Вы находитесь на странице: 1из 10

Page 0.

1
SECTION 0: SPIN ANGULAR MOMENTUM

1. Introduction 2
2. Spin algebra 3
3. Intrinsic spin s = 1/2 4
3.1 Spin states 4
3.2 Matrix representation 5
3.3 Two-component wave functions 6
4. Particles and atoms in a magnetic field 7
4.1 Magnetic dipoles in a magnetic field 7
4.2 Lamor precession 7
4.3 Stern-Gerlach experiment 9












References
The original version of these notes was based on section 5.3 of the book by Mandl and pages 171-184 of the
book by Griffiths (second edition).

Original Version: November 2010
Revision Date: 11 November 2010
Printed: 11 November 2010

Page 0.2
1. INTRODUCTION
1

In classical mechanics, a rigid body can have two types of angular momentum:
Orbital angular momentum, associated with the motion of its centre of mass, given by
l r p =


.
Spin angular momentum, associated with the bodys motion about its centre of mass, given by
s I =


where I is its rotational inertia about the rotation axis and is its rotational frequency about that axis.
As an example, a planet has orbital angular momentum due to its orbital motion about the sun and also spin
angular momentum associated with the spinning of the planet about its own axis.
However, the distinction between spin angular momentum and orbital angular momentum is sometimes a
matter of convenience; in fact the spin angular momentum of a body rotating about an axis is simply the sum
of the orbital angular momenta of all the particles making up the body, in orbital motion about the axis.
Orbital and spin angular momentum also exist in quantum mechanics, but here the distinction between them
is absolutely fundamental. For example, the electron in the hydrogen atom has orbital angular momentum
associated with its motion about the nucleus (described by spherical harmonics). But it also has another form
of angular momentum not linked to its motion in space, and not described by any function of the position
variables (r, , ). This is its intrinsic spin; as the name suggests, this is an intrinsic property of the particle.
Intrinsic spin is in some ways similar to classical spin, and it is sometimes convenient to picture the spin of a
quantum particle as being due to it actually spinning about an axis, with s

pointing along the axis, but this


is very misleading. It must be stressed that intrinsic spin is not caused by any motion of the particle, and it
is not given by the classical expression s I =

.
For example, the electron is believed to be a point particle which cannot therefore spin; and since a point
particle cannot have any internal structure, its spin angular momentum cannot be decomposed into orbital
angular momentum of its constituent parts. Moreover, even if the electron had a finite size, say 1 fm, the speed
of a point on the electrons equator would far exceed the speed of light.
All elementary particles carry intrinsic spin angular momentum s

in addition to their extrinsic orbital


angular momentum l

.
The experimental evidence that led to the discovery of intrinsic spin is described later the Stern-Gerlach
experiment.

1
This introduction is closely based on the introduction to Section 4.4 of the book by Griffith.
Page 0.3
2. SPIN ALGEBRA
The algebra of spin angular momentum s

is completely analogous to that of orbital angular momentum l

.
The fundamental commutation relations are the same:

2
, etc.
, 0 etc.
x y z
z
s s i s
s s
l
=
l
l
l
=
l
l


It follows from the second relation that we can define simultaneous eigenstates ,
s
s m of the operators
2
s and
z
s . These will satisfy:

( )
2 2
, 1 ,
, ,
s s
z s s s
s s m s s s m
s s m m s m
=
=

(1)
where s and m
s
are the quantum numbers associated with the operators
2
s and
z
s respectively.
As with orbital angular momentum, we can define spin raising and lowering operators:

x y
s s is

= (2)
with

( ) ( )
1/2
, 1 1 , 1 .
s s s s
s s m s s m m s m

l
=
l
l
(3)
This is identical in form to the corresponding formula for the raising and lowering operators for orbital angular
momentum.
The commutation relations for angular momentum operators permit the angular momentum quantum
number to take on integral and half-integral values
2
. In the case of orbital angular momentum solution of the
relevant equation for the angular part of the wave function requires that l be a non-negative integer
3
. Since the
spin eigenstates are not functions of (r, , ) and are not solutions of any differential equation, there is no
reason to exclude half-integral values of s and m
s
. Allowed values are therefore:

0, 1/2, 1, 3/2,
, 1, , 1,
s
s
m s s s s
=
=


Every elementary particle has a specific and unchangeable value of s, called its spin. Some examples are given
in the following table.
s = 0 pions
s = 1/2 electrons, protons, neutrons etc
s = 1 photons
s = 3/2 deltas
s = 2 gravitons
In contrast, the orbital angular momentum quantum number l for a given particle can take on any
non-negative integer value, which can be changed by some external interaction.

