

CHAPTER 5. SMALL OSCILLATIONS

72

Figure 5.5: Illustration of a diatomic molecule with one vibrational, three translational, and two rotational degrees of freedom.
Consider the simple diatomic molecule illustrated
in Figure 5.5. This system has six degrees of

freedom but only one non-zero eigen-frequency, k/m. The other ve degrees of freedom are associated
with the translational motion (3 degrees) and rotational motion (2 degrees). To avoid these nonvibrational modes its best to consider the system with zero center of mass velocity and zero total
angular momentum.
Landau presents an interesting analysis, including various examples, but since this is a specialized
topic we wont be covering it in this class.

Damped Oscillations (25)


Mechanical systems with dissipation
Sometimes open systems cannot be formulated using a time-dependent external potential because the
reservoir system does not have a known motion. An important example is friction. We can still use
Lagrangian dynamics by augmenting Lagranges equations as,
(
)
d L
L

= fi (q, q,
t)
dt qi
qi
where f is a generalized external force.
Frictional forces, such as air resistance, are often velocity dependent and zero in the absence of
motion. A common approximation is that the force is linear in velocity, that is,
ffr,i =

k=1

ik qk

()

CHAPTER 5. SMALL OSCILLATIONS

73

This approximation is accurate for the drag force in very low speed ows (e.g., Stokes ow). This friction
force can be expressed as a velocity potential by introducing the Rayleigh dissipation function,
1
ik qi qk
2 i=1
s

Ffr =

ffr,i =

and

k=1

F
qi

Notice that the derivative is with respect to velocity, not position.

Dissipation of Energy
Mechanical energy is not conserved in systems with friction and an important quantity to nd is dE/dt.
Recall the general denition of mechanical energy,
E=

i=1

qi

L
L
qi

The rate of change of energy is,


dE
dt

L d
+
qi
qi
qi i=1 dt
i=1
s

L
qi

dL
dt

The last term may be written as,


d
L(q, q)

dt

s
s

L dqi L dqi
+
qi dt
qi dt
i=1
i=1

s
s

L
L
qi +
qi

q
i
i
i=1
i=1

The terms with q cancel leaving,


dE
dt

(
)
s
d L
L

qi
dt

q
i
i
i=1
i=1
[ (
)
]
s

L
d L
=

qi
dt qi
qi
i=1
=

qi

qi fi

i=1

where the right hand side is the power (work done per unit time) by the external dissipative force. If
the frictional force is linear in velocity then using (*) gives,
( s
)
s

dE
=
qi
ik qk
dt
i=1
k=1

s
s

ik qi qk = 2F

i=1 k=1

Its clear from this expression that the coecients ik give the rate at which mechanical energy is
dissipated.
Landau

writes F as F .

CHAPTER 5. SMALL OSCILLATIONS

74

Systems with damped oscillations


So far the formulation has been for mechanical systems in general. Now we return to the small oscillations approximation. For an isotropic velocity potential, F(q),
after Taylor expanding and keeping
terms of O(q2 ) we have,
s
s
1
Ffr =
ik qi qk
2 i=1
k=1

The damping frictional force is,


ffr,i =

ik qk =

k=1

ik x k

k=1

The equation of motion for a system with just one coordinate is,
m
x + kx = x

()

For the general case it is,


s

k=1

mik x
k +

k=1

kik xk =

ik x k

(i = 1, . . . , s)

k=1

or in matrix form,
M
x + Ax + Kx = 0
For this linear system of equations we look for solutions of the form,
xk (t) = Ak ert
where r is now complex. Without damping r = i but due to damping re[r] < 0 (i.e., exponential decay
in time).
The analysis of the general case is the same as before only now the eigenvalue problem is a bit more
complicated,
(r2 M + rA + K)x = 0
To illustrate the essential features of the solution, lets just do the case where the system only has one
coordinate. The equation of motion (*) can be written as,

where 0 =

x
+ 02 x = 2x
k/m and = /2m. Using the trial solution x(t) = Aert ,
r2 x + 2rx + 02 x = 0

Factoring out the x and r is found from the roots of the quadratic polynomial,
r2 + 2r + 02 = 0
so
r =

2 02

When < 0 (small friction) r is complex with re[r] = and im[r] =


may be written as,
x(t) = aet cos(t + 0 )

02 2 = ; the solution

which is a damped oscillation with a shifted frequency < 0 . When > 0 then both roots are real
and negative and the solution is exponential damping.

CHAPTER 5. SMALL OSCILLATIONS

75

Forced, damped oscillations (26)


Adding a periodic driving force to the one-dimensional damped oscillator equations of motion gives,
x
+ 2x + 02 x =

1
(f cos t)
m

Solving this inhomogeneous ODE (for < 0 ) gives,


x(t) = aet cos(t + ) + b cos(t + )
Notice that the rst term is the homogeneous solution, which decays exponentially so for long times
the solution is dominated by the second term (the particular solution). At long times (specically, for
t 1/) the solution is,
x(t) b cos(t + )
with amplitude and phase,
b

m (02 2 )2 + 42 2
(
)
2
= atan
02 2

Notice that the oscillations are at the driving frequency but the amplitude and phase depend on the
relative magnitudes of the driving and natural frequencies. The maximum amplitude is at resonance
(0 = ) and is nite,
f
b=
2m
which increases as the damping decreases (as 0).

Parametric Resonance (27)

(Lecture 17)
Consider a variant of simple harmonic motion in which the natural frequency is time-dependent,
x
+ 2 (t)x = 0

()

where (t) is a periodic function with period T . This ODE is known as the Hill equation, named after
the American astronomer G. W. Hill, who introduced it in 1886. An example is a simple pendulum
with a vertically
oscillating pivot, which is equivalent to a variable gravitational acceleration, g(t), so
then (t) = g(t)/.
General properties regarding the solution of the Hill equation may be obtained using Floquet analysis.
Given that (t) = (t + T ) means that if x(t) is a solution then so is x(t + T ). Furthermore, since the
Hill equation is linear, x(t) and x(t + T ) can only dier by a constant, that is, x(t + T ) = x(t). The
most general form that has these properties is,
x(t) = t/T (t)

()

where (t) is an arbitrary periodic function (i.e., (t + T ) = (t)).


Since the Hill equation (*) is a second order ODE there are two solutions, x1 (t) and x2 (t) with a
general solution thats a linear combination,
x(t) = A1 x1 (t) + A2 x2 (t)

CHAPTER 5. SMALL OSCILLATIONS

76

Now take (*) for x1 and multiply through by x2 ,


x2 x
1 + 2 (t)x2 x1 = 0
Subtract from this equation the symmetric duplicate, x1 x
2 + 2 (t)x1 x2 = 0, and we have
x2 x
1 x1 x
2 = 0
or

d
(x2 x 1 x1 x 2 ) = 0
dt

which means that


x2 x 1 x1 x 2 = (constant)
The left hand side is called the Wronskian of the solution.
Next we make use of (**) and write
x1 (t + T ) = 1 x1 (t)

x 1 (t + T ) = 1 x 1 (t)

and

and similarly for x2 . Combining the various terms allows us to write,


x2 (t + T )x 1 (t + T ) x1 (t + T )x 2 (t + T ) =

(2 x2 (t))(1 x 1 (t)) (1 x1 (t))(2 x 2 (t))

= 1 2 (x2 (t)x 1 (t) x2 (t)x 2 (t))


Since the Wronskian is a constant then
1 2 = 1
There are two possibilities, either |1 | = |2 | = 1 or or |1 | > 1, |2 | < 1. These may be written as,
1)

1 = ei , 2 = ei

2)

1 = , 2 = 1/

where and are real.


The rst possibility gives the general solution,
x(t) = A1 eit/T 1 (t) + A2 eit/T 2 (t)
Recall that 1 , 2 are periodic functions so x(t) = x(t + T ) and the solution is periodic. We call this
the stable case. The second possibility has as a general solution,
x(t) = A1 t/T 1 (t) + A2 t/T 2 (t)
Since > 1 the magnitude of the rst term increases exponentially in time so we call this the unstable
case or parametric resonance.
To proceed further we need to specify the form of the periodic function (t); well use,

(t) = 0 1 + h cos t
Notice that for the vibrating pendulum example,

and
0 = g/
Without

h = 2 a/g

loss of generality we call the larger of the two values 1 .

CHAPTER 5. SMALL OSCILLATIONS

77

where a is the amplitude of the vibrational displacement so 2 a is the maximum acceleration of the
pivot. It turns out that the largest resonance window occurs when 20 so well write = 20 +.
The equation of motion (*) may be written as,
x
+ 02 [1 + cos{2(0 + 21 )t}]x = 0
which is the Mathieu equation. We look for a general solution of the form,
x(t) = A1 est cos(0 + 12 )t + A2 est sin(0 + 12 )t
where
s = i/T (stable case)

s = ln()/T (stable case)

Note that this solution is not exact because the periodic oscillations are not sine and cosine but rather
Mathieu functions, which are similar but not exactly the same as the trigonometric functions.
After some manipulations (see pg. 82 for some guideposts) we arrive the result,

]
1[ 1
s=
( 2 h0 )2 2
4
We have the unstable solution if s is real so the condition for parametric resonance is,
| 21 h0 | > ||
Recall that h is the dimensionless amplitude of the driving oscillation so as long as thats large enough or
, which is the dierence between the driving frequency, , and the resonance frequency for this window,
20 , then the amplitude grows in time. A similar result may be found for the resonance window around
0 , that is, for = 0 + parametric resonance occurs if,
5 2
24
h 0 < <

1 2
24 h 0

Note that all these results are accurate only when |h| 1 although the results are qualitatively similar
for nite h. Finally, if we include a damping force, with damping coecient , then the condition for
parametric resonance for 20 is,

( 12 h0 )2 42 > ||
Notice that if the damping is suciently strong (if >
occur.

