LectureNote CM2
LectureNote CM2
Undergraduate School - B2
Year 2015-2016
Institute of Physics
10 Dao Tan, Ba Dinh, Ha Noi
2
Recommended text books and
resources
3
Course Syllabus
• Cource Contents:
Class Contents Number of hours Note
1 Mathematical Backgrounds (part 1) 3 Lecture
2 Mathematical Backgrounds (part 2) 3 Lecture
Reminder of Newtonian Mechanics (part 1)
3 Reminder of Newtonian Mechanics (part 2) 3 Lecture
4 The Lagrangian Formalism 3 Lecture
5 Noether’s Theorem and Symmetries 3 Lecture
6 Applications of the Lagrangian Formalism (part 1) 3 Exercise
7 Applications of the Lagrangian Formalism (part 2) 3 Exercise
8 Mid-term test 3
9 The Hamiltonian Formalism (part 1) 3 Lecture
10 The Hamiltonian Formalism (part 2) 3 Lecture
11 Poisson brackets and a relation between CM and QM 3 Lecture
12 Canonical Transformation 3 Lecture
13 Applications of the Hamiltonian Formalism 3 Exercise
4
Contents
1 Mathematical backgrounds 8
1.1 Unit vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.1.1 Unit vectors in Cartesian coordinates . . . . . . . . . . . 8
1.1.2 Unit vectors in cylindrical coordinates . . . . . . . . . . . 8
1.1.3 Unit vectors in spherical coordinates . . . . . . . . . . . . 9
1.1.4 General unit vectors and vector algebra . . . . . . . . . . 9
1.2 Vector differentiation of a scalar field and the gradient . . . . . . . 11
1.3 Conservative vector field . . . . . . . . . . . . . . . . . . . . . . 12
1.4 The vector differential operator and its action . . . . . . . . . . . 12
1.4.1 The divergence of a vector . . . . . . . . . . . . . . . . . 12
1.4.2 Laplacian operator ∇2 . . . . . . . . . . . . . . . . . . . 13
1.4.3 The curl of a vector . . . . . . . . . . . . . . . . . . . . . 13
1.5 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3 Lagrangian formalism 21
3.1 The principle of least action . . . . . . . . . . . . . . . . . . . . 21
3.1.1 Configuration space . . . . . . . . . . . . . . . . . . . . 21
3.1.2 Lagrangian . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.2 Lagrange’s equation under changing coordinate systems . . . . . 23
5
Classical Mechanics II
8 Canonical transformations 45
8.1 The equation of canonical transformation . . . . . . . . . . . . . 45
8.2 Infinitesimal canonical transformations . . . . . . . . . . . . . . . 48
6
Classical Mechanics II
7
Chapter 1
Mathematical backgrounds
8
Classical Mechanics II
as followed:
ρ̂ = cos(ϕ) · x̂ + sin(ϕ) · ŷ
φ̂ = − sin(ϕ) · x̂ + cos(ϕ) · ŷ (1.3)
ẑ = ẑ
We see that ρ̂ and ϕ̂ depend on ϕ and are not constant in direction. Hence when
differentiating or integrating in cylindrical coordinates, these unit vectors them-
selves must also be operated on:
∂ ρ̂
∂ ϕ = − sin(ϕ) · x̂ + cos(ϕ) · ŷ
∂ ϕ̂
∂ϕ = − cos(ϕ) · x̂ − sin(ϕ) · ŷ (1.4)
∂ ẑ
=0
∂ϕ
Similarly with cylindrical unit vectors, spherical unit vectors are functions of θ
and ϕ and are not constant in direction so when differentiating or integrating in
spherical coordinates, these unit vectors themselves must also be operated on.
9
Classical Mechanics II
So we have
εi jk = εki j = ε jki = −ε jik = −εik j = −εk ji
and useful relations between the Levi-Civita symbol and Kronecker delta function:
3
∑ εmnk .εi jk = δmi .δn j − δm j .δni
k=1
3
∑ εm jk .εn jk = 2δmn (1.10)
j,k=1
3
∑ εi jk .εi jk = 6
i, j,k=1
A = kAk · Â = Ai êi
Ai
⇒ Â = · êi
kAk
In the following we list some important properties of vector algebra (example
in three dimensional vector space):
• Equality of vectors:
A = B ⇔ (A1 , A2 , A3 ) = (B1 , B2 , B3 )
A1 = B1
⇔ A2 = B2
A3 = B3
Equal vectors are parallel and have the same length, but don’t necessarily
have the same position.
