Vous êtes sur la page 1sur 63

The Principles of Dynamics

Dr C. T. Whelan.1

Lent 2000.

1 A
LT
EXed by Tom Bentley
ii

Technicalities

Copyright & Disclaimer


2000,
c T. J. Bentley, Pembroke Collge, Cambridge.

Permission is granted to copy and distribute these notes so long as the following con-
ditions are met:
• This page of technicalities is distributed with the notes, in whatever form that
distribution may take.
• The notes remain free. Nothing more than the cost of printing should be charged
for their use or distribution.
• That neither T. J. Bentley, nor C. T. Whelan be held responsible for the conse-
quences of any errors or omissions contained herein.

In particular, it should be noted that these notes are a work in progress. It is antici-
pated that a corrected and expanded set wll be made available in the near future.
Thanks

I would like to thank Dr. Colm Whelan for allowing me to produce these notes, and
for his invaluable help in finding the many errors. Any remaining errors are mine and
corrections should be sent to tjb37@cam.ac.uk.

Books

By far the best book for this course is Goldstein’s Classical Mechanics: It covers all the
major parts of the course in a helpful and clear style. Landau and Lifshitz’s Mechanics
is also quite good, especially for the section on Adiabatic Invariants.

Some Other Points to Note

It is a great shame that this course is so very avoidable in the Mathematical Tripos:
The material it teaches you lies behind vast swathes of theoretical physics and applied
mathematics, not least among which are:
• Quantum Mechanics,
• General Relativity.

However the theory has applications in differential equations, financial modelling, pop-
ulation modelling, and many other areas. In short, if you intend to study these things
later, or if you simply want to understand mechanics at it’s most natural level, this is a
very good course to study.
Contents

0 Introduction: Sin and Death 1

1 From Newton to Lagrange 5


1.1 Summary of Newton’s Laws . . . . . . . . . . . . . . . . . . . . . . 5
1.1.1 Momentum, Work and Forces . . . . . . . . . . . . . . . . . 6
1.1.2 Systems of Particles . . . . . . . . . . . . . . . . . . . . . . 7
1.1.3 Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.1.4 Constraints . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.2 D’Alembert’s Principle and Lagrange’s Equations . . . . . . . . . . . 10
1.2.1 Two preliminary lemmas . . . . . . . . . . . . . . . . . . . . 10
1.2.2 D’Alembert’s Principle . . . . . . . . . . . . . . . . . . . . . 11
1.2.3 Lagrange’s Equations . . . . . . . . . . . . . . . . . . . . . . 12
1.3 Generalisations of Lagrange’s Equations . . . . . . . . . . . . . . . . 14
1.3.1 Velocity Dependent Potentials . . . . . . . . . . . . . . . . . 14
1.3.2 Constraints . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
1.3.3 Lagrange’s Equations for Impulsive Forces . . . . . . . . . . 19
1.3.4 Some Definitions . . . . . . . . . . . . . . . . . . . . . . . . 20
1.4 Hamilton’s Principle . . . . . . . . . . . . . . . . . . . . . . . . . . 20
1.4.1 Hamilton’s Principle . . . . . . . . . . . . . . . . . . . . . . 20
1.4.2 Conservation Laws and Symmetries . . . . . . . . . . . . . . 21

2 Rigid Bodies 25
2.1 Frame of Reference . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.1.1 Rotating Frames . . . . . . . . . . . . . . . . . . . . . . . . 25
2.1.2 Transforming Between Frames . . . . . . . . . . . . . . . . . 26
2.1.3 Euler’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . 27
2.2 The Moment of Inertia . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.2.1 Properties of the Moment of Inertia Tensor . . . . . . . . . . 32
2.3 Spinning Tops . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
2.3.1 Deriving the Lagrangian . . . . . . . . . . . . . . . . . . . . 33
2.3.2 Conserved Quantities . . . . . . . . . . . . . . . . . . . . . . 35
2.3.3 Steady Motion . . . . . . . . . . . . . . . . . . . . . . . . . 36
2.3.4 Stability Investigation . . . . . . . . . . . . . . . . . . . . . 36

3 Hamilton’s Equations, & Onwards to Abstraction 39


3.1 Hamilton’s Equations . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.1.1 An alternative approach . . . . . . . . . . . . . . . . . . . . 39
3.1.2 Cyclic Coordinates and Conservation Theorems . . . . . . . . 41

iii
iv CONTENTS

3.1.3 The principle of Least Action . . . . . . . . . . . . . . . . . 41


3.2 Canonical Transformations . . . . . . . . . . . . . . . . . . . . . . . 44
3.2.1 Canonical Transformations . . . . . . . . . . . . . . . . . . . 44
3.2.2 Generalisation to Higher Dimensions . . . . . . . . . . . . . 46
3.2.3 Poisson Brackets . . . . . . . . . . . . . . . . . . . . . . . . 47
3.3 The Sympletic Condition and CTs . . . . . . . . . . . . . . . . . . . 49
3.3.1 The Special Case . . . . . . . . . . . . . . . . . . . . . . . . 49
3.3.2 The General Case . . . . . . . . . . . . . . . . . . . . . . . . 52
3.4 More on ICTs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
3.4.1 The Hamiltonian as the generator of an ICT . . . . . . . . . . 53
3.4.2 Symmetry and Conserved Quantities . . . . . . . . . . . . . . 54
3.5 The Hamilton-Jacobi Equation . . . . . . . . . . . . . . . . . . . . . 55
3.5.1 ‘Nice’ coordinates . . . . . . . . . . . . . . . . . . . . . . . 55
3.5.2 The principle Function and the Lagrangian . . . . . . . . . . 58

4 Integrable Systems 59
4.1 Integrable Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
4.2 Action-Angle Variables . . . . . . . . . . . . . . . . . . . . . . . . . 59
4.3 Adiabatic Invariants . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
Chapter 0

Introduction: Sin and Death

Synopsis: A short historical introduction to serve as motivation for the course, and to
put the material in context. The underlying Principle of Mechanics as being a Varia-
tional Principle.

Newton’s Laws During the 17th century thinking man was obssed with Sin and
Death. How, if God was perfect could He create a Universe which was not perfect?
Why should there be so much suffering and apparent waste in the world? The answer
was that it was all in some way necessary to lead to the final point: The salvation of
Mankind. That path was determined by the end point.
Newton had given the world a mechanicsal universe, working according to a set of
simple fixed laws and the whole majestic clockwork had no need for a Divine Hand to
drive it.

Some of the greatest minds of the time were seduced into trying to find the under-
lying metaphysical reason for Newton’s Laws: Trying, if not to find the hand of the
creator then at least to find his finger prints on the Cosmos.
Leibnitz, in particlular, was determined to prove that all was for the best in the best
possible world. He felt that the world we live in exhibits:
‘The greatest simplicity in its premises and the greatest wealth in it phe-
nomena.’
Leibitz had 3 major problems with Newtonian Mechanics
1. Occult Virtues: Leibnitz held that Newton had not explained ‘Gravity’ by postu-
lating a ‘Gravitational Force’ - Forces are define in terms of directly measurable
quantities (masses and velocities and their rates of change), ie as a property of
their motion. Leibtitz felt that the underlying mechanism had not been found: He
argued that Newtonian theory was a kinematical one, that is a science of motion.
What he sought was a science of powers. Ie. Dynamics
Leibnitz recognised that energy was conserved in certain mechanical systems
and suggested that a principle of energy conservation might be the underlying
one, from which all Laws of Motion could be derived. He deduced something
like a potential energy function.
2. Action at a Distance: To get round this he postulated an ether of very fine par-
ticles. Much of his ideas on this subject anticipated what we would call a field

1
2 CHAPTER 0. INTRODUCTION: SIN AND DEATH

theory..

3. Absolute Space: How could the stars be treated as an absolute frame of refer-
ence? Leibitz argued that space was not a thing in itself, just a relation between
objects in it. He claimed that all inertial frames should be as good as the next.

Action In this intellectual climate Maupertuis advanced an argument based on God’s


efficiency. He claimed that the Laws of Nature were acted out in a way where the
least possible action was expended. He was unclear as to what ‘action’ was, but is had
something to do with mvs.

The Variational Principle Euler liked the idea and defined the action of a particle
moving from A to B as
Z B
mv ds.
A

He postulated that for any given particle, the path taken was ‘chosen’ so that the action
would be least. Actually, he always assumed the existance of a potential energy func-
tion V (r) from which all forces were derived. Ie in our terms we are dealing with a
conservative force.

To progress further he invented the Calculus of Variations1 , that is, a necessary


condition to extremise of the integral
Z B
dy
F (y, , t) dt
A dt

is that
 
∂F d ∂F
− = 0.
∂y dt ∂ ẏ

We can generalise this to a set of N independent coordinates yn :


Z B
δ F (yn , ẏn , t) dt = 0
A
 
∂F d ∂F
⇒ − = 0, 1 ≤ n ≤ N.
∂yn dt ∂ ẏn

This is called the Variational Principle.


Having shown this Euler was able to show that if we had a conservative system (i.e.
there exists a potential energy function V ) then the path of a particle as deduced from
the variational principle was precisely the same as Newton’s Laws:

Consider
Z B Z t(b) Z t(b)
mv ds = mv 2 dt = 2T dt
A t(a) t(a)
1 See Methods IB if you need to review this - it forms a crucial rôle in this course.
3

m

where T = kinetic energy = 2 ẋ2 + ẏ 2 . If there exists a potential
energy function V then

T + V = constant = E ⇒ L ≡ T − V = 2T − E.

We want to make the integral


Z t(b)
L(x, ẋ, y, ẏ) dt
t(a)

stationary, but the Euler Lagrange equations imply that


 
d ∂L ∂L
− =0
dt ∂ ẋ ∂x
d ∂V
⇒ (mẋ) = − ,
dt ∂x
and similarly for the ys. But these are Newton’s Laws
d
F = −∇V = (mẋ)
dt

So the Variational Principle and energy conservation imply Newton’s Law. Quite clever
maybe, but it does it give us anything new?
Yes: The y’s in the E-L equations are implicitly dependent of any particluar coordinate
system. We used Cartesian coordinates, but there was no reason to do this.

Generalized Coordinates Let us introduce generalized coordinates

{q1 , ..., q3N }

If we have a system of N particles (in 3 dimensions) free from constraints, it has 3N


degrees of freedom, and we can choose to describe the motion in terms of any 3N
independent variables {qi }3N
i=1 . Usually these generalized coordinates will not form a
convenient set of N vectors in R3 .
Example: Planetry Motion. We have a radial force
µm
r2
⇒ the potential function V (r) is radial. If we choose our coordinates to be (r, θ) we
can define
1  
L ≡ T − V = m ṙ2 + r2 θ̇2 − V (r)
2
R
Now if δ L dt = 0 then the E-L equations for θ imply:

∂L d ∂L
− =0
∂θ dt ∂ θ̇
d  2 
⇒ mr θ̇ = 0
dt
⇒ mr2 θ̇ = constant = l.
4 CHAPTER 0. INTRODUCTION: SIN AND DEATH

We’ve derived the conservation of angular momentum, you’ll recall from IA Dynamics
that if m is constant this implies Kepler’s Second Law. Applying the E-L equations for
r implies:

∂L d ∂L
− =0
∂r dt ∂ ṙ
d ∂V
⇒ (mṙ) − mrθ̇2 + =0
dt ∂r

But now the conservation of angular momentum can be used to replace θ̇ to get

l2 ∂V
mr̈ − =−
mr3 ∂r
i.e.
1 l2
 
d
mr̈ = − V +
dr 2 mr2
1 l2
   
d 1 2 d dr
⇒ mr̈ṙ ≡ mṙ = − V +
dt 2 dr 2 mr2 dt
l2
 
d 1 2
⇒ mṙ + + V (r) = 0
dt 2 2mr2

This is the conservation of energy.

The Differences between Analytic and Vectorial Mechanics


• In analytic dynamics the equations of motion can be deduced from a single uni-
fying principle. In vectorial mechanics we have Newton’s Laws

• In vectorial mechanics we look at the motion of the individual particles that make
up the system. In analytic dynamics we treat the system as a whole
• It frequently happens that there are constraints on the system (eg. in a rigid body
we have the constraint that the distances between particles remains fixed) In the
Newtonian point of view the must ascribe forces to these constraints. In analytic
mechanics we don’t care about these forces, it is enough to know the constraints.
Chapter 1

From Newton to Lagrange

Synopsis: A brief recap of dynamics, followed by the development of the Lagrangian


formalism. Lagrange’s Equations must take various forms depending on the nature of
the forces (conservative, derivable from a velocity dependent potential, or even more
general), and Constraints (holonomic, monogenic, etc.). Hamilton’s Principle. Con-
served quantities.

