You are on page 1of 10

Chapter 2

Review of Classical Mechanics

HW #2: 1.7.2, 2.7.3, 2.7.8, 2.8.1, 2.8.2, 2.8.3, 2.8.4.

2.1 Variational principle


In classical mechanics an action of a single particle described by coordinates
q(t) (in configuration space) is
!
S[q] ≡ dt L(q, q̇). (2.1)

which can be thought of as a functional of q(t). Then the classical equations


of motion can be obtained from the most important principle in physics - the
variational principle:
δS[q]
= 0. (2.2)
δq
By Taylor expanding the Lagrangian
∂L ∂L
L(q + δq) ≈ L(q) + δq + δ q̇ (2.3)
∂q ∂ q̇

18
CHAPTER 2. REVIEW OF CLASSICAL MECHANICS 19

we obtain
" #
δS δL
!
0= = dt
δq δq
$
dt (L(q + δq) − L(q))
=
δq
% &
$ ∂L(q) ∂L
dt ∂q
δq + ∂ q̇
δ q̇
=
δq
% % & % & &
∂L(q)
δq + dtd ∂L d ∂L
$
dt ∂q ∂ q̇
δq − dt ∂ q̇
δq
= , (2.4)
δq
or up to a boundary term
" #
∂L d ∂L
− =0 (2.5)
∂q dt ∂ q̇
For example,
1
L = q̇ 2 − V (q) (2.6)
2
gives rise to
∂V
q̈ = −. (2.7)
∂q
In what follows we will consider two more examples:
• Charged particle
1 '' ˙''2 e ⃗
Lem = m '⃗q ' − eφ(⃗q ) + ⃗q˙ · A(⃗
q) (2.8)
2 c
• Two-body problem
1 ' '2 1 ' '2
L = m1 '⃗q˙1 ' + m2 '⃗q˙2 ' − V (|⃗q1 − ⃗q2 |) (2.9)
' ' ' '
2 2
Exercise: Derive equations of motion for theories (2.8) and (2.9) using
variational principle (2.4).

2.2 Field theories


In classical field theory an action for a collection of fields described by
Φi (x0 , x1 , x2 , x3 ) is
!
S[Φ , ..., Φ ] = d4 x L(Φ1 , ∂0 Φ1 , ∂1 Φ1 , ∂2 Φ1 , ∂3 Φ1 , ...)
1 N
(2.10)
CHAPTER 2. REVIEW OF CLASSICAL MECHANICS 20

(which is a slightly more complicated functional) one can still use the varia-
tional principle to obtain N equations of motion
δS
=0 (2.11)
δΦi
for N degrees of freedom. Note that the action is dimensionless which sug-
gests that the so-called Lagrangian density L must have the dimensions

[L] = [Length]−4 = [T ime]−4 = [Mass]4 = [Energy]4. (2.12)

By Taylor expanding the perturbed Lagrangian


∂L i ∂L
L(..., Φi +δΦi , ∂µ Φi +∂µ Φi , ...) = L(..., Φi , ∂µ Φi , ...)+ δ ∂µ Φi
( )
i
δΦ + i
∂Φ ∂ (∂µ Φ )
(2.13)
where

∂µ =
∂xµ
we get (up to the boundary term)
* % &+
∂L ∂L
d4 x δΦi
$
δS ∂Φi
− ∂µ ∂(∂µ Φi )
0= = (2.14)
δΦi δΦi
or " #
∂L ∂L
− ∂µ = 0. (2.15)
∂Φi ∂ (∂µ Φi )
For example, the scalar field Lagrangian
1
L = − ∂µ φ∂ µ φ − V (φ) (2.16)
2
gives rise to an equation of motion
∂V
!φ − = 0, (2.17)
∂φ
where
" #2 " #2 " #2 " #2
µ ∂ ∂ ∂ ∂
! ≡ ∂µ ∂ = − + + + . (2.18)
∂t ∂x ∂y ∂z
For a massless vector field coupled to a conserved current, i.e.
1
L = − Fµν F µν + Aµ j µ (2.19)
4
CHAPTER 2. REVIEW OF CLASSICAL MECHANICS 21