2
See the discussion at the top of page 165 of the textbook by Griffith (second edition).
3
See the discussion around eq. (4.29) of the book by Griffith.
Page 0.4
3. INTRINSIC SPIN S = 1/2
We consider only s = 1/2. Besides being the simplest, non-trivial case it is also the most important, since it
includes all the particles that make up ordinary matter (electrons, neutrons and protons).
3.1 Spin states
For 1/ 2 s = the quantum number m
s
has two possible values, m
s
= 1/2, and there are correspondingly two
eigenstates ,
s
s m of the spin operators:
spin-up 1/2, 1/2 (or

)
spin-down 1/2, 1/2 (or

)
The eigenvalue equations (1) become:
1
2 z
s =
2 2 2
3
( 1)
4
s s s = =
1
2 z
s =
2 2 2
3
( 1)
4
s s s = =
Since and are eigenstates of s
z
with different eigenvalues, they must be orthogonal; we assume they are also
normalised. Therefore

1
0


= =
=

From the general formula (3) for the action of the spin-raising s
+
and spin-lowering s

operators we have:
0 s

= s

=
s

= 0 s

=
It then follows from the definitions (2) of the spin-raising and spin-lowering operators that
2
x
s =


2
y
s i =


2
x
s =


2
y
s i =


The most general spin state of a spin-half particle must be a linear superposition of the eigenstates and .

1 2
a a = (4)
with the coefficients a
1
and a
2
being complex. Taking the scalar product of this expression with and
respectively we find

1
2
a
a


=
=

If you were to measure the z component of the spin you would get /2 with probability
2
1
a or /2 with
probability
2
2
a . Since there are no other possibilities we have

2 2
1 2
1 a a = ,
which is equivalent to saying that must be normalised.
Page 0.5
3.2 Matrix representation
4

For s = 1/2 there are two eigenstates of the spin operators s
2
and s
z
. These must span a two-dimensional vector
space, the spin space, in which they are represented by 2-element column vectors:
1 0
,
0 1

1 1


= =




( ) ( )

The most general spin state (4) is represented by the column vector, or spinor,

1
1 2
2
1 0
0 1
a
a a
a

1 1
1

= =



( )
( ) ( )
(5)
and the spin operators are represented by 2 2 matrices. For example, from

2 2 2 2
1 1 0 0
3 3
and
0 0 1 1
4 4
s s
1 1 1 1



= =





( ) ( ) ( ) ( )

we get

2 2
1 0
3
0 1
4
s
1


=



( )
. (6)
To derive this, substitute
2
a b
s
c d
1


=



( )
into the above equations and carry out the matrix multiplications. The
first equation gives
2
3 4 a = and c = 0 and the second leads to
2
3 4 d = and b = 0.
Similarly, from

1 1 0 0
and
0 0 1 1
2 2
z z
s s
1 1 1 1



= =





( ) ( ) ( ) ( )


we get

1 0
0 1
2
z
s
1


=



( )

(7)
and from

1 0 1
, 0
0 1 0
0 1 0
, 0
1 0 1
s s
s s


1 1 1



= =





( ) ( ) ( )
1 1 1



= =





( ) ( ) ( )

(8)
we find

0 1 0 0
,
0 0 1 0
s s

1 1


= =




( ) ( )
. (9)
Finally, from the definitions
x y
s s is

= it follows that

4
See the book by Mandl, section 5.3.1.
Page 0.6

0 1 0
,
1 0 0
2 2
x y
i
s s
i
1 1


= =




( ) ( )

. (10)
It follows from the matrix representation (7) (10) that:
The operators
2
, , ,
x y z
s s s s are all Hermitian they represent observables which implies that their ei-
genvalues must be real and they must be represented by Hermitian matrices (the matrices are invariant
under complex conjugation plus transposing,
*
qp pq
O O = ).
The operators s

are not Hermitian (in fact s


+
is the Hermitian conjugate of s

, and vice versa) they


cannot therefore represent observables.
For convenience, Pauli introduced the Pauli spin matrices