1
4 h0 )

then parametric resonance does not

Motion in a Rapidly Oscillating Field (30)


Although this section is at the end of the chapter some of the applications of this theory are related to
parametric resonance problems. Consider the forced motion of a particle in one coordinate (q = x); the
equation of motion may be written as,
dU
m
x=
+f
dx
where f is the driving force, which well take to be periodic and of the form (notation here is slightly
dierent from Landau),
f = f cos( t + )
Another

resonance window occurs when 0 ; see Problem 2 on pg. 83

CHAPTER 5. SMALL OSCILLATIONS

78

In the absence of forcing the motion has period T as determined from the potential U (x) (see Section 11).
We take the forcing to be rapid by comparison, that is, T >> T = 2/ . The motion may then be
decomposed into two parts,
x(t) = X(t) + (t)
where X is the slow, large-scale motion and is the small, rapid oscillations. The latter has the property
that,
t+T
(t) =
(t) 0
t

that is, it averages out to zero over the period of a rapid vibration.
After a few manipulations Landau shows that,
(t)

1
f
m2

and the equation of motion for X is,

mX(t)
=

d
Ue (X)
dX

where the eective potential is,


f2
4m2
If the amplitude f of the rapid forcing is a constant then we have no eect because the additional term
is a constant. However, if f (x)(= f (X)) then the motion X(t) is altered. For example, for the simple
pendulum with a vertically oscillating support (see Problem 1, pg. 95),
Ue (X) = U (X) +

When >

f () = ma2 sin
2g/a the inverted state ( = ) is stable (see Figure InvertedPendulumPotential).

Anharmonic Oscillations (28)


Lecture 18
Recall that the small oscillations approximation was formulated by Taylor expanding the Lagrangian
to quadratic order. The natural question is: What if we extend the expansion to include cubic order?
Landau develops this theory for the general case of a system with s generalized coordinates. Since this
gets messy well consider the simpler case of just one coordinate.
In general, the kinetic energy for a closed system takes the form (see (5.5)),
1
a(q)q2
2
Taylor expanding a(q) about the steady state q0 gives,
T =

a(q)

= a(q0 + x) = a(q0 ) + x

da
dq

)
+ ...
q=q0

= m + nx
where m and n are constants obtained from a(q). Note that for s coordinates we have the s-by-s matrix
mik and the s-by-s-by-s tensor nik (Yuck!) Similarly, the potential energy Taylor expands to be,
(
(
(
)
)
)
dU
1 2 d2 U
1 3 1 d3 U
U (q) = U (q0 + x) = U (q0 ) + x
+ x
+ x
...
dq q=q0 2
dq 2 q=q0 3
2 dq 3 q=q0
1
1
= (Discard) + (Equals zero) + kx2 + x3
2
3

CHAPTER 5. SMALL OSCILLATIONS

79

Figure 5.6: Eective potential Ue () for a simple pendulum with a vertically oscillating support (see
Problem 1, pg. 95)
For s coordinates kik is a matrix and ik is a tensor. Using the change of variables into normal
coordinates the matrices mik and kik may be diagonalized but (unfortunately) the tensors dont simplify.
Collecting the results above for the one-variable case our Lagrangian, to cubic order, is,
L(x, x)
=

1
1
1
(m + nx)x 2 kx2 x3
2
2
3

Since
d L
dt x
L
x

=
=

d
(m + nx)x = m
x + nx 2
dt
1 2
nx kx x2
2

The equation of motion is,


m
x + nx 2 =

1 2
nx kx x2
2

which may be written as,


1
x
+ 02 x = x2 x 2
2

()

where 02 = k/m, = /m, and = n/m. Needless to say, the non-linear ODE is dicult to solve.
Well construct a solution by using successive approximation, otherwise known as perturbation
theory. The solution will be assembled in successive pieces,
x(t) = x(1) (t) + x(2) (t) + x(3) (t) + . . .
where x(1) satises the small oscillation equation,
x
(1) + 2 x(1) = 0

CHAPTER 5. SMALL OSCILLATIONS

80

so x(1) (t) = a cos t. Note that in this equation we write instead of 0 because the anharmonic
terms (left hand side in eqn. (*)) produce a frequency shift. For that reason we also do a perturbation
expansion and write,
= 0 + (1) + (2) + . . .
Landau illustrates the application of these perturbation expansions but instead of considering
eqn. (*) he chooses to use,
x
+ 02 x = x2 x3
which is similar to (*) with = , = 0, and Taylor expanding the potential U (q) to order x4 . Putting
in x = x(1) + x(2) , = 0 + (1) , and x(1) = a cos t gives (after dropping third order terms),
x
(2) + 2 x(2)

1
= a2 (1 cos 2t)
2
+20 (1) a cos t
()

The last term in (**) is an un-physical resonance term; if it was present then the amplitude of the
solution would grow in time, as in forced resonant oscillations. This tells us that (1) = 0, that is, there
is no frequency shift to rst order.
The solution of (**) may be obtained using the methods discussed in Section 22; we nd,
x(2) (t) =

a2
a2
cos 2t
+
20
602

The rst term of this solution is the shift in the center of the oscillations due to the asymmetry of the
potential; recall that,
( 3 )
1
d U
=
2m dq 3 q=q0
The second term tells us that the oscillations, to this order, are composed of two frequencies, 0 from
the x(1) (t) contribution and 20 from the x(2) (t) contribution (remember that (1) = 0 so = 0 ).
Going to next order in the perturbation expansion gives,
(
)
3
52
(2) = a2

80
1203
which is a frequency shift; the contribution to the solution is,
( 2
)

1
a3
(3)
cos 3t
x (t) =
1602 302
2
Finally, these results may be applied to the simple pendulum in the medium angle approximation.
In that case = 0 (since the potential is symmetric about = 0) and so the frequency of oscillations
is,
3a2
= 0 + (2) = 0 +
80
and the solution is,
a3
x(t) = x(1) + x(2) + x(3) = a cos t
cos 3t
3202
This solution and its power spectrum are sketched in Figure 5.7.
Some closing (and discouraging) notes on this analysis: For some simple mechanical systems this
perturbation analysis yields somewhat useful results. However, the more common situation when nonlinear contributions appear in the equations of motion is that perturbation analysis fails to be accurate

CHAPTER 5. SMALL OSCILLATIONS

81

Figure 5.7: Sketch of the motion, x(t), and the power spectrum for the anharmonic oscillations of a
simple pendulum.
because the motion becomes chaotic; see Figure 5.8. This occurs for even relatively simple systems such
as the double pendulum. Given this limitation, and the tedious complexity of perturbation analysis,
anharmonic motion is usually treated by alternative methods developed in the eld of dynamical systems.

CHAPTER 5. SMALL OSCILLATIONS

82

Figure 5.8: Sketch of the motion, x(t), and the power spectrum for the anharmonic oscillations of a
typical non-linear mechanical system with chaotic motion, such as the double pendulum.

Chapter 6

Rigid Body Motion


Lecture 19

Angular Velocity (31)


We dene a rigid body as a system of particles for which the distance between the particles is xed,
such as the lattice of atoms in a solid if we neglect the microscopic thermal motion. The motion of a
rigid body is separated into two parts: the motion of the center of mass plus the rotation of the body.
It is useful to consider both an inertial frame of reference (with cartesian coordinates X, Y , and Z)
and a body frame of reference (with cartesian coordinates x, y, and z or x1 , x2 , and x3 ). The latter
is xed to the body so the positions of the particles are constant in time in the body frame; the center
of mass of the rigid body is the origin for the body frame.
Call r the position vector in the inertial frame for an arbitrary point on the rigid body (e.g., one of
the particles). From our denitions,
r=R+r
where R is the position vector for the center of mass and r is the position vector in the body frame
(see Figure 6.1). An innitesimal displacement of that point on the rigid body may be written as the
sum of the displacement of the center of mass plus a rotation of the body, that is, (see Fig. 6.2)
dr = dR + d r
The corresponding velocity associated with this motion is
dr
dR d
=
+
r
dt
dt
dt
Since v = r ,
v =V+r
where V is the translational velocity of the (center of mass) of the body and is the angular velocity
of the body.
If the center of mass is at rest (V = 0) then v and are perpendicular so each particle rotates
(instantaneously) about an axis that passes through the origin and points in the direction . Important
note: Conservation of angular momentum does not imply that is constantrigid body motion turns
out to be more complicated. Finally, Landau goes on to show that if we have the origin of the body
frame located at a point dierent from the center of mass that the angular velocity, , is unchanged
(see (31.3)).
83

CHAPTER 6. RIGID BODY MOTION

84

Figure 6.1: Illustration of a rigid body with coordinates specied in an inertial frame of reference (with
cartesian coordinates X, Y , and Z) and a body frame of reference (with cartesian coordinates x, y,
and z).

CHAPTER 6. RIGID BODY MOTION

85

Figure 6.2: Illustration of the relation between an innitesimal displacement dr and an innitesimal
rotation d (not to scale).

