Rel QM WalterPfeifer
Rel QM WalterPfeifer
An Introduction
by Walter Pfeifer
www.walterpfeifer.ch/
Contents
Contents 1
Preface 2
Index 72
2
Preface
Relativistic quantum mechanics are used to describe high-energy particles and
highly ionised atoms. They give a consistent formalism for spin-½ particles and
provide finer details of atomic and molecular spectra. In short, they are an
important tool of modern physics.
This book deals mainly with the Dirac equation, its properties, its applications and
its limiting cases. A formalism for particles with arbitrary spin and remarks on other
relativistic quantum mechanical equations are given.
This publication is an introduction and is directed towards students of physics and
interested physicists. The detailed developments and the numerous references to
preceding places make it easier to follow. However, knowledge of the elements of
quantum mechanics, relativistic mechanics and electrodynamics is a prerequisite.
In order to relieve the reader, we don't deal with rotations of the coordinate system
and not with Lorentz groups either. We have no renaming of matrices, no
Feynman daggers, no Einstein convention of summation over repeated indices, no
quantum field theory, no second quantisation and no natural units with = = c = 1 .
The system SI (MKSA system) is used without exception.
We use the following symbols: operators are written in bold letters, a three-
dimensional vector is marked with an arrow, two- or four-spinors or -vectors are
underlined and the symbols for matrices are doubly underlined.
3
x Σ x' Σ'
(x,y,z,t)
(x',y',z',t')
v
z z'
t ′ = t , x ′ = x, y ′ = y , z′ = z − vt (1.1.1)
hold between the space-time coordinates of both systems for a given event. If the
velocity v is not negligible with respect to c , i.e. for the relativistic case, a Lorentz
transformation has to be performed as follows
ct − (v / c ) z z − vt
ct ′ = , x ′ = x, y ′ = y , z ′ = . (1.1.2)
1 − (v / c ) 1 − (v / c )
2 2
x 0 = ct , x1 = x, x 2 = y , x 3 = z (1.1.3)
with corresponding expressions for x 0′, ", x 3′. Equations (1.1.2) read now
x 0 − (v / c ) x 3 x 3 − (v / c ) x 0
x 0′ = , x 1′ = x 1, x 2′ = x 2 , x 3′ = . (1.1.4)
1 − (v / c ) 1 − (v / c )
2 2
⎛ 1 −v / c ⎞
⎜ 0 0 ⎟
⎜ 1 − (v / c ) 1 − (v / c )
2 2
⎟
⎜ ⎟
0 1 0 0
a=⎜ ⎟ (1.1.5)
⎜ 0 0 1 0 ⎟
⎜ ⎟
⎜ −v / c 1 ⎟
0 0
⎜⎜ ⎟⎟
⎝ 1 − (v / c ) 1 − (v / c )
2 2
⎠
with elements a00 , a01, ⋅, a03 , a10 , a11, ", a33 , equations (1.1.4) can be written as
3
x i ′ = ∑ aik x k , i = 0,1,2,3, (1.1.6)
k =0
⎛ x 0′ ⎞
⎛ x0 ⎞ ⎜ ⎟
⎜ 1⎟ ⎜ x1′ ⎟
x
and by means of the column vectors x = ⎜ 2 ⎟ and x ′ = ⎜ ⎟ we have
⎜x ⎟ ⎜ x 2′ ⎟
⎜⎜ 3 ⎟⎟ ⎜ ⎟
⎝x ⎠ ⎜ x 3′ ⎟
⎝ ⎠
x′ = ax. (1.1.7)
∂x i ′
= aik . (1.1.8)
∂x k
Of course, motions in the x- or y-direction result in corresponding transformation
formulas. Four-vectors which transform like x are named contravariant. With the
help of (1.1.4) we calculate the gauge
( ) ( ) ( ) ( )
2 2 2 2
x 0′ − x1′ − x 2′ − x 3′
(x )0
2
(1− (v / c ) ) − ( x ) (1− (v / c ) ) −
2 3
2 2
(x ′) − (x ′)
2 2
= 1 2
(1.1.9)
1 − (v / c )
2
( ) − (x ) − (x ) − (x )
2 2 2 2
= x0 1 2 3
.
The first and the last line show that the form of this expression isn't changed by the
Lorentz transformation (1.1.4). The expression (1.1.9) is said to be form invariant.
There are other linear transformations which preserve the Lorentz metric (1.1.9): a
rotation of the coordinate system, the reflection of the space coordinates
( x ′ = − x, y ′ = − y , z′ = −z ) and the time reversal ( t ′ = −t ) . These transformations
and the translational transformation of the type (1.1.2) can be coupled individually
resulting in a transformation with the same Lorentz gauge. That is, all these
transformations form a group: the inhomogeneous Lorentz group.
5
x 0′ + (v / c ) x 3′ x 3′ + (v / c ) x 0′
x =
0
, x = x ′, x 2 = x 2′, x 3 =
1 1
(1.2.1)
1 − (v / c ) 1 − (v / c )
2 2
3
x i = ∑ aik−1x k ′ or x = a x ′ with
−1
i.e.
k =0
⎛ 1 (v / c ) ⎞
⎜ 0 0 ⎟
⎜ 1 − (v / c ) 1 − (v / c ) ⎟
2 2
⎜ ⎟ (1.2.2)
−1 ⎜ 0 1 0 0 ⎟
a =⎜ ⎟,
0 0 1 0
⎜ ⎟
⎜ (v / c ) 1 ⎟
⎜ 0 0 ⎟
⎜ 1 − (v / c )2 1 − (v / c )
2
⎟
⎝ ⎠
∂ ∂
∂ ∂x 3 ∂x 0 (
+ v /c) 3
∂ ∂ ∂x 0 ∂x
= 0 + 3 =
∂x 0′ ∂x ∂x ′ ∂x ∂x ′
0 0
1 − (v / c )
2
∂ ∂
=
∂x ′ 1 ∂x1
(1.2.3)
∂ ∂
= 2
∂x 2′ ∂x
∂ ∂
+ (v / c ) 0
∂
= ∂x ∂x .
3
3′
∂x 1 − (v / c )
2
In short,
∂ 3
∂ ∂x i 3
∂ −1
=∑ = ∑ a (1.2.4)
∂x k ′ i = 0 ∂x ∂x ′ i = 0 ∂x
i k i ik
6
holds, where (1.2.2) has been used. The inverse relation reads
∂ 3
∂ ∂x i ′ 3
∂
∂x k
= ∑
i = 0 ∂x ′ ∂x
i k
= ∑
i = 0 ∂x ′
i
aik .
1 ∂2 1 ∂2 ∂2 ∂2 ∂2 ∂2 3
∂2
c 2 ∂t 2
− ∇ 2
≡ − − − ≡ − ∑
c 2 ∂t 2 ∂x 2 ∂y 2 ∂z 2 (∂x 0 )2 i =1 (∂x i )2
appears. Applying eq. (1.2.3) repeatedly one obtains the Lorentz form invariance
∂2 3
∂2 ∂2 3
∂2
−∑ = −∑ (1.2.5)
(∂x 0′ )2 i =1 (∂x i ′ )2 (∂x ) i =1 ( ∂x )
0 2 i 2
∂
E = i= . (1.2.6)
∂t
Therefore we write
∂ ∂ 1 1
= = E ≡ p0 ,
∂x 0
c∂t i=c i=
E K
where we have defined p 0 = . The x-component of the momentum operator p
c
reads
∂
px = −i= (1.2.7)
∂x
and we have
∂ −1 −1 1
= px ≡ p .
∂x 1
i= i=
∂ ∂
Inserting in (1.2.3) and defining p 2 ≡ −i= , p 3 ≡ −i= , we obtain the following
∂y ∂z
relation between operators
E′ p 0 − (v / c ) p 3 1 p 3 − (v / c ) p 0
≡ p 0′ = , p ′ = p1 , p 2′ = p 2 , p 3′ = . (1.2.8)
c 1 − (v / c ) 1 − (v / c )
2 2
7
⎛ p0 ⎞
⎜ 1⎟
p
That is, the four-vector p = ⎜ 2 ⎟ transforms like x (cf.(1.1.4)). Consequently it is
⎜p ⎟
⎜⎜ 3 ⎟⎟
⎝p ⎠
contravariant. Due to the strong analogy with (1.1.9) the operator expression
(p ) − (p ) − (p ) − (p )
0 2 1 2 2 2 3 2
must be form invariant in a Lorentz transformation, i.e.
(p ′) − (p ′) − (p ′) − (p ′) = ( p 0 ) − ( p1 ) − ( p 2 ) − ( p 3 )
2 2 2 2 2 2 2 2
0 1 2 3
(1.2.9)
E ′2 E2
or 2
′ 2
′ 2
′
− px − py − pz = 2 − px 2 − py 2 − pz 2 .
2
c c
A similar expression comes from the relativistic relation between energy and
momentum (as quantities) of a free particle
E 2 = c 2 p 2 + m02c 4 . (1.2.10)
K
Following the rules of quantum mechanics we substitute E and p by the
K
corresponding operators E and p , (1.2.6) and (1.2.7), and obtain
K
E 2 − c 2 p 2 ≡ E 2 − c 2 ( px 2 + py 2 + pz 2 ) = m02c 4 (1.2.11)
Because the left hand side of (1.2.11) is form invariant, the rest mass m0 is the
same in every inertial system as we expect.
acts on the particle. Our formulas and quantities are written in the system SI. The
K
potentials Φ and A obey differential equations which contain the electric charge
K
density ρ and the electric current density J . The equations read
8
1 ∂ 2Φ ρ
− ∇ 2Φ = , (1.3.3)
c ∂t
2 2
ε0
K
1 ∂2 A K K
− ∇ 2
A = μ J
0 . (1.3.4)
c 2 ∂t 2
The quantities μo and ε 0 are the permeability and the dielectricity respectively of
the vacuum. An additional constraint can be chosen. We take the “Lorenz gage”:
K 1 ∂Φ
div A + 2 = 0. (1.3.5)
c ∂t
In the framework of special relativity it is natural to introduce the contravariant four-
vectors
⎛Φ ⎞
A = ( A0 , A1 , A2 , A3 ) ≡ ⎜ , Ax , Ay , Az ⎟ (1.3.6)
⎝c ⎠
and J = ( J 0 , J 1, J 2 , J 3 ) ≡ (c ρ , J x , Jy , Jz ) (1.3.7)
A0 − ( v / c ) A3 A3 − ( v / c ) A 0
A0′ = , A1′ = A1 , A2′ = A2 , A3′ = , (1.3.8)
1 − (v / c ) 1 − (v / c )
2 2
J 0 − (v / c ) J 3 J 3 − (v / c ) J 0
J 0′ = , J 1′ = J 1 , J 2′ = J 2 , J 3′ = . (1.3.9)
1 − (v / c ) 1 − (v / c )
2 2
∂ 2 A0′ 3
∂ 2 A0′ ρ′ J 0′
−∑ = = 2 = μ0 J 0′ , (1.3.10)
( ) ( ) cε 0 c ε 0
2 2
∂x ′0 i =1
∂x ′
i
1
ε 0 μ0 = . (1.3.11)
c2
With the help of (1.2.5), (1.3.8) and (1.3.9) we obtain from (1.3.10)
9
⎛ ⎞ 0
⎜ ∂ ∂ 2 ⎟ A − (v / c ) A J 0 − (v / c ) J 3
2 3 3
⎜ ( ∂x 0 )2 ∑
− = μ . (1.3.12)
i =1 ( ∂x ) ⎟
0
⎠ 1 − (v / c ) 1 − (v / c )
i 2 2 2
⎝
∂ 2 A3′ 3
∂ 2 A3′
−∑ = μ0 J 3′ (1.3.13)
( ) ( )
2 2
∂x ′
0 i =1
∂x ′i
⎛ ⎞ 3
⎜ ∂ ∂ 2 ⎟ A − (v / c ) A J 3 − (v / c ) J 0
2 3 0
⎜ ( ∂x 0 )2 ∑
− = μ0 . (1.3.14)
i =1 ( ∂x ) ⎟ ( ) ( )
i 2 2 2
⎝ ⎠ 1 − v / c 1 − v / c
⎛ ⎞
⎜ ∂ ∂2 ⎟ 0
2 3
−∑ A = μ0 J 0 , (1.3.15)
⎜ ( ∂x 0 ) 2
i =1 ( ∂x ) ⎟
i 2
⎝ ⎠
which is form invariant with respect to (1.3.10). Multiplying (1.3.12) by v/c and
adding it to (1.3.14) results in
⎛ ⎞
⎜ ∂ ∂2 ⎟ 3
2 3
⎜ ( ∂x 0 )2 ∑
− A = μ0 J 3 (1.3.16)
i =1 ( ∂x )
i 2 ⎟
⎝ ⎠
⎛ ⎞
⎜ ∂2 3
∂2 ⎟ k′
⎜ −∑ 2 ⎟
A = μ0 J k ′ with k = 1, 2 . (1.3.17)
( ) ( )
2
⎜ ∂x 0′ i =1
∂x i′
⎟
⎝ ⎠
When handled in the same way as the preceding equations, it results immediately
in the form invariant form affiliated to the coordinate system Σ .