10
Classical Mechanics II
• Vector addition
A + B = (A1 , A2 , A3 ) + (B1 , B2 , B3 )
(1.12)
= (A1 + A2 , A2 + B2 , A3 + B3 )
• Multiplication by a scalar
λ .A = λ .(A1 , A2 , A3 ) = (λ .A1 , λ .A2 , λ .A3 )
11
Classical Mechanics II
Divergence of V is defined as
∂ ∂ ∂ ∂
∇ · V = ê1 + ê2 + ê3 · (V1 ê1 +V2 ê2 +V3 ê3 ) ≡ êi · V j ê j
∂ x1 ∂ x2 ∂ x3 ∂ xi
∂V j
⇒ ∇·V = (êi · ê j )
∂ xi
From (1.6) we have
∂V j ∂Vi ∂V1 ∂V2 ∂V3
∇·V = δi j = ≡ + +
∂ xi ∂ xi ∂ x1 ∂ x2 ∂ x3
12
Classical Mechanics II
Note that ∇ · V 6= V · ∇:
∂ ∂ ∂ ∂
V · ∇ = Vi = V1 +V2 +V3 .
∂ xi ∂ x1 ∂ x2 ∂ x3
1.5 Exercises
Given differentiable vector fields functions A and B and differentiable scalar field
functions f and g of position (x1 , x2 , x3 ), prove the following relations:
1. ∇ ( f g) = f ∇ g + g∇
∇f,
2. ∇ · ( f A) = f ∇ · A + ∇ f · A,
3. ∇ × ( f A) = f ∇ × A + ∇ f × A,
4. ∇ × (∇
∇ f ) = 0,
5. ∇ · (∇
∇ × A) = 0,
6. ∇ · (A × B) = (∇
∇ × A) · B − (∇
∇ × B) × A,
8. ∇ × (∇ ∇ · A) − ∇ 2 A,
∇ × A) = ∇ (∇
9. ∇ (A · B) = A × (∇
∇ × B) + B × (∇
∇ × A) + (A · ∇ )B + (B · ∇ )A,
13
Classical Mechanics II
10. (A · ∇ )r = A,
11. ∇ · r = 3,
12. ∇ × r = 0,
13. ∇ · (r−3 r) = 0,
15. dϕ = dr · ∇ ϕ + ∂∂tϕ dt
14
Chapter 2
where:
• F(r, ṙ) is the force acting on the particle. In general this force depends on
the position r and the velocity ṙ of the particle.
r = r0 + v.t (2.2)
This is the main content of first law of Newton stating an existent of such an inertial
frame.
An inertial frame is not unique. There are an infinite number of inertial frame.
Let S be an inertial frame. Then there are 10 linearly independent transformations
S → S0 such that S0 is also an inertial frame:
15
Classical Mechanics II
τ = r×F (2.4)
From (2.3) and (2.4) we have
τ = L̇ (2.5)
16
Classical Mechanics II
(2.8) implies that the change in kinetic energy is equal to the work done by the
force acting on the particle. This is the content of the ”work-kinetic energy ”
theorem.
In a special case of conservative force field in which the force depends only on
position r rather than velocity ṙ and the work done is dependent of the path taken.
Thus, for a closed path, the work done vanishes
I
F.dr = 0 ⇔ ∇ × F = 0. (2.9)
We thus can write F in this form:
F = −∇
∇V (r), (2.10)
where V (r) can be understood as the potential. Systems having a potential of this
form can be gravitational, electrostatic and interatomic force-fields.
Substituting (2.10) into (2.8) we have:
Zt2
T (r2 ) − T (r1 ) = − ∇V (r).dr = −(V (r2 ) −V (r1 )), (2.11)
t1
⇒ T (r1 ) + T (r2 ) = V (r1 ) +V (r2 ) ≡ E, (2.12)
where E is the energy of the particle and we see that it is conserved when the
particle is moving in conservative force-fields.
2.1.4 Exercises
• The simple harmonic oscillator: a system moving in one-dimension with
a force proportional to the distance x to the origin: F(x) = −kx, (k is a con-
stant). This force arises from the potential V = kx2 /2. Examine which quan-
tities between momentum, angular momentum and energy are/is conserved?
(ans: momentum is not conserved because F 6= 0, angular momentum is not
defined in one-dimension; energy is conserved).