1.1 Summary of Newton’s Laws


If r is the radius vector of a particle wrt some origin, then the velocity, v is

dr
v=
dt
The linear momentum P is P = mv. Newton’s first two laws imply

d
F= (mv) (1.1)
dt
A reference frame in which (1) holds is called inertial or Gallelian.
Newton’s Third Law:

‘To every action there is an equal and opposite reaction.’

What does this mean? Is it always true?


Suppose we have two particles i and j and suppose i exerts a force Fij on j. Then we
can translate NIII to read Fij = Fji Ie the force j exerts of i is equal and opposite.
This is the ‘weak’ formulation of the law.
If Fij = −Fji and the forces act along the line connecting the particles, we have a
central force. This is the strong form of the law, and it holds for many forces in Nature
eg. Gravity, Electrostatics.
Consider the example of the Biot-Savart Law between moving charges:

1. If we have two charges moving with parallel velocity vectors that are not perpen-
dicular to the line joining the two particles
Then the weak form holds, but not the strong form.

5
6 CHAPTER 1. FROM NEWTON TO LAGRANGE

2. Consider two charges moving instantaneously such that their velocity vectors are
perpendicular
The 2nd charge exerts a non-zero force on the first while experiencing no ‘reac-
tion’ force at all.

1.1.1 Momentum, Work and Forces


If the force acting on the particle is zero then
d
(mv) = 0 ⇒ mv = const = P
dt
Define the angular momentum of a particle about O to be

L=r∧P

Define the moment of the force, or the torque, to be


d
N=r∧F = r∧ (mv)
dt
d
= (r∧mv)
dt
d
= L
dt
So if the total torque is zero the angular momentum is constant/conserved.
The work done by an external force upon a particle in going from A to B is
Z B
WAB = F· ds,
A

but v = ṡ so
Z tB
dv
WAB = m · v dt
tA dt
m tB d 2
Z
= v dt
2 tA dt
m 2 2

= vA − vB
2
⇒ work done = change in kinetic energy.
Definition: If the force field is st. the work done is the same for any path then we
have a conservative system.
This is true iff
I
F· ds = 0

Which implies there exists a function V (r) st. F = −∇V .


Z B
F · ds = −VB + VA
A

which implies WAB = VA − VB = TB − TA ⇒ TA + VA = TB + VB ie. energy is


conserved.
1.1. SUMMARY OF NEWTON’S LAWS 7

1.1.2 Systems of Particles


Suppose we have a system of N particles. We distinguish between the external applied
force and the internal forces between particles.
Newton’s first two laws become
d X
Pi = Fji + Fext
i (1.2)
dt j

(note Fii = 0.) P


Applying NIII in the weak form means that the Fij term cancels.
Define the centre of mass by
P P
mi ri mi ri
R = Pi = i
i mi M
Then
!
d X d X
Pi = mivi
dt i dt i
d2 X 
= mi r i
dt2
d2 R
= M 2
X dt
ext
= Fi
i

⇒P = M Ṙ (1.3)
So the momentum of the system is the same as the momentum of it’s centre of mass.
Note that to get this result we have only required the weak form of NIII.
Now consider the total torque of the system
X X d  
ri ∧ Ṗi = ri ∧ Ṗi
i i
dt
= L̇
X X
= ri ∧ Fext
i + ri ∧ Fji
i ij

But if we now assume the strong form of NIII


ri ∧ Fji + r ∧ Fij = (ri − rj ) ∧ Fji
= 0 (1.4)
So
dL
Next = (1.5)
dt
This is the conservation of angular momentum
Define
r0i = ri − R
⇒ vi = vi0 + V
8 CHAPTER 1. FROM NEWTON TO LAGRANGE

ie. Working in the CoM frame. Now


! !
X X X d X
L= R ∧ mvi + r0i ∧ mi vi0 + mi r0i ∧v+R∧ mi r0i
i i i
dt i

But the
Plast two terms are zero.
Now mi r0i define the radius vector of the centre of mass in a coord system with it’s
origin at the CoM. Ie.
X
mi r0i = 0
i

X X
L = R ∧ mi V + ri ∧ mvi0
i i
X
or R ∧ M V + r0i ∧ mi vi0

The total angular momentum about a point O is the angular momentum of the system
concentrated at the CoM plus the angular momentum about the centre.

1.1.3 Energy
We wish to calculate the work done by all the forces in moving the system from an
initial configuration A to a final one B.

XZ B
WAB = Fi · dsi
i A

XZ B XZ B
ext
= F · ds + Fji · dsi
i A ij A

Recall that Fii = 0 then


XZ B
WAB = mi v̇i · vi dt
i A

XZ B  
1
= d mv 2
i A 2 i
⇒ WAB = TB − TA

Now we want to transform to the CoM frame:

1X
mi V + v0i · V + v0i
 
T =
2 i
1X 1X 2
= mi V 2 + mi v 0 i
2 i 2 i
1 1X 2
⇒T = MV 2 + mi v 0 i
2 2 i
1.1. SUMMARY OF NEWTON’S LAWS 9

If we can assume that the external forces can be derived from a potential energy func-
tion then the first term in ?? can be written
XZ B XZ B
Fext
i · dsi = − (∇i Vi ) · ds
i A i A
B
X
= − Vi


i A

Where we are using V for the potential.


If, now, the internal forces are conservative then the ‘mutual’ forces on the ith and jth
particles can be obtained from a potential function Vij
If the strong form of the action-reaction law holds then Vij can only be a function of
the distance between the ith and jth particles,

Vij = Vij (|ri − rj |)

Fji = −∇i Vij = ∇j Vij = −Fij (1.6)

If the Vij were also functions of the difference of some other variable (e.g. velocity)
then the forces would still be equal and opposite but not necessarily lie along the line
connecting the two particles. If ?? holds then ∇i Vij = (ri − rj ) f where f is some
other function.
Now when all the forces are conservative the second term in ?? becomes a sum over
terms of the form
Z B Z B
− (∇i Vij ) · dsi + (∇j Vij ) · dsj
A A

But, in Cartesian coordinates dsi − dsj = dri − drj ≡ drij , so the term for the i-jth
pair is
Z
(∇ij Vij ) · drij

Then the total work due to the internal forces is


B
Z B
1 X 1 X
− (∇ij Vij ) · drij = − Vij (1.7)
2 2
i,j,i6=j A i,j,i6=j
A

(note the factor of 21 is present because we’re double counting in the indices.)
If both the external and the internal forces can be derived from potentials, and internal
forces are radial, then we can define a total potential energy
X 1X
V= Vi + Vij (1.8)
i
2 ij

such that the total energy is conserved.


The 2nd term in ?? is the internal potential energy of the system. In general Vij
is not constant and can change as the system changes with time. But a special case
10 CHAPTER 1. FROM NEWTON TO LAGRANGE

is... Definition: A rigid body is a system of particles in which the distances |rij | are
constant and cannot vary with time. ie
2 2
|rij | = |ri − rj | = c2ij , const ∀i, j, t

But then
d 2
(rij ) = 0 ⇒ rij · drij = 0
dt
Thus if Newton’s Law holds in the strong form: Fij · drij = 0, and internal forces do
no work.

1.1.4 Constraints
Take as examples:
• Rigid bodies
• Gas molecules in a container
• A particle moving on a solid sphere
Definition:If the condition of constraint is such that it can be written in the form
f (r1 , ..., rn , t) = 0 then we have a holonomic constraint. An example is the rigid body.
Constraints which cannot be written this way are called non-holonomic, eg a gas in a
container. If the constraint contains time explicitly then it is said to be rheonomous, if
it does not it is called scleronomous, eg. bead on rigid wire is subject to the latter type
of constraint, but if the bead is on a moving wire then we have the former type.
Note that this means a holonomic constraint must allow us to eliminate some vari-
ables
P
Very often the constraint can be written i gi (x1 , ..., xn ) dxi = 0 then the con-
straint will be holonomic. If an integrating function exists f (x1 , ..xn ) such that gi =
∂f
∂xi ie the constraint is holonomic only if

∂f gi ∂f ∂gi ∂2f ∂f gj
= gi + f = gj gi + f = (1.9)
∂xj ∂xj ∂xj ∂xi ∂xj ∂xi
Example:The Rolling Disc.

Other examples of non-holonomic constraints are a particle on the sphere, and all
constraints depending on higher derivatives

1.2 D’Alembert’s Principle and Lagrange’s Equations


1.2.1 Two preliminary lemmas
The cancellation of the dots If we have a function x = x(qi , q̇i ) then
X ∂x
ẋ = q̇i
i
∂qi
1.2. D’ALEMBERT’S PRINCIPLE AND LAGRANGE’S EQUATIONS 11

So that
∂ ẋ ∂x
=
∂ q̇i ∂qi

The interchange of the d and the ∂


  X 2
d ∂x ∂ x ∂ X ∂ ẋ ∂ ẍ
= q̇j = q̇j =
dt ∂qi j
∂q i ∂q j ∂ q̇ i j
∂q j ∂ q̇i

So that
 
d ∂x ∂ ẋ
=
dt ∂qi ∂qi

by the cancellation of the dots.

1.2.2 D’Alembert’s Principle


Definition: A virtual displacement of a system refers to a change in the configuration
of the system as a result of an arbitrary infinitesimal displacement δri consistent with
the forces and constraints at time t. The displacement is called virtual so as to dis-
tinguish it from an actual displacement occurring in time during which the forces and
constraints can vary.
Suppose the system is in equilibrium, ie. the total force on each particle is zero, Fi = 0.
Then clearly

Fi · δri = 0

so as not to affect constraints and forces. If we decompose F as Fi = Fext


i + fi then
X X
Fext
i · δri + fiext · δri = 0
i i

We now make the assumption that constraint forces do no work (ie the second term is
zero) under the virtual displacement. Ie we assume we have a rigid body. Then
X
Fext
i · δri = 0 (1.10)
i

This is the Principle of Virtual Work, or what some authors call D’Alemberts Principle,1
ie. The condition for he equilibrium of a system is that the virtual work of the applied
forces is zero..
Consider a system described by n generalised coordinates. Let us assume al constraints
are holonomic. We remark that {qi } may be less in number than the total number 3N
of degrees of freedom of the system (constraints).
Now the work can be done in an infinitesimal displacement will be proportional to the
elements dqi ,
X
dW = Qr dqr ,
r
1 We shall reserve this for a later result
12 CHAPTER 1. FROM NEWTON TO LAGRANGE

Qr is then defined as the generalised force.


Consider now a system of N particles, let Fi be the force on the ith particle, let Pi be
its momentum. From Newton

Fi − Ṗi = 0


X 
Fi − Ṗi · δri = 0
i

where δri is a virtual displacement. But Fi = Fext


i + fi so
X 
Fext
i + fi − Ṗi · δri = 0
i
P
We make the assumption that forces of constraint do no work , ie fi · δri = 0 and
we obtain
X 
Fext
i − Ṗi · δri = 0, (1.11)
i

what we shall call D’Alemberts Principle - this is the dynamic principle of virtual work.

1.2.3 Lagrange’s Equations


continuing from above:

X X
Ṗi · δri = mi r̈i · δri
i i


X ∂ri
= mi r̈i · · δqj
ij
∂qj
X X d   !
∂ri d ∂ri
= mi ṙi − mi ṙi δqi
j i
dt ∂qj dt ∂qj

But as we’ve seen


 
d ∂ri ∂vi dvi dri
= , and =
dt ∂qj ∂qj dq̇j dqj
So
X ∂ri X d  ∂vi

∂vi

mi r̈i = mvi − mi vi
i
∂qj i
dt ∂ q̇j ∂qj
!! !
X d ∂ X1 ∂ X1
= mi vi2 − mi vi2 δqj
j
dt ∂ q̇ j i
2 ∂q j i
2

Now let us make use of the fact that we have holonomic constraints - we can define our
coordinates {qi } such that they form a complete set

ri = ri (q1 , ...qn , t)
1.2. D’ALEMBERT’S PRINCIPLE AND LAGRANGE’S EQUATIONS 13

So
dri X ∂ri dqk ∂ri
vi = = +
dt ∂qk dt ∂t
k
P ∂ri
Hence δri = j ∂qj δqj since δri is indept of time. So we have now that:

X X ∂ri
Fext
i · δri = Fext
i · δqj
i ij
∂qj

We now define the Generalised Forces corresponding to our generalised coords as


X ∂ri
Qj = Fext
i ·
i
∂qj

1
vi2 − KE we an write
P
So using T = 2 i

X  d  ∂T  dT
 
− − Qj δqj = 0
ij
dt ∂ q̇j dqj

which is just D’Alemberts principle again!