Finally we write eq. (1.3.5) with the help of (1.1.3) and (1.3.6) in the system Σ ′
⎛ ∂ ∂ ⎞ ⎛ ∂ ∂ ⎞
⎜ ∂x 0 + (v / c ) ∂x 3 ⎟ A0 − (v / c ) A3 ∂A1 ∂A2 ⎜ ∂x 3 + (v / c ) ∂x 0 ⎟ A3 − ( v / c ) A0
⎜ ⎟ + 1 + 2 +⎜ ⎟ = 0,
⎜⎜ 1 − (v / c )
2
⎟⎟ 1 − (v / c )
2 ∂x ∂x ⎜⎜ 1 − (v / c )
2
⎟⎟ 1 − (v / c )
2
⎝ ⎠ ⎝ ⎠
which results finally in
10
∂ψ =2 2 K
i= = Hψ ≡ − ∇ ψ + V ( r , t )ψ (2.1.1)
∂t 2m
dominates. The complex wave function ψ depends on space and time. Planck's
constant h divided
K by 2π is marked by = . The potential V depends on the
position vector r of the particle (mass m) and t. In (2.1.1) the Hamiltonian H is
defined. This equation contains a differentiation with respect to the time and in ∇ 2
there are second order derivatives with respect to spatial coordinates. This is
inconsistent with the principles of special relativity, where time and spatial
coordinates have the same importance and are interrelated linearly in the
transformation formulas. That is why the Schrödinger equation is not form invariant
in a Lorentz transformation, which is not acceptable in situations where velocities
are not negligible compared with c.
The relativistic equation which Dirac was looking for, had to treat space and time
in a manifestly symmetric fashion. In addition the equation had to be form invariant
in a Lorentz transformation, and the Hamiltonian had to be Hermitian - a condition
which is well-known to guarantee real expectation values for the energy.
For a free particle with the rest mass m0 Dirac's equation reads
K
∂Ψ ( r , t ) ⎛ ∂ ∂ ∂ ⎞ K K
i= + i=c ⎜ α1 + α2 + α 3 ⎟Ψ ( r , t ) − β m0c 2Ψ ( r , t ) = 0 (2.1.2)
∂t ⎝ ∂x ∂y ∂z ⎠
K K
The second term can be written as a scalar product i=cα ⋅ ∇ or with the help of
K K
(1.2.7) as −cα ⋅ p , and using (1.2.6) we have
where p1 means px etc. The operator E acts on the left hand side, and on the
right hand side we have the operator c (α1 p1 + α 2 p 2 + α 3 p 3 + β m0c ) , which is
identical with E , because we demand that (2.1.3) holds for every Ψ . Applying
these operators once more to each side of (2.1.3) we obtain
⎛ 3 ⎞
E 2 Ψ = c 2 ⎜ ∑ α i 2 ( p i ) + ∑ (α iα k + α kα i ) p i p k + m0c ∑ (α i β + βα i ) p i + β 2m02c 2 ⎟Ψ .
3 3
2
⎝ i =1 i ≠ k =1 i =1 ⎠
(2.1.4)
We demand that Ψ satisfies also the relation (1.2.11) like this
12
⎛ 3 ⎞
E 2Ψ = c 2 ⎜ ∑ ( p i ) + m02c 2 ⎟Ψ
2
(2.1.5)
⎝ i =1 ⎠
α i2 = 1, α i β + βα i = 0 for i = 1, 2, 3,
α i α k + α kα i = 0 for i ≠ k = 1, 2, 3, (2.1.6)
β 2 = 1.
In this way, both sides of (2.1.3) are again column vectors with the same
dimension. The operator on the right hand side of (2.1.3) is the Hamiltonian Hf of
the free particle:
⎛ 3 ⎞
H f = c ⎜ ∑ α i p i + β m0c ⎟ (2.1.8)
⎝ i =1 ⎠
Ψ a H fΨ b = ∫Ψ †a H fΨ bdτ (
with Ψ a = ψ a1 *,ψ a2 *," ,ψ an * .
†
) (2.1.9)
( )
†
H fΨ a Ψ b = ∫ H fΨ a Ψ bdτ = ∫Ψ a H f Ψ bdτ ,
† †
(2.1.10)
( HΨ )
†
=Ψ a H
† †
where the well-known mathematical relation a has been used.
Hermiticity of the operator H f means that the expression (2.1.9) equals (2.1.10).
13
Therefore, we have H f = H f
†
i.e. the matrix H f is Hermitian. Because the
coefficients in H f are arbitrary, the matrices α i and β are also Hermitian and
their diagonal matrix- elements are real. Their traces tr (i.e. the sums of the
diagonal elements) vanish for the following reason: According to (2.1.6) one has
α i = −α k α i α k ( i ≠ k ) . The well-known relation concerning the trace of the
product of two matrices, say A and B ,
tr ( AB ) = tr ( BA ) ,
tr α i = 0. (2.1.11)
The same relation is true for β . As will be shown at the end of this section the
{ }
choice of the matrix set α i , β is not unique. We choose β to be diagonal. It's
usual to make up the matrix set by the Pauli spin matrices ( 2 × 2 matrices)
⎛ 0 1⎞ ⎛ 0 −i ⎞ ⎛1 0 ⎞
σ1 = ⎜ ⎟, σ 2 = ⎜ ⎟, σ3 = ⎜ ⎟, (2.1.12)
⎝ 1 0⎠ ⎝ i 0⎠ ⎝ 0 −1⎠
{ }
Mostly, the matrix set α i , β is represented by the following 4 × 4 matrices
⎛0 σi ⎞ ⎛1 0 ⎞
αi = ⎜ ⎟ for i = 1, 2, 3, β =⎜ ⎟, (2.1.15)
⎜σ 0 ⎟⎠ ⎝ 0 −1⎠
⎝ i
⎛0 0 0 1⎞ ⎛ 0 0 0 -i ⎞
⎜ ⎟ ⎜ ⎟
0 0 1 0⎟ 0 0 i 0⎟
α1 = ⎜ , α2 = ⎜ ,
⎜0 1 0 0⎟ ⎜ 0 -i 0 0 ⎟
⎜⎜ ⎟ ⎜⎜ ⎟⎟
⎝1 0 0 0 ⎟⎠ ⎝ i 0 0 0⎠
(2.1.16)
⎛0 0 1 0⎞ ⎛1 0 0 0 ⎞
⎜ ⎟ ⎜ ⎟
0 0 0 −1⎟ 0 1 0 0⎟
α3 = ⎜ , β= ⎜ .
⎜1 0 0 0⎟ ⎜ 0 0 −1 0 ⎟
⎜⎜ 0 −1 0 0 ⎟⎠
⎟ ⎜⎜ 0 0 0 −1⎟⎟
⎝ ⎝ ⎠
14
The matrices are evidently Hermitian and meet the relations (2.1.6) and (2.1.11).
{ }
The matrix set α , β can be turned into an equivalent set α ′ , β ′ by unitary
i { } i
{ }
Again, it can easily be shown that the new matrices α i ′ , β ′ are Hermitian and
meet the relations (2.1.6) and (2.1.11). Therefore Dirac's equation (2.1.3) must
{ }
also hold with the matrix set α i ′ , β ′ . We name the corresponding solution Ψ ′ ,
which satisfies
⎛ 3 ⎞
EΨ ′ = c ⎜ ∑ α i ′ p i + β ′m0c ⎟Ψ ′.
⎝ i =1 ⎠
{ }
If we substitute here α i ′ , β ′ by the expressions of (2.1.17) and multiply the
−1
equation by U from the left, we obtain
⎛ 3 ⎞ −1
( )
E U Ψ ′ = c ⎜ ∑ α i p i + β m0c ⎟ U Ψ ′ .
−1
⎝ i =1 ⎠
( )
That is
Ψ = U −1Ψ ′ or UΨ = Ψ ′ . (2.1.18)
In section 2.3 we will learn that the probability density ρ of the particle is
calculated analogously to classical quantum mechanics, namely
⎛ψ 1 ⎞
⎜ ⎟
ψ
ρ = Ψ Ψ ≡ (ψ 1*,ψ 2 *,ψ 3 *,ψ 4 * ) ⎜ 2 ⎟ .
†
⎜ψ 3 ⎟
⎜⎜ ⎟⎟
⎝ψ 4 ⎠
{ }
That is, for both sets α i ′ , β ′ and {α } , β
i
the density function is identical. One
can say that unitary transformations do not change the physics. We will use this
fact in the next section.
For a particle which is not free, unlike the situation in (2.1.3), but influenced by
K
electromagnetic forces, the electromagnetic potentials Φ and A , (1.3.1), have to
be built in the Dirac equation in the same way as in the Schrödinger equation. The
principle of minimal coupling of the quantum mechanics - or minimal substitution -
orders that the potential energy eΦ (for a charge e) has to be subtracted from the
energy operator E , and to each p i a component −eAi is added. That is, instead
of (2.1.3) we have
⎛ 3 ⎞
( − Φ )Ψ = ⎜ ∑ α i ( p − eA ) + β m0c ⎟Ψ
i i
E e c (2.1.20)
⎝ i =1 ⎠
In section 2.4 it will be shown that the asymptotic form of this equation for v c is
the classical Pauli equation with two-component spinors.
(1 2 3 0
⎦⎥)
⎡E ′− c α ′ p1′ + α ′ p 2′ + α ′ p 3′ − β ′m c 2 ⎤Ψ ′ ( rK′, t ′ ) = 0
⎣⎢
(2.2.1)
Since the coordinate systems are equivalent, the relations (2.1.6) hold also in Σ′
i.e.
β ′2 = α ′2i = 1, α i ′ β ′ +β ′α i ′ =0 for i = 1, 2, 3,
. (2.2.2)
α i ′α k ′ + α k ′α i ′ = 0 for i ≠ k = 1, 2, 3.
α ′†i = α i ′ , β ′† = β ′. (2.2.3)
{ }
interrelated with the corresponding set α i , β by a unitary transformation this way
16
α ′i = U † α i U , β ′ = U† βU for i = 1, 2, 3 with U = U
† −1
(2.2.4)
This is Pauli's theorem, which we give here without proof. Due to (2.1.18) and
(2.1.19) the Dirac equation generates the same density ρ (and the same physics)
{ } { }
for both sets α i ′ , β ′ and α i , β . Therefore, without loss of generality we can
{ }
use the same set α i , β in the Lorentz system Σ′ and in Σ . We shall no longer
differentiate between both sets of matrices:
β ′ = β , α i ′ = α i , i = 1, 2, 3 . (2.2.5)
K
( )
We now turn to the wave function Ψ ′ ( r ′, t ′ ) ≡ Ψ ′ x′ , (2.2.1), where x ′ is given in
( ) ( )
Ψ ′ x′ merely by inserting x x′ in Ψ (see (1.2.1)). However, a combination of
the components ψ 1 ( x ) ,ψ 2 ( x ) ,ψ 3 ( x ) ,ψ 4 ( x ) may occur. Since both the Dirac
equation (2.1.2) as well as the Lorentz transformation (1.1.4) are linear in space-
time coordinates, we demand that the transformation of Ψ ( x ) is linear in the
components of Ψ and make the ansatz
( )
Ψ ′ x′ = S ( a )Ψ ( x ) . (2.2.6)
( )
terms of Ψ ′ x′ like this
( )
Ψ ( x ) = S −1 ( a )Ψ ′ x ′ . (2.2.7)
⎛1 0 0 0⎞
⎜ ⎟
0 1 0 0⎟
α0 ≡ 1≡ ⎜ (2.2.8)
⎜0 0 1 0⎟
⎜⎜ ⎟
⎝0 0 0 1 ⎟⎠
17
⎡ ⎛ ∂ ∂ ∂ ∂ ⎞ ⎤ K
c ⎢i= ⎜ βα 0 + βα 1 + βα 2 + βα 3 ⎟ − β m0c ⎥Ψ ( r , t ) = 0 .
2
(2.2.9)
⎣ ⎝ c ∂t ∂x ∂y ∂z ⎠ ⎦
⎛ 3 ∂ ⎞
⎜ i= ∑ βα μ μ − m0c ⎟Ψ ( x ) = 0 . (2.2.10)
⎝ μ =0 ∂x ⎠
⎛ 3
⎝ μ =0
−1 ∂
∂x
−1 ⎞
⎜ i= ∑ βα μ S ( a ) μ − m0cS ( a ) ⎟Ψ ′ x ′ = 0 .
⎠
( )
Multiplying with S ( a ) from the left and using S ( a ) S (a ) = 1
−1
(unity matrix) we
obtain
⎛ 3
⎝ μ =0
−1 ∂
∂x
⎞
⎜ i= ∑ S ( a ) βα μ S ( a ) μ − m0c ⎟Ψ ′ x ′ = 0 .
⎠
( ) (2.2.11)
∂ 3
∂
∂x μ
= ∑ a ,
ν νμ
ν =0 ∂x ′
(2.2.12)
⎛ 3 ∂ ⎞
⎜ i= ∑ S ( a ) βα μ S ( a ) ∑ aνμ ν − m0c ⎟Ψ ′ x ′ = 0 . ( )
3
−1
⎝ μ =0 ν =0 ∂x ′ ⎠
⎡ 3 ⎛ 3
⎣ μ =0 ⎝ ν =0
−1 ⎞ ∂
⎠ ∂x ′
⎤ ′ ′
⎢i= ∑ ⎜ ∑ S ( a ) βα ν S ( a ) aμν ⎟ μ − m0c ⎥Ψ x = 0 .
