You are on page 1of 32

11.

Angular Momentum Theory, November 20, 2009 1


Chapter 11. Angular Momentum: General Theory
1 Introduction. Angular momentum is intimately related to rotational
motion of a system of particles. It is essential for understanding the infrared
and microwave spectra of a molecule and the properties of the electrons in
atoms.
The spin of an electron or of a nucleus is an angular momentum that has
no classical analog. The spin is accompanied by magnetic moment, which
means that we can change the energy of a particle by acting on it with an
external magnetic eld. NMR and ESR spectroscopies are based on this
ability. Understanding these extremely rich and important spectroscopic
techniques requires a good understanding of angular momentum.
In this chapter we dene the dynamics of angular momentum through
the commutation relations between the operators representing its projection
on the coordinate axes. These commutation relations allow us to determine
the eigenstates of the angular momentum operator and derive the matrix
representation of the relevant operators. The chapters that follow use this
theory to examine the properties of spin, the orbital angular momentum, and
the angular momentum of more than one particle.
2 The denition of the angular momentum operator. In quantum me-
chanics, we encounter two kinds of angular momenta. The orbital angular
momentum is very similar to the angular momentum in classical mechanics.
It describes some aspects of the motion of a particle and it is dened by

L = r p (1)
where r is the position of a particle and p is its momentum. To dene the
operator

L describing the orbital angular momentum in quantum mechanics,


we replace, in Eq. 1, r with

r and p with

p:

L =

r

p (2)
In addition to the orbital angular momentum, the particles studied in
quantum mechanics have an intrinsic angular momentum called spin. This
is not caused by the motion of the particle (e.g. its rotation); it is instead
a property similar to charge or mass. This angular momentum cannot be
11. Angular Momentum Theory, November 20, 2009 2
dened by Eq. 2. Born, Heisenberg, and Jordan
1
noted that the angular
momentum dened by Eq. 2 satises the commutation relations
[

L
x
,

L
y
] = i h

L
z
(3)
[

L
y
,

L
z
] = i h

L
x
(4)
[

L
z
,

L
x
] = i h

L
y
(5)
We take this to be a denition: a vector operator whose components satisfy
Eqs. 35 is an angular momentum. This denition is valid for both the
orbital angular momentum and the spin. A compact way of writing these
expressions, using the Levi-Civitta tensor, is explained in Appendix A11.1. In
this chapter we will use often a variety of expressions involving commutators;
Appendix A11.2 reviews the denition and derives several useful facts.
3

L
2
commutes with

L
x
,

L
y
, and

L
z
. We know from classical mechanics
that the rotational energy is proportional to the square of the angular mo-
mentum vector. Since this is an important quantity, we are naturally led to
examine the operator

L
2
=

L
2
x
+

L
2
y
+

L
2
z
(6)
This operator commutes with

L
x
,

L
y
, and

L
z
.
We start with
[

L
z
,

L
2
] = [

L
z
,

L
2
x
] + [

L
z
,

L
2
y
] + [

L
z
,

L
2
z
] (7)
Any operator commutes with a power of itself, so
[

L
z
,

L
2
z
] = 0 (8)
To calculate [

L
2
x
,

L
z
], we use the relationship, derived in Appendix A11.2
(Eq. 160),
[

A,

B
2
] = [

A,

B]

B +

B[

A,

B],
giving us
[

L
z
,

L
2
x
] = [

L
z
,

L
x
]

L
x
+

L
x
[

L
z
,

L
x
], (9)
1
M. Born, W. Heisenberg, and P. Jordan, Zur Quantenmechanik II, Zeitschrift f ur
Physik 35 (89), 557-615 (published August 1926; received November 16, 1925). English
translation in: B. L. van der Waerden, editor, Sources of Quantum Mechanics (Dover
Publications, 1968, ISBN 0-486-61881-1).
11. Angular Momentum Theory, November 20, 2009 3
followed by Eq. 5, to obtain
[

L
z
,

L
2
x
] = i h

L
y

L
x
+ i h

L
x

L
y
= i h
_

L
y

L
x
+

L
x

L
y
_
(10)
You can show similarly that
[

L
z
,

L
2
y
] = i h
_

L
y

L
x
+

L
x

L
y
_
(11)
Using Eqs. 811 in Eq. 7 gives
[

L
2
,

L
z
] = 0 (12)
In the same way, you can show that
[

L
2
,

L
x
] = 0 (13)
and
[

L
2
,

L
y
] = 0 (14)
Common sense tells us that if Eq. 12 is true, then Eqs. 13 and 14 must also
be true, because there is no physical agent that dierentiates between the
OZ and OX or OY directions.
4 The eigenstates of

L
2
and

L
z
. Since

L
2
commutes with

L
z
, there exists
a set of eigenfunctions of

L
2
that are also eigenfunctions of

L
z
. Because

L
z
does not commute with

L
x
or with

L
y
, a joint eigenstate of

L
2
and

L
z
cannot
be an eigenstate of

L
y
or of

L
z
.
Let us examine in more detail what this means. Since

L
2
and

L
z
commute,
there exists a set of kets |L
z
; , that satisfy the eigenvalue equations for
those two operators:

L
2
|L
z
; , = |L
z
; , (15)

L
z
|L
z
; , = |L
z
; , (16)
is the value that L
2
can have in a measurement. It is therefore related
to the rotational energy. is the value that the projection of the angular
momentum on the Z-axis can take. The symbol L
z
inside the ket reminds us
that this is an eigenket of

L
z
, not of

L
x
or

L
y
.
But

L
2
also commutes with

L
x
, so there must exist a set of kets |L
x
; ,
such that

L
2
|L
x
; , = |L
x
; , (17)

L
x
|L
x
; , = |L
x
; , (18)
11. Angular Momentum Theory, November 20, 2009 4
As long as no electric or magnetic eld acts on the system there is no physical
dierence between the OX and the OZ axes. Therefore, the values that
can take in |L
x
; , are the same as the values it can take in |L
z
; , .
In what follows, we concentrate on the joint eigenstates of

L
2
and

L
z
. At
this point there is no reason to prefer

L
z
over

L
x
or

L
y
.
Suppose that we have performed an experiment that places the system
in the state |L
z
; , . We can measure L
2
and get the value , or we can
measure L
z
and get the value . However, if we measure L
x
, when the system
is in state |L
z
; , , we nd L
x
to be

with the probability


|L
x
; ,

| L
z
; , |
2
When the system is in the state |L
z
; , , we know L
2
and L
z
with certainty
but we only know the probability that L
x
takes a given value. For someone
familiar with quantum mechanics, this is not surprising.