CHAPTER 6. RIGID BODY MOTION

86

Inertial Tensor (32)


To analyze rigid body motion we need to formulate the Lagrangian in a useful form. We start with the
kinetic energy,
N
N

1
1
2
2

T =
m r =
m v
2
2
=1
=1
Note that Landau leaves particle indices implied to make the notation cleaner but Ill explicitly write
them out. From the above,
T

1
m (V + r )2
2
=1

1
=
m [V2 + 2V ( r ) + ( r ) ( r )]
2
=1

The last term may be rewritten as,


( r ) ( r ) = r (( r ) ) = 2 r2 ( r)2
Furthermore, the penultimate term is,
N

(
m V ( r ) =

=1

)
m r

(V ) = 0

=1

since the center of mass is the origin for the body frame. Therefore the kinetic energy may be written
as,
T

{
}
1
m V2 + [2 r2 ( r)2 ]
2
=1
(
) ( 3
)2
)( 3
N
3

1
1
x2,
i xi,
=
V2 +
m
2i
2
2
=1
i=1
i=1

=1

The term in the square brace may be written as,


( 3 3
) ( 3
)( 3
)
( 3
)
3
3
3

2
2
k xk, =
(x, i,k xi, xk, )
i k i,k
x,
i xi,
i k
i=1 k=1

=1

i=1

i=1 k=1

k=1

=1

Collecting the above the kinetic energy of a rigid body may be written as,
1
1
V 2 +
i k Ii,k
2
2 i=1
3

T =

k=1

where
Ii,k =

=1

( 3

)
(x2, i,k

xi, xk, )

=1

is the inertial tensor for a rigid body of N particles. In matrix notation this is,
T =

1
1
V 2 + T (I )
2
2

CHAPTER 6. RIGID BODY MOTION

87

where is a column vector, T is the transposed row vector, and I is a symmetric 3-by-3 matrix. Note
that the inertial tensor for solid bodies may be dened similarly, replacing the sum over particles with
integrals over mass density, , that is,
N

(r) dr

=1

See Problem 2, pg. 102 for some worked examples.


Landau introduces the repeated index summation shorthand (also known as Einstein notation):
When the same index appears twice in a term then summation over that index is implied. Here are
some examples,
Long Hand
Short Hand
3
3

i k Ii,k i k Ii,k
i=1 k=1
3

x2,

x, x, = x2,

=1

In this short hand notation,


1
1
V 2 + i k Ii,k
2
2
While this notation is elegant, in these notes Ill often explicitly write the summation signs, just for
clarity.
Since I is a real, symmetric matrix its always possible to nd a body coordinate system in which
the matrix is diagonal. Specically, the direction vectors for this coordinate system are the eigenvectors
of I and the elements of the diagonal matrix are the eigenvalues of I. In this form, I11 = I1 , I22 = I2 ,
I33 = I3 , and Iik = 0 (i = k). We call this the principal axes system and, in general, well assume that
it has been selected as the body frame.
Rigid bodies fall into three categories:
T =

Spherical Top I1 = I2 = I3
Symmetrical Top I1 = I2 = I3
Asymmetrical Top I1 = I2 = I3
Notice that we call a rotating rigid body a top, like the childrens toy. Also notice that a spherical
top need not be a sphere, for example, it could be a solid cube or any similarly symmetric object. Some
general results regarding the principal axes are:
If the body has a plane of symmetry then two of the principal axes lie in that plane and the third
is perpendicular to the plane.
If the body has an axis of symmetry then one of the principal axes is parallel to that axis.
For a co-linear system (i.e., particles along a line) we take x
as parallel to the particles and the
3
other two axes as perpendicular. In that case, I1 = I2 = m x23, and I3 = 0.

CHAPTER 6. RIGID BODY MOTION

88

Figure 6.3: Geometry of a cylinder (radius a) rolling inside a cylinder (radius R); see Problem 6, pg. 104.

Parallel Axis Theorem


While the dynamics is most conveniently formulated using body coordinates with the origin at the
center of mass we sometimes need the inertial tensor in a dierent set of coordinates. Choosing new
coordinates x the inertial tensor is,

=
Ii,k
m (x, )2 i,k xi, xk, )

Note that this transformation also aects the kinetic energy in a complicated way unless V = 0.
Call a the displacement from the unprimed to the primed coordinate systems, that is, r = r + a. If
the axes in the two systems are parallel (i.e., the system dier only by a displacement with no rotation)
then from the above,

Ii,k
= Ii,k + (a2 i,k ai ak )
which is the general form of the parallel axis theorem. If the displacement a is parallel to one of the

= Ii,i + a2 for i = 2, 3 and Ii,k


= Ii,k
principal axes, say the x1 -axis, then this result simplies to Ii,i
otherwise.

Constrained Rotations
In some cases the motion is simplied because the axis of rotation is constrained. Typically this results
in a simple relation between the angular velocity and the generalized coordinates; a common example
is rolling without slipping.
Example (Problem 6, pg. 104) Consider a cylinder rolling inside of another cylinder (see Figure 6.3).
The center of mass of the rolling cylinder is located at,
x =
y

(R a) sin

= (R a) cos

taking the origin at the center of the outer cylinder. Using as our generalized coordinate the velocity
of the center of mass is,

V = x 2 + y 2 = (R a)

CHAPTER 6. RIGID BODY MOTION

89

Figure 6.4: Decomposition of rolling into a translation plus a rotation about the center of mass.
The angular velocity is obtained by considering the rolling as a rotation about an instantaneous axis of
rotation located at the point of contact so,
=

V
Ra
=

a
a

with the direction of being along the z-axis. The rotational inertia along this (principal) axis through
the center of mass is
1
I = a2
2
(see Problem 2(c) on pg. 102).
The kinetic energy is (see (32.3)),
T

=
=
=

1
1
V 2 + I2
2
2
(
)2
1
1
Ra
2 2
(R a) + I
2
2
2
a
3
(R a)2 2
4

Notice that we may consider rolling as composed of a pure translation (with velocity V plus a pure
rotation about the center of mass (see Figure 6.4). We can also consider it as a pure rotation about the
point of contact, in which case,
T

=
=
=

1 2
I
2
{(
) }2
]
1[
Ra
2
I + a

2
a
3
(R a)2 2
4

Of course we may choose any representation and we should nd the same result.
In the presence of gravity the potential energy is,
U = gy = g(R a) cos
so the Lagrangian is,
L=

3
(R a)2 2 + g(R a) cos
4

CHAPTER 6. RIGID BODY MOTION

90

In the small oscillation approximation we may use the Taylor expansion cos = 1 21 2 + . . . to write,
(
)
1
1 3
(R a)2 2 (g(R a)) 2 + . . .
L=
2 2
2
The frequency of the small oscillations of the cylinder rolling inside a cylinder are,

k
g(R a)
g
=
=
=
3
3
2
m
2 (R a)
2 (R a)
so the oscillations are like those of a pendulum of length 32 (R a). Notice that even in the limit a 0
the period of the oscillations is aected by the rolling, that is, a particle sliding inside the cylinder
moves faster than a particle rolling inside of the cylinder.

Angular Momentum (33)


Lecture 20
We will now consider rigid body rotation and take the center of mass to be at rest (V = 0). In this
case the center of mass will be a xed point and, in general, it will be the only xed point with the rest
of the body in motion with a variable axis of rotation (i.e., the direction of the angular velocity, , will
not be constant in time).
We take the inertial (lab) frame to be coincident with the body frame at time t = 0; the velocity of
each particle on the body is then,
v = r
The total angular momentum for the body is,

M =
m r v

m r ( r )

m [r2 r (r )]

using the BACCAB rule. For any single component of M we may write,
[
]
3

Mi =
m r2 i x,i
xk, k

k=1

k=1

m [r2 ik x,i xk, ]

k Iik

k=1

or in matrix form,
M = I
Now what makes things complicated in the inertial (lab) frame is that although M is constant (in
the absence of external torques) the inertial tensor I changes in the inertial frame since the positions of
the particles change in that frame as the body rotates. For that reason we instead analyze the motion

CHAPTER 6. RIGID BODY MOTION

91

in the body frame where I is xed. The body frame is not an inertial frame and in it M is not constant
(though the magnitude M = |M| is constant in the absence of external torques). To keep the (already
complicated) analysis simple well take the body frame axes aligned with the principal axes of the body
so that I is diagonal.
Example Consider the rotation of an object thats a symmetric top (I1 = I2 = I3 ) such as a disk,
a cylinder, or a football (Fig. 46, pg. 107 in Landau and Figure 6.5 here).
For a symmetric top we can choose the direction of the x1 -axis so that the angular momentum is
initially in the x1 -x3 plane, which means that we may write it in the body frame of reference as

sin

M = M 0 = MM
cos
Since M = I we may nd the angular velocity as = I1 M. Fortunately, in the body frame the
inertia tensor is a diagonal matrix and thats easy to invert,

1/I1
= 0
0

0
1/I2
0

(M/I1 ) sin
M sin
0

=
0
0
0
(M/I3 ) cos
M cos
1/I3

Recall that the axis of symmetry is in the x3 direction so 3 = (M/I3 ) cos is the component of angular
velocity along that axis.
Later well see that its useful to express the result above for the angular velocity as,


pr sin
0
= 0 +

= ax x
3 + pr M
0
pr cos
ax

where
ax

pr

M
M
I1 I3
cos
cos =
3
I3
I1
I1
M
I1

In the inertial frame pr is the rate of precession of the body about M, which is xed in that frame
(again, see Figure 46, pg. 107 and Fig. 6.5 here). Note that if the angular momentum points roughly
in the x3 direction (i.e., if is small) then
ax
I1 I3
I1
=
cos
1
pr
I3
I3
For example, for a thin disk of radius R we have I1 = 41 R2 and I3 = 12 R2 (see Problem 2(c), pg. 102)
so ax /pr = 1/2. Physically, a disk-shaped object, such as a frisbee or a dinner plate, rotating with
an angular momentum (and angular velocity) roughly about its axis of symmetry has a spins about
that axis at half the frequency that it precesses (wobbles). A nice Java applet illustrating the motion
may be found at: http://faculty.ifmo.ru/butikov/Applets/Precession.html .

CHAPTER 6. RIGID BODY MOTION

92

Figure 6.5: Angular velocity in the inertial frame of reference for free rotation of a symmetric top. Note
that in this frame the angular momentum, M, is constant. Figure 46 in Landau.

