Anda di halaman 1dari 53

1

ADVANCED QUANTUM MECHANICS AND


INTRODUCTION TO GROUP THEORY
(PHYS5000) LECTURE NOTES
Lecture notes based on a course given by Roman Koniuk.
The course begins with a discussion on advanced quantum mechanics and then
moves to group theory, Hydrogen, and the Dirac equation
York University, 2012

Presented by:
LATEXNotes by:

ROMAN KONIUK
JEFF ASAF DROR

2012

YORK UNIVERSITY

2
CONTENTS

I. Foundations of Quantum Mechanics


A. Vector Spaces, Dual Spaces, and Scalar Products
1. Vector Spaces
2. Dual Spaces
3. Scalar Products
B. Tensor Product(Outer Product)
C. Operators
D. Matrix Representations
E. Projection Operators
F. Hermitian and Unitary Operators
G. Continuum Basis
H. Postulates of Quantum Mechanics

2
2
2
3
3
3
4
4
4
5
7
7

II. Compatibility, Incompatibility, and Uncertainty

III. Pure States and Mixtures: Density Matrix

IV. Group Theory for Quantum Mechanics


A. Fundamentals
B. Lie Groups
C. Lie Algebra
D. SU (3)

14
14
20
22
26

V. Accidental Degeneracies

27

VI. Dirac Equation


A. Non-relativistic limit of the Dirac equation
B. Covariance of the Dirac equation

31
41
47

VII. Hydrogen

50
I.

FOUNDATIONS OF QUANTUM MECHANICS

A.

Vector Spaces, Dual Spaces, and Scalar Products


1.

Vector Spaces

In this class we will use the Bra and Ket spaces. Ket vector spaces obey the following:
|ai + |bi V
|ai + |0i = |ai
|ai + |ai = |0i
|ai + (|bi + |ci) = (|ai + |bi) + |ci
|ai + |bi = |bi + |ai
|ai V
( |ai) = () |ai

(I.1)
(I.2)
(I.3)
(I.4)
(I.5)
(I.6)
(I.7)

Vectors in this space are linearly independent. In other words


n
X

ai |ii = 0

(I.8)

i=1

ai = 0
Lecture 2 - Jan 6, 2012

(I.9)

3
2.

Dual Spaces

The adjoint of a vector |ai in the vector space defines a bra ha|. Such that if
|ci = |ai + |bi

(I.10)

hc| = ha| + hb|

(I.11)

then

3.

Scalar Products

For any vectors |ai , |bi V we denote ha|bi is the scalar product bra-ket. The inner product is in general a
complex number. i.e.
ha|bi C

(I.12)

Inner products obey the relation

ha|bi = (hb|ai)

(I.13)

The inner product of a vector with itself is a positive definite quantity:


ha|ai 0

(I.14)

ha|ai = 0 |ai = |0i

(I.15)

Further

The Schwartz inequality says that


2

|ha|bi| ha|ai hb|bi

(I.16)

The space of ket vectors and the dual space of bra vectors form a hilbert space. Hilbert space is a complete space.
We define the norm of a vector as
p
(I.17)
|a| = ha|ai
The set of vectors
|1i , |2i , ..., |ni

(I.18)

Form an orthonormal basis.

B.

Tensor Product(Outer Product)



Let a1 represent the state of particle 1 and let a2 represent the state of particle 2. Then the two-particle system
is represented by the tensor product
1 2 1 2 1 2
a a = a a = a a
(I.19)


Any operator 1 only operates on a1 and any operator 2 only operates on a2 . The commutator of operators
that act on different particles must be zero:
 1 2
, = 0
(I.20)

4
C.

Operators

Let
|ai = |a0 i

(I.21)

|ai = |ai

(I.22)

( |ai + |bi) = |ai + |bi

(I.23)

If is a linear operator then

where C and

The product of two operators simply means to carry out the operators in sequence:
|ai = ( |ai)

(I.24)

1 = I = 1

(I.25)

The inverse of an operator is denoted by 1 :

D.

Matrix Representations

Consider |ai and the operator such that


|ai = |bi

(I.26)

the matrix form of has the elements ij = hi| |ji. If we consider the expansions
X
|ai =
ai |ii

(I.27)

|bi =

bj |ji

(I.28)

(I.29)

If |ai is given then you can find ai since


hi|ai = hi|

|ni

= ai
bj =

(I.30)
ai j,i

(I.31)

The matrix elements of the identity operator are


I = hi| I |ji = i,j
E.

(I.32)

Projection Operators

Consider the set of basis vectors |1i , |2, i , ..., |ni. The projection operator of any vector onto the ith basis vector is
given by
Pi = |ii hi|

(I.33)

Pi |ai = hi|ai |ii


= ai |ii

(I.34)
(I.35)

This is clear since

5
Now consider the following
|ai =

ai |ii

(I.36)

hi|ai |ii

(I.37)

|ii hi|ai

(I.38)

Pi |ai

(I.39)

(I.40)

X
i

X
i

X
i

Hence
X

Pi = I =

|ii hi|

The projection operator squared is simply the projection operator since


P 2 = |ii hi| (|ii hi|)
= |ii hi|
=P

(I.41)
(I.42)
(I.43)

hj |Pi | ki = hj |ii hi| ki


= j,i i,k

(I.44)
(I.45)
(I.46)

The matrix representation of Pi is:

This looks like

0 0
..
0 .

0 0

0 0
0 0

0 0 0

0 0 0

1 0 0

.
0 .. 0
0 0 0

(I.47)

The identity is given by


I=

Pi

(I.48)

F.

Hermitian and Unitary Operators

Hermitian operators obey


=

(I.49)

In other words operators for which the operator is equal to the hermitian conjugate.
Unitary operators are operators for which
U = U 1

(I.50)

In quantum mechanics observables are associated with Hermitian operators. For examples
xx

(I.51)

6
~
i x

px

(I.52)

The eigenvalues of Hermitian operators are all real (since they correspond to observables). The states corresponding
to distinct eigenvalues are orthogonal.
Unitary operators are used to transform to another basis (called unitary since they preserve the scalar product).
Proof:
hb|ai = hU b0 |U a0 i


= b0 U U a0

(I.53)

(I.55)

(I.54)

= hb |a i

Hence the scalar product is conserved. Other things that are preserved by unitary transformations are trace, determinant, and algebraic equations involving matrices and vectors.
Any operator in quantum mechanics doesnt take you out of the Hilbert space.
Let
A |ai = |a0 i

(I.56)

Then
U A |ai = U |a0 i

(I.57)

(I.58)

(I.59)
(I.60)

U AU U |ai = U |a i
0

A U |ai = U |a i

where A0 = U AU is called a unitary-similarity transformation.


Typically use a unitary-similarity transformation to diagonalize an operator. We most typically diagonalize the
Hamiltonian (then we can just read off the eigenvalues).
U U = UD

(The subscript D means diagonal)

(I.61)

] The way we solve for U is by solving the characteristic equation (also called the secular equation).
An important unitary operator is the time-evolution operator:
2

U (t) = eiHt/~ = 1

iHt 1 (iHt)
+
+ ...
~
2 ~2 2!

(I.62)

where H is the Hamiltonian. We use this operator to evolve states.


U |(0)i = |(t)i

(I.63)

If the Hamiltonian is diagonal in some basis then


U = eiHt/~

(I.64)
2

1i

Ht 1 (iHt)
+
~
2
~2

(I.65)

This is easy to calculate since for a diagonal matrix

1 0 0

H = 0 ... 0
0 0 m

(I.66)

n1 0 0

Hn = 0 . . . 0
0 0 nm

(I.67)

0
.
..

1

If we are acting on a basis vector =


0 then

..
.

0
t 1
U |i = 1 i
+
~
2
G.



2 t2
+ ...
i 2
~

(I.68)

Continuum Basis

So far we have dealt with the discrete basis. For a continuum basis, such as the position basis. A state of definite
position is given by
|x0 i

(I.69)

x |x0 i = x0 |x0 i

(I.70)

In the continuum basis we know that


Z
|xi hx|

I=

(I.71)

hx|x0 i = (x x0 )

(I.72)

The scalar product of a mix of discrete and continuous states is the wavefunction:
hx|ni = n (x)

(I.73)

|hx|ni| = P (x)
H.

(I.74)

Postulates of Quantum Mechanics

1. The state of a particle is represented by a vector is Hilbert space |i


2. Every observable corresponds to a Hermitian operator
3. Every measurement of the observable corresponding to the operator results only in an eigenvalue
4. The average value of an observable is calculated in quantum mechanics by
h| |i

(I.75)

and we call this the expectation value


5. The time evolution of a state is given by the Schrodinger equation:
i~

d
|(t)i = H |(t)i
dt

(I.76)

Lecture 4 - Jan. 11, 2012


c

z }|i {
In measurement |i may not be in an eigenstate of but one can expand |i = i |i i hi |i. The coefficients hi |i
give the amplitude of finding the system in the eigenstate |i i. The probability of being in this eigenstate is
P

P (i ) = |h|i|

(I.77)

8
Once a measurement is made the state is in that eigenstate (until the time evolution operator acts) collapse of the
wavefunction. We of course have
X
P (i ) = 1
(I.78)
i

For continuous eigenspectra. For the position spectrum:


Z
P (x) dx = 1

(I.79)

all x

By dimensions we know that P (x) is a probability density.


II.

COMPATIBILITY, INCOMPATIBILITY, AND UNCERTAINTY

1. Two variables are compatible if there corresponding operators commute. For example if
|, ] = 0

(II.1)

then and are compatible. In this case we can have simultaneous eigenfunctions. In other words
|i = i |i ;

|i = i |i

(II.2)

This implies exact knowledge of i and i .


2. Two variables are incompatible of their corresponding operators dont commute. In this case you cannot have
simultaneous eigenfunctions. Hence you cannot have exact knowledge of and simultaneously.
The uncertainty is defined as
D
E1/2
2
A = (hA hAii)

(II.3)
1/2

= h(A hAi) (A hAi)i


D
E1/2
2
= A2 2A hAi + hAi
1/2


2
= A2 hAi

(II.4)
(II.5)
(II.6)

This is a measure of how far on average A is away from the average. Noncommutivity is related to uncertainties by:
AB

1
|h[A, B]i|
2

(II.7)

As an example consider the uncertainty of x and p:


1
|h[x, p]i|
2 



1 ~
~


1+
x


2
i x
i
x
1
|hi~i|
2
~

xp

III.

(II.8)
(II.9)
(II.10)
(II.11)

PURE STATES AND MIXTURES: DENSITY MATRIX

Suppose you have an experiment. The beam is 50% spin up and 50% spin down. The state
1
( + )
2

(III.1)

9
is a pure state. It is a linear combination of states but we know the state exactly. Another possibility is literally 50%
of the electrons and 50% are in the down state. How do we differentiate these two cases? We could use
hSz i =

~
z
2

(III.2)

where


 
 

0 i
1 0
,
,
i 0
0 1

(III.3)

In the case of a pure state we calculate this the expectation value by





1
~
1


( + ) Sz ( + ) = (0)
4
2
2
=0

(III.4)

0 1
1 0

(III.5)

In the second case we use


1
1
h| Sz |i + h| Sz |i = 0
2
2
Next we do the same for Sx . For the first case:



1
1
1~
~


( + ) Sx ( + ) =
h2i =
22
2
2
2

(III.6)

(III.7)

For the second case


1
1
h| Sx |i + h| Sx |i = 0
2
2

(III.8)

Hence the two states are distinguishable since we can measure the spin in the x direction and see the difference.