⎦
( ) (2.2.13)
In order to achieve form invariance between (2.2.10) and (2.2.13) the following
relations must hold
∑ S ( a )βα ν S ( a ) aμν
3
−1
= βα μ . (2.2.14)
ν=0
−1
On both sides we multiply by aλμ (cf. (1.2.2)) and sum over μ . By means of
3
∑
μ
−1
aλμ aμν
=0
= δνλ we obtain
S ( a ) βα λ S ( a ) = ∑ aλμ−1 βα μ ,
3
−1
(2.2.15)
μ =0
18
which is the conditional equation for the matrix S ( a ) . It holds also for general
Lorentz transformations including instant rotations of the coordinate system and
even for reflections. In fact, this equation forces the form invariance of the Dirac
equation. We make clear that in reality here one defines a four-component wave
function to be a Lorentz function, if it transforms according to (2.2.7) by means of
the relation (2.2.14).
Here we do not construct the matrix S ( a ) step by step (as does Greiner, 1990,
p.110), but we give S ( a ) directly and check it by means of the system of
equations (2.2.15). For our Lorentz transformation, i.e. the translation, we assert
⎛ γ +1 0 − γ −1 0 ⎞
⎜ ⎟
1 ⎜ 0 γ +1 0 γ − 1⎟ 1
S (a) = ⎜ ⎟ with γ = . (2.2.16)
2 ⎜− γ −1 γ +1 1 − (v / c )
2
0 0 ⎟
⎜ ⎟
⎝ 0 γ −1 0 γ + 1⎠
⎛ γ +1 0 γ −1 ⎞
0
⎜ ⎟
1 ⎜ 0 γ +1 0 − γ − 1⎟
S (a ) =
−1
⎜ ⎟. (2.2.17)
2 ⎜ γ −1 0 γ +1 0 ⎟
⎜ 0 − γ −1 γ + 1 ⎟⎠
⎝ 0
−1
The condition S S = 1 is checked easily. By matrix multiplication the relation
−1 −1
= β S β = β Sβ or β S β = S
† †
S (2.2.18)
S ( a ) βα 0 S (a ) = a
−1 −1
00 βα 0 + a03
−1
βα 3
⎛ γ 0 γv / c 0 ⎞
⎜ ⎟
0 γ 0 −γ v / c ⎟
and for both sides we obtain ⎜ .
⎜ −γ v / c 0 −γ 0 ⎟
⎜⎜ ⎟
⎝ 0 γv / c 0 −γ ⎟⎠
Thus, we have found that the Dirac equation of a free particle remains form
invariant in a Lorentz transformation, and therefore it is relativistic.
We show now that the equation (2.1.20), which contains electromagnetic fields, is
also relativistic. With the aid of (1.2.6) we write (2.1.3) like this
⎛ 0 3 2⎞
⎜ c p − ∑ α i c p − β m0c ⎟Ψ = 0 .
i
(2.2.19)
⎝ i =1 ⎠
⎛ ⎞
( ) α i c ( p i − eAi ) − β mo c 2 ⎟Ψ = 0
3
⎜
⎝
c p 0
− ceA 0
− ∑
i =1 ⎠
(2.2.20)
and obtain
∂Ψ 3
∂Ψ
i=Ψ = −i=c ∑Ψ α i i + m0 c 2Ψ βΨ ,
† † †
(2.3.1)
∂t i =1 ∂x
where substitutions from (1.1.3) have been made. The adjoint of eq. (2.1.2) reads
∂Ψ ∂Ψ
† 3 †
−i= = i=c ∑ α †i + m0 c 2Ψ † β †
∂t i =1 ∂x
i
∂Ψ
†
and α i as mentioned in the context of eq. (2.1.10).
†
Notice the sequence of
∂x i
We multiply the equation from the right by Ψ , take into consideration the
Hermiticity of the Dirac matrices α i = α i , β = β ( † †
) (see the context of (2.1.10))
and obtain
∂Ψ ∂Ψ
† 3 †
-i= Ψ = i=c ∑ i α iΨ + m0c 2Ψ † βΨ (2.3.2)
∂t i =1 ∂x
∂ ∂
( ) ( )
3
i= Ψ †Ψ = −i=c ∑ i Ψ † α iΨ . (2.3.3)
∂t i =1 ∂x
20
We name
4
Ψ †Ψ ≡ ∑ψ *kψ k = ρ (2.3.4)
k =1
Clearly, ρ is positive definite, i.e. real and positive at every position and time.
Furthermore we mark the spatial component
cΨ α iΨ ≡ j i , i = 1, 2, 3 .
†
(2.3.5)
∂ K K
ρ + div j = 0 with j = ( j 1 , j 2 , j 3 ) . (2.3.6)
∂t
This differential equation has the familiar form associated with the conservation of
K
flow of a fluid of density ρ and current density j , in which there are neither
K
sources nor sinks. It is thus reasonable to interpret ρ ( r , t ) as the probability
K K
density and j ( r , t ) as the probability density current of the particle. The density
K K K
ρ ( r , t ) is directly measurable but j ( r , t ) is not, because the uncertainty relation
prohibits to measure simultaneously position and velocity exactly.
We expect that the flow conservation equation (2.3.6) holds in every inertial
system. First we show that ( c ρ , j 1 , j 2 , j 3 ) is a contravariant four-vector (cf.
(1.1.6/7)). Using (2.3.4) and (2.2.8) we write
c ρ = cΨ α 0Ψ ,
†
(2.3.7)
j μ ( x ) = cΨ ( x ) α μΨ ( x ) for μ = 0,1, 2, 3 .
†
(2.3.8)
( )
j μ ′ x′ = cΨ ′ x′ α μΨ ′ x′
†
( ) ( )
where we insert (2.2.6) and obtain
( ) (
j μ ′ x ′ = c S ( a )Ψ ( x ) α μ S ( a )Ψ ( x ) )
†
(2.3.9)
= cΨ
†
( x ) S † ( a ) α μ S ( a )Ψ ( x ) .
With the help of (2.2.18) we write
( )
j μ ′ x′ = cΨ
†
( x ) β S −1 ( a ) βα μ S ( a )Ψ ( x ) . (2.3.10)
( a ) βα S ( a ) ,
3
∑ βα ν aμν
−1
=S μ
(2.3.11)
ν =0
( )
3
j μ ′ x ′ = cΨ
†
( x ) β ∑ βα ν aμνΨ ( x )
ν =0
3 3
(2.3.12)
= ∑ aμν cΨ ( x ) α νΨ ( x ) = ∑ aμν j ( x ) .
ν
†
ν =0 ν =0
With the aid of (2.3.7) we write the flow conservation equation (2.3.6) in the
system Σ′ like this
0=
i′
∂c ρ ′ ( x ′ ) 3 ∂j x
+∑
′
≡∑
( )
3 ∂j
μ′
x′
.
( ) (2.3.13)
c ∂t ′ i =1 ∂x i ′ μ =0 ∂x μ ′
We insert (2.3.12) and (1.2.4) this way
3 3
∂ −1 3 3
∂ 3 ⎛ 3 −1 ⎞ν
0 = ∑∑ λ λμ ∑ μν
a a j ν
( x ) = ∑ λ ∑ ⎜ ∑ λμ μν
a a ⎟j ( x )
μ =0 λ =0 ∂x ν =0 λ = 0 ∂x ν = 0 ⎝ μ = 0 ⎠ (2.3.14)
3
∂ 3 3
∂j ( x ) λ
= ∑ λ ∑ δ λν j ν ( x ) = ∑ λ
.
λ =0 ∂x ν =0 λ = 0 ∂x
K ⎛ 3 K ⎞ K
EΨ ( r , t ) = ⎜ c ∑ α i ( p i − eAi ) + β m0c 2 + eΦ ( r , t ) ⎟Ψ ( r , t ) . (2.4.1)
⎝ i =1 ⎠
K K
Clearly, the potential Φ ( r , t ) is connected with the 4 × 4 unit matrix. Since Ψ ( r , t )
is an eigenfunction for the energy operator we have
K K K
EΨ ( r , t ) = EΨ ( r , t ) ≡ ( E ′ + m0c 2 )Ψ ( r , t ) . (2.4.2)
E ′ mo c 2 . (2.4.3)
22
K
We split the four-component vector Ψ ( r , t ) in two two-component spinors
ϕ and χ like this
K
K ⎛ ϕ (r ,t ) ⎞
Ψ ( r , t ) = ⎜⎜ K ⎟⎟ . (2.4.4)
⎝ χ (r ,t ) ⎠
Using (2.1.15), (2.4.2) and (2.4.4) we write eq. (2.4.1) this way
⎛ϕ ⎞
(E′ + m c ) ⎜ χ ⎟
0
2
⎝ ⎠
⎡ 3 ⎛0 σi ⎞
⎛1 0 ⎞ ⎛ 1 0 ⎞⎤ ⎛ ϕ ⎞
= ⎢c ∑ ( p i − eAi ) ⎜ ⎟ + m0c ⎜
2
⎟ + eΦ ⎜ ⎟⎥ ⎜ ⎟ (2.4.5)
⎜ 0 ⎠⎟
⎢⎣ i =1 ⎝σ i ⎝ 0 −1⎠ ⎝ 0 1 ⎠ ⎥⎦ ⎝ χ ⎠
⎛ϕ ⎞ ⎛σ χ ⎞ ⎛0⎞ ⎛ϕ ⎞
or E ′ ⎜ ⎟ = c ∑ ( p i − eAi ) ⎜ i ⎟ − 2m0c 2 ⎜ ⎟ + eΦ ⎜ ⎟ .
3
⎜σ ϕ ⎟
⎝χ⎠ i =1 ⎝ i ⎠ ⎝χ⎠ ⎝χ⎠
( E ′ − eΦ ) ϕ = c ∑ ( p i − eAi )σ i χ
3
i =1
(2.4.6)
( E ′ − eΦ + 2m c ) χ = c ∑ ( p − eA )σ ϕ .
3
2 i i
0 i
i =1
i =1
first equation. The resulting relation contains the nonrelativistic spinor wave
function ϕ .
E ′ϕ
= c ∑ ( p i − eAi )σ i c ∑ ( p j − eA j ) σ j ϕ + eΦϕ
3 3
1
(2.4.7)
i =1 ⎛ E ′ − eΦ ⎞ j =1
2m0c ⎜ 1 +
2
⎟
⎝ 2m0c 2 ⎠
In addition to (2.4.3) we demand that the potential energy eΦ is also small, i.e.
E ′ − eΦ m0c 2 (2.4.8)
and therefore
−1
⎛ E ′ − eΦ ⎞ E ′ − eΦ
⎜1+ 2 ⎟
≅ 1− .
⎝ 2m0 c ⎠ 2m0 c 2
1 ⎛ E′ ⎞ 3
( ) ( p j − eA j )σ j ϕ
3
E ′ϕ ≅ ⎜1− 2 ⎟ ∑
2m0 ⎝ 2m0c ⎠ i =1
p i
− eA i
σ i ∑
j =1
. (2.4.9)
( pi − eAi )σ i 2meΦc 2 ∑ ( p − eA )σ
3 3
1
+
2m0
∑
i =1 j =1
j j
j
ϕ + eΦϕ
0
We deal with the first expression on the right hand side of (2.4.9)
∑ ( p − eA )( p − eA ) σ σ
3
i i j j
i j
=
i , j =1
∑ ( p p + e A A ) (σ σ )
2
∑(p )σ
3 3
− eA + +σ jσ i
k k 2 i i 2 i j
k i j
(2.4.10)
k =1 i < j =1
{
−e ⎡⎣( p1 A2 + A1 p 2 ) σ 1σ 2 + ( p 2 A1 + A2 p1 ) σ 2 σ 1 ⎤⎦ + cyclic terms . }
Using the relations (2.1.13), (2.1.14) and
∂ K
p i ( A j ") = −i= i (
A j ( r ) ")
∂x
K (2.4.11)
⎛ ∂A j ( r ) ⎞ j K ⎛ ∂ ⎞ j K j K
= −i= ⎜ ⎟" − A ( r ) i= ⎜ i " ⎟ = ( p A ( r ) )" + A ( r ) ( p ")
i i
⎝ ∂x ⎠ ⎝ ∂x ⎠
i
K
, where three points represent a function of r , we obtain
∑ ( p − eA )( p − eA ) σ σ
3
i i j i
i j
i , j =1
2
= ∑ ( p − eA ) 1+ 0
3
k k
k =1
{
−e ⎡⎣ p1 A2 − A2 p1 − ( p 2 A1 − A1 p 2 ) ⎤⎦ σ 1σ 2 + cyclic terms }
2
= ∑ ( p − eA ) 1 − e {⎡⎣ p }
3
k k 1
A2 − A2 p1 ⎤⎦ iσ 3 + cyclic terms (2.4.12)
k =1
2
K
= ∑ ( p k − eAk ) 1 − e [ p× A]3 iσ 3 + cyclic terms { }
3
k =1
2
K K
= ∑ ( p − eA ) 1 − e=∑ ⎡⎣∇ × A⎤⎦ σ
3 3
k k
l l
k =1 l =1
2
= ∑ ( p − eA ) 1 − e= ∑ B σ
3 3
k k
l l
.
k =1 l =1
In the last line equation (1.3.1) has been inserted. For later applications we make
up the following modification of (2.4.12) setting Ak equal to zero
K
∑ σ i pi σ j p j = ∑ ( pk ) ≡ p2 .