CHAPTER 6. RIGID BODY MOTION

93

Equations of Motion for a Rigid Body (34)


To nd the equations of motion for a rigid body we just need two basic results that are already familiar.
Starting in the inertial (lab) frame, we know that the center of mass motion for a rigid body is,

dV
f = F
=
dt

where f is the force on particle and F is the net force. We may also write this in terms of linear
momentum, P, as,
dP
=F
dt
For angular momentum the result is similar,

dM
r f
=K=
dt

where K is the net torque. From our earlier result that M = I,


d
(I) = K
dt
The problem is that these results are only valid in the inertial frame for which the inertial tensor I is
not constant. To get a useful equation of motion for (t) we need to reformulate the problem in the
body frame.

Euler Equations (36)


[Note that were skipping Section 35 for the moment but will return to it in the next lecture.]
Consider an arbitrary vector A with components Ak in the inertial frame and components Ak in
the body frame. In general,
d
d
A= A+A
dt
dt
What this means is that the change of A in the inertial frame may be expressed as the sum of: the
change of A in the body frame plus the change of A due to the rotation of the body.
Since dM/dt = K in the inertial frame then,
d
M+M=K
dt
Writing this in terms of the components of these vectors in the body frame,

M
K
d 1 1 1 1
+

=
M2
2
M2
K2
dt
M3
3
M3
K3
At this point well drop the primes, understanding that well just be working in the body frame.
Since M = I if we align the body frame axes with the principal axes then each component is
Mk = Ik k and,

1
I1 d
(I3 I2 )2 3
K1
dt
I2 d2 + (I1 I3 )3 1 = K2
dt
3
I3 d
(I2 I1 )1 2
K3
dt

CHAPTER 6. RIGID BODY MOTION

94

which are Eulers equations for rigid body rotation.


Example Consider the symmetric top (I1 = I2 = I3 ) in free rotation (K = 0). From the third
equation,
d3
=0
dt
That is, the component of angular velocity along the symmetry axis, x3 , is constant. Notice that for
the other two components,
d1
I3 I2
+
2 3
dt
I1
I1 I3
d2
+
3 1
dt
I2
Since I1 = I2 we may dene
=

= 0
= 0

I3 I1
3
I1

and write the equations above as,


1 + 2

2 1

= 0
= 0

By applying d/dt to the rst equation we have,


1 + 2 = 0

Using the second equation here gives the equation of motion,


1 =0
1 + 2

with a symmetric result for 2 .


If the angular velocity is initially in the x1 -x3 plane (2 = 0 at t = 0) then the solution is
1 (t) = A cos t

2 (t) = A sin t

Comparing with our earlier analysis we see that = ax . The free rotation of a symmetric top in the
body frame is relatively simple: the angular velocity vector performs a rotation around the symmetry
axis with frequency (see Figure 6.6). Notice that the ratio,

I3 I1
I3
=
=
1
3
I1
I1
depends on the geometry of the object. For example, for a thin disk of radius R we have I1 = 14 R2
and I3 = 12 R2 (see Problem 2(c), pg. 102) so /3 = 1.
The next section will help us understand these results, obtained in the body frame, as they apply
to the inertial frame.

CHAPTER 6. RIGID BODY MOTION

95

Figure 6.6: Angular velocity in the body frame of reference for free rotation of a symmetric top. Note
that in the body frame the inertia tensor is constant so the angular momentum, M = I has a similar
rotation about the x3 axis. Compare with the result in the inertial frame (see Figure 6.5).

CHAPTER 6. RIGID BODY MOTION

96

Eulerian Angles (35)


Lecture 21
Recall that weve established the equations of motion (Eulers equations),
I3 I2
d1
+
2 3
dt
I1
d2
I1 I3
+
3 1
dt
I2
I2 I1
d3
+
1 2
dt
I3

= K1
= K2
= K3

However these apply to motion as observed in a non-inertial (body) coordinate system. We now need
a convenient way to transform back to an inertial (lab) frame of reference.
The orientation of a rigid body can be uniquely expressed in terms of three, non-orthogonal angles
,
and are about the Z, x3 ,
(see Figure 47, pg. 110 and Fig. 6.7 here). Notice that the rotations ,
and N (node line) axes, respectively; note that in some cases we are free to align x1 along the N -axis.
We can formulate the components of angular velocity in the body frame, i , in terms of these rotations,
specically,

= sin sin + cos


= sin cos sin

= cos +

Note that since the Euler angles are non-orthogonal 3 = even though is a rotation angle about

the x3 -axis; for example, if 0 then 3 + .


Example Lets repeat the example we did in Section 33, namely the free rotation of a symmetrical
top (I1 = I2 = I3 ). We are free to choose the direction of x1 anywhere in the plane perpendicular to
x3 so for convenience well take it such that (instantaneously) = 0. In that case we have,

=
= sin

= cos +

(Symmetrical Top)

We take the Z-axis of the inertial (lab) frame to be in the direction of the angular momentum, M,
which is xed in that frame. In the body frame the components of M are,
M1 = I1 1 = I1 ,

M2 = I2 2 = I2 sin ,

M3 = I3 3 = I3 ( cos + )

Since we have = 0 the x1 -axis is perpendicular to the Z-axis so we may also write these components
as,
M1 = 0 ,
M2 = M sin ,
M3 = M cos
By comparison we obtain the equations of motion,
= 0
I1 = M
= M cos
I3 ( cos + )

CHAPTER 6. RIGID BODY MOTION

Figure 6.7: Euler angles (see Fig. 46 in Landau)

97

CHAPTER 6. RIGID BODY MOTION

98

or
=
=

0
M
I1
M
M
cos
cos
I3
I1

The rst equation gives = constant, that is, the angle between the axis of the top and the direction
of M is constant. The second equation gives the angular velocity of the precession as = M/I1 = pr .
Finally, we have that = ax , as in the example in Section 33. Finally, we also obtain the same result
as before for,
3

cos +

pr cos + ax

See Fig. 46, pg. 107 in Landau, reproduced in Figure 6.5 in these notes.
Using our expression for the rotational kinetic energy in the principal axis coordinates,
Trot =

1
1
1
I1 21 + I2 22 + I3 23
2
2
2

combined with our relations between the s and the Euler angle rotations we can express T in terms
of the latter. The general expression for the asymmetric top (I1 = I2 = I3 ) is complicated but for the
symmetric top (I1 = I2 = I3 ) its just
Trot

=
=

1
1
I1 (21 + 22 ) + I3 23
2
2
1
1
2
2
2

2
I1 ( sin + ) + I3 ( cos + )
2
2

()

This expression allows us to write the Lagrangian using the Euler angles as the generalized coordinates.
The resulting equations of motion can be quite complicated. In some special cases we may extract a
solution but, in general, the solution requires numerical calculation.
Example Problem 1, pg. 112, analyzes the motion of a symmetric top with the lowest point xed
in space (similar to a childs top); we take that point as the origin of both the body and inertial (lab)
coordinates (see Fig. 6.8). Note that in this case the motion is such that the center of mass is not a
xed point; the constraint is that it remains a constant distance from the origin.
The Lagrangian is
L = Trot + TCM U
where Trot is given by (*). From our choice of coordinate system the angles and are the familiar
spherical coordinate angles so the center of mass kinetic energy is
TCM =

1 2 2 1 2 2 2
+ sin
2
2

and the potential energy is


U = g cos

CHAPTER 6. RIGID BODY MOTION

99

Figure 6.8: Symmetric top spinning with a xed endpoint in a constant gravitational eld (Fig. 48 in
Landau).

CHAPTER 6. RIGID BODY MOTION

100

Figure 6.9: Illustration of possible precession (motion about the Z axis) and nutation (motion in ) for
a symmetric top spinning with a xed endpoint in a constant gravitational eld (Fig. 49 in Landau).
By inspection Trot , TCM , and U do not depend explicitly on either or these two angles are cyclic
coordinates (see pg. 30). For cyclic coordinates the corresponding generalized (angular) momenta are
constant; they are:
p

L
= I3 ( + cos ) = M3

L
= [(I1 + ) sin2 + I3 cos2 ] + I3 cos = Mz

Physically, this result comes from the fact that the torque due to gravity is perpendicular to the Z-axis
and to the x3 -axis so these angular momenta are conserved.
The total energy is also constant and may be written as E = Trot + TCM + U . Combining these
expressions we may eliminate the terms containing and to leave us with a single equation of motion
to solve for (t), namely,
1 2
I + Ue () = E
2 1
where I1 = I1 + 2 . Landau gives the expressions for the eective potential Ue and the eective
(constant) energy E at the bottom of pg. 112. The important result is that this expression is in a form
that allows us to use the method developed in Section 11 to obtain the solution for one-dimensional
systems. The motion oscillates between the turning points max and min with Ue (max ) = Ue (min ) =
E . The resulting motion is illustrated in Fig. 49 in Landau, reproduced here in Figure 6.9.

Asymmetrical Top (37)


We nish our study of rigid body motion by considering the free (torque-less) rotation of an asymmetric
top with I1 < I2 < I3 . Well start with a general result, sometimes called the tennis racket theorem.
A tennis racket is an example of an asymmetrical top with x1 parallel to the length of the racket, x3
perpendicular to the face of the racket, and x2 in the direction perpendicular to those two axes.

CHAPTER 6. RIGID BODY MOTION

101

Figure 6.10: Intersection of a sphere and an ellipsoid for various ratios of E : M 2 .


From conservation of energy, E and of angular momentum, M, we have,
E
M2

1
(I1 21 + I2 22 + I3 23 ) = constant
2
= I12 21 + I22 22 + I32 23 = constant
=

Its useful to express this as


M12
M2
M2
+ 2 + 3
I1
I2
I3
M12 + M22 + M32

= 2E
= M2

which gives us two equations for the three variables M1 , M2 , and M3 . The rst equation is that of an
ellipsoid with semi-major axes,

a = 2EI1
b = 2EI2
c = 2EI3

so the ratio a : b : c is I1 : I2 : I3 . The second equation is that of a sphere with radius R = M .