Lecture 5 - Jan 13, 2011


Every state we have seen thus far has been a pure state. We say that the pure state is coherent while mixed states
are incoherent.
The density matrix is used to handle both pure and
P mixed states. For a mixed state we say the probability(not
amplitude) of being in state |i i is equal to i and i i = 1. Therefore the expectation value for a mixed state is
written as
X
hAi =
i hi | A |i i
(III.9)
i

Important point: there are two averaging procedures in this equation. One is the usual quantum mechanical average
procedure (hi | A |i i). e.g.
Z
hxi = (x)x(x)dx
(III.10)
Z
= P (x)xdx
(III.11)
The other is the classical averaging of multiplying the probability of being in a state by the value of being in that
state. Its important to realize that |i i need not be orthogonal but they must
P be normalized. For example you may
have the states 50% |i and 50% 12 (|i + |i). For a mixed state hAi = i i hi | A |i i can be written as follows
hAi = Tr (A) ;

where =

i |i i hi |

(III.12)

Recall that
Tr(B) =

X
j

hj| B |ji

(III.13)

10
Proof of the relation shown above:
Tr(A) =

hj| A |ji

(III.14)

(III.15)

hj|

XX
j

i |i i hi | A |ji

i hj|i i hi | A |ji

i hi | A

(III.16)

|ji hj|i i

(III.17)

i hi | A |i i

(III.18)

The operator is called the density matrix operator which is defined as


X
=
i |i i h|

(III.19)

Remember that you can write any state |i i as


|i i =

Cj |ji

(III.20)

where here j is labeling the basis state and i is labeling state state in the mixture. We can now expand the density
matrix as
X
=
i |i i hi |
(III.21)
i

XXX
i

i Cji Cki |ji hk|

(III.22)

j,k |ji hk|

(III.23)

j,k

where j,k = i i Cji Cki . Traces are basis-independent quantities (since Tr(ABC) = Tr(BCA)). In particular Tr(A)
is independent of basis. Next consider the trace of :
X X
Tr() =
hj|
i |i i hi |hi
(III.24)
j

XX
j

XX
i

XX
X

i Cj

i hj|i i hi |ji

(III.25)

Ck hk|ji

(III.26)

i Cj Cj

(III.27)

(III.28)

=1

(III.29)

If we have a pure state the density matrix is just the projection operator
X
=
i |i i hi |

(III.30)

= |i i hi |

(III.31)

2 =

(III.32)

Hence in for a pure state

11
For a mixed state
2 6=

(III.33)

Proof: We can go to basis which diagonalizes and then we have

1 0 0

= 0 2 0
..
.
0 0

(III.34)

but for a mixed state all the i < 1 and hence the square of them is less than 1. For example find for case the pure
state |x i = 12 (|z i + |z i). This can be done in several ways
=

i |i i hi |

(III.35)




1

1

( + )
= ( + )
2
2

(III.36)

1
(|i h| + |i h| + |i h| + |i h|)
2

(III.37)
(III.38)

j,k =

i Cji Cki

(III.39)

= Cj Ck
but C1 C1 = C1 C2 = C2 C1 = C2 C2 =

1
2

(III.40)

hence
1
2

j,k =

(III.41)

We use this result to find some expectation values:


hSz i = Tr(Sz )

1~
1
Tr
1
22

~
1
= Tr
1
4

(III.42)
1
1



1
1

1 0
0 1


(III.43)


(III.44)

=0

(III.45)

and
hSx i = Tr(Sx )

1~
1
Tr
1
22

~
1
= Tr
1
4
~
=
2
Assignment: to be handed in on Monday
1. Find for case B (50% and 50% )
2. Evaluate Tr(Sz ) and Tr {Sx }

(III.46)
1
1



1
1

0 1
1 0


(III.47)
(III.48)
(III.49)

12

Detectors

Detectors

3
1
R

3
2

1
G

FIG. 1. The EPR experiment

Lecture 6 - January 20th, 2012


Consider the EPR setup shown in figure 1. The experiment has two observations.
1. 100% of the time if the two detectors are pointing in the same directions then opposite colours flash. 100%
anti-correlation of the colours if the pointers are pointing in the same direction
2. If we pay no attnetion to the direction of the pointers then there is no correlation (completely random)
Classical understanding of the experiment says if we consider observation 1 then the particles know prior to the
experiment what they will choose. This can be though of as the particles having genes. This is shown in table I.
Consider the second row of table I. In total there are 9 possible pointer positions: 11, 12, 13, 23, ... (always true). For
TABLE I. The gene table of the EPR experiment
1 2 3 1 2 3
G G G R R R
G G R R R G
G R G R G R
R G G G R R
R G R G R G
G R R R G G
R R G G G R
R R R G G G

the second gene 11, 22, 33, 12, 21 gives us anti correlation (e.g. G R) while 13, 31, 32, 23 gives us perfect correlation
(e.g. G G). Looking at this gene we see that 95 of the time we get anti correlation. But there is nothing special
about this gene. The only important feature of this gene is that one of the colours is different. In other words wed
get anti-correlation 59 of the time for genes 2, 3, 4, 5, 6, 7. The conclusion is that for 6 out of the 8 genes we get 95 of
the time. For genes 1 and 8(the remaining two) we have perfect anti-correlation. The final conclusion is:
P (anti-correlation) >

5
9

(III.50)

i.e. Bells theorem says that classically there is anti-correlation greater then 59 of the time. The experiment violates
Bells theorem!
We now consider the quantum calculation. We consider the state of an entangled positron electron pair emitted by
a decaying 0 meson. This state is
1
= (|i |i)
2

(III.51)

Comments:
1. This is a pure state.
2.
Jz = 0

(III.52)

13
3.
J 2 = 0

(III.53)

4.
S1z S2z =

~2

(III.54)

5. Define the correlator:


C(O1 , O2 ) p

hO1 O2 i

(III.55)

hO12 i hO22 i

Note if O1 = O2 C = 1.
6. Its easy to see (or show) that we get perfect anti-correlation
C(S1z S2z ) = 1

(III.56)

C(S1x S2x ) = C(S1y S2y ) = 1

(III.57)

7. Further one can show that

The state we are considering is an example of an entangled state. The state cannot be written as a simple product.
This confirms observation 1 since the particles always have perfect anti-correlation. Next we need to show that
quantum mechanics can produce observation 2. Consider the explicit pointer directions shown in figure 2:

120 deg

FIG. 2. The EPR pointer directions

n1 = z

3
x

2
3
n3 =
x

2
n2 =

(III.58)
1
z
2
1
z
2

(III.59)
(III.60)

We measure the correlation of the spins dotted into one of the pointer directions:

0
0
X
X
X


h| S1 n
i S2 n
j |i = h| S1
n
i S2
n
j |i
ij

=0

Lecture 7, January 25th, 2012

(III.61)

(III.62)

14
IV.

GROUP THEORY FOR QUANTUM MECHANICS


A.

Fundamentals

The uses of studying group theory is that you can study the effects of symmetry. A symmetry is if we can perform
some transformation and the system doesnt change. For example if we have a triangle

Then there are 6 symmetry transformations. Three reflection symmetries, 2 rotations (120o and 240o ), as well as the
identity transformation.
Symmetries lead to conservation laws
h| A |bi ha0 | A0 |b0 i = ha| U 1 A0 U |bi

(IV.1)

ha| A |bi = ha0 | A0 |b0 i

(IV.2)

If

then we have an invariance. U 1 AU = A if [A, U ] = 0. For example consider the time translation operator,
U (t) = eiHt/~ . Operators that commute with U will not change with time. For example
[H, U ] = 0 Energy conservation

(IV.3)

Conservation of energy is a result of time translation invariance. Next consider the spatial transformation operator
T (x) = eixp/~

(IV.4)

Commutation with this operator implies momentum conservation. Invariance with respect to rotations leads to
angular momentum conservation. Invariance gauge transformation leads to charge conservation.
U (1) Charge conservation
SU (2) Weak charge conservation
SU (3) Colour conservation

(IV.5)
(IV.6)
(IV.7)

Lecture 8 - January 27th, 2012


Suppose R is the rotation operator and R leaves the Hamiltonian invariant. For example
R |`mi

(IV.8)

leaves all the m states invariant. For example all the states with a given ` invariant (2px , 2py , 2pz ) are degenerate.
The mathematics of symmetry is Group Theory. A group G is a set with a rule for assigning to every ordered pair
a third element satisfying the following:
1. Closure: If f, g G then h = f g G
2. Associativity: For f, g, h G,

f (gh) = (f g)h

3. Identity: There is an identity element e such that ef = f e = f (where f is every element in the set)

15
4. Inverse: Every element f G has an inverse which we call f 1 such that f f 1 = f 1 f = e
A group really is a multiplication table satisfying all these properties. The table specifies g1 g2 g1 , g2 G. If the
group elements are discrete, then we can write the multiplication table.
e g1
e e g1
g1 g1 g2 g1
.
g2 g2 ..
.. ..
..
. .
.

g2
g2
...
..
.

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

...

..

A representation of G is a mapping of the elements of G onto a set of linear operators with the following properties.
1. Let D be the representation
D(e) = I

(IV.9)

where I is the identity operators in the space where the linear operator act.
2.
D(g1 )D(g2 ) = D(g1 g2 )

(IV.10)

In other words the group multiplication law is mapped onto the natural multiplication in the linear space on
which the operators act.
When we write
D(g) |i

(IV.11)

We mean that D is the representation, g is the group element, and |i is the state.
Consider for example the group Z3 .
Def 1. A group is finite if it has a finite number of elements. Otherwise its infinite.
Def 2. The number of elements in a finite group is called the order of the group.
Z3 is of order 3:
e a b
e e a b
a a b e
b b e a
Note that Z3 is commutative (g1 g2 = g2 g1 ).
Def 3. Commutative groups are called Abelian groups
A representation of Z3 is
D(e) = 1
D(a) = e2i/3
D(b) = e4i/3
The dimension of the representation is the dimension of the space on which the representation acts. This representation
is two dimensional. This representation clearly represents the table since
D(a)D(b) = e2i/3 e4i/3 = e2i = 1
The rest of the relations are easy to show. Another representation for the same group

1 0 0
0 0 1
0

D(e) = 0 1 0 ;
D(a) = 1 0 0 ;
D(b) = 0
0 0 1
0 0 1
1

(IV.12)
is

1 0

0 1
0 0

This is a 3 dimensional representation. This representation has a special name. This is called the regular representation. The trick to construct the regular representation is as follows.

16
1. First we identify the orthonormal basis on which the presentation acts with group elements.
|e1 i |ei , |e2 i |ai , |e3 i |bi

(IV.13)

2. Now we can construct the representation with


[D(g)]i,j = hei | D(g) |ej i

(IV.14)

The dimension of any regular representation is equal the number of group elements.
As an example we check D(a):
he1 | D(a) |e1 i = he|ai
=0

(IV.15)
(IV.16)

he1 | D(a) |e2 i = he|bi


=0

(IV.17)
(IV.18)

he1 | D(a) |e3 i = 1

(IV.19)

Do the rest in bundles:


(
1 for e2
he2 | D(a) |e1,2,3 i =
0 for e1 and e2

(IV.20)

Assignments:
Construct the multiplication table for Z2
1. Find the 1 dimensional representation
2. Find the 2D representation
3. Is 2 the regular representation

Lecture 9 - January 30th, 2012


Recall the trick to find the regular representation. The reason it works is because the following
[D(g1 g2 )]i,j = [D(g1 )D(g2 )]
= hei | D(g1 )D(g2 ) |ej i
X
=
hei | D(g1 ) |ek i hek | D(g2 ) |ej i

(IV.21)
(IV.22)
(IV.23)

Di,k (g1 )Dk,j (g2 )

(IV.24)

We can transform the basis. A transformation on the states implies a transformation on the operators:
D(g) D0 (g) = S 1 D(g)S

(IV.25)

The transformed operators (D0 (g)) will have the same multiplication table so we say that D0 and D are equivalent
because they only differ by a trivial change of basis.
Def 4. A representation is reducible if it has an invariant subspace. This means that the action D(g) on any
vector in the subspace is still in the subspace.