• Particle moving under gravity: Consider a particle of mass m moving in
3-dimensions under the gravitational pull of a much larger particle of mass
M. The force is F = −G.M.m.r̂/r2 which arises from the potential V =
−G.M.m/r. Are momentum, angular momentum and energy are conserved?
(Ans: momentum is not conserved because F 6= 0; angular momentum is
conserved because F ∼ r; energy is conserved).
• The damped simple harmonic oscillator: We now include a friction term
so that F = −kx − γ ẋ. Examine the conservation of momentum, angular
momentum, and energy? (Ans: the force is not conservative so energy is not
conserved but lost until the partice stops moving).
17
Classical Mechanics II
where fi j is the internal force which particle jth acts on particle ith and Fi is the
sum of all external forces acting on particle ith .
Summing over (2.14) for all particles we have:
N
∑ Ftot
i = ∑ fi j + ∑ Fi = ∑ (fi j + f ji ) + ∑ Fi (2.15)
i=1 i, j, j6=i i i< j i
According to the the third law of Newton: fi j = −f ji , the first term in (2.15)
vanishes. We denote:
• Total of all external forces acting on the system of particles:
N
Fext = ∑ Fi
i=1
Then the second law of Newton for the whole of system can be written as
Fext = M R̈. (2.16)
(2.16) has the same form as the second law of Newton for a point particle. This
formula implies that the centre of mass of a system of particles acts just as if all
the mass were concentrated there.
18
Classical Mechanics II
We can write the second law of Newton (2.16) for the center of mass as
Fext = P̈.
L = ∑ Li = ∑(ri × pi )
i i
L̇ = ∑(ri × ṗi )
i
= ∑ ri × ( ∑ fi j + Fi )
i j6=i
= ∑ ri × fi j + ∑ ri × Fi
i, j, j6=i i
= ∑ (ri − r j ) × fi j + ∑ ri × Fi
i< j i
τ ext = ∑ ri × Fi ,
i
then we have
L̇ = τ ext .
So if the total external torque τ ext = 0 the angular momentum of the system of
particles is conserved.
19
Classical Mechanics II
(2.18) tells that the kinetic energy splits up into the kinetic energy of the centre
of mass, together with an internal energy describing how the system is moving
around its centre of mass.
Similarly to the case of single particle we have the following:
• The change in total kinetic energy:
Z
T (t2 ) − T (t1 ) = ∑ Fi .dri + ∑ fi j .dri
i i, j,i6= j
where ∇ i = ∂ /∂ ri
In order to get fi j = −f ji and fi j ∝ (ri − r j ) we can require the internal potential to
depend only on the distance between particle ith and particle jth :
so that the external force acting on particle i doesn’t depend on position of other
particles. Hence, similarly to the case of single particle we can prove (exercise)
T +V = const
20
Chapter 3
Lagrangian formalism
∂V
ṗA = − , (3.1)
∂ xA
where pA = mẋA .
3.1.2 Lagrangian
We define Lagrangian L(xA , ẋA ), a function of position xA and velocity ẋA , as
follow:
21
Classical Mechanics II
To describe the principle of least action, we consider all smooth paths xA (t) in
C with fixed end points:
xA (ti ) = xinitial
A
; xA (t f ) = xAfinal
To each path, we assign a number called the action S defined as
Z tf
S[xA (t)] = L(xA (t), ẋA (t))dt. (3.3)
ti
The principle of least action is “The actual path taken by the system is an
extremum of S.”
Proof: Consider varying a given path slightly, so
xA (t) → xA (t) + δ xA (t) (3.4)
Z tf tf
∂L d ∂L ∂L
⇒ δS = A
− A
δ xA dt + A δ xA (3.5)
ti ∂x dt ∂ ẋ ∂ ẋ ti
Since δ xA (ti ) = 0 = δ xA (t f ), the final term vanishes. In order to have δ S = 0
for all variation δ xA (t) we require
∂L d ∂L
− = 0, with A = 1, 2, ..., 3N (3.6)
∂ xA dt ∂ ẋA
These are known as Lagranges equations (or sometimes as the Euler-Lagrange
equations).
To finish the proof, we need only show that Lagranges equations are equivalent
to Newtons ⇒ Exercise!!.