where
Fµν = ∂µ Aν − ∂ν Aµ (2.20)
the equations of motion are
∂µ F νµ = j ν . (2.21)
Another useful example is the so-called Klein-Gordon theory of a complex
field with Lagrangian density
1 1
L = − ∂µ ϕ∂ µ ϕ∗ − m2 ϕϕ∗ (2.22)
2 2
whose equations of motion are given by
!ϕ = m2 ϕ (2.23)
Note that for some time the Klein-Gordon equation(2.23) was regarded as
a relativistic Schrodinger equation (to be derived in the following chapter),
where ϕ was considered to be a wave function. This is however misleading
since the Schrodinger equation must be first order in time, when the Klein-
Gordon equation is second order in time. Nevertheless it is true that in non-
relativistic limit the Klein-Gordon equation reduces to Schrodinger equation.
This can be seen by considering the following ansatz for scalar field
ϕ(x, t) = ψ(x, t) exp(−imt) (2.24)
Then under assumption

m≫ φ (2.25)
∂t
we have
" # " #
∂ ∗∂ 2 ∗ ∂ ∂
ψ ψ−m ψ ψ = im + ∗
ψ −im + ψ − m2 ψ ∗ ψ(2.26)
∂t ∂t ∂t ∂t
" #
∗ ∂ ∂ ∗
≈ im ψ ψ−ψ ψ
∂t ∂t
and this the Lagrangian (2.22) reduces to
1 1
L = im (ψ ∗ ∂0 ψ − ψ∂0 ψ ∗ ) + ∂i ψ∂ i ψ (2.27)
2 2
or after integrating by parts
1
L = imψ ∗ ∂0 ψ − ∂i ψ∂ i ψ ∗ (2.28)
2
whose equation of motion take the form of a Schrodinger equation

i ψ = −∂i ∂ i ψ. (2.29)
∂t
CHAPTER 2. REVIEW OF CLASSICAL MECHANICS 22

2.3 Hamiltonian mechanics


Conjugate variables, Phase space, Hamiltonian, Hamiltonian equations
Consider a system with N degrees of freedom (d.o.f.) whose Lagrangian
is given by some function of 2N variables

L(q1 , ..., qN , q̇1 , ..., q̇N ) (2.30)

Then Hamiltonian function is defined as a Legendre transform of the La-


grangian function,
N
,
H(q1 , ..., qN , , p1 , ..., pN ) ≡ pi q̇i − L(q1 , ..., qN , p1 , ..., pN ) (2.31)
i=1

and the conjugate momenta are defined as


∂L
pi ≡ . (2.32)
∂ q˙i

where i = 1...N. (Note that Hamiltonian depends only on q ′ s and on p’s and
so one must express q̇’s in terms of q ′ s and on p’s using 2.32 and plug the
result into 2.31.)
The states of such system are described by a 2N dimensional vector
(q1 , q2 , ..., qN , p1 , p2 , ..., pN ) in the so-called phase space Γ, where {q1 , q2 , ..., qN }
are the position coordinates and {p1 , p2 , ..., pN } are the momentum coordi-
nates. The corresponding pairs of the phase space coordinates(q1 , p1 ),(q2 , p2 )...
(qN , pN ) are known as conjugate pairs, or conjugate variables.
We have already seen how the variational principle can be used to derive
equations of motion from Lagrangians (i.e. Euler-Lagrange equations), but
it turns out that the same classical evolution can be recovered from the so-
called Hamiltonian-Jacobi equations given by
∂H
ṗi = − (2.33)
∂qi
and
∂H
q̇i = . (2.34)
∂pi
For example, consider a simple harmonic oscillator described by La-
grangian
1
L = q̇ 2 − V (q) (2.35)
2
CHAPTER 2. REVIEW OF CLASSICAL MECHANICS 23

then for variable q the conjugate momentum is


∂L
p≡ = q̇ (2.36)
∂ q̇
and thus
1
H(p, q) = pq̇(p) − q̇(p)2 + V (q)
2
1
= p2 − p2 + V (q)
2
1 2
= p + V (q). (2.37)
2
and equations of motion are
∂V (q)
ṗ = − ∂H
∂q
=−
∂q
∂H
q̇ = ∂p
= p. (2.38)

Exercise: Derive Hamiltonian and then Hamilton-Jacobi equations of mo-


tion for theories (2.8) and (2.9).