17
Every group has a trivial representation, D(g) = 1. e.g. for Z2 ,
D(e) = 1;

D(a) = 1

(1D)

(IV.26)

One can have a two dimensional representation:


D(e) =

1 0
0 1

!
;

D(a) =

1 0
0 1

!
;

(2D)

(IV.27)

A third dimensional representation of Z2 is

1 0 0
1 0 0

(IV.28)
D(a) = 0 0 1
D(e) = 0 1 0 ;
0 1 0
0 0 1




c0
c
0
0


0

Any action onto a vector a a stays inside the subspace. The same goes for the vector 0 0 .
0
0
b0
b
A representation is irreducible if it is not reducible. A representation is completely reducible if it is equivalent
to a representation whose matrix elements have the following form:

D1 (g)
0
...

.
0
(IV.29)
D2 (g) ..

..
..
.
.
...

where Di (g) is irreducible. This is called block diagonal form. A representation in block diagonal form is said to
be the direct sum of the sub representations Dj (g):
D1 D2 ...
e.g. Take our 3-D representation to Z3 and apply the similarity transformation

1 1 1
1

S = 1 w2 w
3
1 w w2
where w e2i/3 . With this process

0
D (e) = 0
0

we find

1 0 0
1 0 0
0 0

1 0 ; D0 (a) = 0 w 0 ; D0 (b) = 0 w2 0
0 0 w
0 1
0 0 w2

(IV.30)

(IV.31)

(IV.32)

If we act on any of the three vectors



0
a
0

;
;
0
b
0
0
0
c

(IV.33)

The vectors will stay in the subspace hence this is an irreducible representation.
In Quantum Mechanics the D(g) are unitary transformations. They map the Hilbert space to an equivalent one.
They reflect the symmetry of the problem if
[D(g), H] = 0

(IV.34)

This means we can always choose the energy eigenstates to transform like irreducible representations of the group.
For example: Y00 doesnt rotate so it transforms under the trivial representation. Y10 transform according to a 3
dimensional representation (since |1, mi can only transform to |1, m0 i). Furthermore the Y2m transforms according to

18
a 5 dimensional representation.
Consider the parity (reflection in a mirror) transformation, P
P2 = e

(IV.35)

Hence
e P
e e P
P P e
There is a trivial reprentation for this group, D(g) = 1 and a non-trivial representation, D(e) = 1 and D(P). If
[D(g), H] = 0

(IV.36)

e.g. 1D parity, x x. Consider the harmonic oscillator potential. This does not means that all the eigenfunctions
are even. However it does mean that the even wavefunctions will transform trivially (they are already even) and the
odd functions will transform according to the non-trivial representation.
Note if two groups have a different physical origin, e.g. Parity and rotations by but have the same multiplication
table we say the two groups are isomorphic.
Assignment (not to hand in): Construct the multiplication table for S3 (the group of permutations of 3 objects).
For notational purposes call the elements, e, a, b, x, y, z

Lecture 10th - February 1st, 2012


Consider the regular representation of Z2
D(e) =

1 0
0 1

!
;

0 1
1 0

D(a) =

!
(IV.37)

We can think of this D(a) as reflection in the y = x line (thinking of it as a 2D space).


This suggests the similarity transformation that will reduce this reducible representation
What about rotating by 45%? Try
1
S=
2

D(a) D (a) = S

1 1
1 1

!
(IV.38)

D(a)S =

1 0
0 1

!
(IV.39)

and
D(a) =

1 0
0 1

!
(IV.40)

Consider the group S3 . The elements are listed below. The notation is as follows the element (3, 1, 2) is element 1
goes to position 2, element 2 goes to position 3 and element 3 goes to position 1.
e = (1, 2, 3)
a = (2, 3, 1)
b = (3, 1, 2)
x = (1, 3, 2)
y = (3, 2, 1)
z = (2, 1, 3)

19
Can think of these as the symmetry group of the equilateral triangle.
e Identity
a, b Rotation
x, y, z Reflection
The multiplication table of this group is as follows

e
b
x
y
z

e
a
b
x
y
z

a
b
e
y
z
x

b
e
a
z
x
y

x
z
y
e
b
a

y
x
z
a
e
b

z
y
x
b
a
e

We lost commutativity and hence this is a non-Abelian group. Notice that the top left corner of the table is closed
within itself. Hence we say that {e, a, b} forms a subgroup. In fact this is the same group as Z3 . Therefore this
subgroup is isomorphic with Z3 . A two-dimensional representation of S3 is
!
!
!
3
3
1
1
1 0

2
2
2
2
D(e) =
;
D(a) =
;
D(b) =
3
3
1
0 1
2 12
2
2

D(z) =

1 0
0 1

!
;

D(x) =

3
2
12

1
2
3
2

!
;

D(y) =

1
2

3
2

23
12

Since this is irreducible there is not similarity transformation that will diagonalize all the matrices in the representation.
It is necessary that at least some of the irreps (irreducible representations) are matrices so that the non-commutativity
(non-Abelian) can hold (numbers always commute!). Heres a 3D rep (reducible representation) of S3

0 1 0
0 0 1
1 0 0

D(b) = 0 0 1
D(a) = 1 0 0 ;
D(e) = 0 1 0 ;
1 0 0
0 1 0
0 0 1

0 1 0

D(z) = 1 0 0 ;
0 0 1

1 0 0

D(x) = 0 0 1 ;
0 1 0

0 0 1

D(y) = 0 1 0
1 0 0

This particular represention (NOT the regular representation) is important because it is the defining representation
for the group (we call this the fundamental representation). It actually implements the permutation on the states.
Note that a subscript on a representation given as
Dn

(IV.41)

means the dimensionality of the representation. However a subscript in denoting the group:
Sm

(IV.42)

means the number of objects the group is acting on. Consider the following where |1i denotes the states which are
the objects
D3 (a) |1i =

3
X

|ki [D3 (a)]k,1 = |2i

(IV.43)

k=1

D3 (a) |2i = |3i =

X
k

|ki [D3 (a)]k,2

(IV.44)

20
D3 (a) |3i = |1i =

|ki [D3 (a)]k,3

(IV.45)

This 3D representation decomposes into a direct sum of the irreducible representations

(
The regular representation of Z2

1 0
0 1

!
,

D3 = D1 D2
!)
0 1
is reducible as
1 0

(IV.46)

D2 = D1 D1

(IV.47)

Consider the following theorems


All of the irreducible representations of a finite Abelian group are 1D. e.g. Z3 , 3D 1 1 1.
If a Hermitian operator H, commutes with all the elements D(g) of a representation of a group G, then you can
choose the eigenstates of H to transform according to irreducible representations of G.
If an irrep appears only once in the Hilbert space every state in the irrep is an eigenstate of with the same
eigenvalue.

Lecture 11th, February 2nd, 2012


We have finished finite (point) groups. One can consider having shapes with more and more sides. Start with
a triangle then square... all the way to a circle. A circle is invariant under any rotation. This is an example of a
continuous group.
A rotation in 2-d can be represented by
!
cos sin
(IV.48)
sin cos
Because the parameter is continuous, the group is said to be continuous. We need to check that this is a group.
This transformation clearly has an identity. Proof that its closed:
!
!
!
cos sin
cos sin
cos cos sin sin cos sin + sin cos
=
(IV.49)
sin cos
sin cos
(cos sin + sin cos ) sin sin + cos cos
!
cos ( + ) sin ( + )
(IV.50)
=
sin ( + ) cos( + )
This group clearly also has an inverse (if = ) you get the identity. This group is clearly Abelian (if we rotate one
way or another way the order doesnt matter). This group is called SO(2). S denotes special, O denotes orthogonal,
and 2 denotes 2D.
Consider multiplication by the complex numbers, ei . There is clearly an identity ( = 0). The other properties are
ei ei = ei(+)
ei ei = 1
This group rotates numbers in the complex plane. This group is called U (1) where U denotes unitary (U 1 = U )
and 1 denotes 1D. U (1) and SO(2) are clearly isomorphic (see figure 3). They have the same multiplication table,
parameterized by 1 real number.

B.

Lie Groups

Consider rotations in 3D, we could characterize these rotations with 3 parameters, e.g. {x , y , z } where each
define a rotation about x, y, z respectively. Alternatively we can use the 3 Euler angles. The group members of

21

SO(2)
Im

U(1)
Im

Re

Re

FIG. 3. SO(2) and U (1) are isomorphic

rotations in 3 dimensions can be written as the set {g()} where is a vector of 3 parameters. If the product
g(00 ) = g()g(0 ) then the group is a Lie group if 00 = (, 0 ) where is an analytic function. The group
elements depend smoothly on the set of parameters. By smooth we mean there is some notion of closeness on the
group such that if two-elements of a group are close together in the space of the group the parameters that describe
the elements are close. Another way to say this in a fancy-shmancy way is to say is that the group is a manifold. A
manifold looks like Rn locally. Thus in the neighborhood of the identity the group elements can be described by a
function of n real parameters. We now parameterize the group by
a ,

a = 1, 2, ...n



g()

=e

(IV.51)
(IV.52)

=0

Then if we find a representation we will parameterize the representation in a way such that


D()
=I

(IV.53)

=0

In some neighborhood of the identity we can Taylor expand (since its analytic) D()
D () = I + ia Xa
here we are using Einstein summation convention. The a are infinitesimal and the

Xa = i
D(a)
a
=0

(IV.54)

(IV.55)

The Xa are matrices for a = 1, 2...n (as many of them as the dimension of the group). These are called generators
of the group.

On the first test, the first question will be to show that


H = E

(IV.56)

Recall our previous discussion. The D() form a group so we can multiply two elements
(I + ia Xa ) (I + b Xb )
In particular we do this. We can write a =

(IV.57)

a
k

where k is large. Consider the following group element



a k
lim I + i Xa = eia Xa
k
k

(IV.58)

Note the summation in the exponents. This is equal to the exponential by a definition of the exponential (not the
Taylor expansion). Hence any finite representation can be written as an exponential.