L
z
and

L
x
do not
commute so the eigenstate |L
z
; , cannot be an eigenstate of

L
x
.
These results are incomprehensible within classical physics, where there
is no physical law to prevent us from knowing with certainty all three pro-
jections of any vector once we know the state of the system (i.e. the values
of the position and of the momentum).
5

L
2
and

L
z
commute with other operators. One diculty in grasping
the physics of angular momentum comes from the fact that in many con-
crete problems,

L
x
,

L
y
, and

L
z
are not the only operators describing the
system. For example, we might examine two particles interacting through a
central force. This means that the potential energy V (r) depends only on
the distance r between the particles and is independent of their orientation
in space. You know two examples of such systems: the hydrogen atom and
the diatomic molecule.
The eigenvalue problem for the Hamiltonian of the hydrogen atom is (see
Metiu, Quantum Mechanics, p. 206)

h
2
2
1
r
2

r
_
r
2

r
_
+

L
2
2r
2
+ V (r) = E (19)
We see that besides the angular momentum squared

L
2
, the equation con-
tains the potential energy V (r) and the radial kinetic energy (the rst term).
You have also learned that the Hamiltonian

H of the hydrogen atom com-
mutes with

L
2
and with

L
z
. This means that

H,

L
2
, and

L
z
have common
11. Angular Momentum Theory, November 20, 2009 5
eigenstates, denoted by |n, j, m. In the state |n, j, m, the energy of the
atom is (see Metiu, p. 300)
E
n
=
e
2
z
2
2(4
0
)
2
h
2
1
n
2
, (20)
the square of the angular momentum is
h
2
j(j + 1), (21)
and the projection of the angular momentum on the z-axis is
hm (22)
In other words, we have

H|n, j, m =
e
2
z
2
2(4
0
)
2
h
2
1
n
2
|n, j, m (23)

L
2
|n, j, m = h
2
j(j + 1)|n, j, m (24)

L
z
|n, j, m = hm|n, j, m (25)
Here j indicates the eigenvalue of

L
2
and m indicates the eigenvalue of

L
z
.
We learn from this example that the eigenstates of

L
2
and

L
z
may also be
eigenstates of other observables that commute with

L
2
and

L
z
(in this case,
the energy). These eigenstates are labeled by additional quantum numbers
(n, in this case); they are |n, j, m, not just |j, m.
A similar situation appears in the case of the diatomic molecules. If we
make the rigid-rotor approximation and the harmonic approximation, the
eigenvalue equation for the energy, Eq. 19 becomes (see Metiu, Chapter 16)

h
2
2
1
r
2

r
_
r
2

r
_
+

L
2
2r
2
0
+
1
2
k(r r
0
)
2
= E (26)
The eigenstates are | = |v, j, m, in which the molecule has the vibrational
energy
E
v
= h
_
v +
1
2
_
(27)
and the values of

L
2
and

L
z
given by Eqs. 21 and 22. Eqs. 24 and 25 hold
for this example also.
11. Angular Momentum Theory, November 20, 2009 6
In general, when

L
2
and

L
z
commute with other operators, their eigen-
states must be eigenstates of those operators. When we need to make this
explicit, these eigenstates will be labeled |a, j, m, where a tells us the value
that the quantities represented by these other operators takes, when the
system is in a pure state |a, j, m and a measurement is made.
6 A more convenient notation. I now change notation by writing
= h
2
j(j + 1) (28)
= hm (29)
|L
z
; , = |L
z
; j, m (30)
The eigenvalue equations Eqs. 15 and 16 become

L
2
|L
z
; j, m = h
2
j(j + 1)|L
z
; j, m (31)
and

L
z
|L
z
; j, m = hm|L
z
; j, m (32)
Setting = h
2
j(j + 1) needs further examination. We must have
0 (33)
because is a value that L
2
can take. It is a minor complication that the
equation = h
2
j(j + 1) has two solutions for j:
j
1
=
1
2
_
1 +
_
1 + 4/ h
2
_
(34)
and
j
2
=
1
2
_
1 +
_
1 + 4/ h
2
_
(35)
Because / h
2
0, both roots j
1
and j
2
are real, j
1
is negative, and j
2
is non-
negative. Remember that is a physical observable while j is an arbitrary
mathematical quantity introduced for convenience. You can verify that the
negative values of j give the same values of as the non-negative ones. We
can choose either for indexing the kets, and we choose
j 0 (36)
Note also that m must be a real number ( hm is a value of L
z
). Because
L
z
is equally likely to be positive or negative, we expect m to be able to take
11. Angular Momentum Theory, November 20, 2009 7
both positive and negative values. Moreover, we guess that if m is an allowed
value, then m is also allowed since the projection of the angular momentum
vector on the OZ axis can be oriented along OZ (m > 0) or opposite to it
(m < 0).
7 A summary of our task. To set up the theory of angular momentum,
we need to do the following.
1. Since most calculations in quantum mechanics are done by representing
operators by matrices, we need to use the kets |L
z
; j, m as a basis set
and nd out how to calculate the matrix elements L
z
; j, m|

O| L
z
; j

, m

for

O =

L
x
and

L
y
.
2. Find the values of j and m allowed by the eigenvalue equations Eqs. 31
32.
To reach these objectives, we only have the commutation relations Eqs. 3
5 and our wits. The strategy is simple but tedious. We know how

L
2
and

L
z
act on |L
z
; j, m. Therefore we must use the commutation relations to
express

L
x
and

L
y
in terms of

L
2
and

L
z
. Determining the allowed values of
j and m is more subtle and cannot be reduced to a few sentences.
8 The ladder operators. In vector calculus, one often describes a vector
{v
x
, v
y
, v
z
} through its spherical components {v
x
+iv
y
, v
x
iv
y
, v
z
}. It might
therefore be useful to take a look at the operators

L
+


L
x
+ i

L
y
(37)
and



L
x
i

L
y
(38)
along with

L
z
and

L
2
. If we nd out how

L
+
and

L

act on |L
z
; j, m, then
it is easy to use

L
x
=

L
+
+

L

2
(39)
and

L
y
=

L
+

2i
(40)
to nd out how

L
x
and

L
y
act on |L
z
; j, m. Working with

L
+
and

L

will
prove to be fruitful.
11. Angular Momentum Theory, November 20, 2009 8
Our strategy relies on commutation relations and I derive here all the
commutations relations that we need. The commutators of

L
+
and

L

with

L
2
and

L
z
are easily derived from the denitions, Eqs. 37 and 38, and the
original commutation relations, Eqs. 35. They are
[

L
z
,

L
+
] = h

L
+
(41)
and
[

L
z
,

L

] = h

(42)
We also have
[

L
2
,

L
+
] = [

L
2
,

L

] = 0 (43)
because

L
2
commutes with both

L
x
and

L
y
, hence with both

L
+
and

L

.
To carry out our program of determining what happens when

L
+
and

L

act on |L
z
; j, m, we need to nd expressions that contain

L
+
and/or

L

in
the left-hand side and only

L
z
and

L
2
in the right-hand side. I derive below
several such expressions. One of them is