The allowed values of M are the curves given by the intersection of the ellipsoid and the sphere; see
Figure 6.10. Note that since 2EI1 M 2 2EI3 that a R c.
When E M 2 /2I3 the curve of possible directions for M is a circle around the x3 axis. Physically,
this means that the angular momentum, in the body frame of reference, varies in time by circling near
the x3 axis (see Fig. 6.11). Since M is xed in the inertial frame this tells us that the body wobbles
(precesses) around the x3 axis as it spins, as weve already seen for the symmetric top. Similarly, when
E M 2 /2I1 the curve is a circle around the x1 axis and the object wobbles around this axis as it spins.
The curves near the x2 axis, however, are very dierent. If M is initially near but not exactly
parallel to the +x2 axis then the resulting motion causes the direction of M to wander signicantly,
reversing direction and pointing near the x2 axis before returning in a loop. Physically, in the inertial
frame the resulting motion is a complicated tumbling; with a tennis racket this tumbling is observed
when you try to ip a tennis racket upward with the face of the racket facing upward (see Fig. 6.12)

CHAPTER 6. RIGID BODY MOTION

Figure 6.11: Tennis racket spinning with angular momentum near the x1 or the x3 axis.

Figure 6.12: Tennis racket tumbling with angular momentum near the x2 axis.

102

CHAPTER 6. RIGID BODY MOTION

103

If we want to go further in the study of the motion of the asymmetric top then we have to solve
the Euler equations, which Landau proceeds to do in the rest of the section. The formulation is rather
complicated and the results are expressed in terms of special functions. For example, the period of the
motion for M in the body frame (or, equivalently for in this frame) is

I1 I2 I3
T = 4K(k)
(I3 I2 )(M 2 2EI1 )
where,

K(k) =
0

/2

du

1 k 2 sin2 u

is a complete elliptic integral of the rst kind and

(I2 I1 )(2EI3 M 2 )
k=
(I3 I2 )(M 2 2EI1 )
Landau goes on to show that in the inertial frame, in general, the rotating object never returns to its
original orientation. We wont study the details of this analysis any further because the calculation
of such dynamics is now done by solving the Euler equations numerically rather than evaluating the
analytical results involving special functions. Nevertheless, its always useful to have exact analytical
results available.

Motion in Non-inertial Frames (39)


Lecture 22
Before you read this section I strongly urge you to watch the classic PSSC lm, Frames of Reference,
featuring Profs. Donald Ivey and Patterson Hume (http://www.youtube.com/watch?v=UXXhlZW1zus).
Landau reminds us that Lagranges equations,
(
)
d L
L

=0
dt v
r
are valid in all frames, not just inertial frames. Weve mostly restricted ourselves to inertial frames
since the Lagrangian is typically simplest in an inertial frame and, thus, so are the resulting equations
of motion. But lets see how the Lagrangian changes when we formulate it in a non-inertial frame.

Non-inertial frame with translational motion


Consider an inertial frame K0 and a non-inertial frame K moving with translational velocity V(t) (see
Figure 6.13). For the moment we only consider the case where frame K is not rotating; note that if V
is constant then K would also be an inertial frame.
Take a particle of mass m moving in a potential U (r); in the inertial frame K0 the Lagrangian is
L(r0 , r 0 , t) =

1
1
m|r0 |2 U (r0 ) = m|v0 |2 U (r0 )
2
2

To transfer from one frame to the other well need the Galilean transformation, v0 = v + V(t). The
Lagrangian in frame K is
1
L (r , r , t) = m|v + V(t)|2 U (r )
2

CHAPTER 6. RIGID BODY MOTION

104

Figure 6.13: Inertial frame K0 and non-inertial frame K ; the latter has translational velocity V(t).

CHAPTER 6. RIGID BODY MOTION

105

Figure 6.14: Inertial frame K0 and non-inertial frame K; the latter has angular velocity (t).
where U (r ) = U (r0 ). We may write,
|v + V(t)|2

= (v + V(t)) (v + V(t))
= |v |2 + 2v V + |V|2

but since
v V =
we have,

d
dV
(r V) r
dt
dt

{
}
1
d
2
dV
2
L = m |v | 2r
+ 2 (r V) + |V| U (r )
2
dt
dt

The last two terms in the kinetic energy may be discarded from the Lagrangian (why?) which gives,
L

=
=

1
dV
m|v |2 mr
U (r )
2
dt
1
m|v |2 + r FK U (r )
2

of the non-inertial frame.


where FK is the eective force on the particle due to the acceleration V

Non-inertial frame with rotational motion


Now consider a rotating frame of reference K with angular velocity that instantaneously overlaps
with the inertial frame K0 at time t = 0 (see Figure 6.14). Note that Landau does the more general
case in which there is both translation and rotation but here well take V = 0.
To formulate the Lagrangian in the non-inertial frame we use,
r0 = r

r 0 = r + r

CHAPTER 6. RIGID BODY MOTION

106

so this Lagrangian for a particle in a potential U is,


L(r, r , t)

=
=

1
m|v + r|2 U (r)
2
1
1
m|v|2 + m(v r) + m| r|2 U (r)
2
2

We may now use this Lagrangian to evaluate the equation of motion in the rotating frame of reference.
Of course, its much more complicated than what we have in the inertial frame (v 0 = U/r0 ); even
= 0) the equation of motion is,
with the simplication that the angular velocity is constant (i.e., if
mv =

U
+ 2mv + m (r )
r

The middle term on the right hand side is the Coriolis force and the last term is the centrifugal force.
Example In Problem 1 of this section Landau evaluates the equations of motion for free fall near
the Earths surface. The rotating frame is an observer standing on Earth, which rotates with angular
velocity e . One nds that an object that falls from a height z = h at the equator lands a distance
y = 13 (2h/g)3/2 g East of its initial point. This result is paradoxical because the surface of the Earth
is moving eastward.
To better understand this result well rst lets look at the problem in a dierent way. Landaus
analysis is done in the rotating, non-inertial frame but lets instead use the inertial frame, in which
angular momentum is conserved. In either frame the problem is dicult to solve exactly so well make
some approximations based on the fact that Earths radius is large.
Consider an object falling
on a at Earth thats moving with constant velocity V (see Figure 6.15).
The time in the air is T = 2h/g since y = h 12 gt2 . In the stationary frame the particle has constant
horizontal West/East velocity vy = V ; note that East is in the j direction.
A particle on the equator of a spherical, rotating Earth (radius Re ) has an angular momentum with
magnitude,
M = mr2
For the falling particle the force of gravity exerts no torque so the angular momentum is constant thus,

m(Re + h)2 e = m(Re + z)2 (z)

where (z)
is the angular velocity when the particle is at height z; note that (h)
= e since the particle
is initially moving with the Earth. Since Re h z,
(
)2 (
)2
h
z
e 1 +
1+
Re
Re
(
)
2(h z)
e 1 +
Re

(z)
=

Note that in this approximation we may keep our picture of the Earth being at.
We now have the West/East velocity,
vy

= r = (Re + z)(z)
e (Re + 2(h z)) = V 2(h z)e

CHAPTER 6. RIGID BODY MOTION

107

Figure 6.15: Geometry for falling on a at Earth moving with velocity V. The correction due to the conservation of angular momentum for a rotating Earth will shift the landing spot by y = 13 (2h/g)3/2 g
East of its initial point.
The displacement resulting from this eastward velocity is,
T
T

y=
vy (z)dt =
V dt
0

but since z h 12 gt2 and T

2(h z(t))dt

2h/g,

= V T

gt2 e dt
0

1
1
= V T ge T 3 = V t ge (2h/g)3/2
3
3
which agrees with Landaus result. Finally, we may generalize this result for latitude other than
the equator ( = 0) by replacing e with e cos . This Coriolis displacement is not very noticeable;
putting in some values we have that for h = 5 meters the displacement y is about 0.2 mm.
Example Lets continue the example above but calculate the eect of the Coriolis force on a falling
object at the equator by the method of successive approximations, as Landau does in Problem 1, pg. 129.
However, while Landau only works out the second order correction, well also work out the third order
perturbation. Specically, we want to nd v(t), the velocity of the falling object, by solving
v = 2v e + g

()

Towards this end well express the velocity as,


v = v1 + v2 + v3 + . . .
The rst order result is obtained by neglecting the Coriolis force, that is,

v 1 = g = g k

CHAPTER 6. RIGID BODY MOTION

108

so

v1 = gt k
since we take the object initially at rest.
To obtain the next order we include the Coriolis term in the acceleration and write (*) as,
v 1 + v 2

2(v1 + v2 ) e + g

2v1 e + g

where weve dropped the v2 e term as being higher order. Since v 1 = g, to second order we have,
v 2

= 2v1 e
(e i)
= 2(gt k)
= 2gte j

so
v2 = gt2 e j
If we integrate in time to get r2 then we recover our expression for the Eastward displacement y from
the previous example.
Lets do one more order of perturbation expansion, specically,
2v2 e

v 3

= 2(gt2 e j) (e i)

= 2gt2 2 k
e

which is an upward acceleration, opposite from gravity. The centrifugal acceleration has a similar
contribution,
v c

= e (r e )
=

2e r
2e r

(centrifugal)

e (e r)

= ac + z2e k
where ac is the centrifugal acceleration at the earths surface. Notice that we used the approximation
that the motion remains in the equatorial plane so e r 0. If we use the approximation for the
falling object that z(t) = h 21 gt2 then
1 gt2 2 k

v c ac + h2e k
e
2
Note that the upward acceleration due to the Coriolis force increases with time while that of the
centrifugal force decreases with time as the object falls.
One nal note: At latitude the third order result for the Coriolis acceleration on a falling object
is,
v

= 2g2e t2 sin cos i


2ge t cos j
+(2g2e t2 cos2 g)k

CHAPTER 6. RIGID BODY MOTION

109

Notice that o the equator theres a deection of the falling motion in the direction of the equator (i.e.,
in the Northern hemisphere the deection is southward, which is the x direction).