22
C.

Lie Algebra

Now in any particular direction the group multiplication is uncomplicated because we can parametrize our element
with
U () = eia Xa

(IV.59)

U (1 )U (2 ) = U (1 + 2 ) = ei(1 +2 )a Xa

(IV.60)

However it is complicated if we go into different directions.


eia Xa eib Xb 6= ei(a +a )Xa

(IV.61)

but we have closure in the group and therefore


eia Xa eib Xb = eia Xa

(IV.62)

In other words the product is equal to some group element. We find that to next to lowest order, i.e. expanding both
sides of the equation above to next to lowest order
1
c = c + c a b fabc
2

(IV.63)

[Xa , Xb ] = ifabc Xc

(IV.64)

where

where the fabc are called the structure constants of the group and [Xa , Xb ] = ifabc Xc is called the Lie Algebra for
he Lie Group. For unitary transformations the fabc are real.
The worry is that this is true to next to lowest order but well need more to characterize group multiplication at
each order. The remarkable thing is that the fabc completely characterizes the group. The matrix generators satisfy
the Jacobi identity
[Xa , [Xb , Xc ]] = 0

(IV.65)

This is true for all cyclic permutations as well. The adjoint representation of an algebra is given by
[Ta ]bc = ifabc

(IV.66)

i.e. the bcth element of the ath generator matrix is equal to ifabc .
The unitary group of order n, U (n) is the group associated with n n unitary matrices. The matrices operate on
n d complex vectors.
U U = I

(IV.67)

To parameterize a complex n n matrix we require 2n2 numbers. The unitarity condition requires n2 constraints on
the matrix and therefore we are left with n2 free parameters. Therefore there are n2 generators.
SU (n) is a subclass of U (n). The S stands for special which means that the determinant of U is +1. In other words

det U = ei
(IV.68)
Because U U = I it implies that

I iX (I + iX) = I

i X X = 0
X=X

(IV.69)
(IV.70)
(IV.71)

In other words the generators are Hermitian. Now we use a general property of matrices:
det A = exp Tr (ln A)

(IV.72)

23
Consider
det (I + ) = 1 + Tr

(IV.73)

TrX = 0

(IV.74)

If we use that det U = 1 we have that

Thus the generators in SU (n) are traceless and Hermitian.


Lecture 13 - February 8, 2012
U (n) is a unitary group. If we have SU (n) then |det U | = 1. U (n) has n2 generators while SU (n) has n2 1.
U (n) and SU (n) have subgroups O(n) and SO(n). These are just the real versions of U (n) and SU (n). These
transformations act on real vectors and are said to correspond to the isometries in n-D (this means that they dont
change the shape of objects). The n2 parameters of U (n) go down to n(n1)
for O(n) (O stands for the orthogonal
2
group). If
|det U | = 1

(IV.75)

for O(n) then we have det U = 1 and this group is called SO(n). The unitarity condition for the orthogonal group
is just
OT O = OOT = 1

(IV.76)

U (n) = SU (n) U (1)

(IV.77)

Note that we often see this notation

It means that U (n) is made up of SU (n) and U (1).


Consider some particular groups.
U (1) : eiQ

(IV.78)

where Q is the charge operator. If we construct a Hamiltonian invariant under this U (1) then it implies that we have
charge conservation. We can generalize the above in the following way
U (1) : ei(x)Q

(IV.79)

The parameter controlling the phase change depends on x. In other words it depends on position in space. Hamiltonians can be constructed that are invariant under U (1) = ei(x)Q . This is called a local or gauge symmetry. In
order to construct such a Hamiltonian we require a gauge field (which predicts the existence of the photon). History
didnt work this way for electromagnetism (the gauge symmetry was found after). Physicists later constructed gauge
Hamiltonians and group was used to predict W + , W , Z, and g.
Consider the SU (2) group. It acts on 2 D complex vectors. The defining representation is 2 D. In other words
two by two matrices. There are n2 1 generators and since n = 2 we have 3 generators. There are three directions
you can go on the group manifold. Recall that the generators are traceless and hermitian. This can be encoded by
writing a general representation as
!
a b
(IV.80)
b a
There are three parameters here: a, Re(b), Im(b). A non-unique but suitable representation are the set of generators
i
Ji =
(IV.81)
2
where
!
!
!
0 1
0 i
1 0
1 =
;
2 =
;
3 =
(IV.82)
1 0
i 0
0 1
and we know the structure constants since
[Ji , Jj ] = iijk Jk
where ijk is the Levi-Cevita symbol. Here we have ijk = fijk .

(IV.83)

24
Def 5. A Casimir operator is an operator which commutes with all of the generators.
For SU (2) we need a Casimir such that
[ , J1 ] = [ , J2 ] = [ , J3 ] = 0

(IV.84)

One such operator is J 2 . The importance of a Casimir operator is that we can construct simultaneous eigenstates of
J 2 and Ji . Its easy to show that (try at home) that
[J3 , J ] = J

(IV.85)

where J = J1 J2 . We define eigenstates |j, mi such that


J 2 |j, mi = j (j + 1) |j, mi

(IV.86)

J3 |j, mi = m |j, mi

(IV.87)

and

Recall from quantum mechanics that J+ and J generate different states within a multiplet. In other words
J3 J |j, mi = (m 1) J |j, mi

(IV.88)

J 2 J |j, mi = j (j + 1) J |j, mi

(IV.89)

Furthermore

and
J |j, mi =

(j m)(j m + 1) |j, m 1i

We use these results to form a two dimensional representation of SU (2) by choosing


!


1 1
1
,
2 2 = 0
!


1 1
0
,
2 2 = 1

(IV.90)

(IV.91)

(IV.92)

We can now find the matrix elements of J3 :




1 1 1 1
1
, J3 ,
=
2 2 2 2
2



1 1 1 1
, J3 ,
=0
2 2 2 2



1 1 1 1
, J3 ,
=0
2 2 2 2



1 1 1 1
1
, Je ,
=
2 2 2 2
2


(IV.93)
(IV.94)
(IV.95)
(IV.96)

Hence
1
J3 =
2

1 0
0 1

Now we find the matrix elements of J+ . Define and as usual. We use the following
s

1 1
J+ |i =
+
(1) = 1
2 2

(IV.97)

(IV.98)

25
h| J+ |i = 0
h| J+ |i = 1
h| J+ |i = 0
h| J+ |i = 0

(IV.99)
(IV.100)
(IV.101)
(IV.102)

J+ =

0 1
0 0

J =

0 0
1 0

(IV.103)

For J = J+
we get

(IV.104)

Its now easy to find J1 and J2 using J = J1 iJ2 :


1
J1 =
2
So we recover Ji =

0 1
1 0

1
J2 =
2

0 i
i 0

!
(IV.105)

i
2 .

Lecture 14, February 10th 2012


We can construct a 2-D representation of SU (2). The SU (2) generators are
are

i
2

and therefore the SU (2) group elements

U = eii i /2

(IV.106)

(here we use summation notation), where i are continuous parameters. Note that in U = eii i /2 is implicit but
it turns out that in SU (2) we can construct the explicit matrix. For convenience we write i 2ni where ni is a
normal vector (determines how much of i is in the direction n
i . With this change in notation we write
U = ein = I + i
n

2 (
n )
+ ...
2

(IV.107)

To simplify this quantity we work out the following


2

(
n ) = (n1 1 + n2 2 + n3 3 )
n21

=
+
=I

n22

n23

(IV.108)

+ n1 n2 1 2 + n1 n3 1 3 + n2 n3 2 3 + n1 n2 2 1 + n1 n3 3 1 + n2 n3 3 2

(IV.109)
(IV.110)

where we used {i , j } , i 6= j. We can now simplify U :


2
n

i 3
+ ...
U = I + i
n
2
2



2
4
3
5
= 1
+
... + i
+
+ ... n

2
4!
3!
5!
= cos + i sin (
n )
 
 

= cos
+ i
n sin
2
2

(IV.111)
(IV.112)
(IV.113)
(IV.114)

Now we work out the matrix


!
0 i
n
= n1
+ n2
+ n3
i 0
!
n3
n1 in2
n3
=
=
n1 + in2
n3
n+
0 1
1 0

!
1 0
0 1
!
n
n3

(IV.115)

(IV.116)

Thus we have explicitly:


cos

+ n3 i sin

in+ sin 2


in sin 2

cos 2 in3 sin

(IV.117)

26
D.

SU (3)

SU (3) is the group of Special unitary transformations acting on 3D complex vectors. For SU (n) there are n2 1
generators and therefore for SU (2) there are 8 generators:
{X1 , X2 , ...X8 }
One possible choice for these generators Xi =

0 1 0

1 = 1 0 0 ;
0 0 0

0 0 1

4 = 0 0 0 ;
1 0 0

0 0 0

7 = 0 0 i ;
0 1 0

(IV.118)

i
2

where the i are the Gell-man matrices:

1 0 0
0 i 0

3 = 0 1 0
2 = i 0 0 ;
0 0 0
0 0 0

0 0 0
0 0 i

6 = 0 0 1
5 = 0 0 0 ;
0 1 0
i 0 0

1 0 0
1

8 = 0 1 0
3
0 0 2

These matrices obey


Tr (i j ) = 2ij

(IV.119)

[Xi , Xj ] = fijk xk

(IV.120)

They obey the Lie algebra relation

However the fijk are not simple as in SU (2).


Nature has an approximate SU (3) global symmetry, flavour:

u

d
s
but nature also has an exact SU (3) local (gauge) symmetry, colour:

q1
2
q
q3

(IV.121)

(IV.122)

Lecture 15, February 13th, 2012


Recall that for SU (2):
cos

+ i (
n ) sin
2
2

(IV.123)

The algebra is defined by


h

j i
k
= iijk
2 2
2
i

(IV.124)

The defining representation is 2D. Consider SO(3):


[Li , Lj ] = iijk Lk

(IV.125)

and the defining representation is 3D. SU (2) and SO(3) have the same Lie Algebra and they are locally isomorphic.
For this to be true the group manifold must have the same dimension (3D). For SU (2)
=0I
= 2 I

27
If you have some angle then and + 2 gives minus the original transformation. This doesnt occur in SO(3) (we
know that since we are familiar with orbital angular momentum).
In fancy shamncy words we say: The groups are locally isomorphic but the global topology of the two groups is
different. In group theory language we say
SO(3)

V.

SU (2)
Z2

(IV.126)

ACCIDENTAL DEGENERACIES

Accidental degeneracies mean there are degeneracies in the spectrum in a QM system that are unexpected. However
there is no such thing as a true accident so this is just a misnomer. We will consider Hydrogen. Consider the
Hamiltonian:
H=

p2
e2

2m
r

(V.1)

We see that this is rotationally invariant (nothing in this problem is picking out a direction). I.e.
[SO(3)(g), H] = 0

(V.2)

where SO(3)(g) is a rotation matrix. Therefore the eigenstates of H transform according to irreducible representations
of SO(3). Also note invariance is O(3) which includes Parity transformations (det = 1). Therefore eigenstates of
the Hamiltonian will also be irreducible representations of Parity. The 1s forms a 1D irrep of SO(3), 2P is a 3D irrep
of SO(3), 3d forms a 5D irrep of SO(3). This again is sloppy language the better way to say this is to say that the
states transform among each other under SO(3) if states transform among each other they are transformed by an
irrep (i.e. block matrix). These are the expected degeneracies.
However in practice we see that the 2s is also degenerate which 2p and 3s, 3p are all degenerate with 3d, etc. These
degeneracies are called accidental. There is way more degeneracy then we expect. There must be more symmetry
then we naively think. Since symmetries produce conservation laws there must be some conservation law we are not
seeing. Before we look at quantum mechanics can we see something in the classical problem? The classical problem
is the Kepler problem where
E=

p2
GmM

2m
r

(V.3)

When we solve this problem we find the orbits are elliptical. The orbitals are shown in figure 4. We also notice that

Ellipses

FIG. 4. The elliptical orbit of the Kepler problem

the angular momentum is conserved and hence the orbit is planar. We summarize the interesting points below:
1. Orbits are ellipses
2. Angular momentum in conserved
3. Orbit is planar