• All the fundamental laws of physics can be written in terms of an action
principle. This includes electromagnetism, general relativity, the standard
model of particle physics, and attempts to go beyond the known laws of
physics such as string theory. For example, (nearly) everything we know
about the universe is captured in the Lagrangian
√ 1 µν
L = g R − F Fµν + ψ̄ 6 Dψ (3.7)
2
where the terms carry the names of Einstein, Maxwell (or Yang and Mills)
and Dirac respectively, and describe gravity, the forces of nature (electro-
magnetism and the nuclear forces) and the dynamics of particles like elec-
trons and quarks.
• There are two very important reasons for working with Lagranges equa-
tions rather than Newtons. The first is that Lagranges equations hold in any
coordinate system, while Newtons are restricted to an inertial frame. The
second is the ease with which we can deal with constraints in the Lagrangian
system.
22
Classical Mechanics II
so
dxA ∂ xA ∂ xA
ẋA ≡ = q̇a + (3.11)
dt ∂ qa ∂t
The change of Lagrange L due to the change of coordinate system:
2 A
∂ L ∂ xA ∂ L ∂ 2 xA
∂L ∂ x
= + q̇b + (3.12)
∂ qa ∂ xA ∂ qa ∂ ẋA ∂ qa ∂ qb ∂t∂ qa
∂L ∂ L ∂ ẋA
= (3.13)
∂ q̇a ∂ ẋA ∂ q̇a
we can prove (???)
∂ ẋA ∂ xA
=
∂ q̇a ∂ qa
Finally we obtain:
∂ L ∂ xA
d ∂L ∂L d ∂L
− = − A (3.14)
dt ∂ q̇a ∂ qa dt ∂ ẋA ∂ x ∂ qa
Equation (3.14) implies that if the choice of coordinates is invertible (i.e: det(∂ xA /∂ qa ) 6=
0), the Lagrange’s equation is invariant with the change between coordinates xA
and qa . So the form of Lagrange’s equations holds in any coordinate system. This
is in contrast to Newton’s equations which are only valid in an inertial frame.
23
Classical Mechanics II
24
Classical Mechanics II
Holonomic Constraints
“Holonomic constraints” are relationships between the coordinates of the form
fα (xA ,t) = 0, α = 1, 2, ..., (3N − n)
Holonomic constraints can be solved in terms of n “generalised coordinates”
qi , i = 1, 2, ..., n. So we have
xA = xA (q1 , q2 , ..., qn )
The system is said to have n degrees of freedom.
We now introduce (3N − n) new variables λα , called “Lagrange multipliers”
and define a new Lagrangian:
25
Classical Mechanics II
26
Chapter 4
27
Classical Mechanics II
Noether’s theorem states that for each such continuous symmetry there exists a
conserved quantity.
Proof:
We have
∂L d ∂ L ∂ Qi
=
∂ s s=0 dt ∂ q̇i ∂ s s=0
Obviously we see that if
∂L
∂s s=0
then this quantity is conserved:
∂ L ∂ Qi
∂ q̇i ∂ s s=0
This is the statement that space is homogeneous and a translation of the sys-
tem by sn does nothing to the equations of motion. Note that these translations
are elements of Galilean group. From Noether’s theorem, conserved quantities
associated with translations can be determined as:
∂L
∑ ∂ ṙi .n = ∑ pi.n.
i=1 i
This quantity is recognized as the total linear momentum of the system in the
direction n. Since n is arbitrary so we can conclude that the total momentum of
the system ∑i pi is conserved.
28
Classical Mechanics II
ri → ri + δ ri
= ri + α n̂ × ri
According to Noether’s theorem, the conserved quantity associated with this sym-
metry is
∂L
.(n̂ × ri ) = n̂.(ri × pi ) = n̂.L (4.2)
∂ ṙi
where L is total angular momentum of the system.
∂L
H = q̇i − L.
∂ q̇i
In the system we are considering, this quantity is total energy.
29
Chapter 5
Applications of Lagrangian
formalism
5.1
If L is a Lagrangian for a system of n degrees of freedom satisfying Lagrange’s
equations, show by direct substitution that
dF(q1 , q1 , ..., qn ,t)
L0 = L +
dt
also satisfies Lagrange’s equations, where F is any arbitrary, but differentiable,
function of its arguments.
5.2
Obtain the Lagrangian equations of motion for a spherical pendulum, i.e. a mass
point suspended by a rigid weightless rod.
5.3
A particle of mass m moves in one dimension such that it has the Lagrangian
m2 ẋ4
L= + mẋ2V (x) −V 2 (x),
12
where V is some differentiable nature of the system on the basis of this equation.