2.4 Liouville’s theorem


Poisson Brackets, Liouville theorem, Classical observables
Now consider a representative point in an infinitesimal 2N-dimensional
hypercube with volume

dΓ = dq1 ...dqN dp1 ...dpN (2.39)

In time δt the representative point moves to a new location

qi′ = qi + q̇i dt ⇒ dqi′ = dqi + dq̇i dt (2.40)


p′i = pi + ṗi dt ⇒ dp′i = dpi + dṗi dt (2.41)

and the volume becomes distorted


" " # # " " # #
′ ∂ q̇1 ∂ q̇1 ∂ ṗ1 ∂ ṗ1
dΓ = dq1 + dq1 + ... dp1 + ... dt ... dp1 + dq1 + ... dp1 + ... dt ....
∂q1 ∂p1 ∂q1 ∂p1
(2.42)
By expanding to the linear order in dt we obtain
N " N "
dΓ′ − dΓ , ∂ q̇i ∂ ṗi
# , #
∂ ∂H ∂ ∂H
= + = − =0 (2.43)
dtdΓ i=1
∂qi ∂pi i=1
∂qi ∂pi ∂pi ∂qi
CHAPTER 2. REVIEW OF CLASSICAL MECHANICS 24

where the second equality comes from Hamiltonian equations of motion (2.33)
and (2.34). The same result can be obtained by looking at the divergence of
the velocity flow v = (q̇1 , ..., q̇N , ṗ1 , ..., ṗN ) vanishes , i.e.
" # N " #
∂ ∂ ∂ ∂ , ∂ ∂H ∂ ∂H
∇·v = , ..., , , ..., ·(q̇1 , ..., q̇N , ṗ1 , ..., ṗN ) = − = 0.
∂q1 ∂qN ∂p1 ∂pN i=1
∂qi ∂pi ∂pi ∂qi
(2.44)
This is the famous Liouville’s theorem which says that the flow in a phase
space of Hamiltonian systems is incompressible, i.e. dΓ = dΓ′.
The Liouville’s theorem can also be expressed in a differential form. Let
us define a multiparticle distribution functionρ(q1 , ..., qN , p1 , ..., pN ) over the
phase space, where

ρ(q1 , ..., qN , p1 , ..., pN )dq1 ...dqN dp1 ...dpN (2.45)

is a probability to find a system in a state between (q1 , ..., qN , p1 , ..., pN ) and


(q1 + dq1 , ..., qN + dqN , p1 + dp1 , ..., pN + dpN ). Then the continuity equation

dρ ∂ρ ∂ ∂
= + (q̇i ρ) + (ṗi ρ) = 0 (2.46)
dt ∂t ∂qi ∂pi
and the incompressibility of flow leads to the Liouville equation
N " #
∂ρ , ∂ρ ∂H ∂ρ ∂H
=− − = − {ρ, H} . (2.47)
∂t i=1
∂qi ∂pi ∂pi ∂qi

where the Poisson brackets are defined as


N " #
, ∂A ∂B ∂A ∂B
{A, B} = − . (2.48)
i=1
∂qi ∂pi ∂pi ∂qi

Note that Liouville equation tells us how probability distributions over the
phase space evolve in time. In quantum mechanics probability distribution
will be defined over vectors in Hilbert space and then (2.47) will be replaced
with quantum Liouville equation
∂ ρ̂ 1
= [Ĥ, ρ̂]. (2.49)
∂t i!
where the commutator is defined as

[Â, B̂] ≡ ÂB̂ − B̂ Â. (2.50)


CHAPTER 2. REVIEW OF CLASSICAL MECHANICS 25

In what follows we will define precisely both ρ̂ (i.e. density matrix) and Ĥ (i.e.
Hamiltonian operator), but for now we should just notice that both Liouville
equations (2.47) and (2.45) look very much alike. In fact the replacement of
Poisson brackets with commutation relation will be a standard operation
i
{ , } → − [ , ]. (2.51)
!
For example, it is trivial to calculate Poisson brackets between two con-
jugate variables
{q, p} = 1 (2.52)
which suggest that the quantum commutation relation between position and
momentum operators should be

[q̂, p̂] = i! (2.53)

which is in agreement with (1.65). Many more example of the replacement


(2.51) will be given in the following chapter.
In Hamiltonian mechanics all of the observables are defined as functions
on the phase space, e.g. q1 - position of the first particle, p5 - momentum of
the fifth particle, etc. Then the ensemble average(or expectation value) of
some observable O(p, q) can be defined as
!
⟨O⟩ = O(q, p)ρ(q, p, t)dqdp. (2.54)