Review for Second Midterm


Lecture 23

Second Midterm
Lecture 24

Chapter 7

Canonical Equations
Hamiltonian Dynamics (40)
Lecture 25
In Lagrangian dynamics we formulate the dynamics based on the Lagrangian,
L(q1 , . . . , qs , qi , . . . , qs , t)
We already introduced the generalized momentum, dened as,
pi =

L
qi

Landau shows that by Legendre transformations one may always construct from the Lagrangian another
function called the Hamiltonian,
H(q1 , . . . , qs , pi , . . . , ps , t)
which also gives the dynamics of the system. The connection between the two may be written as,
H=

pi qi L

i=1

which we recall is the same as the denition of the energy (see Section 6). The dierence is that before
we dened the energy as E(q1 , . . . , qs , qi , . . . , qs , t) whereas the Hamiltonian is a function of ps and qs.
From Lagranges equations we can derive Hamiltons equations (also known as the canonical equations),
d
H
d
H
qi (t) =
;
pi (t) =
dt
pi
dt
qi
Notice that in Lagrangian dynamics the equations of motion are a set of s second order ODEs (e.g.,
qi = . . .) for Hamiltonian dynamics we have a set of 2s rst order ODEs. In Lagrangian dynamics the
initial conditions are qi (t = 0), qi (t = 0) while in Hamiltonian dynamics its qi (t = 0), pi (t = 0). While
the canonical equations have an elegant, symmetric form, it doesnt mean that Hamiltonian dynamics
necessarily leads to a simpler construction of the solution for the dynamics.
Example Consider the Lagrangian for a pendulum (mass m2 ) with a pivot point (mass m1 ) thats
free to slide horizontally (Problem 2, pg. 11),
L = 21 (m1 + m2 )x 2 + 12 m2 2 2 + m2 cos x + m2 g cos
110

CHAPTER 7. CANONICAL EQUATIONS

111

in the coordinates x and . To formulate the Hamiltonian we would write the expression for the energy,
E

L
L
x +
L
x

1
2 (m1

+ m2 )x 2 + 12 m2 2 2 + m2 cos x m2 g cos

and make the change of variable from velocities to momenta, that is,
H(q1 , . . . , qs , pi , . . . , ps , t) = E(q1 , . . . , qs , qi , . . . , qs , t)
To make this change of variable we evaluate the momenta,
px

L
= (m1 + m2 )x + (m2 cos )
x
L
= (m2 2 ) + (m2 cos )x

We may solve these two equations to nd x and in terms of px and p , for example,
x =

px cos p
m1 + m2 (1 cos )

The expression for is similar. Since the expressions for x and are rather complicated its clear that
the resulting expression for H will be messy so this system is easier to treat using Lagrangian dynamics.
The example above may have you questioning the utility of Hamiltonian dynamics. After all, it
seems as if it introduces additional steps to formulate H and yields equations of motion that are not
necessarily any easier to solve. Whats not yet obvious is the exibility thats available in Hamiltonian
dynamics because instead of using s generalized coordinates we have 2s coordinates and momenta with
a exibility as to the choice of the momenta. You might wonder, Once we choose the coordinates and
specify the Lagrangian arent the momenta entirely specied? Yes, but we have freedom in the choice
of our Lagrangian, which means that we have some freedom in the choice of momenta.
Example Consider the two Lagrangians,
L=

1 2
mq
2

L =

1 2
mq + qf
(q)
2

where f is some specied function. You can check that both Lagrangians give the equation of motion
for a free particle in one-dimension since the latter gives,
(
)
d L
d
L
df
=

(mq + f (q)) q
dt q
q
dt
dq
dq df
df
= m
q+
q
dt dq
dq
= m
q
This is not unexpected since the two Lagrangians dier by a term of the form df (q, t)/dt.
Not surprisingly the energy is the same for these two Lagrangians,
E=

1
L
q L = mq2
q
2

E =

L
1
q L = mq2
q
2

CHAPTER 7. CANONICAL EQUATIONS

112

however the generalized momenta for the two Lagrangians are dierent,
p=

L
= mq + f (q)
q

L
= mq
q

p =

q = p/m

q = (p f (q))/m

1 2
p /m
2

H(q, p ) =

or
Finally, the Hamiltonians are,
H(q, p) =

1
(p f (q))2 /m
2

In this case it happens to be that p is more convenient than p but in other situations this freedom of
choice for the momenta will prove very useful.

Routhian (41)
The Routhian is a hybrid that mixes Lagrangian and Hamiltonian dynamics by treating some of the
generalized coordinates by the former and the rest by the latter. For example, suppose our system has
two generalized coordinates, x and . The corresponding Lagrangian and Hamiltonian are of the form,
t)
L(x, , x,
,

and

H(x, , px , p , t) =

L L
+
L
x

We could formulate a Routhian of the form,


R(x, , x,
p , t) =

L
L

in which case wed use Lagranges equations to get the equations of motion for x and Hamiltons
equations to get the equations of motion for . Landau gives an example (symmetric top spinning on
its end) in which the Routhian is simpler than the Lagrangian. Its interesting that such a formulation
is an option but we wont consider it further.

Poisson Brackets (42)


Dene Poisson brackets as
[f, g] =

)
s (

f g
f g

pk qk
qk pk

k=1

By inspection,
[f, g] = [g, f ]

[f, f ] = 0

For arbitrary f (q1 , . . . , qs , p1 , . . . , ps , t),


df
f
=
+ [H, f ]
dt
t

()

This is a consequence of the canonical equations,


dqi
H
=
dt
pi

dpi
H
=
dt
qi

CHAPTER 7. CANONICAL EQUATIONS

113

Notice that the canonical equations themselves may be written as


dqi
= [H, qi ]
dt

dpi
= [H, pi ]
dt

Landau derives several useful identities for Poisson brackets, for example,
[pi , qk ] = i,k
which well use later in this lecture. Poisson brackets serve as an elegant link between classical and
quantum mechanics since (*) applies in both with a suitable denition of the Hamiltonian operator and
of Poisson brackets in terms of the commutation of operators, [f, g] = (i~)1 (f g gf ).

Canonical Transformations (45)


Lecture 26
We have commonly used various types of generalized coordinates. For example, cartesian coordinates
(q1 = x, q2 = y) and polar coordinates (Q1 = r, Q2 = ). The transformation from cartesian to polar is

;
Q2 = atan(q2 /q1 )
Q1 = q12 + q22
This is an example of a point transformation, Qk = Qk (q1 , . . . , qs , t). We know that we can formulate
the Lagrangian as L(q1 , . . . , qs , t), use the transformation q Q to get L(Q1 , . . . , Qs , t), formulate and
solve the equations of motion for the Qs and then transform back Q q. Of course the dynamics are
equivalent but the formulation is simpler in the transformed coordinates (e.g., normal coordinates for
linear oscillations).
We can use point transformations in Hamiltonian dynamics,
qk

Qk

L(q, q,
t)
qk

t)
Pk =
L(Q, Q,
Q k
pk =

But the beauty of Hamiltonian dynamics, as mentioned in the beginning of the chapter, is that we have
greater freedom in our choice of P s and Qs. Specically, we may dene canonical transformations,
Qk = Qk (p1 , . . . , ps , q1 , . . . , qs , t)

Pk = Pk (p1 , . . . , ps , q1 , . . . , qs , t)

Notice that point transformations are a special case of canonical transformations with Qk = Qk (q1 , . . . , qs , t)
and Pk = Pk (q1 , . . . , qs , t).
Freedom is good but not anarchywe need to ensure that our new Hamiltonian H (P, Q, t) describes
the same dynamics as H(p, q, t). Landau derives the result that a canonical transformation is valid if
there exists a generating function, F (q, Q, t), for the transformation such that,
pk
Pk

F (q, Q, t)
qk

=
F (q, Q, t)
Qk

The new Hamiltonian is,


H (P, Q, t) = H(p, q, t) +

F
t

CHAPTER 7. CANONICAL EQUATIONS

114

Example Consider a system with simple harmonic motion; the Lagrangian is


L(q, q)
=

1 2 1 2
mq kq
2
2

so p = L/ q = mq.
The Hamiltonian is,
H(q, p) =

1 2 1 2
p + kq
2m
2

F (q, Q) =

1
q2
km
2
tan Q

Consider the generating function,

so

F
q
= km
q
tan Q
F
1
q2
=
=
km 2
Q
2
sin Q

P
With a little algebra we nd,

2P/ km sin Q

=
2P km cos Q

The Hamiltonian in the new variables is,


H (P, Q)

F
+ H(p, q)
t
1 2 1 2
=
p + kq
2m
2
)2
)2
(
(

1
1
=
2P km cos Q + k
2P/ km sin Q
2m
2

k
=
P = P
m

which is very simple! The canonical equations give,


dQ
H
=
=
dt
P

dP
H
=
=0
dt
Q

Q(t) = t + c1
P (t) = c2

where the constants c1 and c2 are xed by the initial conditions. Described in terms of P and Q the
motion is quite simple. Physically, Q is a generalized angle and P is a generalized amplitude, which is
constant since the ratio (energy)/(frequency) is constant.
An alternative approach to formulating a canonical transformation from a generating function is to
guess a useful transformation and then verify that its indeed a canonical transformation. Landau shows
that a transformation is canonical if and only if,
[Qi , Qk ] = 0

[Pi , Pk ] = 0

[Pi , Qk ] = i,k

CHAPTER 7. CANONICAL EQUATIONS

115

For a system with just one coordinate its only necessary to show that [P, Q] = 1.
Example Consider a system with a single coordinate q and corresponding momentum p. We
introduce the transformation,
Q=q
;
P = p + f (q)
In an example done in Section 40 we showed that L(q, p) and L (Q, P ) produce the same dynamics,
which implies that this must be a canonical transformation. We can verify this by computing the
Poisson bracket,
( )
P Q P Q
df
(0) = 1
[P, Q] =

= (1)(1)
p q
q p
dq
and, trivially, [Q, Q] = [P, P ] = 0 since we only have one coordinate.