3 3
2
(2.4.13)
i , j =1 k =1
24
In order to show clearly the so-called Pauli term, we choose a special case of eq.
(2.4.9) with
K
E′ eΦ ( r )
≅ 0, ≅0. (2.4.14)
2m0c 2 2m0c 2
K ⎡ 1 e= 3 ⎤ K
∑(p − eAk ) + eΦ −
3
E ′ϕ ( r , t ) ≅ ⎢ ∑ Bl σ l ⎥ ϕ ( r , t ) .
k 2
(2.4.15)
⎣ 2m0 k =1 2m0 l =1 ⎦
This is the well-known Pauli equation, which was postulated by W. Pauli in the
framework of classical quantum mechanics, where the last term in (2.4.15) was
derived independently. The expression
= K K K
σ =S ≡S (2.4.16)
2
is the well-known spin operator for two-spinors. With its help the last expression in
(2.4.15) can be written as
e K K
− B ⋅S . (2.4.17)
m0
The two components of ϕ , therefore, describe the spin degrees of freedom in the
nonrelativistic equation for spin-½ fermions. Since spin exists both at low as well
as at high velocities, this implies that the Dirac equation describes fermions with
spin ½. This is another remarkable consequence of this equation.
The last term of (2.4.15) contains the correct value of the magnetic moment
e= J
μ= = 9.2740154 ⋅ 10 −24 , (2.4.18)
2m0 T
which is positive for a positive charge e .The observed magnetic moment of the
electron agrees to about 1%o.
In order to show explicitly the gyromagnetic ratio g = 2 of the electron, we take a
K K
homogeneous magnetic field B . Its vector potential A reads
K 1K K
A = B × r or A1 = ( B2 x 3 − B3 x 2 ) with cyclic permutations.
1
2 2
K K
The relation B = curl A (cf. (1.3.1)) can easily be shown. Supposed the magnetic
K
field (and A ) is feeble, i.e. ( Ak ) and ( p k Ak ) are negligible, the first expression
2
on the right hand side of (2.4.15) reads with the aid of (1.2.7), (2.4.11) and the
K K K K K K
( )
geometric relation B × r ⋅ p = B ⋅ ( r × p )
25
2
1 K2 e
∑(p )
3 3
1
2m0 k =1
k
− eA k
≅
2m0
p −
m0
∑A p
k =1
k k
K e 1 K K K 1 K2 e K K K
=
1
2m0
p2 −
m0 2
(
B×r ⋅p =
2m0
p −
2m0
)B ⋅ (r × p) (2.4.19)
1 K e K K
= p2 − B ⋅L .
2m0 2m0
K K K
Here the well-known angular momentum L = r × p has been used. Inserting
(2.4.19) and (2.4.16) in (2.4.15) we obtain
K ⎡ 1 K2 e K K K⎤ K
E ′ϕ ( r , t ) ≅ ⎢ p − (
L + 2S ⋅ B ⎥ ϕ ( r , t ) )
(2.4.20)
⎣ 2m0 2m0 ⎦
K
The coefficient of S is the gyromagnetic ratio, and it has the correct value 2.
In the second special case we exclude permanent magnetic fields and set
Ai = 0 , (2.4.21)
K 1 ⎛ E′ ⎞ 3 3
K
2 ⎟∑ i ∑
E ′ϕ ( r , t ) ≅ ⎜ 1 − σ p i
σ j p j ϕ (r ,t )
2m0 ⎝ 2m0c ⎠ i =1 j =1
K (2.4.22)
1 3 i eΦ ( r , t ) K K K
3
+
2m0 i =1
∑ σ i
p 2 ∑
2m0c j =1
σ j p j ϕ ( r , t ) + eΦ ( r , t ) ϕ ( r , t ) .
K ⎡ 1 ⎛ E′ ⎞ 3 eΦ 3 2⎤ K
E ′ϕ ( r , t ) = ⎢ ⎜ 1 − 2 ⎟∑ ( p )
k 2
+ 2 2 ∑(
pk ) ⎥ ϕ ( r , t )
⎣ 2m0 ⎝ 2m0c ⎠ k =1 4m0 c k =1 ⎦
K (2.4.24)
i=e 3 ⎛ ∂Φ ( r , t ) ⎞ 3 K K K
2 2 ∑ i ⎜ ⎟ ∑ σ j p ϕ ( r , t ) + eΦ ( r , t ) ϕ ( r , t ) .
− σ j
4m0 c i =1 ⎝ ∂x i
⎠ j =1
K
⎛ ∂Φ ( r , t ) ⎞ 3
⎟ ∑ σ j ( p ")
3
∑ σi ⎜ i
⎝ ∂x
i
i =1 ⎠ j =1
K
2 ∂Φ ( r , t )
( p j ")
3
= ∑σ j
j =1 ∂x j
K K
⎪⎧ ⎛ ∂Φ ( r , t ) 2 ∂Φ ( r , t ) 1 ⎞ ⎪⎫
+ ⎨σ 1σ 2 ⎜ p − p ⎟ + cyclic terms ⎬"
⎝ ∂x ∂x
1 2 (2.4.25)
⎩⎪ ⎠ ⎭⎪
K
∂Φ ( r , t ) j
( { K K
p ") + i σ 3 ⎡⎣grad Φ ( r , t ) × p ⎤⎦ 3 + cyclic terms " }
3
=∑
j =1 ∂x j
K
∂Φ ( r , t ) j K K
( )
3 3
=∑ p " + i ∑ σ k ⎡⎣grad Φ ( r , t ) × p ⎤⎦ k " ,
j =1 ∂x j
k =1
where (2.1.13) and (2.1.14) have been applied. We insert (2.4.25) in (2.4.24)
obtaining
K K
K ⎡⎛ E ′ − eΦ ( r , t ) ⎞ p 2 K ⎤ K
E ′ϕ ( r , t ) = ⎢⎜ 1 − 2 ⎟ + eΦ ( r , t ) ⎥ ϕ ( r , t )
⎢⎣⎝ 2m0c ⎠ 2m0 ⎥⎦
K
= 2e 3 ∂Φ ( r , t ) ∂ K
2 2 ∑
− ϕ (r ,t ) (2.4.26)
4m0 c i =1 ∂x i
∂x i
=e 3
K K K
+
4m02c 2
∑σ
k =1
k
⎡⎣grad Φ ( r , t ) × p ⎤⎦ k ϕ ( r , t ) .
K
Further simplifications can be made, if Φ ( r , t ) is spherically symmetric, i.e.
K
Φ ( r , t ) = Φ ( r , t ) , which results in
K
3
∂Φ ∂ ∂Φ ∂ ∂Φ r
∑
i =1 ∂x ∂x
i i
=
∂r ∂r
, grad Φ =
∂r r
. (2.4.27)
K K K K
The angular momentum operator L = r × p , the spin operator S , (2.4.16), and the
relations (2.4.27) are inserted in (2.4.26) like this
K K
K ⎡ p2 E ′ − eΦ ( r , t ) p 2 ⎤ K
E ′ϕ ( r , t ) = ⎢ − 2
+ eΦ ( r , t ) ⎥ ϕ ( r , t )
⎣ 2m0 2m0c 2m0 ⎦
. (2.4.28)
⎡ e= 2 ∂Φ ( r , t ) ∂ e 1 ∂Φ ( r , t ) K K ⎤ K
+ ⎢− + S ⋅ L⎥ϕ (r ,t )
⎣ 4m0
2
∂r ∂r 2m02c 2 r ∂r ⎦
The first and the third term on the right hand side of (2.4.28) give the nonrelativistic
Schrödinger equation. The last term is the spin-orbit energy, which is well-known
in quantum mechanics, where it was introduced artificially in order to explain the
energy fine structure of atoms. However, using the Dirac equation this term
emerges automatically. It is also very important in nuclear physics (together with
the last but one expression) for the classification of single-particle states of nuclei,
but there the potential is not only of electromagnetic origin.
27
2.5 The Dirac equation for particles with an anomalous magnetic moment
In the nonrelativistic equation (2.4.15),
K ⎡ 1 e= ⎤ K
∑(p − eAk ) + eΦ −
3 3
E ′ϕ ( r , t ) = ⎢ ∑B σ ⎥ϕ (r ,t ) ,
k 2
l l
(2.5.1)
⎣ 2m0 k =1 2m0 l =1 ⎦
the last term represents the interaction energy between the normal magnetic
moment
e=
μ= (2.5.2)
2m0
of the Dirac particle and the magnetic field. On the other hand, experimentally, the
magnetic moment of the proton (rest mass M p ) has been determined to
ep =
μ p = (1 + K p ) with K p = 1.79 and ep = e , (2.5.3)
2M p
(where e is the elementary charge), and for the neutron has been found
en =
μn = K n with K n = −1.91 and en = e . (2.5.4)
2Mn
We maintain that the extended, relativistic Dirac equation for the proton reads
(cf.(2.1.20))
K ⎡ 3 ⎤ K
EΨ ( r , t ) = ⎢c ∑ α i ( p i − eAi ) + β M pc 2 + eΦ ⎥Ψ ( r , t )
⎣ i =1 ⎦
(2.5.6)
e= 3
⎛ Ei ⎞ K
− βK p ∑ ⎜ Σ i Bi − iα i
2M p i =1 ⎝ c ⎠
⎟Ψ ( r , t ) ,
⎛σ 0⎞
where Σ i = ⎜ i ⎟ , and Bi and Ei are the components of the strengths of the
⎜0 σ ⎟
⎝ i ⎠
magnetic and electric external fields respectively. For the neutron an analogous
equation holds, which differs by the absence of the terms eΦ and eAi .
In order to show that (2.5.6) achieves the right nonrelativistic form, we write this
equation explicitly with spinors from (2.4.4)
28
K K
⎛ ϕ (r ,t ) ⎞ 3 ⎛ 0 σi ⎞ ⎛ ϕ (r ,t ) ⎞
E ⎜⎜ K ⎟⎟ = c ∑ ⎜ ⎟ ( p − ep A ) ⎜⎜ K ⎟⎟
i i
⎝ χ (r ,t ) ⎠ ⎝ χ (r ,t ) ⎠
⎜ 0 ⎟⎠
i =1 ⎝ σ i
K
⎡⎛ 1 0 ⎞ ⎛ 1 0⎞ ⎤ ⎛ ϕ (r ,t ) ⎞
+ ⎢⎜ ⎟ M pc + ⎜ ⎟ epΦ ⎥ ⎜⎜ K ⎟⎟
2
(2.5.7)
⎢⎣⎝ 0 −1⎠ ⎝ 0 1⎠ ⎥⎦ ⎝ χ ( r , t ) ⎠
⎛1 0 ⎞ ep = 3 ⎡⎛ σ i 0⎞ ⎛0 σ i ⎞ E ⎤ ⎛ ϕ ( rK , t ) ⎞
−⎜ ⎟ Kp ∑ ⎢⎜
⎜
⎟ Bi − i ⎜
σ i ⎟⎠ ⎜σ
⎟ i ⎥ ⎜⎜ K ⎟⎟
0 ⎟⎠ c ⎥ ⎝ χ ( r , t ) ⎠
⎝ 0 −1⎠ 2M p i =1 ⎢⎝ 0
⎣ ⎝ i ⎦
or using (2.4.2)
⎛ ϕ (r ,t ) ⎞ ⎛σ i χ ⎞ ep = 3 ⎛σ χ ⎞
⎟⎟ ≅ c ∑ ( p − eA ) ⎜⎜
3
E ′ ⎜⎜ i i
⎟ + K pi ∑ Ei ⎜ i ⎟
⎝ χ (r ,t ) ⎠
⎟ 2M pc i =1 ⎜⎝ −σ i ϕ ⎟⎠
i =1 ⎝σ iϕ ⎠
(2.5.8)
⎛0⎞ ⎛ϕ ⎞ e= 3 ⎛ σ iϕ ⎞
− 2M pc 2 ⎜ ⎟ + eΦ ⎜ ⎟ − K p p ∑ Bi ⎜ ⎟.