L
+
= (

L
x
i

L
y
)(

L
x
+ i

L
y
) (used Eqs. 37&38)
=

L
2
x
+

L
2
y
+ i(

L
x

L
y


L
y

L
x
)
=

L
2
x
+

L
2
y
+ i[

L
x
,

L
y
]
=

L
2
x
+

L
2
y
+ i
2
h

L
z
(used Eq. 3)
=

L
2


L
2
z
h

L
z
(used

L
2
=

L
2
x
+

L
2
y
+

L
2
z
) (44)
Similarly

L
+

=

L
2


L
2
z
+ h

L
z
(45)
We can add Eqs. 44 and 45 to obtain

L
+

+

L

L
+
2
=

L
2


L
2
z
(46)
Using

+
=

L

(47)
and

=

L
+
(48)
we can rewrite Eq. 46 as

+

L

L
+
2
=

L
2


L
2
z
(49)
11. Angular Momentum Theory, November 20, 2009 9
9 Find expressions for

L
+
|L
z
; j, m and

L

|L
z
; j, m. For notational
convenience, I will drop the marker L
z
from the eigenkets |L
z
; j, m and
reinstate it only when there is some possibility of confusion.
The fact that

L
2
and

L
+
commute leads to

L
2

L
+
|j, m =

L
+

L
2
|j, m = h
2
j(j + 1)

L
+
|j, m (50)
This equation tells us that if |j, m is an eigenstate of

L
2
with the eigenvalue
h
2
j(j +1) then

L
+
|j, m is also an eigenstate of

L
2
, with the same eigenvalue
h
2
j(j + 1).
Since [

L
z
,

L
+
] = h

L
+
(Eq. 41), we have
_

L
z

L
+


L
+

L
z
_
|j, m = h

L
+
|j, m (51)
Use

L
z
|j, m = hm|j, m in the left-hand side and rearrange the terms:

L
z
_

L
+
|j, m
_
= h(m+ 1)
_

L
+
|j, m
_
(52)
Eq. 52 tells us that

L
+
|j, m is either zero, which is uninteresting, or it is an
eigenstate of

L
z
with the eigenvalue h(m+ 1). This means that

L
+
|j, m = C
+
(j, m)|j, m+ 1 (53)
where C
+
(j, m) is a number that may or may not depend on j and m.
Note that in Eq. 53,

L
+
acts on |j, m and does not aect the value of j
but changes m. This is in agreement with Eq. 50, which says that

L
+
|j, m
is an eigenfunction of

L
2
with the eigenvalue h
2
j(j + 1).
A similar argument starts from [

L
z
,

L

] = h

(Eq. refeq11:3.11.4) and


concludes that

|j, m = C

(j, m)|j, m1 (54)


where C

(j, m) is another, unknown constant. Also,



L

|j, m is an eigen-
function of

L
2
corresponding to the eigenvalue h
2
j(j + 1).
The operators

L
+
and

L

are called ladder operators;



L
+
is called a raising
operator and

L

, a lowering operator. When



L
+
acts on |j, m, it increases
m by 1 and multiplies the result by C
+
(j, m); when

L

acts on |j, m, it
decreases m by 1 and multiplies the result by C

(j, m). The operators



L
+
and

L

do not aect j in |j, m, because both



L
+
and

L

commute with

L
2
and therefore they have common eigenstates with

L
2
. In other words, they
must convert a state in which

L
2
has the value h
2
j(j + 1) into a state in
which

L
2
has the same value h
2
j(j + 1).
11. Angular Momentum Theory, November 20, 2009 10
10 The constants C
+
(j, m) and C

(j, m). To determine C


+
(j, m) or
C

(j, m), we need to derive rst a few helpful results. Consider


| =

A|x (55)
where |x is an arbitrary ket and

A is an arbitrary operator. We have
| =

Ax|

Ax = x|

A


A| x (56)
I used here the property

Ax| z = x|

A

z of adjoint operators (|x and |z


are arbitrary kets).
In addition, if

A|x = a|y (57)


where a is a number, then
| = ay | ay = a

ay | y (58)
Let us use Eqs. 56 and 58 to determine C
+
. To do this I apply Eq. 56 to
|

L
+
|j, m (59)
and obtain
| = j, m|

L

L
+
| j, m (60)
We also have
=

L
+
|j, m = C
+
|j, m+ 1 (61)
Now apply Eq. 58 to get
| = C

+
C
+
j, m+ 1 | j, m+ 1 = C

+
C
+
(62)
I used the fact that j, m+ 1 | j, m+ 1 = 1 (or, if |j, m+ 1 is zero, then so
is C
+
). The two expressions for | must be equal and therefore
j, m|

L

L
+
| j, m = C

+
C
+
(63)
I am very pleased with this equation, for two reasons. First, I obtained
a formula for C
+
. Second, this equation contains

L
+

, which, according to
Eq. 44 is

L
+

=

L
2


L
2
z
h

L
z
11. Angular Momentum Theory, November 20, 2009 11
This expresses

L
+

in terms of

L
2
and

L
z
. I know how

L
2
and

L
z
act on
|j, m and therefore I can readily calculate the matrix element in Eq. 63.
Eq. 63 determines the absolute value of C
+
but not its phase. Since all
physical properties of a ket are independent of a phase factor in front of it,
I can take in C
+
any phase I want. I choose it so that
C
+
(j, m) =
_
j, m|

L

L
+
| j, m (64)
Now use Eq. 44 in Eq. 64:
C
+
(j, m) =
_
j, m|
_

L
2


L
2
z
h

L
z
_
| j, m (65)
Using the eigenvalue equations for

L
2
and

L
z
(Eqs. 3132), we rewrite Eq. 65
as
C
+
(j, m) = h
_
j(j + 1) m(m+ 1) (66)
(we used j, m| j, m = 1).
Finally, putting C
+
(j, m) given by Eq. 66 into Eq. 53 gives

L
+
|j, m = h
_
j(j + 1) m(m+ 1) |j, m+ 1 (67)
Note that

L
+
|j, m is an eigenstate of

L
2
corresponding to the eigenvalue
hj(j + 1) and an eigenstate of

L
z
corresponding to the eigenvalue (m+ 1) h.
However, |j, m is not an eigenstate of

L
+
. It should not be, because

L
+
does not commute with

L
2
or with

L
z
.
To calculate C

(j, m), we follow the reasoning used to nd C


+
(j, m). This
leads to
C

(j, m) = h
_
j(j + 1) m(m1) (68)
and

|j, m = h
_
j(j + 1) m(m1) |j, m1 (69)
We can now calculate (use Eqs. 66, 69, 39)

L
x
|j, m =
1
2
(

L
+
+

L

)|j, m
=
1
2
[C
+
(j, m)|j, m+ 1 +C

(j, m)|j, m1] (70)