28
4. Orbit closes (no precession)
The major axis points in the same direction are all times and has the same magnitude. We are getting hints of another
1
the orbit does not close! In fact in the real solar system the
conserved quantity. If the potential was not 1r but r1+
orbit does not close. They almost do but they in fact dont. The most famous example of this is Mercury. Where
the parahelien precesses. The orbit of Mercury precesses because
1. There are other planets which also exert a force on Mercury (classical Newtonian effect)
2. Einsteins theory of General Relativity modifies the

1
r

potential

Now back to quantum mechanics, we define a new quantity called


M=

p L e2
r

(V.4)

This quantity is a conserved quantity. To show that this is conserved we need to show that [M, H] = 0. However
there is a problem, p and L dont commute and M is not Hermitian. To make this Hermitian we do the following:
M=

e2
1
(p L L p) r
2
r

(V.5)

2
In assignment
 9 4we will show that [M, H] = 0. We will also show that L M = M L = 0 as well as M =
2H
2
2
L +~ +e .

Lecture 16, February 15th, 2012


The vector M is always perpendicular to L. This is shown in figure 5. We have the commutation relations

L
M

FIG. 5. The direction of the vector M

[Li , Lj ] = iijk ~Lk


[Li , Mj ] = i~ijk Mk
~
[Mi , Mj ] 2i HLk ijk

(V.6)
(V.7)
(V.8)

Notice that [Li , Lj ] closes. It is just SO(3). Further note that [Li , Mj ] and [Li , Lj ] close. However [Mi , Mj ] doesnt
close! Since H is there. However
[H, M]
[H, L]

(V.9)
(V.10)

So if we restrict ourselves to subspace that corresponds to a particular eigenvalue of H we can replace H by the
eigenvalue E. If you replace H by a number then the relations all stay within M, L! Therefore in the subspace we
have closure. This implies that we have a Lie Algebra in the subspace. Thus we can write the commutator (its
understood that we are staying in this subspace for which this is true)
[Mi , Mj ] =

2i~
ELk ijk

(V.11)

29
Therefore the algebra of all the Li and Mi close. For convenience we rescale M
 1/2
M
M0 =
2E

(V.12)

Remember that for bound state spectrum we have E < 0. Now r = (r1 , r2 , r3 ) and p = (p1 , p2 , p3 ) satisfy
[ri , pj ] = i~ij

(V.13)

where i, j = 1, 2, 3. Lets extend i, j to {1, 2, 3, 4}. In other words add r4 , p4 . We define


Lij = ri pj rj pi

(V.14)

L23 = r2 p3 r3 p2 = Lx

(V.15)

Mi0 = Li4 = ri p4 r4 pi

(V.16)

For example In this notation

We call

If we do this and still use [ri , pj ] = i~ij then you retrieve all the commutation relations of Li and Mi . Note that
r4 and p4 are NOT time and energy. r4 and p4 are artificial constructs. This is not 4D.
[Li , Lj ] = i~ijk Lk

(V.17)

is the Lie Algebra of SO(3). (L2 3, L31 , L12 ) = (Lx , Ly , Lz ) but (L14 , L24 , L34 ) = (M1 , M2 , M3 ). We know that SO(3)
is a subgroup and we also know that there are 6 generators. We also know that in SO(n) there are n(n1)
generators.
2
In fact we have the Lie Algebra of SO(4). We have rotations in 4D (but its not real 4D, just 3 real dimensions and
1 artificial). We can construct the generators
Lnm = |ni hm| |mi hni

(V.18)

L12 = |1i h2| |2i h1|

(V.19)

Thus

Warning: L12 is not a matrix element. This is the operator. Using the basis


0
1


1
0
|2i = ; ...
|1i = ;
0
0
0
0

(V.20)

we have
By finding the matrix elements we have

L12

1
=
0
0

1
0
0
0

0
0
0
0

0
0

(V.21)

We now find the commutator


[L23 , L31 ] = (|2i h3| |3i h2|) (|3i h1| |1i h3|) (|3i h1| |1i h3|) (|2i h3| |3i h2|)
= |2i h1| |1i h2|
= L12
= Lz

(V.22)
(V.23)
(V.24)
(V.25)

30
L2
L3
eiL12 = 1 + L12 + 2 12 + 3 12
2!
3!

1
0 0 0
2

0
1 0 0
=1+
+

2 0
0 0 1 0
0
0 0 0 1

cos sin 0 0

sin cos 0 0
=

0
0 1 0
0
0 0 1

(V.26)
0
1
0
0

0
0 1
3
0
1 0
+

0 3! 0 0
0
0 0

0
0
0
0

0
0
0
0

0
+ ...
0
0

(V.27)

(V.28)

In a fancy shancy way we can say


SO(4) =

SU (2) SU (2)
Z2

We can define
I=

1
(L + M) ,
2

K=

1
(L M)
2

(V.29)

Thus we get
[Ii , Ij ] = i~ijk Ik

(V.30)

[Ki , Kj ] = i~ijk Kk

(V.31)

[I, K] = [I, H] = [K, H] = 0

(V.32)

and

We have two casimir operators


I 2 = i (i + 1) ~2 ;

K 2 = k (k + 1) ~2

(V.33)

where i, k = 0, 12 , 1, .... There are two Casimir Invarients


C = I2 + K2 =


0
1 2
L +M 2 ;
2

C 0 = I 2 K 2 = L M0 = 0

(V.34)

Hence
I2 = K2

(V.35)

C = I 2 + K 2 = 2I 2 = 2k (k + 1) ~2

(V.36)

which implies that i = k. Therefor the eigenvaleus of

where k = 0, 12 , 1, .... We just said



0
1 2
L +M 2
2

1 2

=
L
M2
2
2E


1 2

=
L
L2 + ~2 + e4
2
2E
1 2
4
=
~
e
2
4E
e4
~2
=

4E
2

C=

(V.37)
(V.38)
(V.39)
(V.40)
(V.41)

31
but C also has values of 2k (k + 1) ~2 , k = 0, 12 , 1, .... Therefore
~2
e4

2k (k + 1) ~2 =
4E
2


4
1
e
~2 2k 2 + 2k +
=
2
4E
4
e
2
~2 (2k + 1) =
2E
1
e4
E=
2 ~2 (2k + 1)2

(V.42)
(V.43)
(V.44)
(V.45)

where 2k + 1 = n where n is an integer. So we can finally write


E=

1 e4
2 ~2 n2

(V.46)

where the degeneracy is n2 .

Lecture 17 - February 27th ,2012


Test 1 consists of (the marks are in brackets)
1. Hydrogen eigenstates (6)
2. EPR (4)
3. Density matrix (6)
4. Groups (18)
Know your definitions
The test is in TEL 0005, from 9:30-11:30.

VI.

DIRAC EQUATION

This is the subject of relativistic quantum mechanics. We first consider the Lorentz group. In SO(3), the quantity
r2 = x2 + y 2 + z 2 is invariant under rotations. In other words
0

Rr2 = r 2 = x 2 + y 2 + z

= r2

(VI.1)

In special relativity the invariant quantity is the length of the 4-vector (ct, x, y, z). In other words
0

c2 t2 x2 y 2 z 2 = c2 t 2 x 2 y 2 z

This quantity measures the space-time distance between space -time events. Notation:

x = (t, r) = (t, x, y, z) = t, ri

(VI.2)

(VI.3)

where from now on we use units such that c = 1. We also have


x = (t, r) = (t, x, y, z) = t, ri

(VI.4)

We call x a contravariant quantity (i.e. a quantity which transforms as the position 4-vector) and we call x a
covariant quantity (quantity that transforms like the derivative of a 4-vector, x ). Therefore we have
x2 = x x

(VI.5)

32
where here Einstein summation notation is used. We can also write
x2 = x x = g x x
where g is called the metric tensor. In special relativity

1 0

0 1
g =
0 0
0 0

(VI.6)

we have
0
0
1
0

0
1

(VI.7)

In general relativity g becomes a dynamical quantity (it satisfies (Einstein) field equations) and may have time
dependence. A Lorentz transformation (LT) is a transformation that leaves the length of 4 vectors invariant. We
write this as
X

= X

(VI.8)

or
X = 1

(VI.9)

we require that
X

= X2

(VI.10)

but we have
X

= X g X

= X g X

(VI.11)

X 2 = g X X

(VI.12)

g = g

(VI.13)

g = g

(VI.14)

g = T g

(VI.15)

Hence we demand that

To rewrite this consider

and

Therefore our condition can be written in matrix notation as


g = T g

(VI.16)

We will specialize to proper LT (LT without parity inversion). We will also specialize to orthochronous LT (LT
without time reversal). These are the physical transformations. Hence we can be in the neighborhood of the identity:
= 1 + 

(VI.17)


g = 1 + T g (1 + )

g = g +  T g + g + 2 T

(VI.18)

We insert this into our condition

(VI.19)

To first order we have T g + g = 0. Alternatively we can write


T = gg

(VI.20)

33
since g 2 = 1. Consider the matrix:

11

21
=
31
41
Then using the equation above we have

11 21

12 22
=
13 23
14 24

31
32
33
34

41
42
43
44

12
22
32
42

13
23
33
43

14
24
34
44

11 12 13 14

21 22 23 24
=
31 32 33 34
41 42 43 44

(VI.21)

(VI.22)

Clearly we have ii = ii and hence each of these is zero. Furthermore 12 = 21 , 13 = 31 , 14 = 41 . The rest
of the relations are 23 = 32 , 42 = 24 , .... Initially we required 16 parameters to specify this matrix. However
these conditions show that there 16 4 3 3 = 6 free parameters. The number of generators of the group are equal
to the number of free parameters. Hence there are 6 generators of the LT group. To specify the generators we use the
conventions that (1, 2, 3, 4) (0, 1, 2, 3) and we called the generators where , = {0, 1, 2, 3}. Note: DO NOT
confuse , with matrix indices, they are labeling a generator.

0 1 0 0
0 0 1 0
0 0 0 1

1 0 0 0
0 0 0 0
0 0 0 0
10 =
20 =
30 =
;
;

0 0 0 0
1 0 0 0
0 0 0 0
0 0 0 0
0 0 0 0
1 0 0 0

0 0 0 0
0 0 0 0
0 0 0 0

0 0 1 0
0 0 0 0
0 0 0 1
12 =
23 =
13 =
;
;

0 1 0 0
0 0 0 1
0 0 0 0
0 0 0 0
0 0 1 0
0 1 0 0
This can be written as

where 

( ) =  

2
is the 4D Levi-Cevita Symbol. One can show that
01 = 10 , 21 = 12

(VI.23)

(VI.24)

The most general infinitesimal LT can be written


1
= 1 + i i0 + i ijk jk
2
This group contains 6 parameters i , i = 1, 2, 3 and i , i = 1, 2, 3. Therefore a finite LT can be written


1
= exp i i0 + i ijk jk
2

(VI.25)

(VI.26)

This is called the Lorentz group, denoted by SO (1, 3).


Lecture 18 - March 2nd, 2012
Consider the LT for i = i = 0 for all i except for 3 .