⎜ −σ χ ⎟
⎝χ⎠ ⎝χ⎠ 2M p i =1 ⎝ i ⎠
Due to (2.4.8) in the nonrelativistic limit the quantity Φ is relatively small. Because
∂Φ
of Ei = − i (cf. (1.3.1)), neglecting time dependence) the electrostatic energy
∂x
=ep =e ∂Φ =
Ei = − p is extremely small, since 2 ⋅ 10 −14 m is a very small
M pc M pc ∂x i
M pc
=ep
length. Therefore Ei can be neglected relative to the energy cp . We obtain
M pc
the following couple of equations
⎛ ep = ⎞
( p i − eAi )σ i χ
3 3
⎜⎜ E ′ − epΦ + K p
2M p
∑B σ i ⎟
i ⎟
ϕ ≅ c ∑ (2.5.9)
⎝ i =1 ⎠ i =1
⎛ e= ⎞
( pi − eAi )σ i ϕ . (2.5.10)
3 3
and ⎜ E ′ − epΦ + 2M pc 2 − K p p
⎜ 2M p
∑ Bi σ ⎟
i ⎟
χ ≅ c ∑
⎝ i =1 ⎠ i =1
ep =
In (2.5.10) the expression K p Bi is much smaller than the energy M pc 2 and
2M p
can be neglected. We insert (2.5.10) in (2.5.9). With the help of (2.4.8) we obtain
∑ ( p − eA )σ ∑ ( p − eA )σ ϕ
3 3
1
E ′ϕ ≅ i i
i
j j
j
2M p i =1 j =1
(2.5.11)
ep = 3
+ epΦϕ − K p
2M p
∑B σ ϕ .
i =1
i i
K ⎡ 1 ⎤ K
∑(p − ep Ak ) + epΦ ⎥ ϕ ( r , t )
3
E ′ϕ ( r , t ) ≅ ⎢ k 2
⎣⎢ 2M p k =1 ⎦⎥ (2.5.12)
ep = K
− (1 + K p )
3
2M p
∑ Bl σ l ϕ ( r , t ) .
l =1
Thus, we have shown that the extension in the Dirac equation (2.5.6) contributes
to the correct magnetic moment in the nonrelativistic limit. From the considerations
in context with (2.1.9) we know that the additional term in (2.5.6) has to be
Hermitian. The summands of the extension in (2.5.6),
ep = ep =
Kp Bi β Σ i and − Kp Ei βα i , (2.5.13)
2M p 2M pc
contain real coefficients and the matrix products β Σ i and iβα i respectively. We
show that these matrices are really Hermitian like this
⎛σ 0 ⎞⎛ 1 0 ⎞
( )
†
βΣ i =Σi β =Σiβ =⎜ i
† †
⎟⎜ ⎟ = βΣ i ,
⎜0 σ i ⎟⎠ ⎝ 0 −1⎠
⎝
⎛0 σ i ⎞⎛ 1
( )
† 0⎞
iβα i = −iα i β = −iα i β = −i ⎜ ⎟⎜ ⎟ = iβα i .
† †
⎜σ 0 ⎟⎠ ⎝ 0 −1⎠
⎝ i
3.1 The wave function of a free particle, helicity, the velocity operator
We start from eq. (2.1.3), which does not contain a potential (i.e. the particle is
free) using (1.2.6)
K
∂Ψ ( r , t ) K ⎛ 3 ⎞ K K
i= ≡ EΨ ( r , t ) = ⎜ c ∑ α i p i + β m0c 2 ⎟Ψ ( r , t ) ≡ Hf Ψ ( r , t ) (3.1.1)
∂t ⎝ i =1 ⎠
∂
It is easy to show that every component of the momentum operator p k = −i= ,
∂x k
(1.2.7), commutes with the Hamilton operator of the free particle,
3
Hf = c ∑ α i p i + β m0c 2 , (3.1.1). Consequently, there exist eigenfunctions with
i =1
defined energy and momentum, as we know from the methods K of quantum
mechanics. The following ansatz is such an eigenfunction of p
K KK ⎛ 3
⎞
ψ ( r ) = ψ 0eip⋅r / = ≡ ψ 0 exp ⎜ i∑ p i x i / = ⎟ . (3.1.4)
⎝ i=1 ⎠
K ∂ ⎛ 3 ⎞
p1ψ ( r ) = −i= 1 ψ 0 exp ⎜ i∑ p i x i / = ⎟
∂x ⎝ i =1 ⎠
(3.1.5)
⎛ 3
⎞
= p1ψ 0 exp ⎜ i∑ p i x i / = ⎟
⎝ i =1 ⎠
K KK
Ψ ( r ,t ) = ψ 0ei( p⋅r -Et ) / = . (3.1.7)
⎛ϕ0 ⎞
ψ0 =⎜ ⎟ (3.1.8)
⎜χ ⎟
⎝ 0⎠
⎛ ϕ01 ⎞ ⎛ χ 01 ⎞
ϕ0 = ⎜ 2 ⎟ and χ 0 = ⎜ 2 ⎟ . (3.1.9)
⎝ ϕ0 ⎠ ⎝ χ0 ⎠
⎛ϕ0 ⎞ K K ⎛ 3 ⎞⎛ϕ0 ⎞ K K
E ⎜ ⎟ ei( p⋅r −Et ) / = = ⎜ c ∑ α i p i + β m0c 2 ⎟ ⎜ ⎟ ei( p⋅r −Et ) / =
⎜χ ⎟ ⎝ i =1 ⎠ ⎜⎝ χ 0 ⎟⎠
⎝ 0⎠
where we have used the expressions (2.1.15). Equation (3.1.11) represents the
following system
3
Eϕ 0 = c ∑ p i σ i χ 0 + m0c 2 ϕ 0
i =1
3
E χ 0 = c ∑ p i σ i ϕ 0 − m0c 2 χ 0
i =1
(E − m c )ϕ
3
or 0
2
0
− c ∑ pi σ i χ 0 = 0
i =1
(3.1.12)
(E + m c ) χ
3
0
2
0
− c∑ p σ i ϕ0 = 0 .
i
i =1
( E − m0c 2 )1
3
−c ∑ p i σ i
i =1
=0
( E + m c )1
3
−c ∑ p σ i i
0
2
(3.1.13)
i =1
(E − m02c 4 ) 1 − c 2 ∑ p i σ i p j σ j = 0 .
3
2
or
i , j =1
K
∑ ( p k ) 1 ≡ p 2 1.
3 3
∑
2
p i j
p σ i
σ j
= (3.1.14)
i , j =1 k =1
E = + m02c 4 + c 2 p 2 (3.1.16)
⎛ ϕ0 ⎞
⎜ 3 ⎟ KK
K N ⎜ ⎟ ei( p⋅r −Et ) / = ,
Ψ ( r ,t ) = c∑ p σ i
i
(3.1.18)
V ⎜ ⎟
⎜⎜
i =1
ϕ ⎟
0⎟
⎝ E + m0c
2
⎠
with which we determine the constant N . The adjoint free particle wave function
reads
⎛ 3
⎞
K ⎜
N * † † i =1
c ∑ σ i pi ⎟ K K
Ψ † ( r ,t ) = ⎜ ϕ 0 ,ϕ 0 ⎟ e −i( p⋅r −Et ) / = . (3.1.20)
V⎜ E + m0c ⎟2
⎜ ⎟
⎝ ⎠
⎛ 2⎛
3
⎞
2
⎞
⎜ c ⎜∑p σ i ⎟i
⎟
2⎜ †
N ϕ 0ϕ 0 + ϕ 0
† ⎝ i =1 ⎠ ϕ 0 ⎟ = 1. (3.1.21)
⎜
⎜ ( E + m0c ) ⎟⎟
2 2
⎜ ⎟
⎝ ⎠
⎛ K ⎞
⎜ c 2 p2
2
⎟ = 1.
N 1+
⎜ ( E + m c 2 )2 ⎟
⎝ 0 ⎠
1+
K
c 2 p2
=
(E + m c ) + E − m c
0
2 2 2 2
0
4
=
2E
(E + m c
0 )
2 2
(E + m c ) 0
2 2 E + m0c 2
E + m0c 2
and N = , (3.1.22)
2E
⎛ ϕ01 ⎞
⎜ 2⎟
ϕ
ψ0 ≅ ⎜ 0 ⎟.
⎜0⎟
⎜⎜ ⎟⎟
⎝0⎠
K
Then, the free particle function Ψ ( r ,t ) in essence is a two-spinor, which we
expect, because Dirac particles are spin ½ particles (section 2.4).
K
We continue to specify the wave function Ψ ( r ,t ) , (3.1.18), i.e. to determine the
spinor ϕ 0 . We introduce the helicity operator
⎛ 3 ⎞
K K
= ⎛σ ⋅ p 0 ⎞ ⎜ ∑
= ⎜ i =1
σ i ⋅ pi 0 ⎟
Sp = ⎜ K K⎟ ≡ ⎟. (3.1.23)
2p⎝ 0 σ ⋅ p⎠ 2 p ⎜ 3
i ⎟
⎜
⎝
0 ∑
i =1
σi ⋅p ⎟
⎠
We show that this operator commutes with the Hamilton operator like this
⎡ 3 ⎤
⎡⎣Hf ,Sp ⎦⎤ = ⎢c ∑ α i p i + β m0c 2 ,Sp ⎥
⎣ i =1 ⎦
K K K K
⎡ ⎛ 0 σ ⋅ p⎞ ⎛1 0 ⎞ 2 ⎛
σ ⋅p 0 ⎞ = ⎤
= ⎢c ⎜ K K ⎟+⎜ ⎟ m0c , ⎜ K K⎟ ⎥
⎣⎢ ⎝ σ ⋅ p 0 ⎠ ⎝ 0 −1⎠ ⎝ 0 σ ⋅ p ⎠ 2 p ⎦⎥
K K K K
⎡⎛ 0 σ ⋅ p ⎞ ⎛σ ⋅ p 0 ⎞⎤ = (3.1.24)
= c ⎢⎜ K K ⎟ ,⎜ K K ⎟⎥
⎣⎢⎝ σ ⋅ p 0 ⎠ ⎝ 0 σ ⋅ p ⎠ ⎦⎥ 2 p
K 2 K 2
c= ⎜ 0
⎛ (σK ⋅ p ) ⎞⎟ c = ⎛⎜ 0 (σK ⋅ p ) ⎟⎞
= − = 0,
2 p ⎜⎜ σK ⋅ pK 2 ⎟⎟ 2 p ⎜⎜ K K 2 ⎟⎟
⎝( ) 0 ⎠ ⎝( σ ⋅ p ) 0 ⎠
whereK (2.1.15) has been used. Apparently, the helicity operator commutes also
with p , for which reason a wave function can be found, which is eigenfunction of
these operators and of Hf .
K K
We construct this function and put the z-axis in the direction of p i.e. p = ( 0, 0,p ) .
K
Then the function Ψ ( r , t ) ,( 3.1.18), reads
⎛ ϕ0 ⎞
K N ⎜ ⎟ i( pz −Et ) / =
Ψ (z) ( r , t ) = ⎜ cpσ z ⎟e .
V⎜ ϕ 0⎟
⎝ E + m0c
2
⎠
Sp = Sp , x + Sp , y + Sp , z
= ⎛σ x 0 ⎞ = ⎛σ y 0 ⎞ = ⎛σ z 0 ⎞
= ⎜ ⎟ px + ⎜ ⎟ py + ⎜ ⎟p .
2p ⎜ 0 ⎟
σx⎠ 2p ⎜ 0 σy ⎟ 2p ⎜ 0 σ ⎟ z
⎝ ⎝ ⎠ ⎝ z⎠
K
If Sp acts on Ψ ( z ) ( r , t ) , only the term Sp ,z contributes. With the given choice of
the coordinate system and with the eigenvalue of pz this operator reduces to
⎛1 0 0 0⎞
= p ⎛σ 3 0 ⎞ = ⎜ 0 −1 0 0⎟
⎟
Sp,z = ⎜ ⎟= ⎜ . (3.1.25)
2 p ⎜⎝ 0 σ 3 ⎟⎠ 2 ⎜ 0 0 1 0⎟
⎜⎜ 0 0 ⎟
0 −1⎟⎠
⎝
⎛ 1⎞ ⎛0⎞
ϕ 0 = ϕ 0+ ≡ ⎜ ⎟ or ϕ 0 = ϕ 0− ≡ ⎜ ⎟ , (3.1.26)
⎝0⎠ ⎝ 1⎠
K
see (3.1.9), the function Ψ ( r ,t ) is an eigenfunction of Sp,z . Namely, we have for
instance
⎛ ϕ 0 ,+ ⎞
⎜ ⎟ N i( pz −Et ) / =
Sp,zΨ E ,p,+ ≡ Sp,z ⎜ cpσ ⎟ e
z
⎜ E + m c 2 0 ,+ ⎟⋅ ϕ V
⎝ 0 ⎠
⎛ 1 ⎞
⎛ 1 0 0 0 ⎞⎜ ⎟
⎜ ⎟⎜ 0 ⎟ N
= 0 −1 0 0 ⎟
= ⎜ e(
i pz − Et ) / =
⎜ cp ⎟
2 ⎜0 0 1 0 ⎟⎜ 2 ⎟ V
⎜⎜ ⎟⎟ ⎜ E + m0c ⎟
⎝ 0 0 0 −1⎠ ⎜ ⎟
⎝ 0 ⎠
⎛ 1 ⎞
⎜ ⎟
⎜ 0 ⎟ N
= =
e(
i pz −Et ) / =
= ⎜ cp ⎟ = Ψ E ,p,+
2⎜ 2 ⎟ V 2
⎜ E + m 0 c ⎟
⎜ ⎟
⎝ 0 ⎠ (3.1.27)
=
with the eigenvalue + . Analogously, the function
2
⎛ 0 ⎞
⎜ ⎟
⎜ 1 ⎟ i pz −Et / =
K N
Ψ E ,p,− ( r ,t ) = ⎜ 0 ⎟e ( )
, (3.1.28)
V⎜ ⎟
⎜ −cp ⎟
⎜ E + m c2 ⎟
⎝ 0 ⎠
36
=
which contains ϕ 0,− , generates the eigenvalue − , if Sp,z acts on it. The
2
normalization condition (3.1.10) is met by the choices (3.1.26) for ϕ 0 .
We interpret the result (3.1.27) and the corresponding one for (3.1.28) as follows.