Since this s true for any virtual displacement and the qj s are independent (holonomic)

 
d ∂T ∂T
− = Qj (1.12)
dt ∂ q̇j ∂qj

Assume we are dealing with a conservative system ie that Fext


i = −∇i V then

X ∂ri X ∂ri
Qj = Fext
i · =− (∇i V) ·
i
∂qj i
∂q j

∂V
ie Qj = − ∂qj
so substitution into equation ?? ⇒
 
d ∂T ∂
− (T − V) = 0
dt ∂ q̇j ∂qj
∂V
But we know that ∂ q̇j = 0 so define the Lagrangian, L by

L=T −V

then
 
d ∂L ∂L
− =0
dt ∂ q̇i ∂qi

Note 1. • This is set of n second order odes, the ∂’s are there only as part of the
notation

• The solution of LEs will involve finding 2n functions eg. at t = 0 qα (0) =


Aα , q̇α = Bα
14 CHAPTER 1. FROM NEWTON TO LAGRANGE

Example:The Spherical Pendulum a particle of mass m which moves under gravity is


attached to a fixed point by a rod of length a. ⇒ It is a particle constrained to move on
a sphere of radius a. In terms of spherical coords (a, θ, φ) with θ measured upwards
from the downward direction the kinetic and potential energies can be written:
1  
T = ma2 θ̇2 + sin2 θφ̇2
2
V = −mga cos θ

So

L=T −V
∂L ∂L
⇒ = ma2 θ̇, = maa sin2 θφ̇
∂ θ̇ ∂ φ̇
∂L ∂L
and = ma2 sin θ cos θφ̇2 − mga sin θ, =0
∂θ ∂φ
So by LEs

ma2 θ̈ − ma2 sin θ cos θφ̇ + mga sin θ = 0


d  2 
and ma2 sin θφ̇ = 0
dt
⇒ sin2 θφ̇ = const

1.3 Generalisations of Lagrange’s Equations


1.3.1 Velocity Dependent Potentials
Suppose there exists a potential U (qi , q̇i ) such that
 
∂U d ∂U
Qj = − +
∂qj dt ∂ q̇j

then we would be able to define a Lagrangian L = T − U and the form of LEs would
be unaltered. U will be called a generalised or velocity dependent potential
Example:Maxwell’s Equations2
1
∇∧E+ Ḃ = 0
c2
∇·D = 4πρ
1
∇ ∧ H − 2 Ḋ = 0
c
∇·B = 0

The force on a charge q is not simply

F = qE = −∇φ
2 Presented here using the auxiliary fields D and H and in Gaussian units: So for those who attended

Electromagnetism we have B = H − 4πM and E = D − 4πP


1.3. GENERALISATIONS OF LAGRANGE’S EQUATIONS 15

but rather
 
1
F=q E+ v∧B
c
E is not the gradient of a scalar function: ∇ · B = 0 ⇒ ∃A st. B = ∇ ∧ A. A is
called the Magnetic Vector Potential
We can write ?? as
 
1 ∂  ∂A
∇∧E+ (∇ ∧ A) = ∇ ∧ E + =0
c ∂t c ∂t
1 ∂A
If we now set E + c ∂t= −∇φ this becomes
 
1 ∂A 1
F = q −∇φ − + (v ∧ [∇ ∧ A])
c ∂t c
Consider:
   
∂Ay ∂Ax ∂Ax ∂Az
(v (∇ ∧ A))x = vy − − vz −
∂x ∂y ∂x ∂x
∂Ay ∂Az ∂Ax ∂Ax ∂Ax ∂Az
= vy + vz + vx − vy − vz − vx
∂x ∂x ∂x ∂y ∂z ∂x
Now we note that
dAx ∂Ax ∂Ax ∂Ay ∂Az
= + vx + vy + vz
dt ∂t ∂x ∂y ∂z
So that
∂ dAx ∂Ax
v ∧ (∇ ∧ A) = (v · A) − +
∂x dt ∂t

    
∂ 1 1 d ∂
Fx = q − φ− A·v − (A · v)
∂x c c dt ∂vx
 
∂u d ∂u
= − +
∂x dt ∂ ẋ

q
U = qφ − A · v
c
So we define

L−T −U

Suppose now that not all the forces are derivable from a potential. We can retain
Lagrange’s Equations in the form
 
d ∂L ∂L
− = Q̃j
dt ∂ q̇j ∂qj
Where L contains the potential of the conservative and velocity dependent forces and
Q̃j represents those forces which cannot be derived from a potential of either type.
16 CHAPTER 1. FROM NEWTON TO LAGRANGE

Suppose now that we have a frictional force, F3 which is proportional to velocity

Fx = −kx vx

Define
1X 2 2 2

Fi = kx vix + ky viy + kz viz
2
(This is known as Rayleigh’s Function)
∂F
Then Fxi = − ∂v x
, or Fi = −∇Fi , and the work done by the ith particle against
friction is

dWi = −Fi · dri = k · v2i dt

The component of the generalised force is


X ∂ri
Q̃j = Fi ·
i
∂qj

And LEs become


 
d ∂L ∂L ∂F
− + =0
dt ∂dotqj ∂qj ∂qj

1.3.2 Constraints
Now let us look at a system which may be rheonomic (time dependent constraints),
non-conservative and non-holonomic.
Consider a system of PN particles with masses mi and positions ri , and accelerations
N
ai . Let Fi = mi ai or 1 (mi ai − Fi ) = 0. for any virtual displacement we have
N
X
(mi ai − Fi ) · δri = 0
1

Define
N
X
δW = F · δri
1

Let us consider a maximal set {qα }α=1 , where all holonomic constraints have been
absorbed.
Suppose the more general (non-holonomic) constraints can be written
n
X
Aβα (qi , t)q̇α + Aβ (qi , t) = 0 (1.13)
i

where β = 1, .., m. Ie. there are n < N constraints.4


We can write ?? as
n
X
Aαβ (q, t) dqα + Aβ (q, t) dt = 0
α=1
3 Notation: For this section F is the frictional force, i is the particle number, and x is the direction
4 ?? Notation here ??
1.3. GENERALISATIONS OF LAGRANGE’S EQUATIONS 17

1
PN
But the kinetic energy of the system is T = 2 i=1 mi ṙi · ṙi , so
n
X ∂ri ∂ri
ṙi = q̇α +
α=1
∂q α ∂t

Define Sα by
 
d ∂T ∂T
Sα = −
dt ∂ q̇α ∂qα
(Remember cancellation of the dots and the interchange of the d and the ∂).

N
X ∂ri
Sα = mi a i ·
i=1
∂qα

Now let δqα satisfy


n
X
Aβα δqα = 0, , β = 1, ..., m
α=1

Remark 1. We are dealing with virtual displacements, ie. terms in dt are lost.
δri are virtual displacements satisfying the constraints
n
X
Sα δqα = δW
α=1

We may write
X
δW = Qα δqα
α
n
X
⇒ (Sα − Qα ) δqα = 0
α=1
6⇒ Sα = Qα

because the δqα are not independent, but are subject to


n
X
Aβα δqα = 0, β = 1, ..., m.
α=1

Now write Sα − Qα = Bα and define

F = (B1 − λ1 A11 − λ2 A21 − ... − λm Am1 ) δq1


+ (B2 − λ1 A12 − λ2 A22 − ... − λm Am2 ) δq2
..
.
+ (Bn − λ1 A1n − λ2 A2n − ... − λm Amn ) δqn

(The λ are Lagrange multipliers, and as such are arbitrary.)


18 CHAPTER 1. FROM NEWTON TO LAGRANGE

Pn Pn
Note 2. F = 0 ∀δqα satisfying α=1 Bα δqα = 0, α=1 Aβα δqα = 0
We have m arbitrary λs, we chose these so that
B1 = λ1 A11 + λ2 A21 + ... + λm Am1
..
.
Bm = λ1 A1m + λ2 A2m + ... + λm Anm
Then
F = (Bm+1 − λ1 A1,m+1 − λ2 A2m+1 − ... − λm Am,m+1 ) δqm+1
..
.
+ (Bn − λ1 A1n − λ2 A2n − ... − λm Amn ) δqn
⇒ F = 0 since each column is zero by the constraints. Hence the equations of motion
are:
n   m
X d ∂T ∂T X
Aβα q̇α + Aβ = 0, − = Qα + λβ Aβα
α=1
dt ∂ q̇α ∂qα
β

This will work for holonomic constraints and for many non-holonomic constraints.
Example: The Rolling Loop
Consider a loop rolling without slipping down an inclined pane. This is actually a
holonomic constraint, but it will still serve to illustrate the principle.
The constraint is
r dθ = dx,
Also it is clear that
V = mg(l − x) sin φ (1.14)
1 1
T = mẋ2 + mr2 θ̇2 (1.15)
2 2
1 1
⇒ L = mẋ2 + mr2 θ̇2 − mg (() l − x) sin φ (1.16)
2 2
One constraint ⇒ one Lagrange Multiplier.
The constraint is of the form
Xn
A1α q̇α = 0
α=1

with A1θ = r, A1x = −1 So Lagrange’s Equations ⇒


mẍ − mg sin φ + λ = 0 (1.17)
mr2 θ̈ − λr = 0 (1.18)
rθ̇ = ẋ (1.19)
where the last equation is the constraint. We have 3 equations for 3 unknowns, θ, x, λ⇒

d
(constraint) (1.20)
dt
⇒ rθ̈ = ẍ ⇒ mẍ = λ (1.21)
g sin φ mg sin φ g sin φ
⇒ ẍ = and λ = and θ̈ = (1.22)
2 2 2r
1.3. GENERALISATIONS OF LAGRANGE’S EQUATIONS 19

⇒ hoop rolls down the plane with half the acceleration it would have if it slipped down
a frictionless plane. The force constraint is λ = mg 2sin φ , and notice that the force of
constraint appears via the Lagrange Multiplier. Example:Atwood’s Machine We have
two masses and a frictionless massless pulley. There is only one independent coordi-
nate, x, the position of the second weight determined by the constraint that the string
has length l.

V = −m1 gx − m2 g(l − x)
1
T = (m1 + m2 ) ẋ2
2
So
1
L=T −V = (m1 + m2 ) ẋ2 + m1 gx + m2 g(l − x)
2
∂L
⇒ = (m1 − m2 ) g
∂x
∂L
and = (m1 + m2 ) ẋ
∂ ẋ
LEs ⇒ (m1 + m2 ) ẍ = (m1 − m2 ) g
 
m1 − m2
or ẍ = g
m1 + m2
Notice
• The force of constraint (tension in the string) appears nowhere. We don’t need
to say anything about it to find the equations of motion
• We can’t deduce anything about it. ie. we cannot determine the tension.
Example:Motion on the sphere
A particle of mass m moving under gravity on a smooth sphere of radius b.
The constraint is x2 + y 2 + z 2 = b2 (this is actually holonomic) ⇒ xẋ + y ẏ + z ż = 0.
We also have that
m 2
ẋ + ẏ 2 + ż 2

T =
2
Define the generalised forces X̃ = 0, Ỹ = 0, Z̃ = −mg, then the equations of motion
are
xẋ + y ẏ + z ż = 0, (1.23)
mẍ = λx, (1.24)
mÿ = λy, (1.25)
mz̈ = −mg + λz (1.26)

1.3.3 Lagrange’s Equations for Impulsive Forces


Consider any dynamical system which moves according to Lagrange’s Equations:
 
d ∂T ∂T
− = Qα
dt ∂ q̇α ∂qα
20 CHAPTER 1. FROM NEWTON TO LAGRANGE

Now integrating w.r.t. t we have:


 t2 Z t2 Z t2
∂T ∂T
− dt = Qα dt
∂ q̇α t1 t1 ∂qα t1

R t2
Think Dirac delta-function Let us assume that as t1 → t2 , Qα → ∞ such that Q̂α = limt1 →t2 t1
Qα dt remains
finite. Then we call the Q̂α the generalised impulsive forces.
In the infinitesimal interval |t1 − t2 | we assume the generalised coordinate don’t change
and the generalised velocities remain finite.Then we write
 
∂T
∆ = Q̂α
∂ q̇α

1.3.4 Some Definitions


Definition:For a holonomic system
 
d ∂T ∂T
= + Q̃i
dt ∂ q̇i ∂qi
∂T
The ∂qi are sometimes called fictitious forces due to our change of coordinates.

Note 3. These are different from the fictitious forces introduced to make a non-inertial
frame appear inertial

Definition: The instantaneous configuration of a system can be described by the n


generalised coordinates q1 , .., qn . This corresponds to a particular point in the Cartesian
hyperspace where the q’s form the coordinate axes. This n-dimensional space is called
configuration space or coordinate space.

Note 4. As time goes on the system point moves in configuration space tracing out the
path of the system. Configuration space is not necessarily the same as physical space

Definition: We say a system is monogenic if all the forces (except the forces of
constraint) are derivable from a generalised potential which may be a function of the
(generalised) coordinates, the (generalised) velocities and the time. For such a system
we have...

