You are on page 1of 21

Chapter 4

Symmetries and Conservation Laws

Chapter contents:

Generalized Momentum

Symmetries and Conservation Laws

Noether’s Theorem

Conservation of Energy

1
Generalized Momentum
Lagrange’s equations: d ∂ L − ∂ L =0
( ) (α=1,2, ... , N )
dt ∂ q̇ α ∂ qα

Definition: Generalized momentum


∂L
pα ≡
∂ q̇ α

It is also called canonical momentum (or conjugate momentum).


Since the generalized coordinates {qα} do not necessarily have the
dimension of length, {pα} also do not need to have the dimension of linear
momentum.
dp α ∂ L
Lagrange’s equations => − =0
dt ∂ qα
2
If the Lagrangian L does not depend explicitly on a given coordinate,
say qk, then the coordinate qk is called a ignorable (or cyclic) coordinate.

∂L dp k
=0 => =0 => p k =constant
∂ qk dt

The generalized momentum pk conjugate to an ignorable


coordinate qk is a constant of motion.

L can still depend on q̇ k even though L does not depend on qk.

Remark:
Being able to find conserved quantities in a system is good because each
conserved quantity can be used to eliminate one degree of freedom,
bringing us closer to the solution.

3
Example:
The Lagrangian for a particle that falls under gravity is
1 2 2 2
L= m( ẋ + ẏ + ż )−mgz (z = vertical direction)
2
The coordinates x and y are ignorable:
∂L ∂L
px≡ =m ẋ=Constant ; p y≡ =m ẏ=Constant
∂ ẋ ∂ ẏ

For the z-direction, we have


∂L ∂L
=m ż , =−m g
∂ ż ∂z
d ∂L ∂L
( )
dt ∂ ż

∂z
=0 => z̈=−g

In this example, the generalized momentum equals to the mechanical


momentum. But note that there are cases where the two momenta do
not equal. 4
Example:
A particle moves around a force center on a 2D plane.
The potential energy depends only on r. Let us
consider the Lagrangian in (x, y) and (r, φ)
coordinates.

Cartesian 1
L= m( ẋ 2 + ẏ 2 )−V (r)
coordinates: 2
2 2
L depends on (x , y , ẋ , ẏ) r= √ x + y

Polar 1 2
coordinates: L= m(ṙ 2 +r φ̇ 2)−V (r)
2
L does not depend on φ => φ is ignorable
∂L 2 2
=> pφ = =m r φ =Constant
∂ φ̇
=> Conservation of angular momentum about O

Note: There is no ignorable coordinate in Cartesian coordinates 5


Example: Charged particle in an electromagnetic field

We have discussed before that the Lagrangian for a charged particle


in an EM field is given by
1 2
L= m v −q Φ+q ⃗ A⋅⃗v
2
1 2 2 2
In Cartesian coordinates: L= m( ẋ + ẏ + ż )−q Φ+q( A x ẋ+ A y ẏ+ A z ż)
2

The scalar and vector potentials are generally functions of (t, x, y, z),
but independent of the velocity.
Generalized momentum:
∂L (similar expressions for py and pz)
px≡ =m ẋ+q A x
∂ ẋ

In vector form: ⃗p =m ⃗v +q ⃗
A

In this case, the generalized (canonical) momentum does not simply


equal to the ordinary mechanical momentum m ⃗v .
6
Symmetries and Conservation Laws
We shall study the connection between symmetry and conservation.
Symmetries: The invariance of a system under transformations
Conservation laws: The existence of constants of motion
Consider a particle moving in a central force field on a x-y plane. The
system has rotational symmetry about the z-axis (out of page). That is,
the physics is the same whether the particle
is at P or P', both positions have the same
distance R away from O.

The system is invariant when the particle


is rotated from P to P'.
[Note: This is true because the potential
energy V is a function of r only]
7
Side note: Rotation matrix
We want to connect the coordinates of P and P’
x '=R cos(φ+θ)
=R cos φ cos θ−R sin φ sin θ
=x cos θ− y sin θ

y '=R sin(φ+θ)
=R sin φ cos θ+R cos φ sin θ
= y cos θ+ x sin θ

The coordinates of the two points are connected by

x ' = cos θ −sin θ x


( )(
y' sin θ cos θ y )( )
Rotation matrix

8
Let us consider an infinitesimal rotation: θ = δθ→0

Keeping to O(δθ) => x '=x− y δ θ ; y '=x δ θ+ y

Original Lagrangian when the particle is at P:


L≡ L(x , y , ẋ , ẏ , t)
After the rotation, the “transformed” Lagrangian becomes:
L '≡L(x ' , y ' , ẋ ' , ẏ ' , t)
=L(x− y δ θ, y+ x δ θ, ẋ− ẏ δ θ , ẏ + ẋ δ θ , t)

As δθ→0, we can Taylor expand L' :


∂L ∂L ∂L ∂L 2
L '=L+ (− y δ θ)+ (x δθ)+ (− ẏ δ θ)+ ( ẋ δ θ)+O(δ θ )
∂x ∂y ∂ ẋ ∂ ẏ
∂L ∂L ∂L ∂L
=L+ x
[(
∂y
−y
∂x
+ ẋ)(
∂ ẏ
− ẏ
∂ ẋ )]
δ θ+O(δ θ )
2