Find the equation of motion for x(t) and describe the physical nature of the system
on the basic of this equation.
30
Classical Mechanics II
5.4
Obtain the equation of motion for a particle falling vertically under the influence
of gravity when frictional force obtainable from a dissipation function kv2 /2 are
present. Integrate the equation to obtain the velocity as a function of time and
show that the maximum possible velocity for fall from rest is v = mg/k.
5.5
The Lagrangian for a particle in an electromagnetic field is
q
L = T − qΦ(r(t)) + A(r(t)).v(t).
c
The electromagnetic field is invariant under a gauge transformation of the scalar
and vector potential given by
A → A + ∇Ψ(r,t),
1 ∂Ψ
Φ → Φ− ,
c ∂t
where Ψ is arbitrary (but differentiable). What effect does this gauge transforma-
tion have on the Lagrangian of a particle moving in the electromagnetic field? Is
the motion affected?
31
Classical Mechanics II
32
Classical Mechanics II
• Write lagrangian of the system in terms of separation r12 and the center of
mass R = (m1 r1 + m2 r2 )/(m1 + m2 )
• Find equation of motion and then the effective potential of the system.
B = ∇×A
1 ∂A
E = −∇φ −
c ∂t
where c is the speed of light. The Lagrangian describing motion of the particle is
1 2 1
L = mṙ − e φ − ṙ.A
2 c
33
Chapter 6
34
Classical Mechanics II
Assume we have an arbitrary function f (x, y) of two variables x and y. Its total
derivative is
∂f ∂f
df = dx + dy. (6.3)
∂x ∂y
Using f (x, y) we can define a new function which depends on three variables x, y
and u as follows:
g(x, y, u) = u x − f (x, y), (6.4)
and the total derivative of g can be calculated as
∂f ∂f
dg = u dx + x du − dx − dy. (6.5)
∂x ∂y
If u is a specific function of x and y:
∂f
u(x, y) = , (6.6)
∂x
then (6.5) becomes:
∂f
dg = x du − dy. (6.7)
∂y
We can invert (6.6) to get x as a function of u and y which is denoted as x(u, y).
So we can obtain g as a function of only u and y as follows:
(6.8) is the Legendre transform. It helps to transform one function f (x, y) to an-
other function g(u, y) where u = ∂ f /∂ x. In this transformation, we have not lost
any information. Vice versa, we can always recover f (x, y) from g(u, y) by
∂g ∂g ∂f
= x(u, y) , = , (6.9)
∂u y ∂y u ∂y
then
∂g
f= u−g (6.10)
∂u
will take us back to the original function.
35
Classical Mechanics II
define the Hamiltonian to be the Legendre transform of the Lagrangian L(qi , q̇i ,t)
with respect to the variables q̇i :
n
H(qi , pi ,t) = ∑ pi q̇i − L(qi , q̇i ,t), (6.11)
i=1
where q̇i is eliminated from the right hand side in favor of pi by this relation:
∂L
pi = ≡ pi (qi , q̇i ,t). (6.12)
∂ q̇i
q̇i = q̇i (qi , pi ,t) can be get back by inverting (6.12).
Consider the variation of H(qi , pi ,t):
∂L ∂L ∂L
dH = (q̇i d pi + pi d q̇i ) − dqi + d q̇i + dt
∂ qi ∂ q̇i ∂t
∂L ∂L
= q̇i d pi − dqi − dt. (6.13)
∂ qi ∂t
On the other hand, dH can be written as
∂H ∂H ∂H
dH = dqi + d pi + dt. (6.14)
∂ qi ∂ pi ∂t
Equating terms of (6.13) and (6.14) we have
∂H
ṗi = −
∂ qi
∂H
q̇i =
∂ pi
∂L ∂H
− = (6.15)
∂t ∂t
(6.15) are Hamilton’s equations. We see that by transforming the Lagrangian
formalism to the Hamiltonian formalism we replace n second order differential
equations for qi to 2n first order differential equations for pi and qi . Technically,
for solving equations, this is not helpful but we will see, conceptually, this is very
helpful.