, defined by

1
2
s =

, (11)
so that from (7) and (10) we get

0 1 0 1 0
, ,
1 0 0 0 1
x y z
i
i

1 1 1


= = =




( ) ( ) ( )
. (12)
3.3 Two-component wave functions
5

As already discussed, a general spin state (4) can be represented
1 2
a a =
To give the complete wave function of the particle we must include its spatial part; this is done by making the
coefficients functions of position:

( ) ( ) ( )
( ) ( )
( )
( )
1 2
1 2
1
2
,
1 0
0 1
r s r r
r r
r
r

=
1 1


=




( ) ( )
1

=




( )


It follows that the probability that the particle will be measured in the volume element
3
d r

at position r

in
the spin-up state is ( )
2
3
1
r d r

(and similarly for the spin-down state). Therefore
( ) ( )
2
2
3 3
1
i
i
P r d r r d r
=
=



is the probability that the particle is in the volume element
3
d r

at position r

with either spin.


Note that, in the absence of a spin-dependent interaction (such as a magnetic field), the energy levels of a
spin-half particle will have an additional two-fold degeneracy due to spin.

5
See section 5.3.2 of the book by Madel; this is not discussed by Griffith (but see footnote 7 on page 210).
Page 0.7
4. PARTICLES AND ATOMS IN A MAGNETIC FIELD
4.1 Magnetic dipoles in a magnetic field
Classically, a particle of mass m carrying a charge q that moves in an orbit with angular momentum L


constitutes a current loop. It therefore creates a magnetic dipole moment which is easily shown to be

2
q
L
m
=

.
For example, for an electron

2
e
L
m
=

.
These equations are also valid in quantum mechanics, with

and L

being quantum-mechanical vector


operators.
Classically, a spinning, charged particle constitutes a magnetic dipole, with its dipole moment proportional to
its spin angular momentum:
s =

(13)
where is called the particles gyromagnetic ratio. It is easily shown
6
that, for an object with identically
distributed charge and mass,

2
q
m
= .
However, quantum mechanically, it follows from Diracs relativistic quantum theory that, although the
relationship s =

is still valid, for an electron the gyromagnetic ratio (13) is given by
e
m
= ,
i.e. double the classical value.
When a magnetic dipole is placed in a magnetic field B

it experiences a torque
B

.
This tends to align it with the field. The potential energy associated with the torque is
B


So in quantum mechanics the Hamiltonian of a charged particle at rest in a magnetic field is
H B s =


(14)
(For a particle at rest there is no orbital angular momentum to interact with the field, and of course no kinetic
energy term.)
4.2 Lamor precession
7

Consider a spin-half particle at rest in a static, uniform magnetic field parallel to the z-axis:

0

B B k =

.

6
For a reference, see footnote 30 on page 178 of the book by Griffith.
7
This is Example 4.3 of the book by Griffith.
Page 0.8
The matrix form of the Hamiltonian (14) is

0
0
1 0
0 1
2
z
B
H B s

1


= =



( )

. (15)
The eigenstates of the Hamiltonian are clearly those of s
z
:
with energy
0
/2 E B

=
with energy
0
/2 E B

=
This shows that the energy is lowest when the dipole moment is parallel to the field (as it is classically). Note
that the suffix on E indicates whether the particle is in the spin-up (+) or spin-down () state.
The time-dependent Schrdinger equation for an arbitrary spin state with spinor is
i H
t

.
Since the Hamiltonian (15) is time independent, the general solution of this equation can be expanded in terms
of the stationary states
8


( ) [ [
[ [
[ [
1 2
1 0
2 0
exp / exp /
exp /2
exp /2
t a iE t a iE t
a i B t
a i B t


l
=
l
1


( )

(16)
The constants a
1
and a
2
are determined by the initial conditions and normalisation:

( )
1
2
2 2
1 2
0
1
a
t
a
a a

= =


( )
=

We assume that these constants are real
9
and put

1 2
cos , sin
2 2
a a

= =
where is a fixed angle determined by the initial conditions. Then the general solution (16) becomes
( )
[ [
[ [
0
0
cos exp /2
2
sin exp /2
2
i B t
t
i B t