9
Invariance of the physical system under the transformation means
L '=L
∂L ∂L ∂L ∂L
As δθ is arbitrary, we must have
( x
∂y
−y
∂x
+ ẋ)(
∂ ẏ
− ẏ
∂ ẋ )
=0

Using the Lagrange’s equations, we obtain


d ∂L d ∂L ∂L ∂L
[ ( ) ( )] (
x
dt ∂ ẏ
−y
dt ∂ ẋ
+ ẋ
∂ ẏ
− ẏ
∂ ẋ )
=0

d ∂L ∂L z-component of the angular


=> ( x
dt ∂ ẏ
−y )
∂ ẋ
=0 momentum

d
(x p y − y p x )=0 => J z =x p y − y p x = Constant
dt

Rotational invariance of the system (i.e., the Lagrangian)


implies the conservation of angular momentum.
Note that we obtain this conclusion even without specifying the details
of the system! 10
Noether’s Theorem
The connection between rotational invariance and conservation of
angular momentum is in fact a result of a more general theorem due
to a mathematician Emmy Noether which we shall prove below.

Consider an infinitesimal continuous transformation of the generalized


coordinates defined by
q ' α=q α +ϵ K α (q1 ,... , q N )≡q α +δ q α (α=1,... , N )

where ε is a small parameter and each Kα can be a function of all


coordinates. Note that we have
q̇ ' α= q̇α +ϵ K̇ α ≡ q̇ α +δ q̇ α

dK α ∂ Kα
where K̇ α ≡ =∑ q̇ β
dt β ∂ qβ
11
If the Lagrangian of the system is invariant under the transformation, we
have
∂L ∂L
0=δ L=∑
α ∂ qα
δ qα +(∂ q̇ α
δ q̇ α )
L=L(q α , q̇ α ,t)

1st order change in L under


the transformation vanishes ∂L ∂L
=∑
α [ ∂ qα
(ϵ K α )+
∂ q̇ α ]
(ϵ K̇ α )

Using the Lagrange’s equation for the first term on the right hand side,
we obtain
d ∂L ∂L
ϵ∑
[( )
α dt ∂ q̇ α
K α+
∂ q̇α
K̇ α =0
]
d ∂L
ϵ
dt (∑α
K =0
∂ q̇ α α )
∂L
I (q α , q̇α )≡∑ K α =∑ pα K α = Constant
α ∂ q̇ α α

If L is invariant under a continuous transformation, there exists a


12
conserved quantity associated with that symmetry.
Remark:
It is not always possible to choose the best coordinates to pick out
ignorable coordinates and identify conserved quantities. But if a symmetry
in the Lagrangian exists, then Noether’s theorem guarantees that there
must exist a corresponding conserved quantity.

Example: Rotational symmetry


As discussed before, an infinitesimal rotational transformation is
given by x '=x− y δ θ ; y '=x δ θ+ y

Comparing to q ' α=qα +ϵ K α (q1 , ... , q N )


=> ϵ=δ θ ; K x =− y ; K y =x

I (q α , q̇ α )≡∑ p α K α
α

= px K x+ p y K y
=x p y − y p x
=J z z-component of the angular
momentum
13
Noether’s theorem => J z = Constant
Conservation of Energy
Consider a system described by a Lagrangian L(q1,…,qN, t). Let us take
the time derivative of L:
dL ∂ L dq α ∂ L d q̇α ∂ L
α ( )
dt
= ∑ +
∂ q α dt ∂ q̇α dt
+
∂t
d ∂L ∂ L d q̇ α ∂ L We have used the Lagrange’s

α [( ) ]
= ∑ dt ∂ q̇ α
q̇ α +
∂ q̇ α dt
+
∂t
equations to rewrite the first
term

d ∂L ∂L
=∑
dt
α
( q̇
∂ q̇ )+α
∂t α

=>
d
dt (∑
α
q̇ α
∂L
∂ q̇ α )
−L =−
∂L
∂t
14
Definition: Hamiltonian

∂L
H≡ ∑
α ∂ q̇α

q̇ α −L=
α
p q̇ −L
α α Recall:
Generalized momentum
∂L
pα ≡
dH ∂L ∂ q̇ α
=> =−
dt ∂t

If L does not depend explicitly on time t (i.e., ∂L/∂t = 0), but only implicitly
through the time variations of generalized coordinates and velocities, then
the Hamiltonian function H is a constant of motion (dH/dt = 0).

Remark:
Here we simply consider H as a function of N independent variables qα
and their time derivatives (and possibly with time t). More formally, H
should be treated as a function of 2N independent variables (qα, pα) and
time t. We will come back to this issue when we discuss the formulation
of Hamiltonian dynamics. 15
Question: What is the physical meaning of H?

For those of you who have heard about the Hamiltonian function H, you
might have a naive understanding that H stands for the total energy of
the system. But here we are going to show that this is not necessarily
true! We shall discuss under what conditions does H stand for the total
energy.