6.1.3 Examples
A particle in a potential
A particle moving in a potential in 3-dimensional space. The Lagrangian is simply
1
L = mṙ2 −V (r) (6.16)
2
36
Classical Mechanics II
37
Classical Mechanics II
∂H
ṙ =
∂p
1 e
= p− A (6.25)
m c
ṗ = −∂ H/∂ r can be expressed in terms of components as:
∂H
ṗi = −
∂ ri
∂φ e e ∂Aj
= −e + pj − Aj (6.26)
∂ ri cm c ∂ ri
?? Exercise: Derive the Lorentz force which is has the same form as derived
using the Lagrangian formalism !
∂H
ṗ = = 0. (6.28)
∂q
38
Classical Mechanics II
Analogous to the case of Lagrangian formalism, we require that the end points of
qi are fixed:
δ qi (t1 ) = 0 = δ qi (t2 ). (6.32)
δ pi can be free at the end points t = t1 and t = t2 . So we see that, in this context,
qi and pi are not symmetric in the Hamiltonian formalism as we expect.
Thus, from (6.31), if we look for extrema δ S = 0 for all δ qi and δ pi we get
Hamilton’s equations:
∂H
q̇i =
∂ pi
∂H
ṗi = − (6.33)
∂ qi
6.4 Exercises
39
Chapter 7
• Anti-commutative:
{ f , g} = −{g, f }
• Linearity:
{α f + β g, h} = α{ f , h} + β {g, h}
• Leibniz rule:
{ f g, h} = f {g, h} + { f , h}g
• Jacobi identity:
40
Classical Mechanics II
We see that the Poisson bracket {, } satisfy the same algebraic structure as matrix
commutator [, ], and the differentiation operator d. This is related to Heisenberg’s
and Schrödinger’s point of view on quantum mechanics respectively.
The relationship between classical mechanics and quantum mechanics is also
shown via these expressions:
{qi , q j } = 0
{pi , p j } = 0
{qi , p j } = δi j (7.2)
df ∂f
= { f , H} + (7.3)
dt ∂t
An interesting consequence of (7.3) is that if we can find a function I(p, q)
(doesn’t depend on time explicitly) which satisfy
{I, H} = 0 (7.4)
then I(p, q) is a constant of motion. I and H are said to be Poisson commute. Here
are some examples of this consequence:
{pi , H} = 0
This means that {I, J} is also a constant of motion. We say that the constants
of motion form a closed algebra under the Poisson bracket.
41
Classical Mechanics II
Runge-Lenz vector
The Runge-Lenz vector was proposed by Hermann, Bernoulli, Laplace and Pauli.
It is defined as
1
A = p × L − r̂, (7.9)
m
where r̂ = r/krk. This vector satisfies these relations:
A·L = 0 (7.10)
and
2 p2 1
{Li , A j } = εi jk Ak , {Ai , A j } = − − εi jk Lk (7.11)
m 2m r
If the Hamiltonian has a fork
p2 1
H= − (7.12)
2m r
then we have
2H
{Ai , A j } = − εi jk Lk (7.13)
m
and
{H, A} = 0. (7.14)
We see that the Hamiltonian with −1/r potential has another constant of motion
A.
42
Classical Mechanics II
∇ · B = 0. (7.17)
(7.17) states that any flux that enters a region must also leave. This means that
there can be no magnetic monopole.
A monopole would have a radial magnetic field:
r
B=g . (7.18)
r3
It’s easy to point out that (7.18) does not satisfy (7.17). This means magnetic
monopoles are forbidden by Maxwell’s equations.
Now we can consider the motion of an electron in the background of a monopole.
We don’t have a gauge potential A to set up the Lagrangian. But we can use the
Poisson brackets (7.16) which contain only magnetic field.
Consider the generalized angular momentum
ge
J = mr × ṙ − r̂ (7.19)
c
where r̂ = r/r. Using (7.16) we can show that the Hamiltonian given at (7.15) and
J satisfy
{H, J} = 0. (7.20)
This implies that J is a constant of motion. Since J is conserved, we can look
at r̂ · J = −eg/c to learn that the motion of an electron in the background of a
magnetic monopole lies on a cone of angle cos θ = eg/cJ pointing from the vector
J.
43
Classical Mechanics II
7.4 Exercises
44
Chapter 8
Canonical transformations
x = (q1 , . . . , qn , p1 , . . . , pn )T (8.1)
With this notation, the Hamilton’s equations are written in the form:
∂H
ẋ = M . (8.3)
∂x
So far we know in the Lagrangian formalism, the Lagrange’s equations are in-
variant under a change of the coordinate system. In the Hamiltonian formalism
we will do the same and let’s see whether the Hamilton’s equations are invariant
under a change of the coordinate system.