1.4 Hamilton’s Principle


1.4.1 Hamilton’s Principle
The motion of the system from time t1 to t2 is such that the integral
Z t2
I= L(qi , q̇i , t) dt,
This is where we started t1
the course: The Variational
Principle. Only now we’re has a stationary value. We write this as δI = 0.
in coordinate space, and
we’re seeking only a sta-
tionary value, not necessar-
ily a minimum.
1.4. HAMILTON’S PRINCIPLE 21

Remarks 1. 1. We’ve seen that if L = T − U and we have holonomic constraints


then
 
d ∂L ∂L
− =0
dt ∂ q̇i ∂qi
⇒ Newton’s Laws, and also Maxwell’s Equations.
2. Recall that
Z  
d ∂f ∂f
δ f (yi , yi0 , x) dx = 0 ⇐⇒ − =0
dx ∂yi0 ∂yi
ie: The variational principle ⇐⇒ Lagrange’s Equations. (assuming the qi ’s are
independent).
If we have non-holonomic constraints then
X
Aβα dqα + Aβ dt = 0, β = 1, ..., m
α

And that using Lagrange multipliers we can write


 
d ∂L ∂L X
− = λβ Aβα (1.27)
dt ∂ q̇α ∂qα
β
X
Aαβ q̇α + Aβ = 0 (1.28)
α

ie. n + m equations for n + m variables, {qi }ni , {λi }m


1 .

Note 5. From now on we shall assume the holonomicy of the constraints

1.4.2 Conservation Laws and Symmetries


If L is not a function of a given qi then
 
d ∂L ∂L
=0⇒ = A.
dt ∂ q̇i ∂ q̇i
That is, A is a constant of the motion, it is conserved. eg. Suppose L = 12 mṙi2 then
∂L dL ∂L
∂ q̇i = dẋi = mẋi and because ∂xi = 0 we have mẋ = const Ie Newton’s First
Law, that momentum is conserved. We say that momentum is a conjuagte variable to
position.
Definition: It will be convenient to define generalized momenta by
∂L
pi = .
∂ q̇i
Note that pi need not be an ordinary momentum. Suppose that the Lagrangian is inde-
pendent of time
dL ∂L X ∂L X ∂L dq̇i
= + q̇i + (1.29)
dt ∂t i
∂qi i
∂ q̇i dt
X d   ∂L 
= q̇i (1.30)
i
dt ∂ q̇i
!
X ∂L
⇒ q̇i − L = const (1.31)
i
∂ q̇i
22 CHAPTER 1. FROM NEWTON TO LAGRANGE

So we define
X
H= Pi q̇i − L
i

(in Cartesian coordinates H = i mi ṙi2 − T − V = T + V = energy).


P
Consider a generalised coordinate qj for which a change in dqj represents a translation
of the entire system (eg. qj is the CoM). Now vi is claerly idependent of the origin of
∂vi ∂T
coordinates ⇒ ∂q j
= 0 and hence ∂q j
= 0. Suppose also that we have a conservative
system, ie
 
d ∂T ∂V X ∂ri
= ṗj = − = Qj = · Fi
dt ∂ q̇j ∂qj i
∂qj

Now considering the effect of the infinitesimal dqj (translation of the system along
some axis)

ri (qj ) −→ ri (qj + dqj )

∂ri ri (qj + dqj ) − ri (qj )


≡ lim (1.32)
∂qj dqj →0 dqj
dn̂
= . (1.33)
dqj

Where n̂ is a unit vector in the direction of the translation.

X ∂ri X
Q̃j = · Fi = n̂ · Fi = n̂ · F
i
∂qj i

Now suppose that qj does not appear in V (and hence in L). Then
1X
T = mi ṙi2
2 i

and
X ∂ri
pj = mi ṙi · (1.34)
i
∂ q̇j
X ∂ri
= mi ṙi · (1.35)
i
∂ q̇j
X
= mi · ri · n̂ (1.36)
i
!
X
= n̂ · mi vi (1.37)
i

Now since qj is not in L ⇒


X
Q̃j = 0 ⇒ F · n̂ = 0 ⇒ n̂ · mi vi = const
i
1.4. HAMILTON’S PRINCIPLE 23

arned however: Never A variable in generalised coordinates which does not appear in the Lagrangian is said
e an ignorable coor- to by cyclic or ignorable.
e We have seen that if the Lagrangian is independent of translation in a given direction
n̂ there is no force in this direction and the momentum component is conserved.
Suppose qj is a cyclic coordinate, and dqj corresponds to a rotation of the system about
some axis. Now, just as before, we will argue that a rotation of the coordinate system
∂T
cannot affect the magnitude of the velocities ie. ∂q j
= 0. We are assuming that qj is
∂L ∂V
ignorable so ∂qj = 0, hence ∂qj = 0.
 
d ∂T ∂V
= pj = −
dt ∂ q̇j ∂qj

Now the deriivative has a different meaning.


The change dqj must correspond to an infinitesimal rotation keeping the magnitude
of ri fixed ie

|r(qj )| = |ri (qj + dqj )|

So

| dri | = |ri sin θ dqj | (1.38)



∂ri
⇒ = |ri sin θ| (1.39)
∂qj

Let n̂ indicate a unit vector defining the axis about which we rotate.

∂ri
= n̂ ∧ ri
∂qj

(since dri ⊥ ri and n̂). So


X ∂ri
Qj = Fi · (1.40)
i
∂qj
X 
= ˆ ri
Fi · n ∧ (1.41)
i
X
= n̂ · (ri ∧ Fi ) (1.42)
i
!
X
= n̂ · Ni , the torque on the ith particle (1.43)
i

So PQj = 0 ⇒ n̂ · N = 0, where N is the total torque.. But this ⇒ pj = const =


n · i mi ri ∧ vi = n · L
We deduce that qj (rotation about n̂) is ignorable ⇒ zero torque ⇒ angular momentum
is conserved.
Summary 1. We have reviewed IA Dynamics, and seen how Lagrange’s equations
are equivalent to Newton’s Laws. It should be apparent, however, that they offer a
more powerful approach to finding and solving the equations of motion: The equations
themselves are easy to find; Conserved quantities are immediately apparent.
24 CHAPTER 1. FROM NEWTON TO LAGRANGE
Chapter 2

Rigid Bodies

Synopsis: This chapter is an in-depth application of the Lagrange formalism as devel-


oped in the previous chapter. We study rotating frames of reference, Eulerian angles,
the Moment of Inertia Tensor and go on to investigate rigid body rotation, in particular
the motion of a symmetrical spinning ‘top’.

2.1 Frame of Reference


2.1.1 Rotating Frames
Let OXY represent a fixed (inertial) frame. Let Oxy be similar. Then
i · i = j · j = I · I = J · J = 1, and i · j = I · J = 0.
The relationship between the frames is
i = I cos θ + J sin θ, j = −I sin θ + J cos theta,
So that
di
= (−I sin θ + J cos θ) θ̇ = jθ̇ (2.1)
dt
dj
= (−I cos θ − J sin θ) θ̇ = −iθ̇ (2.2)
dt
So some general vector r = XI + Y J = xi + yj ⇒
dr
= ẋi + ẏj + xi̇ + y j̇ (2.3)
dt
= ẋi + ẏj + xθ̇j − y θ̇i (2.4)
Now we define ω = θ̇k, so that
dr ∂r
= +ω
dt ∂t
Where the first term is from ẋi + ẏj and the second from xωj − yωi. Now this is clearly
true of any vector so
 
d ∂
= + ω∧ (2.5)
dt ∂t

25
26 CHAPTER 2. RIGID BODIES

For example F = mv̇, where m is constant so (??) ⇒


 
∂v
F = m +ω∧v (2.6)
∂t
 

= m + ω∧ (ṙ + ω ∧ r) (2.7)
∂t
∂2r ∂r dω
= m 2 + 2ω ∧ + ∧ r + ω ∧ (ω ∧ r) (2.8)
∂t ∂t dt
So
∂2r ∂r dω
m 2
= F − 2ω ∧ − ∧ r − ω ∧ (ω ∧ r)
∂t ∂t dt
And you see we’ve shown that the Coriolis Force is 2mv ∧ ṙ and the Centrifugal Force
is mω ∧ (ω ∧ r).
We interpret this as saying that if we wish to pretend a non-inertial frame is inertial, we
must invent ‘fictitious forces’.

2.1.2 Transforming Between Frames


A rigid body with N particles can have at most 3N degrees of freedom. This number
2
will be reduced by the constraint rij = c2ij , fixed. You need at most 6 coordinates: To
establish the position of one particle in the body we need three coordinates. Call this
particle 1. Now to fix the position of particle 2 we need only two coordinates (it must
lie on a sphere centred on particle 1 and with radius cij ). If we take a third particle, 3,
we can only rotate the axis joining particles 1 and 2, this is the final degree of freedom.
All other particles are uniquely fixed.
The relation between 2 Cartesian frames can be written

(x1 , x2 , x3 ) → (x01 , x02 , x03 ); x 7→ x0 = Ax.

Then, for a transformation preserving length

x · x = Ax · Ax (2.9)
X
⇒ ajk aik = δij (2.10)
k
ie. AAT = AT A = I (2.11)
⇒ det A = ±1, (2.12)

Let us assume that {x0i } = {xi } at time t = 0.

Remark 2. det A = −1 cannot be achieved by any rigid change of coordinate axis.


We will assume det A = +1 always.
We can transform from a given Cartesian frame to another with the same origin by
at most three rotations.
Suppose we start with X, Y, Z. Then rotate through φ counter-clockwise about the Z
axis.
Then rotate c.clockwise through θ about the ξ axis:
2.1. FRAME OF REFERENCE 27

Finally rotate c.clockwise through ψ about ρ0


This defines our new axes x0 , y 0 , z 0 Let us formalise this in matricies
 
cos ψ sin ψ 0
B = − sin ψ cos ψ 0 (2.13)
0 0 1
 
1 0 0
C = 0 cos θ sin θ  (2.14)
0 − sin θ cos θ
 
cos φ sin φ 0
D = − sin φ cos φ 0 (2.15)
0 0 1
Then the entire rotation has the form A = BCD, or
 
cos ψ cos φ − cos θ sin φ sin ψ cos ψ sin φ + cos θ cos φ sin ψ sin ψ sin θ
A = − sin ψ cos φ − cos θ sin φ cos ψ − sin ψ sinφ + cos θ cos φ cos ψ cos ψ cos θ
sin θ sin φ − sin θ cos φ cos θ

2.1.3 Euler’s Theorem


The most general displacement of a rigid body with one fixed point is a rotation about
some axes.
At any instant the orientation of such a body can be specified by an orthogonal trans-
formation A(t), for simplicity we assume A(0) = I. A(t) will be a cts. function of
time. The transformation will be a rotation if
1. The transformation leaves one direction unchanged. (the axis about which it
rotates).
2. The magnitude of vectors are left unchanged
Note: (2) follows from the orthogonality condition.

Proposition 1. The real orthogonal transformation specifying the physical


motion of a system with one fixed point always has eigenvalue +1.
Proof
(A − λI) r = 0 has a solution iff (2.16)
det (A − λI) = 0 (2.17)
⇒ (A − I) AT = I − AT

(2.18)
28 CHAPTER 2. RIGID BODIES








 
















Figure 2.1: Eulerian Angles: First rotate about the z-axis, then about the (new) ξ-axis,
and finally about the z 0 -axis.
2.2. THE MOMENT OF INERTIA 29

So det(A−I)−det AT = det(I −AT ). But (I −A)T = I −AT , hence det(A−I) =


det(I − A) using det AT = det A.
For any n × n matrix we have that det(−B) = (−1)n det(B), so det(A − I) =
det(I −A) = − det(A−I)⇒ det(A−I) = 0. Hence the eigenvalues is λ = +1. .
Now we can transform A st.
 
λ1 0 0
XAX −1 =  0 λ2 0 
0 0 λ3

where the λi are the eigenvalues of A. This ⇒ det A = λ1 λ2 λ3 and λi = 1 for some
i=1, 2, 3. Suppose that λ3 = 1 then λ1 = λ∗2 and |λ1 | = |λ2 | = 1, since A is a rotation.
Then there are three cases:

1. λ1 = λ2 λ3 ⇒ A = I, a ‘rotation by 2π’.

2. λ1 = λ2 = −λ3 ⇒ rotation through π.

3. λ1 = eiφ , λ2 = e−iφ then


 −φ   
e 0 0 cos φ sin φ 0
similar
 0 e−iφ 0 −→ − sin φ cos φ 0
0 0 1 0 0 1

ie. rotation about the z-axis in the new frame. This proves Euler’s Theorem.

Note 6. 1. For
 iφ 
e 0 0
A= 0 e−iφ 0
0 0 1

we have mathrmtr(A) = 12 cos φ, and remember that the trace is the same for
similar matricies.

2. The sense of direction of the rotation is not yet well defined, since if λ is an
eigenvalue, so too is −λ. Ie. if x is an eigenvector the Ax = λx then x is also
an eigenvector with the same eigenvalue.