Hamilton-Jacobi equation (47)


The previous section discussed how a canonical transformation q, p Q, P may be dened in terms of
a generating function, F (q, Q, t). Specically,
pk =

F
qk

Pk =

F
Qk

H (P, Q, t) = H(p, q, t) +

F
t

An alternative generating function for this transformation is (q, P, t) for which,


pk

Qk

H (P, Q, t) =

(q, P, t)
qk

(q, P, t)
Pk

H(p, q, t) +
t

(1)
(2)
(3)

Well see in a moment why this form of the generating function is useful.
Suppose we could nd a transformation for which the Hamiltonian was,
H (P, Q, t) = 0
In that case the canonical equations would give,
d
H
Qk (t) =
=0
dt
Pk
d
H
Pk (t) =
=0
dt
Qk

Qk = constant

Pk = constant

In other words all the Qs and P s become cyclic variables. This is a special transformation and Landau
writes the generating function for it as = S(q, P, t).
How might we nd this transformation? From (*3) we know that if H = 0 that implies that the
generating function obeys the equation,
S
+ H(q, p, t) = 0
t

CHAPTER 7. CANONICAL EQUATIONS

116

Furthermore, using (*1) we have,


S
S
+ H(q,
, t) = 0
t
q

(HJ)

which is a partial dierential equation for S(q, P, t) known as the Hamilton-Jacobi equation.
Before going to the rst example, a few notes regarding the Hamilton-Jacobi equation:
Since the transformed momenta Pk are constant well write Pk = k so S(q, P, t) = S(q, t; ).
The generating function S turns out to be the action, that is,

S = L(q(t), q(t),

t) dt
where L is the Lagrangian. This result comes from the fact that,
d
S(q, t) =
dt

S S dqi
+
t
qi dt
i=1
s

S
+
pi qi
t
i=1

H +

pi qi

From (1)
From (HJ)

i=1

So why do we need the Hamilton-Jacobi equations if we can just nd S as the integral of the
Lagrangian? The problem is that to solve that indenite integral we need to know q(t) as a
function of time, which means we already know the motion.
When the Hamiltonian does not depend explicitly on time then H(q, p) = E = constant. In that
case we may write,
S(q, t; ) = S (q; ) Et
where S is Hamiltons characteristic function.
Partial dierential equations are typically much harder to solve than ordinary dierential equations. To make things worse, the Hamilton-Jacobi equation is often non-linear. One of the more
successful approaches, however, turns out to be separation of variables. The idea is to use a trial
solution of the form,
S = S1 (q1 ; 1 ) + . . . + Ss (qs ; s ) + St (t; t )
Furthermore when energy is conserved we have St (t; t ) = Et.
Lecture 27
Example To illustrate the formulation of a solution for the Hamilton-Jacobi equation well use it to
solve the 2D simple harmonic oscillator. Clearly, this can be easily solved using Lagrangian mechanics
so this is cracking a walnut with a sledgehammer. The Lagrangian is,
L(q1 , q2 , q1 , q2 ) =

1
1
1
m(q12 + q22 ) k1 q12 k2 q22
2
2
2

CHAPTER 7. CANONICAL EQUATIONS

117

The corresponding Hamiltonian is,


H(q1 , q2 , p1 , p2 ) =

1
1 2
1
(p + p22 ) + k1 q12 + k2 q22
2m 1
2
2

The Hamilton-Jacobi equation is


[(
)2 (
)2 ]
1
S
S
1
1
S
+
+ k1 q12 + k2 q22 +
=0
2m
q1
q2
2
2
t
Using separation of variables trial solution,
S = S1 (q1 ; 1 ) + S2 (q2 ; 2 ) Et
the HJ equation is,
{

1
2m

dS1
dq1

)2

1
+ k1 q12
2

{
+

1
2m

dS2
dq2

)2

1
+ k2 q22
2

}
E =0

The key to using separation of variables is to notice that the equation above may be written as:
{Function of q1 } + {Function of q2 } + E = 0
This means that each of the rst two terms must equal a constant, that is,
1
2m
1
2m

(
(

dS1
dq1

)2
)2

1
+ k1 q12
2

dS2
1
+ k2 q22
dq2
2
1 + 2 E

= 1
= 2
= 0

The equation for S1 is easy to solve; well re-write it as,


(

dS1
dq1

)2

1
= 2m(1 k1 q12 )
2

so

1
2m(1 k1 q12 ) dq1
2

1
S1 =
2m(1 k1 q12 ) dq1
2
dS1 =

or

Integrating we get,
1
S1 (q1 ) = q1
2

1
2m(1 k1 q12 ) +
2

(
)
m
m1
1 asin 2
k1
k1

Of course the expression for S2 is similar. Note that sometimes its more convenient to leave the solution
in integral form.
Weve seen that the HJ equation yields the canonical transformation that yields constant Qk and
Pk . But how can we use S(q, P, t) to nd qk (t)? We have already chosen Pk = k and were free to

CHAPTER 7. CANONICAL EQUATIONS

118

set Qk = k . Note that these constants will later be set by the initial conditions, such as xing the
amplitude and phase. From the denition of the generating function,
Qk =

S(q, P, t)
Pk

k =

S(q, t; )
k

Now we can solve this equation to get q(t). If S formulated using separation of variables then,
k =

Sk
E
(S1 (q1 ; 1 ) + . . . + Ss (qs ; s ) Et) =

t
k
k
k

Remember that E depends on the s; in the example above we had E = 1 + 2 .


Example Returning to our example,
1 =

S
1

=
=

=
Solving for q1 (t) gives

S1
E

t
1
1

2m
dq1

t
2
1 12 k1 q12
(
)]
[
m
2
k1

acos
q1
t
2
k1
21

q1 (t) =

21
cos
k1

k1
t+
m

k1
1
m

with a similar result for q2 (t). Notice that the constant is connected to the amplitude and is linked
to the phase, that is,

21
k1
q1 (t) = A1 cos(1 t + 1 )
;
A1 =
;
1 =
1
k1
m

with 1 = k1 /m. From this 1 = 12 k1 A21 so


E = 1 + 2 =

1
1
k1 A21 + k2 A22
2
2

which is expected from the form of the Hamiltonian.


Clearly the Hamilton-Jacobi method is overkill for the 2D harmonic oscillator but it proves useful for
more complicated problems in celestial mechanics especially when used in conjunction with perturbation
methods.

Example As a second example of nding the motion using the Hamilton-Jacobi method lets consider the motion of a particle under a central force. The motion is in a plane and the Hamiltonian
is,
(
)
1
1 2
2
H(r, , pr , p ) =
pr + 2 p + U (r)
2m
r

CHAPTER 7. CANONICAL EQUATIONS

The HJ equation is,


1
2m

[(

S
r

)2

1
+ 2
r

119

)2 ]
+ U (r) +

S
=0
t

Well solve this using the separation of variables trial solution,


S = Sr (r; r ) + S (; ) Et
which gives,
1
2m

[(

dSr
dr

)2

1
+ 2
r

dS
d

)2 ]
+ U (r) E = 0

To collect all the terms that depend on r we multiply this equation by 2mr2 and write it as,
{ (
} {(
)2
)2 }
dSr
dS
2
2
r
+ 2mr (U (r) E) +
=0
dr
d
This equation has the form,
{Function of r} + {Function of } = 0
so it must be that,
(
)2
S
= = r = constant

The generating function has the general property that,


pk =

S
qk

In this example,
S
dS
=
= constant

d
In other words, angular momentum is constant since is a cyclic variable. From this we have S = p
and = p2 .
Turning now to nding Sr , we have
(
)2
dSr
r2
+ 2mr2 (U (r) E) = p2
dr
p =

or

dSr
dr

)2
=

p2
+ 2m(E U (r))
r2

which may be solved in integral form as,

Sr =

p2
r2

+ 2m(E U (r)) dr

Finally, we may assemble the pieces and compute the motion using,
r =
With a little algebra we get,

p
S
=
[Sr + p Et]
r
r p

=
r2

p dr

+ 0
2
2
p /r + 2m(E U (r))

CHAPTER 7. CANONICAL EQUATIONS

120

which is the same as (14.7) for the motion of a particle in a central eld with (constant) angular
momentum M = p . Landau works through a more general case of this example on pg. 151 with
U = a(r) + b()/r2 .