Assuming we have this transformation of the coordinate system defined at
(8.1):
xi → yi (x), i = 1, . . . , 2n. (8.4)
45
Classical Mechanics II
We have:
∂ yi
ẏi = ẋ j (8.5)
∂xj
∂ yi ∂ H ∂ yl
= M jk
∂xj ∂ yl ∂ xk
∂H
= (Ji j M jk Jlk )
∂ yl
Collating all indices we have
∂H
ẏ = (JMJ T ) (8.6)
∂y
where
∂ yi
Ji j =
∂xj
is the Jacobian of the transformation of coordinates. We see that Hamilton’s equa-
tions are invariant under any transformation of coordinates whose Jacobian M
satisfies:
JMJ T = M (8.7)
or
∂ yi ∂ yl
M jk = Mil (8.8)
∂xj ∂ xk
The Jacobian M that satisfies this condition is said to be symplectic and a trans-
formation of coordinates with a symplectic Jacobian is said to be a canonical
transformation.
We have this theorem relating canonical transformation with a symplectic Ja-
cobian:
Theorem:
The Poisson bracket is invariant under canonical transformations. Conversely,
any transformation which preserves the Poisson bracket structure so that
is canonical.
46
Classical Mechanics II
Proof
• Firstly, we show that the Poisson bracket is invariant under canonical trans-
formations. Consider two functions:
∂ f ∂g ∂ f ∂g
{ f , g} = − (8.10)
∂ qi ∂ pi ∂ pi ∂ qi
∂f ∂g
= Mi j
∂ xi ∂xj
∂f ∂f
= Jki (8.11)
∂ xi ∂ yk
Subtituting (8.11) into (8.11) we have
∂f ∂g
{ f , g} = Jki Mi j Jl j (8.12)
∂ yk ∂ yl
If the transformation (8.11) is canonical where (8.7) is hold then
∂f ∂g
{ f , g} = Mkl . (8.13)
∂ yk ∂ yl
This emplies that Poisson brackets are invariant under a canonical transfor-
mation.
• Conversely, we will write the Jacobian using notation (qi , pi ) (in the old
co-ordinates) and (Qi (q, p), Pi (q, p)) as:
∂ Qi /∂ q j ∂ Qi /∂ p j
Ji j = (8.14)
∂ Pi /∂ q j ∂ Pi /∂ p j
47
Classical Mechanics II
48
Classical Mechanics II
This integration is evaluated for all paths q(t) with fixed end points
f inal
qi (0) = qinitial
i , qi (T ) = qi . (8.27)
Then the true path taken is an extremum of the action or δ S = 0.
Now we consider the action evaluated only along the true path qclassical
i and
define f inal
initial,qi ,T
W (qi ) = S[qclassical
i ]. (8.28)
From the Lagrangian formalism we know how to get the variation of the action
S caused by varying qi (t):
Z T T
∂L d ∂L ∂L
δS = dt − δ qi (t) + δ qi (t) (8.29)
0 ∂ qi dt ∂ q̇i ∂ q̇i 0
49
Classical Mechanics II
If we evaluate this on the classical path (qi (t) satisfying Euler-Lagrange’s equa-
tion) and qi (t) is fixed at the initial point (δ qi (0) = 0) then we have
∂L
δS = δ qi (t) (8.30)
∂ q̇i t=T
f inal
On the other hand, from (8.28) we have (if we consider only a variation of qi )
∂W f inal
δ S = δW = f inal
δ qi (8.31)
∂ qi
So we get
∂W ∂L f inal
f inal
= = pi (8.32)
∂ qi ∂ q̇i t=T
8.5 Exercises
Exercise 1
• A system moving in one-dimensional space described by the Hamiltonian:
p2 1
H= − 2. (8.38)
2 2q
50
Classical Mechanics II
Exercise 2
Show directly that the transformation
1
Q = log sin p , P = q cot p (8.42)
p
is canonical.
Exercise 3
Show directly for a system of one degree of freedom that the transformation
αq2 p2
αq
Q = arctan , P= 1+ 2 2 (8.43)
p 2 α q
Exercise 4
Solve the problem of the motion of a point projectile in a vertical plane, using
the Hamiltonian-Jacobi method. Find both the equation of the trajectory and the
dependence of the coordinates on time, assuming the projectile is fired off at time
t = 0 from the origin with the velocity v0 , making an angle α with the horizontal.
51