Similarly (use Eqs. 66, 69, 40)

L
y
|j, m =
1
2i
[C
+
(j, m)|j, m+ 1 C

(j, m)|j, m1] (71)


11. Angular Momentum Theory, November 20, 2009 12
11 The matrix elements of

L
+
,

L

,

L
x
,

L
y
in the basis set |j, m. Now
that we know how

L
+
and

L

act on |j, m (Eqs. 67 and 69), we can calculate


the matrix elements
j

, m

|

L
+
| j, m = h
_
j(j + 1) m(m+ 1) j

, m

| j, m+ 1
= h
_
j(j + 1) m(m+ 1)
j

,m+1
(72)
The only non-zero matrix elements are those for which m

= m+1 and j

= j.
Therefore only
j, m+ 1 |

L
+
| j, m = h
_
j(j + 1) m(m+ 1) (73)
diers from zero.
The matrix elements of

L

are
j

, m

|

L

| j, m = h
_
j(j + 1) m(m1)
j

,m1
(74)
and only
j, m1 |

L

| j, m = h
_
j(j + 1) m(m1) (75)
diers from zero.
Since (see Eq. 39)

L
x
=

L
+
+

L

2
we have
j

, m

|

L
x
| j, m =
h
2
(C
+
(j, m)
j

,m+1
+C

(j, m)
j

,m1
) (76)
Only the matrix elements
j, m+ 1 |

L
x
| j, m =
h
2
C
+
(j, m) (77)
and
j, m1 |

L
x
| j, m =
h
2
C

(j, m) (78)
dier from zero.
Using (see Eq. 40)

L
y
=

L
+

2i
11. Angular Momentum Theory, November 20, 2009 13
with Eqs. 72 and 74 leads to
j

, m

|

L
y
| j, m =
h
2i
(C
+
(j, m)
j

,m+1
C

(j, m)
j

,m1
) (79)
Only the matrix elements
j, m+ 1 |

L
y
| j, m =
h
2i
C
+
(j, m) (80)
and
j, m1 |

L
y
| j, m =
h
2i
C

(j, m) (81)
dier from zero.
This covers all the matrix elements used in applications. However, we
still dont know what values j and m can take, since we do not know the
eigenvalues of

L
2
and

L
z
. Next, we start working to nd them.
The values of j and m
We have decided to use the eigenstates |j, m of

L
2
and

L
z
as a basis set
for other operators relevant to angular momentum, and we have managed
to calculate their matrix elements. The remaining task is to determine the
allowed values of j and m.
12 For a given j, m has a highest and a lowest value. A classical inter-
pretation of angular momentum tells us that if

L
2
is xed then

L
z
must have
an upper and a lower bound. Indeed the projection of a vector on an axis
can at most be equal to the length of the vector: the projection has the
upper bound and the lower bound . Since quantum mechanics is not
classical mechanics, it will be reassuring to prove that, for a given j, m has
an upper and a lower bound.
Recall that (Eqs. 46 and 49)

L
+

+

L

L
+
2
=

L
2


L
2
z
(82)
and

+

L

L
+
2
=

L
2


L
2
z
(83)
11. Angular Momentum Theory, November 20, 2009 14
From properties of the inner product, we know that for any ket |x and
any operator

A, we have x|

A


A| x 0 (with equality only when

A|x = 0).
Taking the matrix element of the operator in Eq. 83 with the non-zero ket
|j, m gives
j, m|

L

| j, m +j, m|

L

L
+
| j, m = 2 h
2
(j(j + 1) m
2
) (84)
Since the left-hand side cannot be negative, we conclude that
j(j + 1) m
2
0 (85)
This means that if j is xed then m has both a lower bound and an upper
bound. We dont know what the bounds are but we know that they exist, so
let us denote them by m

and m
u
.
This gives us two new equations to play with:

L
+
|j, m
u
= 0 (86)

|j, m

= 0 (87)
To derive them, observe that, according to Eq. 53,

L
+
|j, m
u
= C
+
(j, m
u
)|j, m
u
+
1. Since no value of m can be larger than m
u
, the state |j, m
u
+ 1 can-
not exist and the ket is zero (which proves Eq. 86). One arrives at Eq. 87
similarly.
Now let us act with

L

on Eq. 86 and with



L
+
on Eq. 87, to obtain

L
+
|j, m
u
= 0 (88)

L
+

|j, m

= 0 (89)
Why do I do that? Because we know that (Eq. 44)

L
+
=

L
2


L
2
z
h

L
z
;
the right-hand side of this equation contains only operators that act on |j, m
according to known rules, which means that we can calculate
0 =

L

L
+
|j, m = (

L
2


L
2
z
h

L
z
)|j, m
= ( h
2
j(j + 1) h
2
m
2
u
h
2
m
u
)|j, m
u
(90)
This implies that
j(j + 1) m
2
u
m
u
= 0 (91)
11. Angular Momentum Theory, November 20, 2009 15
Similarly, using (see Eq. 45)

L
+

=

L
2


L
2
z
+ h

L
z
in Eq. 89 gives
j(j + 1) m
2

+m

= 0 (92)
Solving Eq. 92 for j(j + 1) and using the result in Eq. 91 gives
m
u
(m
u
+ 1) = m

(m

1) (93)
This equation has two solutions
m
u
= m

(94)
and
m
u
= m

1 (95)
The second root is excluded by the fact that m
u
m

by denition.
We see therefore that m
u
= m

. Symmetry requires the possible pro-


jections of

L along OZ to be equal to and of opposite sign to the possible
projections in the direction opposite to OZ. Eq. 94 agrees with this require-
ment.
13 The values of m and j. We have learned a lot but we still dont know
which values j and m can take. Here we settle this question.
Let us consider a ket |j, m
0
where m
0
is an allowed value of m for the
given j. Acting on that ket with

L
+
repeatedly, we obtain

L
+
|j, m
0
= C
+
(j, m
0
)|j, m
0
+ 1,

L
2
+
|j, m
0
= C
+
(j, m
0
)C
+
(j, m
0
+ 1)|j, m
0
+ 2,
etc. After some number n
u
of steps, this procedure must generate a ket
proportional to |j, m
u
. If it did not, then we could apply

L
+
indenitely,
which contradicts the fact that m has a largest value, namely m
u
. Therefore
we must have
m
u
= m
0
+n
u
(96)
where n
u
is a non-negative integer. Similarly, applying