Time-Dependent Perturbation Theory (Goldstein 12.2)


Lecture 28
The set of exactly solvable problems in classical mechanics is rather limited (if you disregard the trivial
variations such as those found in freshman physics books). For many important problems of interest
(e.g., Is the moons orbit stable?) we have to resort to either numerical methods or perturbation theory.
In this section we consider the latter.
We had a taste of perturbation theory in Section 28, in evaluating anharmonic corrections to simpler
harmonic motion, and in Section 39, to determine the deection due to Coriolis forces. We now apply
perturbation theory in Hamiltonian dynamics, where is particularly powerful. We also have the added
benet that the Hamiltonian formulation connects nicely to perturbation theory in quantum mechanics.
Well start with the unperturbed Hamiltonian, H0 (q, p, t) for which its assumed that weve already
solved the Hamilton-Jacobi PDE to get the generating function S(q, P, t). Recall that the canonical
transformation dened by S makes the P s and Qs constants so weve written them as P = , Q =
and S(q, t; ). Since perturbation expansions are already a little messy Ill present it for s = 1 but the
extension to multiple coordinates and momenta is not intellectually challenging.
Now write our perturbed Hamiltonian as
H(q, p, t) = H0 (q, p, t) + H(q, p, t)
Idea now is to use the same canonical transformation that was formulated from the Hamilton-Jacobi
equation for H0 for the perturbed Hamiltonian H. Since S is not the solution of the HJ equation for
H we should probably label it as but well stick with Goldsteins notation and just remember that
its the S that comes from H0 .
Transforming the perturbed Hamiltonian from (q, p) to (Q, P ) we have,
H (Q, P, t) =

H(Q, P, t) +

S
t

= H0 (Q, P, t) + H(Q, P, t) +

S
t

= H(Q, P, t)
Note that Goldstein writes H as K. Of course the Q and P will no longer be constant in time but we
can nd equations for their motion using the canonical equations,
dQ
H
=
dt
P

dP
H
=
dt
Q

At this point weve not made any approximations but weve not made much progress either since these
canonical equations are probably just as dicult to solve as those for H(q, p, t).
We now introduce the perturbation expansion,
Q(t) =

Q0 + Q1 (t) + Q2 (t) + . . .

P (t) =

P0 + P1 (t) + P2 (t) + . . .

CHAPTER 7. CANONICAL EQUATIONS

121

where Q0 = and P0 = , that is, for the unperturbed problem. The perturbation expansion for the
canonical equations is


H
H
dQ1
dP1
=
=

;
dt
P Q0 ,P0
dt
Q Q0 ,P0


dQ2
H
dP2
H
=
;
=
dt
P Q1 ,P1
dt
Q Q1 ,P1
...
Goldstein writes this in a compact form as,

(
)

Pn
n = J
H
;
n =
Qn

n1

(
;

J=

0 1
1 0

Example To illustrate the perturbation technique well solve a problem that doesnt require it
but has all the elements. Furthermore, since the exact solution is known we can compare it to our
perturbation solution. The unperturbed Hamiltonian well use it that of a free-particle,
H0 =

p2
2m

The corresponding Hamilton-Jacobi equation is


( )2
1
S
S
+
=0
2m q
t
It is easily solved and we nd,
S = q

2
t
2m

where

S
=P
q

The transformed coordinate is,


Q==

S
S

=
=q t
P

or

t+
(Unperturbed)
m
Notice that since q = /m and Q = we may write this as,
q=

Q = q qt

Furthermore, P = = mq = p so for a free particle the transformation (q, p) (Q, P ) is a Galilean


transformation to the moving inertial frame of reference (with velocity q)
for the coordinate Q while
keeping the momentum P as dened in the original frame of reference.
We now introduce the perturbation,
1
m 2 q 2
2
We know the exact solution since this is just the Hamiltonian for simple harmonic motion with frequency
. In the transformed coordinates the perturbation is,
(
)2
1
H(Q, P, t) =
m 2
t+
2
m
(
)2
1
1
m 2
Pt + Q
=
2
m
H(q, p, t) =

CHAPTER 7. CANONICAL EQUATIONS

122

Lets see how this looks if we do not use perturbation theory. The canonical equations are,
(
)
dQ
H
1
2
=
= t
Pt + Q
(1)
dt
P
m
(
)
H
dP
1
=
Pt + Q
= m 2
(2)
dt
Q
m
Although this looks nasty to solve if we apply d/dt to the second equation, after a few manipulations
we get,
d2 P
= 2 P
dt2
so, as expected, we have simple harmonic oscillator.
But lets see how perturbation theory solves this problem. To rst order (*1) and (*2) are written
as,
(
)
1
dQ1
= 2 t
P0 t + Q0
dt
m
(
)
dP1
1
= m 2
P0 t + Q0
dt
m
These are easy to solve but to simplify the algebra a bit lets take the initial condition q(0) = 0 which
means that Q0 = 0. In that case,
dQ1
dt
dP1
dt

1
P0 2 t2
m

= 2 P0 t

so
Q1
P1

1 2
P0 t3
3m
1
= 2 P0 t2
2

Finally, we assemble the rst-order perturbed motion as,


q(t)

=
=
=
=

1
P (t)t + Q(t)
m
1
(P0 + P1 )t + (Q0 + Q1 )
m(
)
(
)
1 2
1 2
1
2
3
P0 P0 t t +
P0 t
m
2
3m
(
)
P0
1
t 3 t3
m
6

Comparing this with the Taylor expansion of the exact solution for simple harmonic motion,
q(t) =
=

P0
sin t
m (
)
P0
1 3 3
t t + . . .
m
3!

demonstrates the correctness of our perturbation solution for small values of t 1.

CHAPTER 7. CANONICAL EQUATIONS

123

Goldstein has more examples and extensions. He also discusses how the real power of this method
shines in celestial mechanics.

Infinitesimal Transformations
Lecture 29
Consider an innitesimal canonical transformation,
Qi = qi + dqi

Pi = pi + dpi

()

The generating function for this transformation may be written as,


(q, P ) =

qi Pi + (q, P ) = 0 (q, P ) + (q, P )

i=1

where 0 (q, P ) is the null transformation. Using (45.8) we can verify that 0 (q, P ) leaves the coordinates
and momenta unchanged,
pj

qi

0
=
Pi =
ij Pi = Pj
qj
qj
i=1
i=1

Qj

Pi

0
=
qi =
ij qi = qj
Pj
Pj
i=1
i=1

The innitesimal transformation gives,


pj

Qj

0
= Pj +
qj
qj
0

= qj +
Pj
Pj

Since the transformation is innitesimal with Pj = pj + O() we have (q, P ) = (q, p) + O() so,

qj

+ O(2 )
= qj +
pj

pj

= Pj +

Qj
and we may drop the higher order term.
Comparing this result with (*) gives,

pj

dpi =

dqi

=
dt
pj

dpi
=
dt
qj

dqi =
Taking = dt gives,

qj

This should remind you of the canonical equations,


H
dqi
=
dt
pj

dpi
H
=
dt
qj

CHAPTER 7. CANONICAL EQUATIONS

124

so we see that the generating function,


= 0 + dtH

(+)

produces the transformation,


(

qi (t)
pi (t)

)
dqi
dt
dt
(
)
dpi
dt
pi (t + dt) = pi +
dt
qi (t + dt) = qi +

That is, the Hamiltonian produces the generating function for the transformation that takes us from
the present to the future.
Finally, from the Hamilton-Jacobi equation,
S
= H
t
if we identify (+) as the Taylor expansion of in time then

=H
t
so = S. So we may think of the action, S, as producing the transformation that takes us from the
future back to the (constant) initial state. Landau presents this result by a dierent approach (using
Poisson brackets) at the end of Section 45.

Liouvilles Theorem (46)


A useful concept from Hamiltonian dynamics is that the state of a system is uniquely dened by a
point (q, p) in phase space. The dynamics is dened by a curve in phase space and if the Hamiltonian is
time-independent then curves in phase space cannot cross (why?). Figure 7.1 illustrates the dynamics
in phase space for a simple pendulum. Unfortunately since phase space has 2s dimensions its dicult
to draw pictures for s > 1.
Suppose that instead of a single initial condition we consider a set of initial conditions that occupy
a volume in phase space, say all the points between (q, p) and q + q, p + p). The volume of this
region of phase space is,
p+p q+q
V =
dp dq
p

We may also write this using the density of states, (p, q),

V =
(p, q) dp dq

Note: for multi-dimensional systems these integrals are over all phase space. In our example = 1
between (q, p) and q+q, p+p) and zero elsewhere. Landau derives an important result: For canonical
transformations p, q P, Q the Jacobian is one. That is,


p p
P Q
dp dq = q q dP dQ = dP dQ
P Q

CHAPTER 7. CANONICAL EQUATIONS

125

Figure 7.1: Trajectories in phase space for a simple pendulum. Curves near point A are small angle
oscillations; curves near B are large angle oscillations. Curves near point C are trajectories for which
the pendulum spins round and round.

CHAPTER 7. CANONICAL EQUATIONS

126

which may be derived using Poisson brackets. From this result we have, in general,


V =
(p, q) dp dq =
(P, Q) dP dQ

Since we may always nd a canonical transformation (using the Hamilton-Jacobi equation) that makes
P (t) = constant, Q(t) = constant this means that V is also constant. In words, during the course of
the dynamics of a system the volume of states is conserved in phase space.
The conservation of phase space volume leads to a paradox in statistical mechanics. By the theory of
ensembles, specically in the micro-canonical ensemble, a closed system at thermodynamic equilibrium
has equal probability of being in any allowed state of phase space. The probability of being in a state
between (p, q) and q + q, p + p) is
p+p q+q
P(p, q) =

dp dq

dp dq

The paradox is that if we initialize a system in a given region of phase space (i.e., not in thermodynamic
equilibrium) then in time the systems dynamics has to evolve (p, q) such that at thermodynamic
equilibrium (p, q) = constant for all points in phase space. But how can an initial state, with volume
V0 , evolve such that it expands to ll all of phase space while not changing volume?!
The resolution of the paradox is closely related to Georg Cantors counterintuitive proof that a line
segment has the same number of points (equal cardinality) as the unit square. To illustrate how this
can be possible Peano introduced the notion of space-lling curves, such as the Hilbert curve shown
in Figure 7.2. The Hilbert curve passes arbitrarily close to every point on the unit square and similar
constructions may be formulated in higher dimensions. Similarly, a space-lling volume Vs may be
constructed such that every point in a larger volume V0 is arbitrarily close to a point in Vs .

CHAPTER 7. CANONICAL EQUATIONS

127

Figure 7.2: Six iterations of the Hilbert curve construction, which becomes a space-lling curve in the
limit of innite iterations.

Anda mungkin juga menyukai