We assign φ with A and −φ with Ā(= A−1 ), and use a right hand rule.
Similarly we have Chasles Theorem: The most general displacement of a rigid body is
a translation plus a rotation.
This suggests that the 6 coordinates needed could well be the 3 cartesian coordinates
to fix the body in space and then the 3 Eulerian angles.

2.2 The Moment of Inertia


We know that the total kinetic energy of the system can be written

mv 2
= + T 0 (θ, φ, ψ)
2
We assume here that θ̇
Ie. the sum of translational and rotational energies. The total angular momentum about etc. are not independent of
θ, φψ.
30 CHAPTER 2. RIGID BODIES

some point O is
X
L = R ∧ Mv + r0i ∧ P0i
i

(again the ang. mom. of the body concentrated at CoM plus the ang. mom. about the
CoM.) The essence of rigid body motion is that all the particles that make up the body
move and rotate together. When a rigid body moves with one point stationary then the
total ang. mom. about that point is
X
L= mi (ri ∧ vi )
i

with ri and vi given wrt the fixed point.


Since ri is fixed relative to the body the velocity vi wrt. the space arises solely from
the rotation
∂ri
vi = + ω ∧ ri
∂t
And
X
L = mi (ri ∧ (ω ∧ ri ))
i
X
mi ωri2 − ri (ri · ω)

=
i
2 2
  P P P  
Lx Pi (ri − xi ) P− i m2i xi yi2
im − Pi mi xi zi ωx
ie. Ly  =  − Pi mi xi yi im i (r i − y i ) − i m i z i y i   ωy 
2 2
P P
Lz − i mi xi zi − i mi zi yi i mi (ri − zi ) ωz

The ang. mom. vector is related to the ang. mom. by the linear transformation
  
Ixx Ixy Ixz ωx
L = Iyx Iyy Iyz  ωy 
Izx Izy Izz ωx

The diagonal terms are called the moments of inertia while the off-diagonal terms are
called the products of inertia. In the case of a continuous mass distribution we would
replace the sums by integrals in the obvious way
Z
Ixy = − ρ(r)xy dτ
V

Notation: Sometime we will make use of the notation (x1 , x2 , x3 ) = (x, y, z). In this
notation
Z
Iij = ρ(r)(r2 δij − xi xj ) dτ
V

So then

L = Iω

where I is the moment of inertia tensor. A second rank tensor.


2.2. THE MOMENT OF INERTIA 31

Remark 3. Sometimes one makes use of ‘dyads’. A dyad, ab ≡ aT b 6≡ ba is the


outer product of two vectors, a and b: It is a 2nd rank tensor. Don’t confuse it with the
inner product, a · b ≡ abT
In this language
X
I= mi (ri δij − ri rj )
i
X
L=I·ω = mi (ri ω − ri (ri ω))

X1
T = mi vi2 (2.19)
i
2
X1
= mi vi (ω ∧ ri ) (2.20)
2

= · mi (ri ∧ vi ) (2.21)
i
2
= ω·L (2.22)
1
= ω · Iω (2.23)
2
Let n be a unit vector in the direction of rotation ie.
ω2 T 1
ω = ωn ⇒ T = n In = Iω 2
2 2
Then we say that I is the moment of inertia about the axis n of rotation. Now let us
consider the vector ri ∧ n. It’s magnitude will be the perpendicular from the axis of the
rotation. So
2T X mi
I= 2 = (vi · vi ) (2.24)
ω ω2
X mi
= (ω ∧ ri ) · (ω ∧ ri ) (2.25)
ω2
X
= mi (n ∧ ri ) · (n ∧ ri ) (2.26)
I the moment of inertia about an axis is the sum over all the particles in the body, of
the product of the masses times their perpendicular distance from the axis.

Let the vector from the origin, O, to the CoM be R. Let the radius vector from O
and R be ri and ri0 respectively. Then the moment of inertia about an axis a is
X 2
Ia = mi (ri ∧ n) (2.27)
2
X
= mi ((ri0 + R) ∧ n) (2.28)
X 
2
2
X X
= mi (R ∧ n) + mi (ri0 ∧ n) + 2mi (R ∧ n) · (ri0 ∧ (2.29)
n)

mi , the total mass of the system, and mi ri0 = 0, by the defini-


P P
If we write M =
tion of the CoM we have that
2
Ia = Ib + M (R ∧ n)
This is called the parallel axis theorem. It states that the moment of inertia about a
given axis is the same as the MoI about a parallel axis going through the CoM + the
MoI of the CoM wrt. the original axis.
32 CHAPTER 2. RIGID BODIES

Moment of Inertia


4 M a + 12 M l

a2 + b2
2
2
2
2
1
2

10 M a
5Ma
3Ma
2Ma

Ma
2
2
2
1

3
1
3 M
1

through centre, k side of length 2c


through centre, ⊥ symmetry axis
symmetry axis

symmetry axis

symmetry axis
diameter
diameter
Axis

radius a, length l

radius a, length l

radius a, height l
sides 2a, 2b, 2c
Dimension(s)
radius a
radius a

Hollow Cylinder
Hollow Sphere
Solid Cylinder

Solid Cuboid
Solid Sphere

Solid Cone
Body

Table 2.1: Some common moments of inertia

2.2.1 Properties of the Moment of Inertia Tensor


The Moment of Inertia Tensor has the following properties:
1. It is symmetric, Ixy = Iyx .
2. All its values are real ⇒ real eigenvalues.
3. Together these imply that it is self-adjoint.
Lemma: All eigenvalues of I are real and it’s eigenvectors are mutually orthogonal.
We know that I can be put in diagonal form. The axes corresponding to this diagonal
Notation: Ii is an eigen- form are known as the principle axes and the diagonal elements I1 , I2 , I3 (ie. the
value and I is the identity eigenvalues of the tensor) are the principle moments of inertia. They satisfy
matrix
det(I − Ii I) = 0
2.3. SPINNING TOPS 33

mi yi2 + zi2 ≥ 0 ⇒ I1 , I2 , I3 ≥ 0. Now consider an inertial frame
P
Now Ixx =
B whose origin is at a fixed point of a rigid body (on a system of space axes S with the
origin at the centre). For an axis fixed in the body
   
dL ∂L
= +ω∧L=N
dt S ∂t B

ie. Ni = ∂L∂t + sumjk εijk ωj Lk . The angular momentum components are Li = Ii ωi ,


i

the principle moments of inertia are time independent:

dωi X
Ii + εijk ωj ωk Ik = Ni
dt
jk

These are Euler’s Equations. In full

N1 = I1 ω̇1 − ω2 ω3 (I2 − I3 ) (2.30)


N2 = I2 ω̇2 − ω3 ω1 (I3 − I1 ) (2.31)
N3 = I3 ω̇3 − ω1 ω2 (I1 − I2 ) (2.32)

2.3 Spinning Tops


2.3.1 Deriving the Lagrangian

Figure 2.2: Eulerian Angles as applied to the spinning top

Consider a symmetric top with one point fixed. We will used a body fixed set
(x, y, z). One of the principle axes will be the z-axis, as fixed in the body.
Since one point is fixed - the Eulerian angles are all we need to describe the body.
θ gives the inclination of the z-axis about the vertical. φ measures the azimuth at the
top of the vertical and ψ measures the rotation angle of the top about it’s own z-axis.
The general infinitesimal rotation associated with ω can be considered the result of 3
rotations:
34 CHAPTER 2. RIGID BODIES

1. φ̇ = ωφ about Z (space frame)

2. θ̇ = ωθ about ξ 0

3. ψ̇ = ωψ about z (body frame)

Now ω is parallel to the space-fixed axis Z ⇒ to put it in terms of the body frame we
need to apply the orthogonal transformation A = BCD, with

(ωφ )x = φ̇ sin θ sin ψ (2.33)


(ωφ )y = φ̇ sin θ cos ψ (2.34)
(ωφ )z = φ̇ cos θ (2.35)

Now the direction of ω θ coincides with the ξ 0 axis. So the components of ω θ wrt. the
body fixed axes is given by applying B:

(ωθ )x = θ̇ cos ψ (2.36)


(ωθ )y = −θ̇ sin ψ (2.37)
(ωθ )z = 0 (2.38)

No transformation is needed for ωψ since it is already about z.


Adding the components of ω = ωφ + ωψ + ωθ we get
 
θ̇ sin θ sin ψ + θ̇ cos ψ
ω = φ̇ sin θ cos ψ + θ̇ sin ψ 
φ̇ cos θ + ψ̇

Hence the body is symmetric.

1  1
T = I1 ωx2 + ωy2 + I3 ωz2 (2.39)
2 2
1  2  1  2
2 2
= I1 θ̇ + φ̇ sin θ + I3 ψ̇ + φ̇ cos θ (2.40)
2X 2
V = − mi ri · g = −M R · g (2.41)
i
= M gl cos θ (2.42)

where l is the distance from the fixed point to the CoM, and the angles are Eulerian
Angles. Hence

1  2  1  2
L= I1 θ̇ + φ̇2 sin2 θ + I − 3 ψ̇ + φ̇ cos θ + M gl cos θ.
2 2
 
The φ and ψ are cyclic. Hence pψ = ∂∂Lψ̇ = I3 ψ̇ + φ̇ cos θ = const. = I3 ω3 . And
pφ = ∂∂Lφ̇ = I1 sin2 θ + I3 cos2 θ ṗhi + I3 ψ̇ cos θ = const. = I1 b, and we define


I1 a = I3 ω3 .
So the two constraints of the motion pψ and pφ can be expressed in terms of a and b.
2.3. SPINNING TOPS 35

2.3.2 Conserved Quantities


The total energy is given by
1  2  I
3
E =T +V = I1 θ̇ + φ̇2 sin2 θ + ω32 + M gl cos θ
2 2
Now

I3 ψ̇ = I1 a − I3 φ̇ cos θ (2.43)

If we substitute in for pφ we get

I − 1φ̇ sin2 θ + I1 a cos θ = I1 b (2.44)

Then equations (??) and (??) ⇒


b − a cos θ
ψ̇ = (2.45)
sin2 θ
and
I1 a cos θ (b − acosθ)
ψ̇ = − (2.46)
I3 sin2 θ

Now ω3 = II13a is a constant of the motion. It is (sometimes denoted n and) called the
spin.
Define E 0 = E − ωI33 2, another constant. We can write
2
I1 θ̇2 I1 (b − a cos θ)
E0 = + 2 + M gl cos θ
2 | 2 sin θ {z }
Ṽ (θ)

Or
I1 θ̇2
E0 = + Ṽ (θ)
2
This looks like a one dimensional problem, with an effective potential Ṽ (θ). Making
the change of variable u = cos θ we have:
I2 2 I1 2
E 0 (1 − u2 ) = (b − au) + M glu 1 − u2

u̇ +
2 2
2E 0 2M gl
And letting α = I1 and β = I1 then

u̇2 = (1 − u2 )(α − βu) − (b − au)2

Hence
Z u1 ()
du
= p
u1 () (1 − u )(α − βu) − (b − au)2
2

Unfortunately this integral is elliptic, and the solutions for θ, φ, ψ are in terms of
elliptic integrals.
36 CHAPTER 2. RIGID BODIES

If we look at the equations of motion, as derived from Lagrange’s Equations we have


that
 
I1 θ̈ − I1 φ̇2 sin θ cos θ + I3 sin θ φ̇ + φ̇ cos θ − M gl sin θ = 0 (2.47)
d   
I1 φ̇ sin2 θ + I3 cos θ ψ̇ + φ̇ cos θ = 0 (2.48)
dt
d   
I3 ψ̇ + φ̇ cos θ = 0 (2.49)
dt
This last reveals

ψ̇ + φ̇ cos θ = const. = ω3 = ‘spin’

And so we can write

I1 θ̈ − I1 φ̇2 sin2 θ cos θ + I3 ω3 φ̇ sin θ − M gl sin θ = 0 (2.50)


⇒ I1 θ̇2 I1 φ̇2 sin2 θ + 2M gl cos θ = const = F (2.51)

This is the conservation of energy. And finally the middle equation gives

I1 φ̇ sin2 θ + I3 ω3 cos θ = const. = D

The motion due to the change in θ is called nutation, and the motion due to change in
φ is called precession.

2.3.3 Steady Motion


Steady motion has θ = α = const. ⇒ φ̇ = const. = Ω, say. Now provided we have
θ 6= 0 we have:

I1 Ω2 cos α − I3 ω3 Ω + mgl = 0

⇒ a pair of real distinct precessional angular velocities Ω1 and Ω2 , provided ω32 >
4I1 M gl cos α. Ie. for sufficiently fast spin, ω, about the axis of symmetry the top can
perform steady motion, with θ = α, with 2 possible precessional velocities, ω1 and ω2 .