L

to |j, m
0
some n

times produces a ket proportional to |j, m

, so that
m

= m
0
n

(97)
11. Angular Momentum Theory, November 20, 2009 16
with n

a non-negative integer. From Eqs. 96 and 97, it follows that


m
u
m

= (m
0
+n
u
) (m
0
n

) = n
u
+n

n (98)
where n is a non-negative integer. Since we already know that m
u
= m

(Eq. 94), we conclude that


2m
u
= n (99)
We also know that (see Eq. 91)
j(j + 1) = m
2
u
+m
u
= m
u
(m
u
+ 1) (100)
This has two solutions
j = m
u
(101)
and
j = m
u
1 (102)
Eq. 102 is not compatible with the facts that m
u
0 and j 0, so we can
eliminate it as irrelevant.
Now combining m
u
= n/2 (Eq. 99) with Eq. 101 gives
j =
n
2
(103)
We have obtained a startling result: j can be either a non-negative integer
(if n is even) or a half-integer (if n is odd). In addition, the largest value of
m is j and (since m

= m
u
) the smallest value of m is j. Therefore m
can take the 2j + 1 values j, j + 1,. . . , j 1, j.
The theory of orbital angular momentum (see Metiu, Chapter 14, p. 220),
based on the equation

L = r p, tells us that half-integer values of j are not
allowed for orbital angular momentum. However, by using the commutation
relations to dene the angular momentum, we have found that j can be ei-
ther integer or half-integer. We know of no additional condition that we can
use to conclude that half-integers are not allowed in general. Therefore the
postulate that the electron has an intrinsic angular momentum (the spin)
with j = 1/2 is in harmony with conclusions derived from the commutation
relations.
14 Summary. This lengthy derivation led us to the following results.

L
2
and

L
z
have common eigenstates |

L
z
; j, m for which

L
2
|

L
z
; j, m = h
2
j(j + 1) |

L
z
; j, m (104)
11. Angular Momentum Theory, November 20, 2009 17
and

L
z
|

L
z
; j, m = hm|

L
z
; j, m. (105)
j can be either a non-negative integer or a positive half-integer (that is, of
the form integer/2). m can take 2j + 1 discrete values
m = j, j + 1, j + 2, . . . , j 2, j 1, j (106)
For example, if j =
3
2
then m can take 2
3
2
+ 1 = 4 values, namely
m =
3
2
,
3
2
+ 1,
3
2
1,
3
2
=
3
2
,
1
2
,
1
2
,
3
2
If j = 3 then m takes the 2 3 + 1 = 7 values 3, 2, 1, 0, 1, 2, 3.
We have also found that

L
x
|

L
z
; j, m =
1
2
[C
+
(j, m)|

L
z
; j, m+ 1 +C

(j, m)|

L
z
; j, m1] (107)
and

L
y
|

L
z
; j, m =
1
2i
[C
+
(j, m)|

L
z
; j, m+ 1 C

(j, m)|

L
z
; j, m1] (108)
with
C
+
(j, m) = h
_
j(j + 1) m(m+ 1) (109)
and
C

(j, m) = h
_
j(j + 1) m(m1). (110)
Note that |j, m = 0 if m [j, j].
The matrix elements needed in computations are

L
z
; j

, m

|

L
z
|

L
z
; j, m =
jj

mm
hm (111)

L
z
; j

, m

|

L
2
|

L
z
; j, m =
jj

mm
h
2
j(j + 1) (112)

L
z
; j

, m

|

L
x
|

L
z
; j, m =
jj

_
C
+
(j, m)
m

,m+1
+C

(j, m)
m

,m1
2
_
(113)

L
z
; j

, m

|

L
y
|

L
z
; j, m =
jj

_
C
+
(j, m)
m

,m+1
C

(j, m)
m

,m1
2i
_
(114)
Note that all these matrix elements are zero if j = j

. This means that


in the basis set |

L
z
; j, m, the matrices of

L
x
and

L
y
consist of smaller
matrices strung along the diagonal; there are no non-zero matrix elements
j, m|

Oj

, m

in which j = j

if

O =

L
x
,

L
y
,

L
+
, or

L

or any function of
them.
11. Angular Momentum Theory, November 20, 2009 18
The eigenvalue problems for angular momentum operators
15 Introduction. We have seen several times that kets are very useful for
theoretical analysis, but practical calculations often use the representation of
kets by vectors and of operators by matrices. In this section I review briey
the general theory and then apply it to the case of j = 3 as an example.
Let us assume that on physical grounds we have decided that a certain
physical system can be described by a ket space generated by the orthonormal
basis set |
1
, |
2
, . . . , |
N
. An arbitrary state | is then represented as
| =
N

i=1
|
i

i
|
N

i=1
|
i
c
i
(115)
and an arbitrary operator

O by

O =
N

i=1
N

k=1
|
i

i
|

O|
k

k
|
N

i=1
N

k=1
|
i
O
ik

k
| (116)
Since we are assumed to know |
i
, i = 1, 2, . . . , N, knowing the vector
= {c
1
, c
2
, . . . , c
N
} (117)
is equivalent to knowing | (because of Eq. 115). Eq. 116 tells us that we
know how to operate with

O if we know the matrix elements
O
ik

i
|

O|
k
(118)
The basis set kets |
i
are represented by the vectors

1
= {1, 0, . . . , 0}
.
.
.

N
= {0, 0, . . . , 1}

(119)
and therefore
=
N

i=1
c
i

i
(120)
The eigenvalue problem

O| = o| is represented by the eigenvalue problem
for the matrix O

k
O
ik
c
k
= o c
i
,
11. Angular Momentum Theory, November 20, 2009 19
which in vector-and-matrix notation is written as
O = o (121)
or

O
11
O
1N
.
.
.
.
.
.
O
N1
O
NN

c
1
.
.
.
c
N

= o

c
1
.
.
.
c
N

(122)
The solution of Eq. 122 provides N eigenvectors
() {c
1
(), . . . , c
N
()}, = 1, . . . , N (123)
and N eigenvalues o(1), . . . , o(N).
From these we can calculate the N eigenkets |(1), . . . , |(N) by using
|() =
N

k=1
c
k
()|
k
, = 1, 2, . . . , N (124)
16 The matrix-and-vector representation for angular momentum problems.
It is natural to use for angular momentum calculations the basis set |j, m.
In this basis set, the matrix elements of

L
z
,

L
2
,

L
x
, and

L
y
are given by
Eqs. 111114. When looking at these equations, we notice that all matrix
elements in which j = j

are zero. The matrices of the operators



L
2
,

L
z
,

L
x
,
and

L
y
are therefore block diagonal:









.
.
.

All matrix elements other than the dots are zero.