In (2.4.16) the spin operator for two-spinors is given. We generalize it for four-
spinors like this
⎛1 0 0 0⎞
K ⎜ ⎟
K = ⎛σ 0 ⎞ = ⎜ 0 −1 0 0⎟
S= ⎜ K ⎟, for example Sz = (3.1.29)
2⎝0 σ ⎠ 2 ⎜0 0 1 0⎟
⎜⎜ ⎟
⎝0 0 0 −1⎟⎠
= K
Using (2.5.6) we formulate Sz = Σ z . We write the quadrate of S with the help of
2
(2.1.13) this way
⎛ 3 2 ⎞
K 2 =2 ⎛ σ
K 2
0 ⎞ ⎜ ∑
= ⎜ i =1
2
σi 0 ⎟
=2 3
S = ⎜ ⎟ = ⎟ = 1 (3.1.30)
4 ⎜⎝ 0 σK 2 ⎟⎠ 4 ⎜ 3
2⎟ 4
⎜ 0
⎝
∑
i =1
σi ⎟
⎠
K
or S 2 = =2 ⋅ s ( s + 1) ⋅ 1 with s = 1 / 2 . (3.1.31)
This means in analogy with quantum mechanics that the Dirac particles have spin
=s = = / 2 , which we have found already in section 2.4. Because Sp,z , (3.1.25),
agrees with Sz , (3.1.29), we see that in (3.1.27) the spin = / 2 of the particle is
K
directed along the z-axis or along p . In the state (3.1.28) it looks backward. In
section 3.6 we will see that the (massless) neutrinos show a special behaviour of
the helicity.
We now deal with the velocity operator and its expectation value. In
nonrelativistic quantum mechanics there is a correspondence between the time
derivative of an operator O (which does not depend explicitly on time) and its
commutator with the Hamilton operator like this
dO 1
= [O , H ] . (3.1.32)
dt i=
We extend this theorem (Heisenberg equation) over the relativistic theory. We
formulate it with the position vector as operator
K
dr 1 K
= [ r , Hf ] , (3.1.33)
dt i=
which represents the velocity operator. We calculate the right hand side of (3.1.33)
using (2.1.8)
37
K
dr c ⎡ K 3 = ∂ ⎤
= ⎢r , ∑ α i + β m0c 2 ⎥
dt i= ⎣ i =1 i ∂x i
⎦
.
⎡K 3 ∂ ⎤
= − ⎢r , ∑ α i i ⎥
⎣ i =1 ∂x ⎦
K
The first component of r yields
dx 1 ⎡ ∂ ⎤ ⎛ ∂ ∂ ⎞
= −c ⎢ x 1 ,α 1 1 ⎥ = −cα 1 ⎜ x 1 1 − 1 − x 1 1 ⎟ = cα 1
dt ⎣ ∂x ⎦ ⎝ ∂x ∂x ⎠
K
i.e.
dr
dt
K K
(
= cα with α = α 1 ,α 2 ,α 3 . ) (3.1.34)
This form of the velocity operator is not so strange as it looks. To show this, we
calculate the average or expectation value of the velocity. In analogy with
nonrelativistic quantum mechanics we write it like this
K
K dr † K K K
v ≡ = ∫Ψ ( r ,t ) cαΨ ( r ,t ) dV , (3.1.35)
dt V
K
which is, in fact, the integral over the current density j (cf. (2.3.5)). We restrict
ourselves on particles moving in the z-direction. Due to (3.1.27) and (3.1.22) the
four-spinor wave function reads
K
Ψ E ,p,+ ( r ,t )
⎛ 1 ⎞ ⎛ 1 ⎞
⎜ ⎟ ⎜ ⎟
⎜ 0 ⎟ ⎜ 0 ⎟
N i( pz − Et ) / = E + m0c 2 . (3.1.36)
= ⎜ cp ⎟e = ⎜ cp ⎟ ei( pz −Et ) / =
V⎜ 2 ⎟ 2EV ⎜ 2 ⎟
⎜ E + m0c ⎟ ⎜ E + m0c ⎟
⎜ ⎟ ⎜ ⎟
⎝ 0 ⎠ ⎝ 0 ⎠
K
The z-component of v in (3.1.35) reads
⎛ 1 ⎞
⎛0 0 1 0 ⎞⎜ ⎟
⎜ ⎟ 0
E + m0c ⎛
2
cp ⎞ ⎜0 0 0 −1⎟ ⎜ ⎟
vz = ⎜ 1, 0 , , 0 ⎟ c ⎜ cp ⎟ ∫ dV
2EV ⎝ E + m0c 2 ⎠ ⎜ 1 0 0 0 ⎟⎜ 2 ⎟V
⎜⎜ ⎟ E + m0c ⎟ (3.1.37)
⎝ 0 −1 0 0 ⎟⎠ ⎜⎜ ⎟
⎝ 0 ⎠
E + m0c 2 cp c2p
= 2 c = .
2E E + m0c 2 E
m0v m0c 2
p= , E= (3.1.38)
1 − (v / c ) 1 − (v / c )
2 2
38
as we expect.
A1 = A3 = 0, A2 = Bx and Φ = 0 . (3.2.1)
∂A2 ∂A1
Bx = By = 0. Bz = − =B. (3.2.2)
∂x ∂y
K ⎡ 3 ⎤ K
EΨ ( r ,t ) = ⎢c ∑ α i ( p i − eAi ) + β m0c 2 ⎥Ψ ( r ,t ) . (3.2.3)
⎣ i =1 ⎦
In analogy with (3.1.12) we can write (3.2.3) as a pair of equations containing two-
spinors like this
(E − m c )ϕ = c ∑σ ( p − eAi ) χ
3
2 i
0 i
i =1
(3.2.5)
(E + m c ) χ = c ∑σ ( p − eA ) ϕ .
3
2 j j
0 j
j =1
( E 2 − m02c 4 )ϕ = c 2 ∑ ( pi − eAi )( p j − eA j ) σ i σ j ϕ
3
i , j =1
(3.2.6)
⎡ 3 ⎤
= c ⎢ ∑ ( p k − eAk ) 1 − e=∑ Bl σ l ⎥ ϕ ,
3
2 2
⎣ k =1 l =1 ⎦
( ) 2 ⎡
⎢ ∑ ( p ) − ep A − eA p + e A ( ) ⎤
3 2
k 2 ( 2) ( 2) ( 2) ( 2) ( 2)
E 2
− m0
2 4
c ϕ = c 2
− e=Bσ z ⎥ ϕ .
⎣ k =1 ⎦
( ) ( ) ( )
p ( 2) A( 2) " = p ( 2) A( 2) " + A( 2) p ( 2) " = 0 + Bx p ( 2) " and obtain ( )
(E 2
− m02c 4 ) ϕ = c 2 ⎡⎣ px 2 + py 2 + pz 2 − e2Bxpy + e 2B 2 x 2 − e=Bσ z ⎤⎦ ϕ . (3.2.7)
ϕ = f (x)e ( )
i ky y + kz z
(3.2.8)
⎡ d2 ⎤
(E 2
− m02c 4 ) f ( x ) = c 2 ⎢ − =2 2 + = 2k y2 + = 2k z2 − e2Bx =k y + e 2B 2 x 2 − e=Bσ z ⎥ f ( x ) or
⎣ dx ⎦
E2 ⎡ ⎤
− m02c 2 − = 2k z2 ⎢ 2
⎛ ⎞
2
⎛ = k ⎞
2
⎥
2 1 d e
c f ( x ) = ⎢− + ⎜⎜ B ⎟⎟ ⎜ x − y
⎟ − σ z ⎥ f ( x ) . (3.2.9)
e =B ⎢
e dx
B
2
⎝ = ⎠ ⎝ eB ⎠ ⎥
⎣ = ⎦
e ⎛ =k ⎞ ⎡ d2 eB d 2 ⎤
ξ= B⎜x − y ⎟, ⎢i.e. = ⎥ (3.2.10)
= ⎝ eB ⎠ ⎣ dx 2 = dξ 2 ⎦
⎛ E2 ⎞
a = ⎜ 2 − m02c 2 − =2k z2 ⎟ / ( e=B ) (3.2.11)
⎝c ⎠
⎛ d2 ⎞
a f (ξ ) = ⎜ − 2 + ξ 2 − σ z ⎟ f (ξ ) (3.2.12)
⎝ dξ ⎠
⎛1 0 ⎞
σ z f (ξ ) ≡ ⎜ ⎟ f (ξ ) = μ f (ξ ) . (3.2.13)
⎝ 0 −1⎠
⎛ f (ξ ) ⎞
f (ξ ) = ⎜ +1 ⎟ with μ = +1 and
⎝ 0 ⎠
(3.2.14)
⎛ 0 ⎞
f (ξ ) = ⎜ ⎟ with μ = −1 .
⎝ f−1 (ξ ) ⎠
Taking into account both solutions (3.2.14) the equation (3.2.12) is summarised
like this
⎛ d2 2⎞
⎜ 2 − ξ ⎟ fμ ( ξ ) = − ( a + μ ) f μ ( ξ ) , (3.2.15)
⎝ dξ ⎠
fμ ,n (ξ ) = Ce −ξ Hn (ξ ) .
2
/2
(3.2.16)
a + μ −1
n= = 0,1, 2," . (3.2.17)
2
From this formula and from (3.2.11) we obtain
Hence, the energy E is quantized! In order to see the order of magnitude of the
quanta, we calculate the basis quantum of E 2 − m02c 4 for an electron in the field
of 2 Tesla. Due to (3.2.18) we get for μ = n = 1, k z = 0 :
E 2 − m02c 4 = 2c 2 =eB
= 2 ⋅ ( 3 ⋅ 108 ) 1.05 ⋅ 10−341.6 ⋅ 10−19 2 J 2
2
(3.2.19)
= (15.4 eV ) ,
2
which is tiny in view of technical applications. On the other hand, we calculate the
order n for an energy 10 MeV = E 2 − m02c 4 by means of (3.2.19):
E 2 − m02c 4
n= = 4.22 ⋅ 1011
2c =eB
2
i.e. the order of the function (3.2.16) is very high, and the energy steps are
relatively small so that practically every similar energy value can be realized.
With the aid of (3.2.5), (3.2.1), (3.2.8), (3.2.13) and (3.2.10) the spinor χ is
expressed by ϕ like this
41
χ=
c
E + m0c 2 1
(
σ p x + σ 2 ( p y − eBx ) + σ 3 p z ϕ )
⎛ ⎞
+ σ 2 ( =k y − eBx ) + =k z μ ⎟ ϕ
c d
= 2 ⎜
−i=σ 1 (3.2.20)
E + m0c ⎝ dx ⎠
c ⎛ ⎛ d ⎞ ⎞
= ⎜ − =eB ⎜ iσ + σ ξ ⎟ + =k μ ⎟ϕ .
E + m0c 2 ⎝ ⎝ dξ
z
⎠
1 2
⎠
)
e −ξ / 2H n (ξ ) = e −ξ / 2 ( −ξ H n (ξ ) + 2nH n −1 (ξ ) )
(3.2.21)
=e −ξ 2 / 2
(ξ H (ξ ) − H (ξ ) ) .
n n +1
a i( k y + k z ) ⎛ H (ξ ) ⎞
, ϕ μ =+1 = Ce y z e −ξ / 2 ⎜ n
2
n= ⎟ (3.2.22)
2 ⎝ 0 ⎠
E + mo c 2 ⎝ 1⎠
i( k y + k z )
Cce y z ⎛ 1⎞
+ = k μ e −ξ 2 / 2
H ( ξ ) ⎜ ⎟ (3.2.23)
E + mo c 2
z n
⎝0⎠
i( k y + k z )
Cce y z −ξ 2 / 2 ⎛ =k z ⋅ 1 ⋅ H n ( ξ ) ⎞
= e ⎜ ⎟.
E + mo c 2 ⎜ −i =eB ⋅ 2n ⋅ H (ξ ) ⎟
⎝ n −1 ⎠
a
For μ = −1, n = − 1 one obtains
2
⎛ 0 ⎞
ϕ −1 = Ce ( )
i ky y + kz z
e −ξ
2
/2
⎜ ⎟,
⎝ H n ( ξ ) ⎠
i( k y + k z )
(3.2.24)
Cce y z −ξ 2 / 2 ⎛ i =eBH n +1 (ξ ) ⎞
χ −1 = e ⎜⎜ ⎟⎟ .
E + m0c 2 ⎝ − =k z ⋅ 1 ⋅ H n ( ξ ) ⎠
From classical and relativistic mechanics we know that, if the momentum of the
electron is perpendicular to the magnetic field ( pz = =k z = 0 ) , the particle circles in
a plane perpendicular to the field. We set the position vector in the x-direction ,
through which the momentum is directed parallel to the y-axis.
42
j 2 (ξ ( x ) ) = cΨ (ξ ) α 2Ψ (ξ )
†
⎛ 0 σ 2 ⎞⎛ϕ ⎞ (3.2.25)
= c (ϕ*, χ * ) ⎜ ⎟ ⎜ ⎟ = c ⎡⎣ϕ * σ 2 χ + χ * σ 2 ϕ ⎤⎦ .