2.3.4 Stability Investigation


Let γ(θ) = −I1 φ̇2 sin θ cos θ + I3 ω3 φ̇ sin θ − M gl sin θ and the stability condition
γ(α) = 0. Now putting θ = α +  where  is small, we have

I1 00 + γ 0 () = 0 ∼ γ(α + ) ≈ γ 0 (α)

If we can now show that γ 0 () > 0 then this equation will reduce to SHM: The steady
point is stable. Now we know that

(D − I3 ω3 cos θ)
φ̇ = (2.52)
I1 sin2 θ
2
⇒ I1 γ(θ) sin3 θ = − cos θ (D − I3 ω3 cos θ)
+I3 ω3 sin2 θ (D − I3 ω3 cos θ) − I1 M gl sin θ (2.53)
2.3. SPINNING TOPS 37

Now differentiating both sides and putting γ(α) = 0, we have

I1 γ 0 (α) sin3 α = sin α (D − I3 ω3 cos α) (2.54)


−2I3 ω3 cos α sin α (D − I3 ω3 cos α) (2.55)
+2I3 ω3 cos α sin α (D − I3 ω3 cos α) (2.56)
+I32 ω32 sin2 α − 4I3 M gl sin2 α cos α (2.57)

So that

D − I3 ω3 cos α = I1 Ω sin2 α (2.58)


M gl
I3 ω3 = I1 Ω cos α + (2.59)

Thus
 2
0 M gl
I1 γ (α) = I12 Ω2 − 2I1 M glcosα + (2.60)

γ 0 (α)
⇒ 00 + =0 (2.61)
I1?????
So that
2
γ 0 (α)
 
M gl 2M gl (1 − cosα)
= Ω− +
I1 I1 Ω I1

Therefore γ 0 (α) > 0 ∀α 6= 0, and so the motion is SHM about the stable point.
38 CHAPTER 2. RIGID BODIES
Chapter 3

Hamilton’s Equations, &


Onwards to Abstraction

Synopsis: The alternative Hamiltonian formalism is developed. Momenta and Coor-


dinates become indistinguishable. Canonical transformations/Poisson brackets allow
us to reformulate the Hamiltonian: By finding a ‘good’ set of coordinates solving the
equations of motion becomes trivial (Hamilton-Jacobi theory). Conserved quantities
and symmetries are related. Liouville’s Theorem provides an important link to Statis-
tical/Continuum mechanics.
Notation 1. Until now we have used Pi , or P for momenta - whether or not they were
generalised momenta. However, central to the Hamiltonian method of doing things
is the concept that coordinates and momenta are viewed equally: Henceforth we shall
write pi for the momenta of a system with coordinates qi . (The transformed coordinates
shall then be written as Pi and Qi .)

3.1 Hamilton’s Equations


3.1.1 An alternative approach
So far we have formulated everything in terms of the n independent variables qi . We
have, in effect, treated q̇i as distinct variables, independent of the qi of which they are
∂L
the time derivative. For example we have used ∂q i
to mean the derivative wrt. qi with
∂L
qj 6= qi and q̇j held constant, and the symbol ∂ q̇i has been taken to mean the derivative
wrt. q̇i with all q̇j and all qj held constant.
We could work in a space define by qi , q̇i , t but it will introduce a greater symmetry if
we work with the conjugate momenta
∂L
pj = .
∂ q̇j
So, in the Lagrangian for-
The generalised momenta pj are said to be conjugate to the qj and the quantities are
mulation when a variable
said to be canonical variables. Definition: The Hamiltonian of the system is defined
is absent from L, its conju-
by
gate variable is conserved.
X
H(qi , pi , t) = q̇i pi − L(qi , q̇i , t)
i

39
40 CHAPTER 3. HAMILTON’S EQUATIONS, & ONWARDS TO ABSTRACTION

Then
∂H ∂H ∂H
dH = dqi + dpi + (3.1)
∂qi ∂pi ∂t
X ∂L ∂L ∂L
= (q̇i dpi + pi dq̇i ) − dqi − dqi − (3.2)
i
∂ q̇ i ∂q i ∂t
 
d d ∂L ∂L
so that dH = i q̇i dpi − ṗi dqi − ∂L
P
But dt pi = dt ∂ q̇i = ∂q i ∂t dt so equating
variables we have
dH ∂H ∂L ∂H
q̇i = , ṗi = − , − =
dpi ∂qi ∂t ∂t
These are the canonical equations of Hamilton. They are a set of coupled partial dif-
ferential equations.
Note 7. Features:
• Hamilton’s Equations are first order. Lagrange’s Equations were second order.
• Hamilton’s Equations are in 2n variables qi and pi , Lagranges equations were
in the n constraints qi . We now need to determine 2n constants.


∂H dH X ∂H dqi ∂H dpi X ∂H ∂H dH ∂H
=0⇒ = + ⇒ − =0
∂t dt i
∂q i dt ∂p i dt i
∂q i ∂p i dq i ∂pi

⇒ H is a constant of the motion.


Example 1. One Dimensional Motion
Consider one dimensional motion and suppose the existence of a potential V (x) st
d
F = − dx V (x). Then
1
L = mv 2 − V (x) (3.3)
2
1
H = mv 2 − mv 2 + V (x) (3.4)
2
= T + V = Total Energy (3.5)
Notation: here q is the Example 2. Electromagnetic Field
charge, not a generalised Consider a small (non-relativistic) particle in an EM field:
coordinate 1 q
L = T − U = mv 2 − qφ + A · v (3.6)
2 c
1X q
= mẋi ẋi + Ai ẋi − qφ, (3.7)
2 i c
in Cartesian coords. The generalised momenta are given by
q
pi = mẋi + Ai
c
Then
X 1  q 2
H = pi − Ai + qφ (3.8)
i
2m c
1  q  2
⇒H = p − A + qφ (3.9)
2m c
3.1. HAMILTON’S EQUATIONS 41

3.1.2 Cyclic Coordinates and Conservation Theorems


Observe
 
d ∂L ∂H
ṗj = =− .
dt ∂ q̇j ∂qj
So if a coordinate is absent from the Hamiltonian the corresponding (conjugate) gen-
eralised momentum is conserved.
If the generalised momentum pj is absent from the Hamiltonian then
∂H
= 0 ⇒ q̇j = 0
∂pj
So qj is conserved.

3.1.3 The principle of Least Action


In coordinate space we have
Z t2
δ L(qi , q̇i , t) dt = 0,
t1

so we should be able to write


Z T2 X !
δ pi q̇i − H(pi , qi , t) dt = 0.
t1 i

But let us stop here and think: This equation is implicitly in phase space. We have to
think about what we mean by an ‘independent variation’: Since we derived Lagrange’s
Equation (in n variables) be assuming δqi = 0, and by slight of hand we now have 2n
variables (qi , pi ), and so by writing the above we are assuming δqi = 0 and δpi = 0.
In the ∆-variation:

• The varied path over which the integral is evaluated may end at different times
from the ‘correct path’
• There may be a variation in the coordinates at the end points.
Consider the family of varied paths defined by
qi (t, α) = qi (t, 0) +αηi (t)
| {z }
‘true path’
42 CHAPTER 3. HAMILTON’S EQUATIONS, & ONWARDS TO ABSTRACTION

So α is an infinitesimal parameter which goes to zero for the ‘exact’ path. The ηi do
not necessarily have to vanish at the end points, t1 and t2 .
All that is required of the ηi is that they are continuously differentiable. We are inter-

ested in finding the ∆-variation on the action integral, ie:


Z t2
∆ L dt
t1

which we define by
Z t2 Z t2 +∆t2 Z t2
∆ L dt = L(α) dt − L(0) dt,
t1 t1 +∆t1 t1

where L(α) means the integral is evaluated along the path α = α, and L(0) means the
integral is evaluated along the path α = 0, it the physical path. The variation is clearly
composed of two parts:
1. The part arising from the change in the limit of integration, which to first order
infinitesimals is
L(t2 )∆t2 − L(t1 )∆t1

2. The part which comes from the change in the integrand along the varied path1
Z t2
δL dt
t1

So
Z t2 Z t2
∆ L dt = L(t2 )∆t2 − L(t1 )∆t1 + δL dt
t1 t1

Looking at the third term


Z t2 XZ t2     t2
∂L d ∂L ∂L
δL dt = − δqi dt + δqi
t1 i t1 ∂qi dt ∂ q̇i ∂ q̇i t1

................. Now
Z t2
δ f (qi , pi , q̇i , ṗi , t) dt = 0
t1
1 Take care: δq1 (t1 ) and δq2 (t2 ) 6= 0 necessarily
3.1. HAMILTON’S EQUATIONS 43

implies (via the Euler-Lagrange Equations) that


 
d ∂f ∂f
− = 0 (3.10)
dt ∂ q̇i ∂qi
 
d ∂f ∂f
− = 0 (3.11)
dt ∂ ṗi ∂pi
P
So if we identify f (qi , pi , q̇i , ṗi , t) = i pi q̇i − H(qi , pi , t) we have

∂H
ṗj = − (3.12)
∂qj
∂H
q̇j = (3.13)
∂pj

Ie. We have recovered Hamilton’s Equations.

Remarks 2. The two Variational principles

1. We have two forms of Hamilton’s principle:


Rt
• In Coordinate Space δ t12 L(qi , q̇i , t) dt = 0, and we require only that
δqi = 0
Rt P
• In phase Space δ t12 pi q̇i − H(qi , pi , t) dt = 0, and we require that
δqi = δpi = 0. Here we treat qi , pi as independent variables.

Both principles give us Hamilton’s Equations


P
2. pi q̇i − H(qi , pi , t) is independent of ṗi and in our derivation we need the δ-
terms at the endpoints to vanish, so that we can dispose of the surface terms
when we integrate by parts. (Because our integrand is independent of ṗi we
don’t actually make use of the condition δpi = 0.
At no stage in the variational derivation do we make use of our original defining
equation

∂L
pi = .
∂ q̇i

That is to say that neither of the coordinates qi , pi is more fundamental.

3. Suppose F (q, p, t) isP


an arbitrary twice diffable function of p, q Then if we add
dF
dt to the integrand pi q̇i −H+ dF
dt then the variational principle is unaltered.

We can apply Lagrange’s Equations


Z t2 X ∂L t2

δL dt = δqi
t1 ∂ q̇i t1

The δqi refers to the variation in qi at the original end points:


Z t2 X t2
∆ L dt = (L∆t + pi δqi )

t1 t1
44 CHAPTER 3. HAMILTON’S EQUATIONS, & ONWARDS TO ABSTRACTION

Now
∆qi (t2 ) = qi (t2 + ∆t2 , α) − q1 (t2 , 0) (3.14)
= qi (t2 + ∆t2 , 0) − qi (t2 , 0) + αηi (t2 + ∆t2 ) (3.15)
So to first order in α and ∆t2 we have
∆qi (t2 ) = q̇i (t2 )∆t2 + δqi (t2 )
Hence
Z t2 X t2
∆ L dt = (L∆t − pi q̇i ∆t + pi ∆qi ) (3.16)

t1 t1
X t2
= pi ∆qi − H∆t (3.17)

t1

We now make the following assumptions:


∂L ∂H
1. The only systems we consider are those such that ∂t = ∂t = 0.
2. The variation is st. H is conserved on the varied path.

3. The varied paths are st. ∆qi = 0 at the end points.


Remark 4. The varied path might even describe the same path in configuration space
as the actual path, the difference is in the speed the system point transverses the curve.
Now given the above qualifications we have
Z t2
∆ L dt = −H(∆t2 − ∆t1 )
t1

But under the same cosiderations


Z t2 XZ t2
L dt = piq̇i dt − H(t2 − t1 )
t1 t1

and
Z t2 X
∆ pi q̇i dt = 0
t1 | {z }
‘action’
P
Remark 5. In older books the quantity pi q̇i is called the action. For us, however
the action is L.

3.2 Canonical Transformations


3.2.1 Canonical Transformations
Canonical Transformations (henceforth CTs) are those that leave the Hamiltonian struc-
ture of the system invariant. Suppose we started with a Hamiltonian H(q, p, t) satisfy-
ing Hamilton’s Equations:
Z
dH ∂H
q̇i = , ṗi = − ⇐⇒ δ (pq̇ − H) dt = 0
dpi ∂qi
3.2. CANONICAL TRANSFORMATIONS 45
exclude the triv-
case of the ‘scale Suppose there exist functions P (p, q, t), Q(p, q, t) and an associated Hamiltonian, K(P, Q, t),
formation’ where not necessarily the transformation of H into the new coordinates, st.
λi Pi , qi = λi Qi for ∂K ∂K
Z
scalars λi . = Q̇, = −Ṗ ⇐⇒ δ (P Q̇ − K) dt = 0
∂P ∂Q
We ask, how are the new coordinates related to the old ones? Consider the function F
defined such that P dQ − K dt + dF = p dq − H dt. Ie. we want
dF = p dq − P dQ + (K − H) dt
Let us take the particular case where F = F1 (q, Q, t) then we have that
∂F1 ∂F1 ∂F1
dF1 = dq + dQ + dt (3.18)
∂q ∂Q ∂t
∂F1 ∂F1 ∂F1
⇒ = p, = −P, =K−H (3.19)
∂q ∂Q ∂t
So
Z t2 Z  
t
δ (pq̇ − H) dt = δ P Q̇ − K dt + δF1 |t21
t1

And so the transformation is Canonical.