We have labeled the rows and the columns in the following order:
11. Angular Momentum Theory, November 20, 2009 20
j = 0, m = 0
j = 1, m = 1
j = 1, m = 0
j = 1, m = 1
etc.
The state with j = 0 generates a 1 1 block, the states with j = 1 generate
a 3 3 block, . . . ; in general, the states having an arbitrary j generate a
(2j + 1) (2j + 1) block. This break-up into blocks makes this basis set
particularly simple to use in calculations. Physically it signies that states
with dierent values of j are not coupled by any of the operators

L
2
,

L
z
,

L
x
,
and

L
y
.
Because of this block property, we can treat physical problems involving
a specic j within a reduced space generated by the basis set
|j, j, |j, j + 1, . . . , |j, j 1, |j, j (125)
17 The case j = 1. Let us solve some problems for j = 1. The basis set
is
|1, 1, |1, 0, |1, 1 (126)
Since we know that j = 1, the rst index is superuous and we use instead
the notation | 1, |0, |1. The matrix representing an operator

O in this
subspace is
O =

1 |

O| 1 1 |

O| 0 1 |

O| 1
0 |

O| 1 0 |

O| 0 0 |

O| 1
1 |

O| 1 1 |

O| 0 1 |

O| 1

(127)
A look at Eqs. 111 and 112 tells us that the matrix elements 1, m|

L
2
| 1, m

and 1, m|

L
z
| 1, m

are zero if m = m

. Therefore the corresponding matri-


ces are diagonal:
L
2
= h
2
(1)(1 + 1)

1 0 0
0 1 0
0 0 1

(128)
11. Angular Momentum Theory, November 20, 2009 21
and
L
z
=

h 0 0
0 0 0
0 0 h

(129)
No surprise here. The basis set consists of eigenstates |j, m of

L
2
and

L
z
and therefore the matrices of these operators, in that basis set, are diagonal
and the diagonal elements are the eigenvalues of the operator.
The matrices corresponding to

L
x
and

L
y
are more interesting. Let us
look at

L
x
, and use Eq. 113 together with Eqs. 109 and 110. I calculate, as
an example, 1 |

L
x
| 1. We have
m

= 1 |

L
x
| m = 1 = 0 (130)
because (see Eq. 113)

,m+1
=
1,1+1
=
1,0
= 0
and

,m1
=
1,11
=
1,2
= 0
Another example is (use Eq. 113)
m

= 1 |

L
x
| m = 0 =
C
+
(1, 0)
1,0+1
+C

(1, 0)
1,01
2
=
C

(1, 0)
2
(131)
with
C

(1, 0) = h
_
1(1 + 1) 0(0 1) = h

2 (132)
Together these give
1 |

L
x
| 0 =
h

2
(133)
Patiently calculating all the matrix elements leads to
2
L
x
=

0
h

2
0
h

2
0
h

2
0
h

2
0

(134)
2
The top row is m

= 1, the middle row is m

= 0, and the bottom row is m

= 1;
the left column is m = 1, the middle column is m = 0, and the right column is m = 1.
In all entries j = 1.
11. Angular Momentum Theory, November 20, 2009 22
These matrix elements were calculated in WorkBook11.Angular Momentum.nb.
The matrix must be Hermitian because

L
x
represents an observable, and
therefore we need only calculate the matrix elements in the lower-left triangle;
the ones in the upper-right triangle are then obtained from m

|

O| m =
m|

O| m

. In addition, the s in Eq. 113 tell us that the diagonal elements


(that is, those with m

= m) are zero and that in the lower-left triangle, only


0 |

L
x
| 1 and 1 |

L
x
| 0 are nonzero.
Exercise 1 Show that for j = 1
L
y
=
h

0 i 0
i 0 i
0 i 0

(135)
Is the matrix Hermitian?
Exercise 2 Verify that the matrices representing the operators satisfy the
commutation relations
L
x
L
y
L
y
L
x
= i hL
z
L
y
L
z
L
z
L
y
= i hL
x
L
z
L
x
L
x
L
z
= i hL
y
and that
L
2
= L
2
x
+L
2
y
+L
2
z
18 The eigenvalues and eigenvectors of

L
x
for j = 1. Now that we
have a matrix representation of

L
x
, we can calculate the eigenvalues and
eigenvectors of this operator. We need to know them because they appear
in some problems in spectroscopy. I solved the eigenvalue problem in Cell 6
of the Mathematica le WorkBook11. Angular momentum.nb. The results
are (Cell 6 of WorkBook11):
eigenvalue eigenvector
h {
1
2
,
1

2
,
1
2
}
0 {
1

2
, 0,
1

2
}
h {
1
2
,
1

2
,
1
2
}
11. Angular Momentum Theory, November 20, 2009 23
The eigenvectors are normalized. I remind you that in this basis set the
eigenvectors of L
z
are {1, 0, 0}, {0, 1, 0}, and {0, 0, 1}.
We are pleased to see that L
x
has the same eigenvalues as L
z
. Since no
force that breaks spherical symmetry acts on the system, there is no preferred
direction in space. The possible values of the projection of the vector

L on
the x-axis must be the same as those of the projection on the z-axis.
Exercise 3 Show that L
y
has the same eigenvalues as L
z
.
The eigenstates of

L
x
in the basis set |L
z
; 1, m are (see WorkBook11, Cell
6)
eigenvalue eigenket
h |L
x
; 1, 1 =
1
2
|L
z
; 1, 1
1

2
|L
z
; 1, 0 +
1
2
|L
z
; 1, 1 (136)
0 |L
x
; 1, 0 =
1

2
|L
z
; 1, 1 +
1

2
|L
z
; 1, 1 (137)
h |L
x
; 1, 1 =
1
2
|L
z
; 1, 1 +
1

2
|L
z
; 1, 0 +
1
2
|L
z
; 1, 1 (138)
I have had to augment the notation with a new index: |L
z
; j, m are eigenkets
of

L
z
, and |L
x
; j, m are those of

L
x
.
If we manage to place the system in the state |L
x
; 1, 1, a measure-
ment of L
x
is guaranteed to give the result h. However, if we measure L
z
,
when the system is in the state |L
x
; 1, 1, we obtain the value h with the
probability
P
1
|L
z
; 1, 1 | L
x
; 1, 1|
2
=
_
1
2
L
z
; 1, 1 | L
z
; 1, 1
1

2
L
z
; 1, 1 | L
z
; 1, 0 +
1
2
L
z
; 1, 1 | L
z
; 1, 1
_
2
=
1
4
(139)
To get this, I used the orthonormality condition L
z
; 1, m| L
z
; 1, m

=
mm
.
Similarly, the probability P
0
that L
z
is zero is given by
P
0
|L
z
; 1, 0 | L
x
; 1, 1|
2
=
1
2
(140)
11. Angular Momentum Theory, November 20, 2009 24
and the probability P
1
that L
z
is 1 by
P
1
|L
z
; 1, 1 | L
x
; 1, 1|
2
=
1
4
(141)
P
1
+P
0
+P
1
= 1, which is a must.
This result is consistent with the fact that

L
z
does not commute with

L
x
and therefore there is no state in which the value of both is known with
certainty.
We can also calculate the average value of L
z
,
L
z
L
x
; 1, 1 |