And then using the Liouville equation (2.47) we obtain


N " #
d⟨O⟩ ∂ρ(q, p, t) ∂ρ ∂H ∂ρ ∂H
! ! ,
= O(q, p) dqdp = − O − dqdp (=2.55)
dt ∂t i=1
∂qi ∂p i ∂p i ∂qi
N " " # " ##
∂ ∂H ∂ ∂H
! ,
=− ρ O − O dqdp =
i=1
∂qi ∂pi ∂pi ∂qi
N " " ##
∂O ∂H ∂O ∂H ∂ ∂H ∂ ∂H
! ,
= ρ(q, p) − −O − dqdp =
i=1
∂qi ∂pi ∂pi ∂qi ∂qi ∂pi ∂pi ∂qi
= ⟨{O, H}⟩.
(2.56)

Once again (2.56) through identification (4.27) is related to the Heisenberg


equation that we shall derived in the next chapter

dÔ i* +
= − Ô, Ĥ . (2.57)
dt !
CHAPTER 2. REVIEW OF CLASSICAL MECHANICS 26

2.5 Canonical transformations


To transform from one set of coordinates to another set of coordinates one
must specify two 2N functions of 2N variables each, i.e.

(⃗p, ⃗q ) → (⃗p′ (⃗p, ⃗q), ⃗q ′ (⃗p, ⃗q)) (2.58)

Such transformations are called canonical if the equations of motion (in terms
of new variables) remain Hamiltonian. To see what restriction this puts on
the transformations we write
, " ∂q ′ ∂H ∂H ∂qi′
#
′ i
q̇i = {qi , H} = −
j
∂qj ∂pj ∂qj ∂pj
, ∂p′ ∂H
"
∂H ∂p′i
#
′ i
ṗi = {pi , H} = − (2.59)
j
∂qj ∂pj ∂qj ∂pj

but
∂H(⃗p′ , ⃗q ′ ) , " ∂H ∂q ′ ∂H ∂p′k
#
k
= + ′
∂qi k
∂qk′ ∂qi ∂pk ∂qi
∂H(⃗p′ , ⃗q ′ ) , ∂H ∂q ′ ∂H ∂p′k
" #
k
= + (2.60)
∂pi k
∂qk′ ∂pi ∂p′k ∂pi

and so
, " ∂q ′ " ∂H ∂q ′ ∂H ∂p′k
#
∂qi′
"
∂H ∂qk′ ∂H ∂p′k
##
i k
q̇i′ = + ′ − +
j,k
∂qj ∂qk′ ∂pj ∂pk ∂pj ∂pj ∂qk′ ∂qi ∂p′k ∂qi
, " ∂H ∂H ′ ′
#
′ ′
= {q , q } + {q , p } (2.61)
k
∂qk′ i k ∂p′k i k
N "
∂qi′ ∂H ∂qk′ ∂H ∂p′k ∂qi′ ∂H ∂qk′ ∂H ∂p′k
, " # " ##
ṗ′i = + ′ − +
∂qj ∂qk′ ∂pj ∂pk ∂pj ∂pj ∂qk′ ∂qi ∂p′k ∂qi
j,k
, " ∂H ∂H ′ ′
#
′ ′
= ′
{p ,
i k q } + ′
{p ,
i kp } . (2.62)
k
∂qk ∂p k

Therefore in order for equations of motion to remain Hamiltonian we must


have the following relations between new variables

{qi′ , qk′ } = {p′i , p′k } = 0


{qi′ , p′k } = δik . (2.63)
CHAPTER 2. REVIEW OF CLASSICAL MECHANICS 27

and then these variables are called canonical and transformations (2.58) are
called canonical.
Now consider infinitesimal canonical transformations (which are in fact
canonical)

∂g
qi → qi′ + ε
∂pi
∂g
pi → p′i − ε (2.64)
∂qi

generated by some function (called generator of transformations)

g(⃗q, ⃗p). (2.65)

If the the Hamiltonian function is invariant under such transformation


, ∂H " ∂g # ∂H " ∂g #
′ ′
0 = H(⃗p , ⃗q )−H(⃗p, ⃗q ) = ε + −ε = ε {H, g} (2.66)
i
∂qi ∂pi ∂pi ∂qi

and thus g is conserved


ġ = {g, H} = 0. (2.67)
Another important result about canonical transformations which leave Hamil-
tonian invariant is that such transformation can be used to obtain new solu-
tions p′ (t) and ′ q(t) from old solutions p(t) and q(t).

You might also like