16
Side note: Kinetic energy of a system of particles
Before studying the Hamiltonian in more details, we need to express
the total kinetic energy in a form that will be useful to the discussion.
Np
Total kinetic energy: 1 2
T= ∑ mi v i
i =1 2

For a holonomic system, we have ⃗r i =⃗r i (q1 ,... , q N ,t)


N
∂ ⃗r i ∂ ⃗r i
=> ⃗v i ≡⃗r˙ i = ∑ q̇ α +
α=1 ∂ q α ∂t

N N
r r ∂ ⃗r i ∂ ⃗r i
2 ˙ ˙
(
v i =⃗r i⋅⃗r i = ∑
∂ ⃗ i

α=1 ∂ q α
q̇ α +
∂ ⃗ i
⋅ ∑
)(
∂ t β=1 ∂ q β
q̇β +
∂t )
∂ ⃗r i ∂ ⃗r i ∂ ⃗r i ∂ ⃗r i ∂ ⃗r i ∂ ⃗r i
=∑ ∑ ⋅ q̇ α q̇ β +2 ∑ ⋅ q̇ β + ⋅
α β ∂ qα ∂ qβ β ∂ q β ∂t ∂t ∂t

17
The kinetic energy can be divided into 3 contributions:
Np
1 2
T= ∑ 2
mi v i =T 0 +T 1 +T 2
i=1

Np
2
where 1 ri
∂⃗
T 0≡ ∑
i=1
2
mi
∂t ( ) T 0 : independent of q̇α

N Np

T 1≡ ∑ ∑
β=1
( i=1
mi
∂⃗ri ∂⃗

ri

∂ qβ ∂ t β ) T 1 : linear in q̇α

N Np

T 2≡
1
2
∑ ∑
α ,β=1
( i=1
mi
∂⃗ri ∂ ⃗

ri
)q̇ q̇
∂ q α ∂ qβ α β
T 2 : quadratic in q̇ α

Note: If the constraints do not depend on time t explicitly, then the


transformation ⃗r i → q α do not depend on t explicitly and ∂ ⃗r i /∂ t =0 :
T 0 =T 1=0
18
In this case, T is always quadratic in q̇α (T =T 2).
N
∂L
Recall: Hamiltonian H≡ ∑ ∂ q̇ λ
q̇ λ −L
λ=1

Let us assume the following conditions:


(1) The potential energy V is velocity independent:
V =V (q α , t) Explicit time dependence is allowed

(2) The transformations ⃗r i =⃗r i (q 1 ,... , q N ) do not involve time t explicitly:


=> T =T 2 (qα , q̇ α) No explicit time dependence

∂L
=> = ∂ (T 2−V )= ∂ (T 2)
∂ q̇ λ ∂ q̇ λ ∂ q̇ λ
Np

[ ( ]
N
1 ∂ ⃗r i ∂ ⃗r i
= ∂ ∑ ∑ m ⋅
∂ q̇ λ 2 α ,β=1 i=1 i ∂ q α ∂ qβ
q̇α q̇ β
) Kronecker delta:

δα β ≡
1 if α=β
{
0 if α≠β
N Np
∂ ⃗r i ∂ ⃗r i
1
= ∑ ∑ mi ( ⋅ )
( δ q̇β + q̇α δ λ β)
2 α ,β=1 i=1 ∂ q α ∂ q β λ α 19
N Np N Np
∂ ⃗r i ∂ ⃗r i ∂ ⃗r i ∂ ⃗r i
=>
∂L 1
=
∂ q̇ λ 2
∑∑
β=1
(
Np
i=1
mi ⋅
∂ q λ ∂ qβ )
q̇ β +
1
∑∑
2 α=1 ( i=1
mi ⋅
∂ qα ∂ qλ )q̇ α

N
∂ ⃗r i ∂ ⃗r i
= ∑∑
α=1
( i=1
mi ⋅
∂ qα ∂ q λ )
q̇ α

The Hamiltonian becomes


N
∂L
H≡ ∑ ∂ q̇ λ
q̇ λ −L
λ=1
N Np

= ∑ ∑
λ , α=1
( i=1
mi ⋅
∂ qα ∂ qλ )
∂ ⃗r i ∂ ⃗r i
q̇α q̇ λ −(T 2−V )

=2 T 2 −(T 2−V )
=T 2 +V =T +V (T =T 2 under our assumption )

20
Conclusion:
If the potential energy V(qα, t) does not depend on q̇α and the
transformations ⃗r i =⃗r i (qα ) do not involve time t explicitly, then the
Hamiltonian function is equal to the total energy:
H =T +V
Furthermore, if V does not depend on time t explicitly, then L also does
not involve t explicitly, and hence ∂L/∂t = 0. In this case, we have

dH ∂L
=− =0
dt ∂t
=> Total energy is conserved

Caution:
The conditions for the conservation of H are in principle different from
those that identify H as the total energy.
It is possible that H is conserved, but not equal to the total energy. We
may also have situations where H = T+V is the total energy, but is not
conserved. 21

You might also like