L
z
| L
x
; 1, 1
when the system is in the state |L
x
; 1, 1. Using Eq. 136 in the right-hand
side (for |L
x
; 1, 1) gives
L
x
; 1, 1 |

L
z
| L
x
; 1, 1 = L
x
; 1, 1 |

L
z
_
1
2
_
| L
z
; 1, 1
L
x
; 1, 1 |

L
z
_
1

2
_
| L
z
; 1, 0
+L
x
; 1, 1 |

L
z
_
1
2
_
| L
z
; 1, 1
Using Eq. 136 again, as well as L
z
; j, m

|

L
z
| L
z
; j, m =
mm
hm, leads to
L
z
=
_
1
2
_
( h)
_
1
2
_

_
1

2
_
(0) +
_
1
2
_
( h)
_
1
2
_
= 0 (142)
This is the same as
L
z
=
1

m=1
hmP
m
= ( h)
1
4
+ (0)
1
2
+ ( h)
1
4
= 0, (143)
as it should be This result is physically reasonable. Since no forces break
spherical symmetry, there is no bias to prefer positive values of L
z
over
negative values. The average is zero because the value h is as probable as
h.
19 The energy of the electron in the hydrogen atom exposed to a magnetic
eld. When we discuss the hydrogen atom, I will show that when the atom
11. Angular Momentum Theory, November 20, 2009 25
is exposed to a magnetic eld B, the energy of its electron changes. The
Hamiltonian is

H =

H
0
+
e
2m
e

L B (144)
where m
e
is the mass of the electron, e is the proton charge, and

H
0
is the
Hamiltonian of the atom in the absence of the eld.
I am ignoring other terms in the Hamiltonian since I only want to illus-
trate how we handle this kind of problem. The question is: what happens to
the energy of the electron when it is exposed to the magnetic eld? I answer
rst the question for the case when I was smart enough to pick the z-axis
along B. In this case B = {0, 0, B} and

L B =

L
z
. The Hamiltonian is

H =

H
0
+
e
2m
e

L
z
B,
where B is the magnitude of the eld B. Since

L
z
and

H
0
commute, they
have joint eigenstates, denoted by |n, j, m, where n gives the energy E
n
in
the absence of the eld.
I want to know what happens to the energy of the states |2, j, m when
the eld is on. To answer this question, I will use these states as a basis set.
Therefore I have

H|n, j, m =

H
0
|n, j, m +
eB
2m
e

L
z
|n, j, m
=
_
E
n
+
eB
2m
e
m
_
|n, j, m (145)
We see that |n, j, m are eigenstates of the Hamiltonian dened by Eq. 144.
The energy spectrum is given by
E
n
+
eB
2m
e
m (146)
As noted above, E
n
is the energy of the atom in the absence of the eld:
E
n
=
Ry
n
2
, (147)
where Ry is the Rydberg constant. For the case n = 2, in the absence of the
eld, the atom has four degenerate states: |2, 0, 0, |2, 1, 1, |2, 1, 0, and
11. Angular Momentum Theory, November 20, 2009 26
|2, 1, 1. If the eld is on, the energy of these states changes to that given by
Eq. 146 and depends on the quantum number m. The states are
state |n, j, m energy change
|2, 0, 0 Ry/4 unchanged
|2, 1, 1 Ry/4 (eB/2m
e
) h decreased
|2, 1, 0 Ry/4 unchanged
|2, 1, 1 Ry/4 + (eB/2m
e
) h increased
The energy-level diagram changes, when the eld is turned on, from having
four states of the same energy E
2
to having three energy levels, E
2
+eB h/2m
e
,
E
2
, and E
2
eB h/2m
e
. The level having the energy E
2
is doubly degenerate:
the states |2, 0, 0 and |2, 1, 0 have energy E
2
.
20 Physical consequences. In the absence of the eld, an atom excited in
any one of the levels of energy E
2
will emit a photon of frequency
(E
2
E
1
)/ h =

Ry
4

_

Ry
1
__
/ h =
Ry
h
_
4 1
4
_
(148)
When the atom is in a magnetic eld, the E
2
level splits into three levels
as shown in Fig. 1. In the absence of the magnetic eld, the atom excited
in a state with n = 2 emits a photon of frequency
2
. When the eld is
present, three emission frequencies,
1
,
2
, and
3
are possible. Not all of
them will be observed because of selection rules that render the rate of some
transitions equal to zero.
3
21 Using a dumb choice of axis. We decided to use as a basis set the
eigenstates |L
z
; n, j, m of

L
z
. Then we wisely decided that the z-axis coin-
cides with the vector B, so that the interaction energy between the electron
and the eld became (e/2m
e
)B

L
z
. The Hamiltonian

H =

H
0
+(e/2m
e
)B

L
z
commutes with

L
z
and with

L
2
and has |L
z
; n, j, m as eigenstates. This
makes the matrix of

H, in the chosen basis, diagonal. The eigenvalue prob-
lem is then trivial.
3
The spectrum of the hydrogen atom in a magnetic eld will be discussed in detail in
a future chapter. For a complete discussion we need to take into account electron spin,
proton spin, and some relativistic eects, which add terms to the Hamiltonian in addition
to the ones used here.
11. Angular Momentum Theory, November 20, 2009 27
E
2,1
E
2,0
E
2,-1
E
1,0
|2,1,1
|2,1,0 and |2,2,0
|2,1,-1
|1,0,0
Figure 1: The energy levels of a hydrogen atom in a magnetic eld for n = 1
and n = 2
11. Angular Momentum Theory, November 20, 2009 28
What would happen if we were not so clever and decided to take the
x-axis along B? We have B = {B, 0, 0} and the Hamiltonian is

H =

H
0
+ (e/2m
e
)B

L
x
(149)
The basis set is still |L
z
; n, j, m.

H no longer commutes with

L
z
and the
basis-set kets are no longer eigenstates of

H. The matrix of

H, in the basis
set |L
z
; n, j, m, is no longer diagonal.
The matrix elements are given by
L
z
; n, j, m|

H | L
z
; n, j, m

=
Ry
n
2
+
eB
2m
e
L
z
; n, j, m|

L
x
| L
z
; n, j, m

(150)
For n = 2, we have j = 0 and m = 0, or j = 1 and m = 1, 0, 1. We have
listed the equations giving the matrix elements L
z
; 2, j, m|

L
x
| L
z
; 2, j, m

in
14 (see Eqs. 113, 109, 110). The matrix H corresponding to the Hamiltonian
in Eq. 149 is (see WorkBook11)
|2, 0, 0 |2, 1, 1 |2, 1, 0 |2, 1, 1
|2, 0, 0
|2, 1, 1
|2, 1, 0
|2, 1, 1