F1 acts as a bridge between the old and then new coordinates. Half the variables are
from the old and half from the new. For example suppose F1 = qQ then p = Q,
P = −q and K = H.
Now there is no need for F , the generating function, to be a function of q, Q and t. for
example consider F2 = F2 (q, P, t) and define F by
F = F2 (q, P, t) − QP
Then
dF
pq̇ − H = P Q̇ − K (3.20)
dt
dF2 (q, P t)
= −Ṗ Q − K + (3.21)
dt
But
d ∂F2 ∂F2 ∂F2
F2 (q, P, t) = dq + dP + dt (3.22)
dt ∂q ∂P ∂t
∂F2 ∂F2 ∂F2
⇒ p= , Q= , K=H+ (3.23)
∂q ∂P ∂t
We can also devise generating functions which are mixed in the sense that they depend
only on p, Q, t or p, P, t Define F = qp + F3 (p, Q, t) then similarly from above we
will have
∂F3 ∂F3 ∂F3
q=− , P =− , K=H+
∂p ∂Q ∂t
And, finally, defining F = qp − QP + F4 (p, P, t) we get
∂F4 ∂F4 ∂F4
q=− , Q= , K=H+
∂p ∂P ∂t
These four generating functions are all we need to describe a specific change.
46 CHAPTER 3. HAMILTON’S EQUATIONS, & ONWARDS TO ABSTRACTION

  




 

Figure 3.1: The functional dependence of the generating functions F1 to F4 .

3.2.2 Generalisation to Higher Dimensions


Now we can generalise this to n-dimensions:
∂F ∂F
Take F1 (qi , Qi , t), pi = , Pi = − (3.24)
∂qi ∂Qi
X ∂F2 ∂F2
then F = F2 (qi , Pi , t) − Qi Pi ⇒ pi = , Qi = (3.25)
∂qi ∂Pi
etc. (3.26)

In the general case,the generating function does not have to conform to one of the four
general types for all degrees of freedom.
Example 3. Let us consider the Harmonic Oscillator
p2 kq 2 1 k
= m p2 + m2 ω 2 q 2 , with ω 2 =

H= +
2m 2 2 m
Let us take

p = f (P ) cos Q (3.27)
F (P ) sin Q
q = (3.28)

We require
F 2 (P )  F 2 (P )
H=K= cos2 Q + sin2 Q =
2m 2m
So this choice has made Q cyclic.
We want to find the function f as yet unspecified, such that the transformation is canon-
ical. To do this we observe
p
= mω cot Q (3.29)
q
⇒ p = mωq cot Q (3.30)

Now suppose there exists a function of the type F1 (q, Q) then


∂F1
p= (q, Q) = mωq cot Q
∂q
3.2. CANONICAL TRANSFORMATIONS 47

The simplest solution to this is

mωq 2
F1 = cot Q
2
And then we must have that
r
∂F1 mωq 2 2P
P =− = 2 ⇒q= sin Q
∂Q 2 sin Q mω

but p = F (P ) cos Q ⇒ f (P ) = 2mωP . And we have that

H = K = ωP

K is cyclic in Q ⇒ Ṗ = 0⇒ P = const. This in turn ⇒ K = const = E. And hence


P =E ω.
Then
∂K
Q̇ = = ω ⇒ Q = ωt + α
∂P
Where alpha is a constant determined from the initial conditions. We can now invert
the transformation to obtain
r
2E
q= sin(ωt + α)
mω 2
2
In going from P = 2mωqsin2 Q
to q we have used the positive square root. We could
have used the negative root: The only difference would have been a trivial difference
of π in the phase:Transformations need not be single valued

3.2.3 Poisson Brackets


Let f (p, q, t) be any function of q, p. Then

df ∂f dq ∂f dp ∂f
= + +
dt ∂q dt ∂p dt ∂t
∂H
But q̇ = ∂p and ṗ = − ∂H
∂q so we can write:

df ∂f ∂H ∂f ∂H ∂f
= − + (3.31)
dt ∂q ∂p ∂p ∂q ∂t
∂f
= [f, H] + (3.32)
∂t

where [f, H] ≡ ∂f ∂H ∂f ∂H
∂q ∂p − ∂p ∂q . This is a Poisson Bracket.
Definition: If f = f (qi , pi , t) and g = g(qi , pi , t) with 1 ≤ i ≤ n, then we define
the Poisson Bracket by
X ∂f ∂g ∂f ∂g
[f, g]q,p = −
i
∂qi ∂pi ∂pi ∂qi

It has the following properties:


48 CHAPTER 3. HAMILTON’S EQUATIONS, & ONWARDS TO ABSTRACTION

• [f, f ]q,p = 0

• (antisymmetry) [f, g]q,p = − [g, f ]q,p

• (linearity) [af + bg, h]q,p = a [f, h]q,p + b [g, h]q,p .

• [pi , H]q,p = ṗi and [qi , H]q,p = q̇i .

∂H ∂H
Proof. To see this we use q̇i = ∂pi and ṗi = − ∂qi
so that

X ∂pi ∂H ∂pi ∂H
[pi , H]q,p = − = ṗi (3.33)
j
∂qj ∂pj ∂pj ∂qj
X ∂qi ∂H ∂qi ∂H
and [qi , H]q,p = − = q̇i (3.34)
j
∂qj ∂pj ∂pj ∂qj

similarly.

• [qi , qj ]q,p = [pi , pj ]q,p = 0.

Proof. Observe
X ∂qi ∂qj ∂qi ∂qj
[qi , qj ]q,p = − (3.35)
∂qk ∂pk ∂pk ∂qk
k
X
= δij − δij = 0 (3.36)

• [qi , pj ]q,p = δij

Proof.
X ∂qi ∂pj ∂qi ∂pj
[qi , pj ]q,p = − (3.37)
∂qk ∂pk ∂pk ∂qk
k
X
= δik δjk (3.38)
= δij = − [pj , qi ] (3.39)

∂H
• If ∂t = 0 the Hamiltonian is conserved.

dH
Proof. Since dt = [H, H]+ ∂H
∂t , then if
∂H
∂t =0⇒ dH
dt = 0 and the Hamiltonian
is conserved.

• (The Jacobi Identity): [f, [g, h]] + [g, [h, f ]] + [h, [f, g]] = 0.

Proof. Left to example sheet


Antisymmetry, Linearity and
the Jacobi property define
what is called a Lie Algebra,
wherein the Poisson Bracket
is the ’product’. Other Lie
algebras include the vector
product and the matrix com-
mutator. The QM correspon-
dence principle says that:
1
[f, g] → i~ (f g − gf ) and
this only works because both
3.3. THE SYMPLETIC CONDITION AND CTS 49

3.3 The Sympletic Condition and CTs


3.3.1 The Special Case
We are looking for conditions that make a given transformation canonical. Let us begin
by considering some restricted canonical transformations: Those where time does not
play an explicit part. In terms of the generating function F we have that ∂F
∂t = 0 so
that K = H. Then

Q = Q(q, p), P = P (q, p)

So we have
X ∂Qi ∂Qi
Q̇i = q̇j + ṗj (3.40)
j
∂qj ∂pj
X ∂Qi ∂zH ∂Qi ∂H
= − (3.41)
j
∂qj ∂pj ∂pj ∂qj

On the other hand we may invert the transformation to get qj = qj (Q, P ) and pj =
pj (Q, P ) as
∂H X ∂H ∂pj ∂H ∂qj
= +
∂Pi j
∂p j ∂P i ∂q j ∂Pi

Then the transformation is canonical only if


∂H
Q̇i = (3.42)
∂Pi
   
∂Qi ∂pi
⇒ = (3.43)
∂qj q,p ∂Pj Q,P
| {z } | {z }
Q as a function of q, p p as a function of Q, P
   
∂Qi ∂qi
and =− (3.44)
∂pj q,p ∂Pj Q,P

In the same way, by considering Ṗi we find that


       
∂Pi ∂pj ∂Pi ∂Qj
=− , =
∂qj q,p ∂Qi Q,P ∂pj q,p ∂Qi Q,P

Let us further restrict our attention to a 2 dimensional phase space (q, p) and consider
the transformation to Q, P , then define
! 
  ∂Q ∂Q 
0 1 ∂q ∂p ∂q Q ∂p Q
J= , M = ∂P ∂P =
−1 0 ∂q ∂p
∂q P ∂p P

Consider
!
∂Q ∂Q 
T ∂q ∂p ∂p Q ∂p P
M JM = ∂P ∂P (3.45)
∂q ∂p
−∂q Q −∂q P
 
0 [Q, P ]q,p
= (3.46)
[P, Q]q,p 0
50 CHAPTER 3. HAMILTON’S EQUATIONS, & ONWARDS TO ABSTRACTION

But consider
∂Q ∂P ∂P ∂Q
[Q, P ]q,p = − ,
∂q ∂p ∂q ∂p
and in particular for our transformation ∂q Q = ∂P p etc. So
[Q, P ]q,p = ∂P p∂Q q − ∂P p∂Q q = [q, p]Q,P
Now if it were that [u, v]q,p = [u, v]Q,P , then it would follow that the condition for the
restricted transformation to be canonical would be that
M JM T = J.
We will now show that this is indeed the condition that the transformation is canonical:
We will show that a CT leaves Poisson brackets invariant. So it is unnecessary to
write [f, g]q,p , since the quantity is the same no matter what canonical coordinates we
choose.
∂ξi
Notation 2. We shall write Mij = ∂ηj where
  

q1 Q1
 ..   .. 
 .   . 
     
 qn  Qn  0 I
η=
 P1  ,
 ξ=
 P1  ,
 J=
    −I 0
 .   . 
 ..   .. 
Pn Pn
Ie. both are column vectors in a 2n-dimensional space, and where I is the n×n identity
matrix. This might seem a bit odd, but fundamentally, the (qi , pi ) are our coordinates:
Why should we treat them differently?
theorem 1. A transformation q, p −→ Q, P is canonical iff
M T JM = M T M T = J.
This is called the Sympletic Condition.
Remark 6. We prove the result first in the case when we have a restricted CT, ie. no
explicit time dependence, and then go on to prove it for the general case
It might be useful, at each
Let us prove the special case:
step, to write out the ma-
tricies and vectors explic-
itly. Proof. First, it follows from Hamilton’s Equations that
∂H
η̇ = J
∂η
The elements of ξ are the Qi and Pi , and these are functions of the qi and pi , ie. of η
so
2n
X ∂ξi
ξ˙i = η̇j (3.47)
j=1
∂η j

⇒ ξ̇ = M η̇ (3.48)
3.3. THE SYMPLETIC CONDITION AND CTS 51

Now, by the inverse transformation H can be considered as a function of ξ and η, so

∂H X ∂H ∂ξj
=
∂ηi j
∂ξj ∂ηi

Or in vector/matrix notation
∂H ∂H
=M
∂η ∂ξ

Now
∂H ∂H
ξ̇ = M η̇ = M J = M JM T
∂η ∂ξ

But
∂H
ξ̇ = J
∂ξ

from Hamilton’s Equations. Hence

M JM T = J

By noting that we could just as well have gone from Q, P to p, q we must also have

M T JM = J.

This proves the special case: That a restricted CT will be canonical if the symplectic
condition holds. The reverse condition holds, and fortunately the proof works in reverse
too.

Remark 7. In our example we saw that


  [Q, P ]q,p = 1 is precisely
T 0 [Q, P ]q,p the same as the statement
M JM =
[Q, P ]q,p 0 that the Jacobian of the CT
is 1, whichR in turn im-
We have just proved, however, that plies that dq dp is an in-
  variant of the transforma-
T 0 1 tion.??? This is one of
M JM = J =
−1 0 the Poincarre integral in-
variants
And hence [Q, P ]q,p = − (P, Q)q,p = 1, but it was a trivial property of the Poisson
Bracket that [Q, P ]Q,P = 1 = − [Q, P ] so we clearly have that

[Q, P ]Q,P = [Q, Q]q,p = [P, P ]q,p = [P, P ]Q,P = 0

This also holds in the general case, so that

[Qi , Pj ]q,p = [Qi , Pj ]Q,P = δij (3.49)


[Qi , Qj ] = [Pi , Pj ] = 0 (3.50)
52 CHAPTER 3. HAMILTON’S EQUATIONS, & ONWARDS TO ABSTRACTION

3.3.2 The General Case


We have yet to show that M JM T = JM T JM is a necessary and sufficient condition
for a CT in the general case of an arbitrary CT. Ie one with time dependence.
IN order to do this it will be necessary to introduce the important idea of an infinitesimal
contact/canonical transformation (or an ICT for short).
P
Lemma 1. The generating function F2 = qi Pi generates the identity transforma-
tion.