⎜σ 0 ⎟ χ
⎝ 2 ⎠⎝ ⎠
Cc ⎛ − =eB 2nH n −1 ⎞
e −ξ ( H n , 0 ) ⎜
2
j2 = ⎟
E + m0c 2
⎝ 0 ⎠
) ⎛⎜⎝ iH
0 ⎞
+
Cc
E + m0c 2 (
e −ξ 0,i =eB 2nH n −1
2
⎟
n ⎠
(3.2.26)
Cc
=eB ⋅ 4n ⋅ e −ξ H n (ξ ) H n −1 (ξ ) .
2
=−
E + m0c 2
n Hn (ξ )
0 1
1 2ξ (3.2.27)
2 4ξ 2 − 2
3 8ξ 3 − 12ξ .
For n = 1 we obtain
Cc
j 2 ,n =1 = =eB ⋅ 4 ⋅ e −ξ 2
2ξ ,
E + m0c 2
1 = p
x= + y . (3.2.28)
2 eB eB
E 2 − m02c 4
py = = 2= eB . (3.2.29)
c
Inserting (3.2.29) in (3.2.28) we obtain
3 2= 3 py
x= = . (3.2.30)
2 eB 2 eB
43
p
r = , (3.2.31)
eB
which is near to the characteristic value (3.2.30).
3.3 Dirac equation with central potential. Parity and total angular
momentum
We introduce a central potential V ( r ) into the Hamilton form of the Dirac equation
(2.1.8) like this
3
H = c ∑ α i p i + β m0c 2 + V ( r ) 1 (3.3.1)
i =1
P = β P0 , (3.3.2)
P 2 = β P0 2 = 1.
2
∂
[H,P ] = −c ∑ ⎡⎢α i ⋅ i= ⎤
3
,P ⎥
i =1 ⎣ ∂x i
⎦
3
⎛ ∂ ∂ ⎞
= −ci=∑ ⎜ α i β i P0 − βα i P0 i ⎟ (3.3.3)
i =1 ⎝ ∂x ∂x ⎠
( ∂
)
3
= −ci=∑ α i β + βα i P0 = 0,
i =1 ∂x i
∂ ∂
where we have used (2.1.6) and P0 = − i P0 . Due to the vanishing
∂x i
∂x
commutator (3.3.3) there are eigenfunctions of the Hamiltonian which are also
eigenfunctions of P with a definite parity.
44
K
b) In quantum mechanics the operator J of the total angular momentum appears
as follows
K K K
J = L+S. (3.3.4)
K
The spin operator S is given in (3.1.29), and from section 2.4 we know that the
K
angular momentum operator L reads
K K K
L = r ×p. (3.3.5)
We now calculate
K K K
⎡⎣H , J ⎤⎦ = ⎡⎣H , L ⎤⎦ + ⎡H ,S ⎤ . (3.3.6)
⎣ ⎦
K
First we take the x-term of ⎡⎣H , L ⎤⎦ , i.e.
⎡ 3
⎤
[H ,Lx ] = ⎡⎣H ,ypz − zpy ⎤⎦ = ⎢c ∑ α i p i + β m0c 2 + V ( r ),ypz − zpy ⎥ (3.3.7)
⎣ i =1 ⎦
Obviously, the term β m0c 2 does not contribute. The spherical potential yields in
(3.3.7)
⎛ ∂ ∂ ⎞ ⎛ ∂V ( r ) ∂r ∂V ( r ) ∂r ⎞ ⎛ ∂ ∂ ⎞
⎜y − z ⎟V ( r )Ψ = ⎜ y −z ⎟Ψ + V ( r ) ⎜ y − z ⎟Ψ
⎝ ∂z ∂y ⎠ ⎝ ∂r ∂z ∂r ∂y ⎠ ⎝ ∂z ∂y ⎠
∂V ( r ) ⎛ 2z 2y ⎞ ⎛ ∂ ∂ ⎞
= ⎜y −z ⎟Ψ + V ( r ) ⎜ y − z ⎟Ψ
∂r ⎝ r r ⎠ ⎝ ∂z ∂y ⎠
we obtain
⎡ 3 ⎤
⎢c ∑ α i p , Lx ⎥Ψ = c ⎡⎣α 2 p + α 3 p , Lx ⎤⎦Ψ
i ( 2) (3 )
⎣ i =1 ⎦
⎡ ∂ ∂ ∂ ∂ ⎤
= −c = 2 ⎢α 2 +α3 ,y − z ⎥Ψ
⎣ ∂y ∂z ∂z ∂y ⎦
(3.3.9)
⎡∂ ∂⎤ ⎡∂ ∂ ⎤
= −c = 2 α 2 ⎢ ,y ⎥Ψ + c =2 α 3 ⎢ ,z ⎥Ψ
⎣ ∂y ∂z ⎦ ⎣ ∂z ∂y ⎦
⎛ ∂ ∂ ⎞
= −c = 2 ⎜ α 2
⎝ ∂z
−α3 ( 2)
( (3)
⎟Ψ = ci= α 3 p − α 2 p Ψ .
∂y ⎠
)
This expression represents [H , Lx ] . Cyclic permutations in (3.3.9) generate the
analogous expressions for ⎡⎣H , Ly ⎤⎦ and [H , Lz ] .
i.e. the last commutator in (3.3.10) vanishes. We write the preceding commutator
K
with the x-component of S :
⎡3 ⎤
( ) (
c ⎢ ∑ α i p i , S x ⎥ = cp ( ) α 2 S x − S x α 2 + cp ( ) α 3 S x − S x α 3
⎣ i =1 ⎦
2 3
)
= ( 2) ⎛⎜ ⎛ 0 σ y ⎞ ⎛σ 0 ⎞ ⎛σ x 0 ⎞⎛ 0 σ y ⎞⎞
=c p ⎜ ⎟⎜ x ⎟−⎜ ⎟⎜ ⎟⎟
2 ⎜ ⎜σ 0 ⎟ ⎝⎜ 0 σ x ⎠⎟ ⎝⎜ 0 σ x ⎠⎟ ⎜ σ y 0 ⎟⎟
⎝⎝ y ⎠ ⎝ ⎠⎠
= (3) ⎛ ⎛ 0 σ z ⎞ ⎛ σ x 0 ⎞ ⎛σ x 0 ⎞⎛ 0 σ z ⎞⎞
+c p ⎜⎜ ⎟⎜ ⎟−⎜ ⎟⎜ ⎟⎟
2 ⎜ ⎜σ 0 ⎟⎠ ⎜⎝ 0 σ x ⎟⎠ ⎜⎝ 0 σ x ⎟⎠ ⎜⎝ σ z 0 ⎟⎠ ⎟⎠
⎝⎝ z
=⎛ 2 ⎛ 0 σ yσ x − σ xσ y ⎞ ⎛ 0 σ zσ x − σ xσ z ⎞ ⎞
= c ⎜ p( ) ⎜ ⎟ + p(3) ⎜ ⎟⎟
2⎜ ⎜σ σ − σ σ ⎟ ⎜σ σ − σ σ ⎟
⎝ ⎝ y x x y
0
⎠ ⎝ z x x z
0 ⎠ ⎟⎠
=⎛ ⎛ 0 2iσ z ⎞ ⎛ 0 2iσ y ⎞ ⎞
= c ⎜ − p( ) ⎜ (3)
⎟+ p ⎜ ⎟ ⎟ i.e.
2
⎜ 0 ⎠⎟ ⎜ 2iσ 0 ⎟⎟
2⎜
⎝ ⎝ 2iσ z ⎝ y ⎠⎠
⎡3 ⎤
(
c ⎢ ∑ α i p i ,Sx ⎥ = −ci= p ( 2) α 3 − p ( 3) α 2 ,
⎣ i =1 ⎦
) (3.3.12)
where (2.1.15) and (2.1.14) have been used. Inserting (3.3.8), (3.3.9), (3.3.11) and
(3.3.12) in (3.3.6) yields
46
[H , J x ] = 0 (3.3.13)
⎛ ∂ ∂ ⎞ ⎛ ∂ ∂ ⎞
Since P0 ⎜ y − z ⎟ = ⎜y − z ⎟ P0 , the first commutator on the right hand
⎝ ∂z ∂y ⎠ ⎝ ∂z ∂y ⎠
= ⎛ 1 0 ⎞ ⎛σ x 0 ⎞
side of (3.3.15) vanishes. Because β Sx = ⎜ ⎟⎜ ⎟ = Sx β , the last term
2 ⎝ 0 −1⎠ ⎜⎝ 0 σ x ⎟⎠
in (3.3.15) disappears also and we have
K
[P , J x ] = 0 and ⎣⎡P , J ⎦⎤ = 0 . (3.3.16)
K
Due to (3.3.3), (3.3.14) and (3.3.16) the operators H , P and J commute mutually.
Therefore, there are energy eigenfunctions, which are simultaneously
eigenfunctions of the parity operator and of the total angular momentum operator.
This property will be used in the next section.
3.4 Separation of the variables for the Dirac equation with central potential
We are looking for the solution of the Dirac equation with the Hamiltonian (3.3.1)
⎛ 3 ⎞ K K
⎜ c ∑ α i p + β m0c + V ( r ) ⎟Ψ ( r ) = EΨ ( r ) .
i 2
(3.4.1)
⎝ i =1 ⎠
K
As in (3.2.4) or in (2.4.4) the four-spinor Ψ ( r ) is split up in two two-spinors like
this
K
K ⎛ ϕ (r ) ⎞
Ψ ( r ) = ⎜⎜ K ⎟⎟ . (3.4.2)
⎝ χ (r ) ⎠
K
In the preceding section is shown that there are energy eigenfunction Ψ ( r ) which
are also eigenfunctions of the parity operator P , i. e.
K K K K
K ⎛ ϕ ( r ) ⎞ ⎛ 1 0 ⎞ ⎛ ϕ ( −r ) ⎞ ⎛ ϕ ( −r ) ⎞ ⎛ ϕ (r ) ⎞
PΨ ( r ) = β P0 ⎜⎜ K ⎟⎟ = ⎜ ⎟ ⎜⎜ K ⎟⎟ = ⎜⎜ K ⎟⎟ = λ ⎜⎜ K ⎟⎟ (3.4.3)
⎝ χ ( r ) ⎠ ⎝ 0 −1⎠ ⎝ χ ( − r ) ⎠ ⎝ − χ ( − r ) ⎠ ⎝ χ (r ) ⎠
The eigenvalue λ is +1 or -1. From (3.4.3) follows that
47
K K K K
for λ = +1 : ϕ ( −r ) = ϕ ( r ) , − χ ( − r ) = χ ( r )
K K K K (3.4.4)
and for λ = −1 : ϕ ( −r ) = −ϕ ( r ) , − χ ( −r ) = − χ ( r )
K
i.e. the internal parity of ϕ ( r ) corresponds to λ (it is positive if λ = +1 etc.). The
K K
spinors ϕ ( r ) and χ ( r ) have opposite internal parities.
K
On the other hand, (3.3.14) reveals that Ψ ( r ) is an eigenfunction of J z and J 2 . As
in quantum mechanics the corresponding eigenvalues read =m and = 2 j ( j + 1) .
K
This is true for each of the four components of Ψ ( r ) .
In atomic and nuclear physics such functions are formed by coupling spherical
K
harmonics Yl ,ml ( r / r ) with spin eigenfunctions xs ,ms . As we have seen in section
3.1, Dirac particles have spin s = ½. The coupling is performed by means of
Clebsch Gordan coefficients ( l ml s ms | jm ) resulting in the function
K K
Ω jml ( r /r ) = ∑ ( l m ,s = 1/ 2,m
ml ms
l s | j m )Yl ml ( r /r ) x1/ 2 ms . (3.4.5)
⎛ 1⎞ ⎛0⎞
x 1 / 2,1 / 2 = ⎜ ⎟ , x 1 / 2 ,−1 / 2 = ⎜ ⎟ . (3.4.6)
⎝0⎠ ⎝ 1⎠
j +m
(l = j − 1 / 2,ml = m − 1 / 2,1 / 2,ms = 1 / 2 | jm ) =
2j
j −m
(l = j − 1 / 2,ml = m + 1 / 2,1 / 2,ms = −1 / 2 | jm ) =
2j
(3.4.7)
j − m +1
( l = j + 1 / 2,ml = m − 1 / 2,1 / 2,ms = 1 / 2 | jm ) = −
2j + 2
j + m +1
(l = j + 1 / 2,ml = m + 1 / 2,1 / 2,ms = −1 / 2 | jm ) = .
2j + 2
K
K 1 ⎛⎜ j + m Yl = j −1 / 2,ml =m −1 / 2 ( r / r ) ⎞⎟
Ω jm,l = j −1 / 2 ( r / r ) = K , (3.4.8)
2 j ⎜ j − m Yl = j −1 / 2 ,m =m +1 / 2 ( r / r ) ⎟
⎝ l ⎠
K
K 1 ⎛⎜ − j − m + 1Yl = j +1 / 2,ml =m −1 / 2 ( r / r ) ⎞⎟
Ω jm,l = j +1 / 2 ( r / r ) = K . (3.4.9)
2 j + 2 ⎜ j + m + 1Yl = j +1 / 2 ,m =m +1 / 2 ( r / r ) ⎟
⎝ l ⎠
K
The internal parity of Ylml ( r / r ) depends on l like this
K K K
P0 Ylml ( r / r ) = Ylml ( −r / r ) = ( −1) Ylml ( r / r ) .
l
(3.4.10)
Since the l-values of the expressions in (3.4.8) and (3.4.9) differ by 1, these
functions have opposite internal parity. Therefore they are able to represent the
K K K
r / r -dependence of ϕ ( r ) and χ ( r ) set out in (3.4.4). Consequently, we assign
K
the ordinal numbers j ,m,l = j − 1 / 2 to ϕ ( r ) and the numbers j ,m,l ′ = j + 1 / 2 to
K
χ ( r ) . An inverse assignment is also possible, but it does not bring new results as
we shall see at the end of this section.