Ry/4 0 0 0
0 Ry/4
eB
2me
h/

2 0
0
eB
2me
h/

2 Ry/4
eB
2me
h/

2
0 0
eB
2m
e
h/

2 Ry/4

The eigenvalues of H were calculated in Cell 7 of WorkBook11. They are


Ry/4, Ry/4, Ry/4
eB
2m
e
h, and Ry/4 +
eB
2m
e
h. They are exactly the
same as the values obtained when we took the z-axis along B. They must
be, because the coordinate system and its axes are only in our heads. The
atom does not know about them. The energies are measurable quantities and
if they depend on our choice of axes, the theory would be erroneous. Note
however that the eigenvectors of H with B taken along OX dier from the
eigenvectors of H with B taken along OZ. This is alright. The eigenvectors
are not measurable quantities.
You see here a conspicuous feature of quantum mechanics. The mathe-
matics depends on the choice of basis set and coordinate axes, but the physics
does not.
11. Angular Momentum Theory, November 20, 2009 29
Appendix A11.1. A compact form of the commutation relations
It is sometimes useful to write the commutation relations Eqs. 35 in a
more compact form. To do this we write the components of the vector

L as

L
1
,

L
2
, and

L
3
, with

L
1


L
x
,

L
2


L
y
, and

L
3


L
z
. Then we can write
[

,

L
j
] = i h
3

k=1

jk

L
k
, = 1, 2, 3 (151)
Here
jk
is the Levi-Civitta tensor. Since this symbol appears often in
physics, it is worth explaining it in detail. To do so, we need to introduce
the concept of permutation. We start with an ordered list of symbols, such
as {1, 2, 3}, or {A, B, C} or {m, , 3}. A permutation is an operation
that changes the order of the symbols in the list. For example, suppose the
permutation
1
acts on {1, 2, 3} to produce {1, 3, 2}. We can write this as

1
{1, 2, 3} = {1, 3, 2}.
1
has nothing to do with the numbers 2 and 3; it
simply exchanges the second and third objects in the list. That is, it takes
the object in position 2 and places it in position 3, and it takes the object in
position 3 and places it in position 2. Because of this,

1
{A, B, C} = {A, C, B} and
1
{m, , 3} = {m, 3, } (152)
A transposition is a permutation that interchanges the position of two
elements in the list.
1
is a transposition, but

2
{1, 2, 3} = {2, 3, 1} (153)
is not. Any permutation can be written as a succession of transpositions
that has the same eect on a list. For example, we can reach {2, 3, 1} from
{1, 2, 3} by two transpositions:
{1, 2, 3} {2, 1, 3} {2, 3, 1} (154)
This decomposition of a permutation into successive transpositions is not
unique. The permutation
2
in Eq. 153 is also equivalent to the following
sequence of transpositions:
{1, 2, 3} {1, 3, 2} {3, 1, 2} {3, 2, 1} {2, 3, 1} (155)
This takes four transpositions, while Eq. 154 took only two, but the outcome
is the same
2
.
11. Angular Momentum Theory, November 20, 2009 30
One can prove the following theorem: the number of transpositions through
which a given permutation is expressed is either odd or even. In other words,

2
might be expressed using two or four or six transpositions, but never using
one, or three, or ve. A permutation that can be decomposed into an even
number of transpositions has the signature +1 and is called even; one that
can be decomposed into an odd number of transpositions has the signature
1 and is called odd.
Mathematica has a function Permutations[list] that generates all per-
mutations of the objects in the list. For example, the lists generated by
Permutations[{1,2,3}] are
Permutations Signature
{1,2,3} +1
{1,3,2} 1
{2,1,3} 1
{2,3,1} +1
{3,1,2} +1
{3,2,1} 1
The Mathematica function Signature[a] gives the signature of the permu-
tation that creates the list a from the ordered version of the list a.
We can now return to the Levi-Civitta symbol. It is dened by

ijk
=
_
signature[{i, j, k}] if i = j and i = k and j = k
0 otherwise
(156)
Thus
112
= 0,
123
= 1,
213
= 1, etc.
In many books, Eq. 151 is written as
[

,

L
j
] = i h
jk

L
k
with the understanding that repeated indices are summed over. This is called
Einsteins convention.
Exercise 4 Show that the cross-product w = v u of vector algebra can be
written as
w
i
=
3

j=1
3

k=1

ijk
v
j
u
k
, i = 1, 2, 3
11. Angular Momentum Theory, November 20, 2009 31
Appendix A11.2. Commutator algebra
This chapter relies heavily on commutators so I collect here some in-
formation about them. These facts are useful in many areas of quantum
mechanics.
From the denition of the commutator [

A,

B] =

A

B

B

A, it is obvious
that
[

A,

B] = [

B,

A] (157)
and
[

A,

B +

C] = [

A,

B] + [

A,

C] (158)
It is also easy to prove that
[f(

A),

A] = 0 (159)
if

A is the operator of an observable and f is an arbitrary function.
Not so obvious, but still easy to prove, is
[

A,

B

C] = [

A,

B]

C +

B[

A,

C] (160)
Indeed, we have (from the denition of a commutator)
[

A,

B

C] =

A

B

C

B

C

A (161)
Since (from the denition)

A

B = [

A,

B] +

B

A, we can rewrite Eq. 161 as
[

A,

B

C] = [

A,

B]

C +

B

A

C

B

C

A,
from which it follows that
[

A,

B

C] = [

A,

B]

C +

B(

A

C

C

A)
= [

A,

B]

C +

B[

A,

C]
This is what we wanted to prove.
If we take

C =

B in Eq. 160, we have
[

A,

B
2
] = [

A,

B]

B +

B[

A,

B] (162)
11. Angular Momentum Theory, November 20, 2009 32
Then taking

C =

B
2
in Eq. 160:
[

A,

B
3
] = [

A,

B

B
2
] = [

A,

B]

B
2
+

B[

A,

B
2
] (use Eq. 160)
= [

A,

B]

B
2
+

B([

A,

B]

B +

B[

A,

B]) (use Eq. 160)
= [

A,

B]

B
2
+

B[

A,

B]

B +

B
2
[

A,

B]
Repeating this procedure, we obtain
[

A,

B
n
] =
n1

s=0

B
s
[

A,

B]

B
ns1
(163)
For fun, lets apply this to

A = x and

B = p, for which we know that
[ x, p] = i h

I (164)
where

I is the unit operator. We have
[ x, p
n
] =
n1

s=0
p
s
[ x, p] p
ns1
=
n1

s=0
p
s
i h

I p
ns1
= i h p
n1
+ p(i h) p
n2
+ + p
n1
(i h)
= i h(n p
n1
) = i h
p
n
p
Exercise 5 Show that if
f( p) =

n=0
f
n
p
n
where f
n
, n 0, are numbers, then
[ x, f( p)] = i h
f( p)
p
Exercise 6 Show that
[

A, [

B,

C]] + [

B, [

C,

A]] + [

C, [

A,

B]] = 0

You might also like