1
(3 ) = exp
3 3jk jk
2


1
= exp
{21 3 + 12 3 }
2


1
= exp
{23 21 }
2
= exp (12 3 )

(VI.27)
(VI.28)
(VI.29)
(VI.30)

34
we have
0
0
0
0

0
0
1
0

0
1
0
0

0
0

0
0

0 0

0 0
0
0

0
0
1
0

0
1
0
0


0
0

0 0
=
0 0
0
0

12 =

0
2
(12 ) =
0
0

0
0
1
0

0
1
0
0

(VI.31)

0
1
0
0

0
0
1
0

0
0

(VI.32)

Since this is diagonal this is easy

0
2
(12 ) =
0
0

0
0
1
0

0
= 12
0
0

0
1
0
0

0
4
(12 ) =
0
0

0
1
0
0

0
0
1
0

0
0
0
0

(VI.33)

(VI.34)

5
and then the sequence repeats,i.e. 12
= 12 etc. We can now find explicitly:

2 2
3 3
(3 ) = I + 12 + 12
+ 12
+ ...
2
3!

0
0
0
3

2
1 2
+ 3! + ...
0
=
3
2
0 + 3! + ....
1 2
0
0
0

1 0
0
0

0 cos sin 0
=

0 sin cos 0
0 0
0
1

(VI.35)

0
0

(VI.36)

(VI.37)

We see that 3 gives a rotation about the 3 axis. Since there is nothing special about 3 , 1 and 2 give us rotations
about the x and y axes. If we set all the i = 0 then we have the group SO (3). Now lets set all parameters to zero
except 1 .
(1 ) = exp (1 10 )
Recall that

10

1
=
0
0

1
0
0
0

0
0
0
0

0
0
0
0

0
1
0
0

0
0
0
0

0
0
0
0

Hence we have

2
10

0
=
0
0

(VI.38)

35
and hence
3
10
= 10

Hence we have
3
2
2
3
(10 ) +
(10 ) + ...
() = I + 10 +
2
3!

2
3
1 + 2 + ... + 3 + ... 0 0
3
2

+ 3 + ... 1 + 2 + ... 0 0
=

0
0
1 0
0
0
0 1

cosh sinh 0 0

sinh cosh 0 0
=

0
0
1 0
0
0
0 1
Hence under the action of the Lorentz Transformation, (1 ) we see that


t
cosh sinh 0 0
t0

0
x sinh cosh 0 0 x

0 =
y 0
0
1 0 y
z
0
0
0 1
z0

(VI.39)

(VI.40)

(VI.41)

(VI.42)

Explicitely we have
t0 = cosh t + sinh x
x0 = sinh t + cosh x
y0 = y
z0 = z
If we define = tanh1 v which is called the rapidity. This gives
v = tanh
v 2 = tanh2
= 1 sec h2
1
=1
cosh2
cosh2 =

1
1 v2

tanh =

v/c
sinh
=p
cosh
1 v 2 /c2

We define = cosh . We can write

and we get
t0 = (1 + x)
x0 = (1 + ct)
We see that we have a subgroup of boosts. So all together the Lorentz group has 3 rotations and 3 boosts. If we
choose a more familiar parameterization
i
U () 1 + J
2

(VI.43)

36
where we relate these J s to our angular momentum operators by
Ji =

1
ijk Jjk
2

(VI.44)

and we define
Ki = Ji0

(VI.45)

[Ji , Jj ] = iijk Jk
[Ji , Kj ] = iijk Kk
[Ki , Kj ] = iijk Jk

(VI.46)
(VI.47)
(VI.48)

This gives

This forms the algebra of SO(3, 1)

Lecture 19 - March 5th, 2012


We have finished with the Lorentz group. Now we move on to the Dirac equation. The Schrodinger equation is given
by
 2

p
+ V = E
(VI.49)
2m
(K + V ) = E
(VI.50)
where K is the kinetic energy operator, V is the potential energy operator and E is the energy operator. This
p2
equation is Galiliean covariant but obviously not Lorentz covariant. This is because E = 2m
+ V is not the relativistic
Energy-momentum (dispersion) relation. The relativistic energy-momentum relation is what we get from constructing
a Lorentz-invariant quantity. This quantity is p2 , where p is the Lorentz 4-vector equal to p = p0 , p = (E, p).
p2 = p p = g p p

(VI.51)

p2 = E 2 p2 = m2 E 2 = p2 + m2

(VI.52)

Thus
E = m2 + p2

1/2

(VI.53)

If we demand the correct dispersion relation then we would have


i~

1/2

= m2 ~2 2

t


4
2

+ ...
=m 1
2m 8m4

(VI.54)
(VI.55)

Even though this equation works many problems are encountered using this method and this route was not historically
taken. Alternatively we can not taken the square root:
E 2 = m2 + p2

where E 2 = i~ t

2

(VI.56)

2
= ~ t
2 . Recall that p = ~ x2 . We now introduce useful notation that says

 = g  

(VI.57)


 + m2 = 0

(VI.58)

with this we have

This is called the Klien Gordan equation. Solutions to this equation take the form
i(krEn t)

n = Ne

where En =

m2 + kn2 . However there two problems with this solution

(VI.59)

37
Some states turn out to have a negative norm.
This generates states with negative energy
The origin of the negative energy can traced back to the fact that we have the square of the energy. Is there an
equation that is first order in time AND first order in space so that one can have manifest Lorentz symmetry? This
is a fancy way of saying symmetry between time and space. With this in mind Dirac wrote

= H = ( p + m)
t
where and are at this point two unknowns. However we do know that
i~

(VI.60)

E 2 = p2 + m2

(VI.61)

(we demand this to be true). We now implement this on our equation


2


2
= ( p + m)
i~
t

= p2 + m2

(VI.62)
(VI.63)

This requires
2

( p + m) = p2 + m2
2

(i pi + m) = p2 + m2
2

i pi j pj + i pi m + i pi m + m = p + m

(VI.64)
(VI.65)
(VI.66)
(VI.67)

This tells us that and cant be just numbers because in order to get the two terms first order in p to be zero we
require that
i + i {i , } = 0

(VI.68)

Continuing we can write the left hand side as


i2 p2i +

1
{j , i } pi pj + {j , } mpj + 2 m2
2

(VI.69)

Now demanding that this equal p2 + m2 we must have


i2 = 1;

{i , j } = 2ij ;

{i , } = 0;

2 = 1

We have the following Lemmas:


1. and i are traceless:
i + i = 0
i = i
i = 2 i
i2

i =
Tr (, ) = Tr (i )
Tr (i ) = Tr ()

(VI.70)
(VI.71)
(VI.72)
(VI.73)
(VI.74)
(VI.75)

Hence the i are traceless (its similar to prove that is traceless) but also we have i2 = 1 and 2 = 1.
2. Now we can diagonalize the i and . If we have a diagonal matrix that the square is equal to the identity we
have

x
1 0 0 0

x
0 1 0 0
(VI.76)

x 0 0 1 0
x
0 0 0 1
This requires that we have eigenvalues of 1.

38
3. The matrices are n n where n is even (since the eigenvalues must add to 0 and they are 1)
The dimensionality of the matrices is even. Is it possible for n to equal 2? There are n2 independent Hermitian matrices
(n2 parameters). However the identity is not traceless so it cant be a candidate for i or . Thus subtracting the
identity gives n2 1 matrices. Thus we have n2 1 = 3 possible independent matrices. However we need at least 4
matrices to have one for all of i , . n = 4 gives n2 1 = 15 matrices which is more then enough.
Conclusion: We can find 4 4 independent Hermitian matrices that satisfy all our requirements. Since n = 4
gives many matrices we can choose which we want to use. We make the choice:
!
!
1 0
0 i
=
;
i =
(VI.77)
0 1
i 0
where each of these entries is a 2 by 2 entry.

Lecture 20, March 7th, 2012


As it stands this equation doesnt look manifestly covariant (though it is). We define

= , i

(VI.78)

where 0 = and i = i . Thus


0

1 0
0 1

!
;

0 i
i 0

!
(VI.79)

This gives the anticommutation relation:


{ , } = g

(VI.80)

This gives



d
m =0
i
dx

(VI.81)


Note that has the form 2 , but it is not a four-vector! It is a four-spinor. We will learn later how transforms
3
4
under rotations and boosts (i.e. LT).
Recall that the Klien-Gordan equation had problems. One problem was there were states with negative norm and
the other problem was that it had states of negative energy. We shouldnt have negative states because if there were
then electrons would fall to these negative energy states emitting energy which we dont observe.
Consider the following
i

b = Hb i b = a Hb
t
t

(VI.82)

and
i

= (Ha ) i a b = (Ha ) b
t a
t

(VI.83)

subtracting these equations gives


i

h


a b ia i i + i b
t

(VI.84)

where the arrows denote direction over which they act. This gives
i




a b + ii a i b = 0
t

(VI.85)

39
i

 0 
b = 0
x a

(VI.86)

 
a b = 0
x

(VI.87)

If we define 0 we have
i
Comparing with the known equation
have

x J

= 0, this is just the continuity equation! Recall in electromagnetism we


J = (, J)

(VI.88)

= J
t

(VI.89)

In electromagnetism we have

Integrating over some volume

Z
dV =

JdV
Z

Q = J dA
t
S

(VI.90)

(VI.91)

Back to Quantum mechanics. If charge was conserved before what is conserved now? What is this object

(VI.92)

This is the probability current. Since this satisfies the continuity equation probability is conserved. Therefore norm
is conserved. The Dirac equation can be written in covariant form as



i m = 0
(VI.93)
x
This is the free Dirac equation (no potential), written in covariant form. To make it non-free we need to add a potential.
However the idea of a potential is non relativistic. The way we add interactions is by adding a four-potential:
 



i eA m = 0
(VI.94)
x
where A = (, A). Note that the Dirac equation
! is mainly good for electromagnetism. We try free-Dirac equation

. and are 2 1 column vectors. refer to positive and negative


solutions of the form p (r, t) = Np eprEp t

energy solutions Ep > 0. We substitute the positive energy solution into the free-Dirac equation:
!




prEp t
i
m Np e
(VI.95)
x

Lecture 21 Missed a lecture

Lecture 22 - March 12th, 2012


We have the Dirac equation


i
m =0
x

(VI.96)

40
!

(not but + , where p x = p x = p0 x0 p x = Et p x. We


We try the solution (r, t) = N e

found a solution terms of arbitrary but we require


ipx

E+m

(VI.97)

For normalization we require


Z

d3 r =

=N

!
N2

(VI.98)

!
L3

= N 2 L3 +

(VI.99)

(VI.100)
!