( )
. (17)
And the normalisation requirement is automatically satisfied. We now calculate the expectation value of the
operator s

as a function of time for this state (17). For the x component:



( ) ( )
0
0 0
0

/2
/2 /2
/2
cos
0 1
2
cos sin
1 0
2 2 2
sin
2
x x
i B t
i B t i B t
i B t
s t s t
e
e e
e

=
1


1

1
=

( )
( )


( )



8
See eq. (2.17) of the textbook by Griffith.
9
See the footnote on page 180 of the book by Griffith.
Page 0.9
leading to
( )
0
sin cos .
2
x
s B t =

(18)
Similarly, for the y component
( )
0
sin sin
2
y
s B t =

(19)
and for the z component
cos
2
z
s =

. (20)
Taken together, (18), (19) and (20) imply that the expectation value s

is tilted at a constant angle to the


z-axis, and precesses about the field direction with frequency given by

0
B = . (21)
This is called the Lamor frequency.
Classically when a magnetic dipole is placed in a magnetic field it is the spin vector s

itself that precesses


about the field
10
with the frequency (21).
A semi-classical representation of the precession of the spin is shown in the following diagram.

Note that this analysis also applies to any neutral atom with magnetic dipole moment s =

placed in a
uniform field
0

B B k =

.
4.3 Stern-Gerlach experiment
11

If a magnetic dipole is placed in an inhomogeneous magnetic field, there is a force

( )
F B =


acting on the dipole, in addition to a torque.

10
This is a specific example of Ehrenfests theorem which says that quantum-mechanical expectation values obey classical laws.
11
See Example 4.4 of the book by Griffith for a more detailed and rigorous discussion.
Page 0.10
Consider a beam of atoms each with magnetic dipole moment

which is passing through a magnetic field.


The field varies in the z direction which is transverse to the direction of the beam. It follows that there will
be a net force acting on each atom which has a z component

z
z z
B
F
z

.
This causes the beam to be deviated by an amount that is proportional to the z component of the dipole
moment

and therefore to the z component of the angular momentum.


Stern and Gerlach (1921) designed an experiment to test whether or not the orbital angular momentum L

of
individual atoms is quantised, as had been suggested a few years earlier by Bohr. They passed a beam of
neutral silver atoms through an inhomogeneous field, as described above.
If angular momentum is not quantised, then the distribution of atoms hitting the detector should be
continuous in the z direction.
On the other hand, if angular momentum is quantised, the beam should be split by the field into
components, corresponding to the allowed components of L

along the z axis.


They found that the beam was split into two components, equally spaced on either side of the original beam
direction. This therefore confirmed the postulate of Bohr.
However, the experimental results imply that there are only two possible orientations of the magnetic dipoles
with respect to the z axis and hence only two distinct quantised values of the z component of the atoms
angular momentum. But, according to the quantum theory of angular momentum (developed a few years after
the initial experiment was carried out), the quantum number l has to be integer, implying that is has an odd
number of components, namely (2l + 1). This suggests that the angular momentum of the atoms is 1/2.
To explain this apparent discrepancy (and others mentioned below)
12
, Uhlenbeck and Goudsmit (1925)
postulated that the atoms were in a state of orbital angular momentum l = 0 and the angular momentum 1/2
that was measured was an intrinsic property of the outermost electron of the atom its intrinsic spin angular
momentum (all the other electrons having their angular momenta coupled to zero).
A formal description of intrinsic spin, as discussed in the previous sections, was provided by Pauli (1927), who
was able to incorporate spin into non-relativistic quantum mechanics through the introduction of a few
additional postulates he was thereby able to correctly predict the magnitude of the magnetic effect. A proper
relativistic quantum approach to the interaction of an electron with a magnetic field, developed by Dirac
(1928), predicts the existence of intrinsic spin with s = 1/2 for the electron; the description of spin follows
naturally from the relativistic approach.
With the postulation of spin, several other magnetic-related phenomena were explained, for example the
anomalous Zeeman Effect and the fine structure seen in the spectrum of hydrogen and other atoms.




12
Pauli had already suggested the previous year, in order to explain the observed emission spectra of alkali metals, that the electron
had a two-valued degree of freedom. This allowed him to formulate the Pauli exclusion principle.

Вам также может понравиться