Proof.
In fact CTs form a group:
∂F2 ∂F2
The identity is canonical, = pi ⇒ pi = Pi , and = qi ⇒ qi = Qi
the inverse of a CT is a ∂qi ∂Pi
CT, two successive CTs is
a CT, and the product is
associative. We only re-
Now let us consider the generating function
quire that the CTs are an-
alytic functions of cts pa- X
rameters in order to have a F = qi Pi + εG(qi , Pi , t)
Lie Group.
where G is an arbitrary diffable function and ε is infinitesimally small. Then

∂F ∂G
Qi = = qi + ε (3.51)
∂Pi ∂Pi
∂F ∂G
pi = = Pi + ε (3.52)
∂qi ∂qi

Now take
∂G
δpi = Pi − pi = −ε (3.53)
∂qi
∂G
δqi = Qi − qi = ε (3.54)
∂Pi

Now G(qi , Pi (q, p)) = G(qi , pi + εf (q, p)) ⇒ (to first order in ε)

∂G
δqi = ε (3.55)
∂pi
∂G
δpi = −ε (3.56)
∂qi
∂G
⇒ δη = εJ (3.57)
∂η
∂ξ 2

Now M = ∂η = I + ∂η δη = I + εJ ∂∂ηG2 So the second derivative is a square matrix:
 2  2
 2 T 2
∂ G
∂η 2 = ∂η∂i ∂η
G
j
. Now M = I − ε J ∂∂ηG2 , but ∂∂ηG2 is symmetric, and J is
ij
antisymmetric, ⇒

∂2G
MT = I − ε J.
∂η 2
3.4. MORE ON ICTS 53

And hence
∂2G ∂2G
   
M T JM = I + εJJ I − ε J (3.58)
∂η 2 ∂η 2
∂2G ∂2G
= J + εJ 2
J − ε 2 J + O(ε2 ) (3.59)
∂η ∂η
Thus to first order

M T JM = J

⇒ for any ICT the symplectic condition holds.


Now consider the CT ξ = ξ(η, t), this evolves ctsly as time increases from some initial
value. Let G = H(q, p, t). Then
∂H
δqi = dt= q̇i dt = dqi (3.60)
∂pi
∂H
δpi = − dt = ṗi dt = dpi (3.61)
∂qi
Thus the Hamiltonian acts as the generator of an ICT which corresponds to the evolu-
tion in time of the system −→ symplectic condition holds.
The continuous evolution of the transformation ξ(η, t) from ξ(η, t0 ) to ξ(η, t) can be
built up as a succession of ICTs in steps of dT , so if η(t0 ) → ξ(t0 ) is canonical as
ξ(t0 ) → ξ(t) is canonical we must have that η(t0 ) → ξ(t) is canonical.
It can be shown that the product of two successive CTs is a CT (question 1, prob-
lem sheet 4). So the symplectic condition holds in general. In the course of our
argument we have seen that

[Pi , Qj ] = [pi , qj ] = −δij,

where the Poisson bracket is evaluated wrt. the canonical set. It can be shown that if
u, v are arbitrarily diffable functions of q, p then

[u, v]q,p = [u, v]Q,P

Ie. all Poisson brackets are invariant. (question 2, problem sheet 4).

3.4 More on ICTs


3.4.1 The Hamiltonian as the generator of an ICT
As we’ve seen, an ICT is a special case of a transformation that is a cts function of
a parameter. If the parameter is small enough to be treated as a function of a first
order infinitesimal then the transformation between canonical variables differ only in
infinitesimals, ie.

ξ = η + δη

with the change being given in terms of the generator G through the equation
∂G(η)
δη = εJ .
∂η
54 CHAPTER 3. HAMILTON’S EQUATIONS, & ONWARDS TO ABSTRACTION

Now
T
∂ηi ∂u
[η, u] = J (3.62)
∂η ∂η
∂u
= J (3.63)
∂η
⇒ δη = ε [η, G] (3.64)

Now consider an ICT in t whose generator is the Hamiltonian:

δη = dt [η, H] (3.65)
η̇ = [η, H] (3.66)

If the motion of the system in a time interval dt can be described by an ICT generated
by the Hamiltonian −→ The motion of the system from t0 to t can be generated by
a single contact transformation equivalent to an infinite sequence of infinitesimals, all
generated by the Hamiltonian.
We can view the Hamiltonian as the generator of an ICT (and consequently a CT)
which describes the motion of the system with time.
A solution of the problem of finding the canonical transformation which relates coor-
dinates and momenta at time t = 0 to their value at t = t is equivalent to solving the
physical problem. There are two views:
• The passive View We regard the transformation from ξ −→ η as mapping from
one phase space to a new phase space. So

u(p, q) −→ U (P, Q),

ie. the functional form will change, but not the value.
• The Active View We regard the transformation as a mapping within one phase
This is the same active
space - a point transformation. This time the functional form remains the same,
view that is used to define
but the value changes. For example, the evolution of the system in time, as
the Lie Derivative of a ten-
generated by the Hamiltonian.
sor field
If we are working in the active sense then we can talk about the change in the function
u under a CT (cf. the passive view, which has u(p, q) = U (P, Q).)

3.4.2 Symmetry and Conserved Quantities


Suppose we have an infinitesimal transformation generated by G(q, p, t), so that
∂u ∂u ∂G
u(η + δη) − u(η) = δη = ε J
∂η ∂η ∂η
then since η = (q1 , ..., qn , p1 , ..., pn )T the above becomes
X ∂u ∂u
u(η + δη) − u(η) = δqi + δpi (3.67)
∂qi ∂pi

X ∂u ∂G 
∂u ∂u
= ε − (3.68)
∂qi ∂pi ∂pi ∂qi
= ε [u, G] (3.69)
3.5. THE HAMILTON-JACOBI EQUATION 55

So assuming that G is not an explicit function of time, we can ask ‘does K = H?’ Well

δH = ε [H, G]

dG ∂G
But dt = [H, G] + , so
∂t
|{z}
=0

dG Compare this to Heisen-


= 0 ⇐⇒ [H, G] = 0
dt berg’s formulation of QM.

The symmetry properties of the system are equivalent to the conservation laws.
The statement now includes all constants, not just the conjugate momenta to cyclic
variables.
Example 4. Suppose qi is cyclic. Ie. the Hamiltonian is independent of qi , and will be
clearly invariant under an ICT which involves qi alone. The equations of transforma-
tion would be

δqj = εδij (3.70)


δpj = 0 (3.71)

And then G = pi , so
dpj
∆H − ε [H, pj ] = 0 ⇒ =0
dt

3.5 The Hamilton-Jacobi Equation


3.5.1 ‘Nice’ coordinates
Having done all this abstract theory, we can now reap some benefits. We have two
approaches to solving problems:
1. If the Hamiltonian is conserved then we can transform to a new set of canonical
coordinates, all of which are cyclic. Then the integration of the new set becomes
trivial.
2. We can seek a CT from (q, p) at t = t to those at t = 0. Under such a transfor-
mation the equations linking (q, p) with the new (q0 , p0 ) are the solutions to the
problem.
Let us consider the first approach:
We can automatically require that our new variables are constant in time if the trans-
formed Hamiltonian is zero, for then from Hamilton’s Equations
∂H
= 0 ⇒ Q̇i = 0 ⇒ Qi = const
∂Pi
and similarly for the Pi . Now we know that the transformed Hamiltonian K is related
to the old Hamiltonian H by the equation
∂F
K=H+ ,
∂t
56 CHAPTER 3. HAMILTON’S EQUATIONS, & ONWARDS TO ABSTRACTION

where F is our generating function. So if we have the new Hamiltonian being zero, we
must satisfy

∂F
H+ =0
∂t
∂F2
It proves convenient to take F = F2 (q, P, t). For then we have, using pi = ∂qi ,

 
∂F2 ∂F2
H qi , ,t + =0
∂qi ∂t

This is the Hamilton-Jacobi Equation. It is a pde for the desired generating function in
the (n + 1) variables q1 , ..., qn , t.
Suppose there exists a solution

S = S(q1 , ..qn , α1 , ..., αn+1 , t).


| {z }
constants

This constitutes a complete solution of the differential equation, and is called Hamil-
ton’s principle Function. One constant is redundant: S itself only appears in the
Hamilton-Jacobi Equation via it’s derivatives wrt. qi and t, so wlog. we can add a
constant to S and the H-J equation will still hold.
To this end was absorb αn+1 into S and look for a solution

S = S(q1 , ..qn , α1 , ..., αn , t).


n
Once we have the n constants {αi }i=1 the solution will be complete.
We are at liberty to take the n constants to be the new (constant) momenta Pi , that is
we can set

Pi = αi .

Now recall, for F2 -type transformations

∂F2 ∂F2
pi = , Qi = ,
∂qi ∂Pi
but this implies

∂S
pi = (qi , αi , t) ,
∂qi
We can evaluate the constants of motion ito. our initial conditions at t = 0. Ie. we find
∂S
the values of qi simply by calculating ∂αi
at t = 0. And Qi is given by

∂S
Qi = βi = (q, α, t)
∂αi
?????
Example 5. Consider the Harmonic Oscillator Hamiltonian with unit mass
1 2 √
p + ω 2 q 2 = E, with ω = k

H=
2
3.5. THE HAMILTON-JACOBI EQUATION 57

∂S
If we set p = ∂q (thus assuming an F2 -type transformation) we obtain the H-J equation
" 2 #
1 ∂S 2 ∂S
+ ωq + =0
2 ∂q ∂t

We notice that S depends on time only in the last term of the H-J equation, so let’s try
S(q, α, t) = W (q, α) − αt. Then
" 2 #
1 ∂W 2 2
+ω q = |{z}
α
2 ∂q
− ∂S
∂t

And so the H-J equation implies that H = α, and we naturally associate α with the
energy.
Now
r
√ Z ω2 q2
W = 2α dq 1 − (3.72)

r
√ Z ω2 q2
S = 2α dq 1 − − αt (3.73)

But
Z
∂S −1/2 dq
β = = (2α) q − t, (3.74)
∂α 2 2
1 − ω2αq
r

q = sin ω (t + β) (3.75)
ω2
∂S ∂W √
p = = = 2α cos ω (t + β) = q̇ (3.76)
∂q ∂q

The initial conditions at t = 0 are given by (q0 , p0 ). If we square the equations for q
and p we get

2α = p2 + ω 2 q 2 = p20 , ω 2 q02 ,

and the other usual trick is


ωq0
= tan ωβ
p0

And we naturally identify β with the phase angle of the oscillator. Doing a bit more
algebra:
r
√ Z
ω2 q2
S = 2α dq 1 − − αt (3.77)

Z
= 2α cos2 ω (t + β) dt − αt (3.78)
Z
1
= 2α cos2 ω (t + β) − dt (3.79)
2
58 CHAPTER 3. HAMILTON’S EQUATIONS, & ONWARDS TO ABSTRACTION

And the Lagrangian of the problem is


1 2
p − ω2 q2

L = (3.80)
2  
1
= 2α cos2 ω (t + β) − (3.81)
2

I.e, in this case we have


Z
S= L dt

3.5.2 The principle Function and the Lagrangian


The above example furnished us with an interesting relation between the principle func-
tion S and the Lagrangian L. Does this hold in general?
∂S ∂S
We have pi = ∂q i
, and also Qi = βi = ∂αi
(q, α, t). This last equation can be inverted
to write q as

qj = qj (α, β, t) .

Then after differentiation in the first equation we can substitute qj to obtain

pi = pi (α, β, t)

And so
dS X ∂S ∂S X
= q̇i + = pi q̇i − H = L.
dt ∂qi ∂t
We can write this general result as
Z
S= L dt + const.
Chapter 4

Integrable Systems

Synopsis: This chapter applies some of the results of the Hamiltonian formalism, as
developed in Chapter 3, as well as investigating the important area of Adiabatic Invari-
ants: Completely Integrable systems.

4.1 Integrable Systems


What is an integrable system?

4.2 Action-Angle Variables


Let us consider a system with one degree of freedom and assume it is conservative, so

H(q, P ) = α1 .

Suppose we are interested in emphperiodic systems: There are two things we could
mean by this:
1. The path of the system describes a closed loop in phase space. For example
vibration. This will happen when both p and q are periodic function of time, and
have the same frequency.
2. The path of the system is periodic in phase space, for example a pendulum going
all the way round it’s fixed point. Then we will have tha tp(q) = p(q + q0 )
Definition: We define the angle variable J, by
I
J = p dq,

where the contour integral is taken around one peroid of the system, whether it is of
type (1) or (2).

4.3 Adiabatic Invariants

59

Vous aimerez peut-être aussi