( p)

= N 2 L3

(p)2
(E+m)2

but ( p) = p2 (see notes from previous lecture). Thus we have


Z

d r = N L

(VI.101)

(E + m)

Inserting in this result gives

p2

p2
.
(E+m)2

(E + m)

(VI.102)

Since is just a number we have


2

d r = N L
3

=N L

(E + m) + p2

(E + m)

2E 2 + 2Em

(E + m)

= N 2 L3

2E

E+m

1/2

(VI.103)

(VI.104)
(VI.105)

We will choose = 1. Thus we require


N=

E+m
2EL3

(VI.106)

It is conventional to choose
u (p, s) =

1/2

where

1
0

!
and

1/2

0
1

E+m

p
E+m

(VI.107)

!
. Note that more explicitly we have

s
0

u (p, s) = E + m
0
0

0
s
0
0

s
s
p
(p

ip
)
x
y
E+m z
E+m
s
s

E+m
pz
E+m (px + ipy )
0
0

(VI.108)

Putting this together we have


+
p,s
=

1
2EL3

u(p, s)epx

(VI.109)

41
similarly we have

p,s
(x) =

p
E+m

1
2EL3

v(p, s)eipx

(VI.110)

where v(p, s) = E + m
(i2 s ). Notice that the zero momentum solutions take the form of
1


0
0
0
1


0
0 1 0
.
, , , and
1
0 0 0
0
1
0
0
Consider the following operation
c = C
where C = i2 =

0 2
i2 0

(VI.111)

!
and we have C 2 = 1. Other properties of C are
C = i 2 0
C = C 1
C = C
C = CT
CC 1 =
C C 1 = T

Consider the Dirac equation





= i ii eAi + eA0 + m
t
with A as the vector potential and in non-manifestly covariant form.
i

(VI.112)




= i ii eAi + eA0 + m
(VI.113)
t



(VI.114)
i = i ii eAi + eA0 m c
t



i = i ii + eAi eA0 + m c
(VI.115)
t
This is the original Dirac equation but e e. c satisfies the Dirac equation but with opposite charge. We also

+
= p,s
(x).
find that Cu (p, s) = v (p, s). Therefore we have Cp,s
c
p,s describes an antiparticle with positive energy with momentum and spin. Diracs interpretation (which was
eventually overtaken by QFT) is that electrons satisfy the Pauli exclusion principle so he says that the world has all
the negative energy states filled. Dirac says that c correspond to holes in the negative energy sea. Pair creation
can be thought of as knocking out an electron in a negative electron state into the positive energy states.
i

Lecture 23, March 16th, 2012

A.

Non-relativistic limit of the Dirac equation

One reason to do this is to make contact with the Schrodinger equation and the Pauli equation. The Pauli equation
is just the Schrodinger equation with two component spinors.
This will give us intuition on what a 4-spinor is. Recall that we have
+ =

1
2EL3

u (p, s) eipx

(VI.116)

42
where u(p, s) =

E+m

Dp s
E+m

!
. Consider the v 0 limit. In this limit we have p 0 and p = (m, 0). Hence

1
1
1
2m
(s = ) =
u (0, s) eimt =
2
2mL3
2mL3

1

1
0 imt
= e
L3 0
0
+

1
0

1
0

!
eimt

(VI.117)

(VI.118)

while we have

+ s =

1
2

0
1

1
2

1
=
L3

1 imt
e
0
0

(VI.119)

Similarly the negative energy solutions are




1
1

s=
2m
=
2
2mL3

0 1
1 0

1
0

1
eimt =
L3

0 imt
e
0
1

s =

1
0 imt
=
e
L3 1
0

(VI.120)

(VI.121)

This is the spinors at rest. To gain more insight on the problem we need to boost these spinors. We dont yet know
how to do this since we only know the transformations in the fundamental representation. However in 1950 Folby
and Woothuysen came up with a technique to the non-relativistic reduction properly. They showed that you could
p
expand the spinors in a m
expansion.
We call the large components of this expansion and the small components (this is clear by the results above, for
p
electrons the components, E+m
are 0). What Folby and Woothuysen found was to find a way to write the spinors
in terms of s in terms of Pauli spinors.
This trick is difficult. We start with a practice problem. Consider you have the following Hamiltonian H =
x Bx + z Bz (a H that just involves pauli matrices). The Hamiltonian is a 2 2 matrix. This Hamiltonian would
describe Pauli spinors interacting with the magnetic field. Also consider the unitary transformation
U = eiy /2 = cos

+ iy sin
2
2

H 0 = U HU 1 = eiy /2 Heiy /2

 


= cos + iy sin
H cos y sin
2
2
2
2

= cos2 (x Bx + z Bz ) + i sin cos ([y , x ] Bx + [y , z ] Bz ) + sin2 (y x y Bx + y z y Bz )


2
2
2
2
= x (cos Bx sin Bz ) + z (cos Bz + sin Bx )

(VI.122)

(VI.123)
(VI.124)
(VI.125)
(VI.126)

x
If tan = B
Bz then the coefficient of x is zero. Hence we have diagonalized H. We have decoupled the upper and
lower components.

43
We now do a more difficult problem (however still not the full problem). Consider the free Dirac equation (no
vector potential).
H = p + m

(VI.127)

Here we want to decouple the large and small components of the Dirac equation (due to low velocity limit). Recall
that
!
0 i
i =
i 0
Hence in block diagonal form we have
H=

0 p
p 0

!
+

1 0
0 1

!
m

(VI.128)

It turns out that eis where is = p(p) does the trick. This can be written as
eis = cos (|p| ) +

p
sin (|p| )
|p|

(VI.129)

If we transform H then we have


H 0 = U HU 1
= (m cos 2 (|p| ) + |p| sin (|p| ))
if we choose tan (|p| ) =

|p|
m

(VI.130)
(VI.131)

then we have
p
H 0 = m2 + p2

(VI.132)

(VI.133)

E
Lecture 24 - March 19th, 2012
We are not responsible for the Foldy-Wouthuysen online but we are responsible for the toy example, Dirac free
example, and the March 19th (today) example.
We now continue with FW. The general case is
H = (p eA) + m + e

(VI.134)

where A0 = and Ai = . So far we were able to completely decouple large and small components of the Dirac
equation but only in the free case (i.e. A = 0). In the general case this is impossible to all orders in |p|
m (this tells us
how relativistic the system is). However Foldy-Wouthuysen tells us how to decouple to any desired finite order in |p|
m.
Recall we are interested in decoupling so that we have an equation which is Schrodinger-like for the Pauli 2-spinor
. We could continue with a systematic expansion through FW (which is done in the posted notes). Instead of
developing a general systematic approach, we will use hindsight to efficiently get to the next level of desired accuracy.
First lets define two quantities:
T =Em

(VI.135)

(roughly speaking this is the kinetic energy). Further we define


V = eA
(just to simplify notation). We can now write the Dirac equation in first order with a Hamiltonian by
!
!
!


0
T
= m + V + (p V) + m
=H

(VI.136)

(VI.137)

44
Consider

p
m

U = U = A +
where A =

2 p2
m2

(VI.138)

with as an undetermined parameter.





p
A + p
m
m
2
A
2
= A2 2 ( p + p) + 2 ( p)
m
m

ij
z }| {
2
z
}|
{
Api
1

= A2 2 +
i + i + 2 pi pj i j + j i
m
m
2

UU =

A +

= A2 +

2 p2
m2

(VI.139)

(VI.140)

=1

(VI.141)

Hence U is unitary. We have


!

UT

0
0

=H

U HU
= H0

!
(VI.142)

!
0
0

!
(VI.143)

(VI.144)

Note that we used T 0 = U (E m)U 1 = T We now find H 0 = U HU 1


U (m) U 1 = m

(VI.145)

We go on now
U V 0 U 1 = AV 0 A +

 2

AV 0 p pV 0 A + 2 pV 0 p
m
m

(VI.146)

The third term is


U (p V) U 1 = A (p V) A +

(A (p V) p)
m

(VI.147)

and the last term is


U mpU 1 = mA2 + 2A p

2 p2
2m2

(VI.148)

Remember our goal is to limit off-diagonal terms. To lowest order the off-diagonal terms all come from the (p V)
term.
0
Hof
f diag = (p V) + 2 p

(VI.149)

We now tune to = 21 . Our new off diagonal part of the Hamiltonian is


0
Hof
f diag = V

(VI.150)

0
T 0 = H11
+ V 0

(VI.151)

The coupled equations become

45
T 0 = V0 2m 0

(VI.152)

The key step is to say that T 0 is must smaller then 2m 0 (since we are in the nonrelativistic limit). This allows us
to write
0 =

p 0

2m

(VI.153)

We have decoupled these equations to some order. The error we are making is higher-order in
are decoupled and we have one equation:
0
T 0 = H11
0 +

|p|
m.

Now the equations

V V 0

2m

(VI.154)

where
0
H11
=V0

2
pV 0 p
1
p2
p4
p2 0
0 p
V

V
+
+

(p

p)
+

(p

V))

8m2
8m2
4m2
2m
8m3
2m

One can simplify this equation and you find that (and we put e = V 0 and eA = V)

2
2
(p eA)
e
p4
e
V2
0
=
+ e
B
+e
+
([] p)
H11 +
3
2
2m
2m
2m
8m
8m
8m2

(VI.155)

(VI.156)

Lecture 25 - March 21st, 2012


Recall we are doing the non-relativistic reduction of the Dirac equation. However we are not doing a full reduction
as we are keeping important relativistic effects. We are doing it systematically so that we keep all relativistic effects
to some order in |p|
m . Foldy-Woothuysen shows how you to do this in general. We will just do this to lowest order.
However in practice for hydogen our results will be highly accurate, namely up to 4 .
Recall we were taking the Dirac equation and applying a unitary transformation
U = A +
where A =

2 p2
m2 .

p
m

(VI.157)

This transformation let us to this coupled system


0
T 0 = H11
0 + V 0
T 0 = V0 2m 0

(VI.158)
(VI.159)

If we have T  2m then
0
T 0 = H11
0 +

V V 0

2m

(VI.160)

where
0
H11
=V0

p2 0
p2
pV 0 p
1
p4
p2
V V0
+
+
( (p V( p)) + p (p V))

(VI.161)
2
2
2
3
8m
8m
4m
2m
8m
2m

We simplify this term by term


(p V) p + p ( (p V)) = 2p2 p V V p
2

= 2p p V V p i (p V) i (V p)
2

= (p V) + p V [ V]

(VI.162)
(VI.163)
(VI.164)

where we have used A B = ijk Aj Bk and square brackets,[ ] means that the operators inside act only inside. Our
potential could have been any vector potential. We note choose the electromagnetic force:
V = eA

(VI.165)

46
This gives (with B = A)
2

(p V) p + p ( (p V)) = (p V) + p2 V 2 e B

(VI.166)

This shows you that the electric interacts with the magnetic field with the expression e B! Thus this actually
predicts that the electron has a magnetic moment. Continuouing with the reduction we have




p2 V 0 + V 0 p2 = p0 V 0 + 2 pB 0 p + 2V 0 p2
(VI.167)
and


pV 0 p = pV 0 p + V 0 p2


= pV 0 p + V 0 p2





= pV 0 p + i pV 0 p + V 0 p2

(VI.168)
(VI.169)
(VI.170)

so we see that
 2 0



 pV 0 p
p V
i pV 0 p
1
2 0
0 2
=
p V +V p +
+

8m2
4m2
8m
4m2

(VI.171)

If we now input in
p = i
V = eA
V 0 = e
then we have
0
H11

 2 
2

V2
(p eA)
e
p4
e
+
=
+ e
B
+e
+
([] p)
2m
2m
2m
8m3
8m2
8m2

if we assume that is spherically symmetric then we have = (r) and thus =


(1)

H11

r d
r dr

(VI.172)

then we finally have

(5)

(6)
z }| { (2) z (1)
}| { z (4)
}|
{
}| { z  }|2 { z
2
4
z}|{

e
(p eA)
p
3 d
=
+ e

B+e
+
L
2m
8m3 2m
8m3
4m2 r dr

(VI.173)

The first term can be rearranged: (we assume that A2 is small)


2

(p eA)
p2
=
2m
2m
p2
=
2m
p2
=
2m
p2
=
2m
p2
=
2m
p2
=
2m

+
+

e
(p A + A p)
2m
ie
( A + A )
2m
ie
A
m
ie
(r B)
2m
ie
B (r )
2m
e
BL
2m

(VI.174)
(VI.175)
(VI.176)
(VI.177)
(VI.178)
(VI.179)

p2
2m

e
is the kinetic energy and 2m
B L is the Zeeman term. Thus we have obtains a field-orbit coupling term without
explicitly putting it in again! Now consider the second term. This is the non-relativistic potential. This is just the
Coloumb potential for hydrogen

e =

e2
r

(VI.180)

47
The third term
p4
8m3

(VI.181)

is just the next order relativistic correction to the kinetic energy. Now consider the fourth term:

e
e
B= SB
2m
m
e
= SB
m

(VI.182)
(VI.183)

This is a prediction that the gyromagnetic ratio of the electron is equal to 2(g = 2)! The full Zeeman effect can be
e
written 2m
B (L + 2S). Now consider the fifth term:
 2 2

e2
= 2 E
(VI.184)
e
2
8m
8m
e2
=
(R)
(VI.185)
8m
2

(with = er ). This term is called Zetterbeuwegung. The sixth term is


e d
SL
4m2 r dr

(VI.186)
LT

This is spin-orbit coupling which is because the electron sees (, 0, 0, 0) (0 , A0 ) under a Lorentz transformation.

Lecture 26 - March 23rd, 2012


Missed this lecture.

Lecture 27 - March 25th, 2012

B.

Covariance of the Dirac equation

We are trying to prove that


{ (p eA ) m} = 0



p0 eA0 m 0 = 0

(VI.187)

p0 = p

(VI.189)

(VI.188)

Contravarient vectors transform as

and the controvarient vectors transform as


p0 = 1

(VI.190)

We have
0

p2 = p p0 = p p

(VI.191)

0 = S

(VI.192)

We insert in

S is the spinor representation of the Lorentz group. After inserting and S into preed equation we get
S 1 () S() =

(VI.193)

48
For Lie groups, to find the generators we only have to satisfy this equation to infinitesimal order and then its true to
all orders
1
= 1 + i i0 + i ijk jk
2

(VI.194)

we have 6 parameters (i , i ) and 6 generators i0 , jk (note these are NOT indices but labels!). Whatever S() is in
infinitesimal form it will look like
1
S() = 1 + i Bi + i ijk Rjk
2

(VI.195)

where i and ijk are the same parameters as for . At this point we dont know what the generators of the spinor
representation are, Bi , Rjk . We insert infinitesimal forms into S 1 S = . Last time we finished class at (he
left this as an exercise)

[Bi , ] = (i0 )

[Rjk , ] = (jk )

(VI.196)
(VI.197)

The new exercise is to verify that


1 0 i
Boosts
2
1
= j k Rotations
2

Bi =

(VI.198)

Rjk

(VI.199)

This appears non-covariant. This can be written in covariant form


=

i
[ , ]
2

(VI.200)

where 0i = 2iBi and ij = 2iRij (these are not the Pauli matrices! We are just using as some matrix). Note
however that this is not a convenient form of the generators. Bi are the generators of boosts for spinors and Rjk are
the generators for rotations for spinors. Lets introduce some more notation.
!
j 0
(VI.201)
j j =
0 j
and
0 1
1 0

!
(VI.202)

We now have (using the matrix we had before)


1
i
2
i
= jk` 5 `
2

Bi =

(VI.203)

Rjk

(VI.204)

A boost in Dirac-space is

2
1
+
+ ...
2
2
2
 
 

+ sinh
= cosh
2
2

S( ) = 1 +

(VI.205)
(VI.206)
(VI.207)

where is the rapidity, tanh() = vc ,


at rest we ahve

E
m

= cosh(), cosh( 2 ) =

eimt
0,s (x) =
L3

s
0

E+m
2m ,

and sinh( 2 ) =

eimt
=
u (0, s)
2mL3

E+m
2m

p
E+m

. For a spinor

(VI.208)

49
remember that

0
0

!
(VI.209)

and therefore

cosh + sinh =
2
2
Lets apply S( ) to

Xs
0

sinh 2
cosh 2

cosh 2
sinh 2

!
(VI.210)

!
.

S( )

Xs
0

 
 

= cosh
+ sinh
2
2
 
!
cosh 2 X s
=
s
sinh 2
!
r
1
E+m


=
s
p
E+m
2m
!
r
E+m
1
=
s
p
2m
E+m

Xs
0

!
(VI.211)

(VI.212)

(VI.213)

(VI.214)

we have retrieved that solutions for arbitrary momentum. Now consider rotations in Dirac space.
5

S( ) = e(i/2)

= cos i 5 sin
2
2

(VI.215)

S( )u(p, s)

(VI.217)

(VI.216)

Consider the rotation below

Recall from our study of SU (2) we know how Pauli spinors transform under rotations.


0

s
s = S()s = cos i sin
2
2
Lets pick to rotate about the z axis. Back to our rotation in Dirac space we have
!

0
1
S( )u(p, s) = E + m S()pS () s

(VI.218)

(VI.219)

E+m

(check this at home) where if we are specializing to rotations in the z axis we have
S() = cos

+ iz sin
2
2

To simplify S( )u(p, s) we need to simplify the following







S() pS () = cos iz sin


p cos + z sin
2
2
2
2

= cos2 p i sin cos (z p pz ) + sin2 z pz


2
2
2
2
= p0

(VI.220)

(VI.221)
(VI.222)
(VI.223)

where p0 = (pz cos py sin , py cos + px sin , pz ). Note that we skipped about a half page of steps in the last step.
This is also left as an exercise.

50

Lecture 28 - March 28th, 2012


For the test:
Some qualitative questions
In accidental degeneracy prove some commutation relation or identities
Nothing too long, but take a good look at the assignment with M
Short-ish proofs from the Dirac equation, similar to as done in the notes
Formula sheet will be posted
VII.

HYDROGEN

One can obtain exact solutions to the hydrogen Dirac problem. The Dirac equation takes the form
 



i eA = 0
x

(VII.1)

where A = (e, 0) with eA = e


r . We have chosen the rest frame of the proton. It no longer looks covariant.
Thats okay as long as you stay in that frame. In this case the Dirac equation becomes (exercise) (where we have
used = 0 and 2 = I).


e2

E = ii i + m

(VII.2)
x
r

. V = V (r), thus just like the Schrodinger equation


One should ask where is the time dependence? It is in 0 i t

(r, t) = (r)eiEt/~

(VII.3)

The time dependence in the wave function is just a phase rotation ( doesnt change with time). Derivative term
just gives the energy term on the left, where E is the constant (an eigenvalue). Now we have
L = r p = i~ (r )
1
S = 5
2
 i 
L , H = [iijk rj k , i` ` ] = ijk j k 6= 0

(VII.4)

This is another exercise. Note that one should memorize Li = iijk rj k , useful expression. That means that L is
not conserved! (L is not a good quantum number).


 1 5

Si, H =
cai , ij j = ijk j k 6= 0
2

(VII.5)

(exercise) This is bad news! We lost all our good quantum numbers. However the good news is that the commutator
 i

L + Si, H = 0
(VII.6)
Hence the total angular momentum is conserved. J is a good quantum number.
3
1
S 2 = S i S i eq =
4
4

(VII.7)

( The magnitude of |S| is constant. This tells us we can label states by good quantum numbers, J 2 , Jz , S 2 , L2 . All
these operators commute with H and therefore we can have simultaneous eigenfunctions of J 2 , Jz , S 2 , L2 . Now we
write the 4-spinor as
!
F (r)
(r) =
(VII.8)
G(r)

51
We introduce the following new object yjm (
r).




1 1
1
, |j, mi Y`,m 12 + `, m +
yjm (r) = `, m
2
2 2



1
1 1

|j, mi Y`,m+ 21 (r)


,

2
2 2

Equivalently one often uses the notation:








1 1 1
1 1 1
yjm (r) = `, m , , j, m Y`,m 12 + `, m + , , j, m Y`,m+ 12 (r)
2 2 2
2 2 2

(VII.9)

(VII.10)

These are just the Clebsh-Gordan coefficients. It is useful to define an operator,


K = ( L + 1)

(VII.11)

K`,m
= kyj,m

(VII.12)

since

where
(
j + 21 if j = ` 21

k=
j + 12
if j = ` +

(VII.13)

1
2

We can write the solution as


k
j,m
(r)

k
fjk (r)yj,m
(
r)
k
k
igi (r)yj,m
(
r)

!
(VII.14)

Note that this needs to be verified. At this point we have solved the angular problem. This is equivalent to (r) =
R(r)Y`,m (, ) but not the radial problem. Notice the angular part is solved with the familiar spherical harmonics.

Lecture 29 - March 30th, 2012


To check this solutions we need the following simplifications to our expressions (prove these)

L
+ i r
r
r
k
k
Lyj,m (
r) = (k + 1) yj,m (
r)
i = i r

k
i yj,m
(
r) = i

k
yj,m
(
r)

+i

k
ryj,m
(
r)

k
yj,m
(
r)

(k + 1) k
yj,m (
r)
r

Separation of variables is successful and we find find two coupled linear 1st order homogenous equations


g k (r) 1 k k
e2
k
f k (r)

g (r)
Ef (r) = m
r
dr
r


e2
df k (r) 1 + k k
Eg k (r) = m
g k (r) +
+
f (r)
r
dr
r
To solve these solutions we rescale our solutions (just as in the S.E. solutions)
f
g
fR = , gR =
r
r
p
2
2
= m E r
r
mE
=
m+E

(VII.15)
(VII.16)
(VII.17)
(VII.18)

(VII.19)
(VII.20)

52
In natural units we call e2 = . This gives two equations:


k
1
dfR
+ fR
+
gR = 0
d



dgR
k
e2
gR 
fR = 0
d

(VII.21)
(VII.22)

We solve with a power series:

fR =
g =

An n e

(VII.23)

Bn n e

(VII.24)

n=0

X
n=0

Substitute forms in. There is a non trivial solution if = k 2 m2 . These differential equations have a continuous
set of solutions. However to enforce normalizability we must force the solutions to terminate the infiite series so that
solutions are well behaved. This discrete subset are our physical solutions. We can force termination by demanding
2 + 2N

1 2 = 0


(VII.25)

where N is an integer. As always this is where the quantization of E comes from (comes from the quantization of ).
Amazingly one can solve this problem in terms of confluent hypergeometric function.
1/2

En,j = m
1

2
q





2
j + 12 2 j + 12
n2 + 2 n j + 21

where we have now translated from k j. We can expand terms of (since in Hydrogen
En,j

m2
m4
=m

2n2
2n4

n
n+

1
2

(VII.26)

p
m

is of order ,

p
).
m


(VII.27)

Note that we have lost a lot of our degeneracies. What about real hydrogen? there also have a proton with a spin,

. This produces a magnetic moment,


=

ge
S
2m

(VII.28)

This creates a spin -spin interaction which splits the 1S1/2 state into two states. The triplet and single states. This
is the famous 21cm line. The energy difference is very simple and its given by
p
E = m 1 2 .
(VII.29)
This is exact in the Dirac equation. The 1s1/2 state from the Dirac equation is

1
0

12 ,s=+ 12 = A0 1 e
1
cos
i
1
sin e

(VII.30)

53
where = 1 2

1/2

. Notice that even the ground state has angular dependence! We also have

12 ,s= 21

= A0 1 e 1
sin ei
1
cos

(VII.31)

We can write this in more convenient notation:


cos
sin ei

!
=

z
x
+ i
y

(VII.32)

and
r=

z
x
i
y
x
+ i
y
z

(VII.33)

Hence
r

1
0

cos
sin ei

(VII.34)

This gives
12 ,s= 21 = A0 1 e

1
r

1/2

(VII.35)

Anda mungkin juga menyukai