We make the following ansatz
K K
ϕ ( r ) = ig ( r ) Ω jm,l = j −1 / 2 ( r / r )
K K . (3.4.11)
χ ( r ) = −f ( r ) Ω jm,l ′= j +1 / 2 ( r / r )
3 ⎛ 0 σ i pi ⎞ ⎛ ϕ ⎞ ⎛ 1 0 ⎞ ⎛ϕ ⎞ ⎛ϕ ⎞ ⎛ϕ ⎞
c∑ ⎜ ⎟⎜ ⎟ + ⎜ ⎟ m0c 2 ⎜ ⎟ + V ( r ) ⎜ ⎟ = E ⎜ ⎟ , (3.4.12)
⎜
i =1 σ p
⎝ i
i
0 ⎠⎟ ⎝ χ ⎠ ⎝ 0 −1⎠ ⎝χ⎠ ⎝χ⎠ ⎝χ⎠
3
c ∑ σ i p i ϕ − m0c 2 χ + V χ = E χ . (3.4.14)
i =1
K K
c ∑ σ i p i ϕ ( r ) = ( E + m0c 2 − V ( r ) ) χ ( r )
3
(3.4.15)
i =1
makes clear that the expression on the left hand side of (3.4.15) is eigenfunction of
K
J 2 and J z ( J commutes with V ( r ) as shown in section 3.3) with the same value
j as χ , furthermore it has the same parity and s-value as χ . Because these
K
properties concern the angle- and spin dependent part of χ ( r ) , i.e.
K
Ω jm,l ′= j +1 / 2 ( r / r ) , the following proportionality must hold
3
K
The angle- and spin dependent part of c ∑ σ i p i ϕ ( r )
i =1 (3.4.16)
K
is proportional to Ω jm,l ′= j +1 / 2 ( r / r ) .
∂ dg ( r ) ∂r dg ( r ) ∂ (x ) + (x ) + (x )
1 2 2 2 3 2
dg ( r ) x i
g ( r ) = = =
∂x i dr ∂x i dr ∂x i dr r
we obtain
3
K
∑σ p ϕ (r )
i =1
i
i
(3.4.17)
dg ( r ) 3 xi K 3
δ K
==
dr i =1
∑ σ i Ω jm,l = j −1 / 2 ( r / r ) + =g ( r ) ∑ σ i i Ω jm,l = j −1 / 2 ( r / r ) .
r i =1 δx
The first sum on the right hand side of (3.4.17) can be written like this with the help
of (2.1.12) and (3.4.8)
⎛ j +m K ⎞
⎜ Yl = j −1 / 2 ,ml =m −1 / 2 ( r / r ) ⎟
3
xi K 1 ⎛ x3 x1 -ix 2 ⎞ ⎜ 2j ⎟
∑σ i Ω jm,l = j −1 / 2 ( r / r ) = ⎜
r ⎝ x1 + ix 2 −x3 ⎠ ⎜
⎟ ⎟
i =1 r j −m K
⎜
⎜ Yl = j −1 / 2,ml =m +1 / 2 ( r / r ) ⎟⎟
⎝ 2j ⎠
⎛ x3 j + m K ⎛ x1 x 2 ⎞ j − m K ⎞
⎜ Yl = j −1 / 2 ,ml = m −1 / 2 ( r / r ) + ⎜ −i ⎟ Yl = j −1 / 2 ,ml =m +1 / 2 ( r / r ) ⎟
r 2j ⎝ r r ⎠ 2j
= ⎜⎜ ⎟.
⎟
⎜ ⎛⎜ x +i x ⎞⎟ j + mYl = j −1 / 2,m =m −1 / 2 ( rK / r ) − x j −m K
1 2 3
⎜ r Yl = j −1 / 2,ml = m +1 / 2 ( r / r ) ⎟⎟
⎝⎝ r ⎠ 2 j r 2 j ⎠
l
(3.4.18)
50
⎛ j − m +1 K ⎞
⎜− Yl = j +1 / 2,ml =m −1 / 2 ( r / r ) ⎟
2j + 2
− ⎜⎜ ⎟.
⎟ (3.4.19)
⎜ j + m + 1 Yl = j +1 / 2,m =m +1 / 2 ( r / r ) ⎟
K
⎜ 2j + 2 ⎟
⎝ ⎠
l
We now investigate the last sum in (3.4.17). First, we will show that (3.4.20) holds
3
xk
also with exchanged l-values. For that, we apply the operator ∑ σ k on both
k =1 r
sides of (3.4.20) and we obtain
3
xk 3 xi K 3
xk K
∑σ
k =1
k ∑
r i =1
σ i Ω jm,l = j −1 / 2 ( r / r ) = −∑ σ k Ω jm,l = j +1 / 2 ( r / r )
r k =1 r
(3.4.21)
3
xi xk 3
( x ) =1 k 2
∑
i ,k =1
σ i
σ
r k r
= 1∑k =1 r2
(3.4.22)
and consequently
K 3
xk K
Ω jm,l = j −1 / 2 ( r / r ) = −∑ σ k Ω jm,l ′= j +1 / 2 ( r / r ) . (3.4.23)
k =1 r
3
K 3 3
xk K
∑
i =1
σ i p Ω jm,l = j −1 / 2 ( r / r ) = −∑ σ i p ∑ σ k Ω jm,l ′= j +1 / 2 ( r / r ) .
i
i =1
i
k =1 r
(3.4.24)
Making use of (2.1.13) and (2.1.14) we deal with the general double sum which
K K
contains arbitrary vectors A and B :
51
∑σ
i ,k =1
i
Ai σ k Bk
( )
3
= ∑ σ i Ai Bi + σ 1σ 2 A1B2 + σ 2 σ 1A2B1 + cyclic permutations
2
i =1
3
(3.4.25)
= ∑ Ai Bi 1 + iσ 3 ( A1B2 − A2B1 ) + cyclic permutations
i =1
K K
( )
3 3
= ∑ Ai Bi 1 + i∑ σ j A × B
j
i =1 j=1
in analogy with (2.4.12). Using (3.4.25) and (1.2.7) we transform the expression
(3.4.24):
3
K
∑σ p Ω
i =1
i
i
jm,l = j −1 / 2 (r / r)
K
⎛ 3 i xi 3
⎛K r⎞ ⎞ K
= −⎜∑ p + i∑ σ j ⎜ p × ⎟ ⎟ Ω jm,l ′= j +1 / 2 ( r / r ) (3.4.26)
⎜ i =1 r ⎝ ⎟
r ⎠j ⎠
⎝ j=1
3
δ xi K 1 3 K
= i=∑ i Ω jm,l ′= j +1 / 2 ( r / r ) + i ∑ σ i Li Ω jm,l ′= j +1 / 2 ( r / r ) ,
i =1 δ x r r i =1
K K K K K
where L = r × p = − p × r , as per (3.3.5), has been inserted. We continue
transforming (3.4.26):
3
K i 3
K
∑ σ i pi Ω jm,l = j −1 / 2 ( r / r ) =
i =1 r
∑σ L Ω
i =1
i i jm,l ′= j +1 / 2 (r / r)
3 K 3
x i 2x i K 3
xi δ K
+i= Ω jm,l ′= j +1 / 2 ( r / r ) + i=∑ Ω jm,l ′= j +1 / 2 ( r / r ) + i= ∑ Ω jm,l ′= j +1 / 2 ( r / r )
i =1 −2r i =1 r δ x
3 i
r
xi ∂ 3
The last term vanishes by the following reason: the sum ∑
i =1 r ∂x
i
is the scalar
K K
product of the unity position vector r / r and the gradientK ∇ . Therefore, it is the
projection of the gradient vector onto the direction of r and it results in the
δ
component of the gradient. Because Ω jml ′ ( r ′ / r ) depends only on directions
δr
∂ K
and not on r, there results Ω jml ′ ( r / r ) = 0 . Finally we have
∂r
3
K i⎡ 3
⎤ K
∑ σ i pi Ω jm,l = j −1 / 2 ( r / r ) = r ⎢2= + ∑ σ i Li ⎥ Ω jm,l ′= j +1/ 2 ( r / r ) .
i =1 ⎣ i =1 ⎦
(3.4.27)
In order to transform the last summand in (3.4.27) we use (3.3.4) with the two-
spinor operator (2.4.16) and form
K 2 ⎛ K = K ⎞2 K2 ⎛ = K ⎞2 K K
J = ⎜ L + σ ⎟ = L + ⎜ σ ⎟ + =σ ⋅ L . (3.4.28)
⎝ 2 ⎠ ⎝2 ⎠
We obtain
52
K K 3 ⎛ K 2 K2 ⎛ = K ⎞2 ⎞
=σ ⋅ LΩ jml ′ ≡ =∑ σ i Li Ω jml ′ = ⎜ J − L − ⎜ σ ⎟ ⎟ Ω jml ′
⎜ ⎝ 2 ⎠ ⎟⎠
i =1 ⎝ (3.4.29)
⎡ 3⎤ ⎛ 3⎞
= ⎢ j ( j + 1) − l ′ ( l ′ + 1) − ⎥ =2 Ω jml ′ = ⎜ − j − ⎟ =2 Ω jml ′ ,
⎣ 4⎦ ⎝ 2⎠
where we have set the eigenvalues mentioned in context with (3.4.5) and the
relation l ′ = j + 1 / 2 (see (3.4.11)). Inserting (3.4.29) in (3.4.27) yields
3
i= ⎛ 3⎞ i= ⎛ 1⎞
∑σ p Ω
i =1
i
i
jm,l = j −1 / 2 = ⎜ 2 − j − ⎟ Ω jm,l ′= j +1 / 2 = − ⎜ j − ⎟ Ω jm,l ′= j +1 / 2 .
r ⎝ 2⎠ r ⎝ 2⎠
(3.4.30)
(3.4.31)
dg ( r ) K g (r ) K
= −= Ω jm,l ′=+1 / 2 ( r / r ) + = ( j − 1 / 2 ) Ω jm,l ′=+1 / 2 ( r / r ) ,
dr r
which confirms the hypothesis (3.4.16). Analogously one derives
3
K
∑σ p χ (r )
I =1
i
i
(3.4.32)
df ( r ) K f (r ) ⎛ 3⎞ K
= −i= Ω jm,l = j −1 / 2 ( r / r ) + i= ⎜ − j − ⎟ Ω jm,l = j −1 / 2 ( r / r ) .
dr r ⎝ 2⎠
Now, the expressions (3.4.11) and (3.4.32) are inserted in (3.4.13) and (3.4.31)
into (3.4.14). The angular spinors Ω can be omitted. We obtain the following
differential equations for the radial functions f and g
df ( r ) f (r ) ⎛ 3⎞
−i=c − i=c ⎜ j + ⎟ = i ( E − m0c − V ( r ) ) g ( r )
2
dr r ⎝ 2⎠
(3.4.33)
dg ( r ) g (r ) ⎛ 1⎞
− =c +=c ⎜ j − ⎟ = − ( E + m0c − V ( r ) ) f ( r ) .
2
dr r ⎝ 2⎠
dF ( r ) F (r )
=c − =cκ + ( E − m0c 2 − V ( r ) ) G ( r ) = 0
dr r
(3.4.35)
dG ( r ) G (r )
=c + =cκ − ( E + m0c 2 − V ( r ) ) F ( r ) = 0 ,
dr r
53
If one makes an inverse choice for l and l ′ (see (3.4.11) and the preceding
paragraph) i.e. if one takes l = j + 1 / 2 and l ′ = j − 1 / 2 , as one can show, in
(3.4.35) only κ changes sign. In the next section we will see that the energy of the
bound states depends on κ and therefore it is independent of the relative choice
of l and l ′ . We stress
κ = j + 1/ 2 . (3.4.36)
3.5 Solution of the radial equations for a Dirac particle in a Coulomb field
We solve the coupled radial equations for a Dirac particle in a Coulomb potential
and determine the energy values for the bound states. The potential is given by a
point nucleus of charge Ze and a particle of charge −e and reads
Ze 2 Ze 2 As
V (r ) = − ≡− with ε 0 = 8.8542 ⋅ 10 −12
4πε 0 r r Vm
. (3.5.1)
e2
and e ≡ 2
4πε 0
dG ( r ) G ( r ) ⎛ m0c 2 + E Ze 2 ⎞
= −κ +⎜ + ⎟ F (r )
dr r ⎝ =c =cr ⎠
(3.5.2)
dF ( r ) F ( r ) ⎛ m0c 2 − E Ze 2 ⎞
=κ +⎜ − ⎟G (r ) .
dr r ⎝ =c =cr ⎠
dG ( r ) G (r ) ⎛ Zα ⎞
= −κ + ⎜ α1 + ⎟ F (r )
dr r ⎝ r ⎠
(3.5.4)
dF ( r ) F (r ) ⎛ Zα ⎞
=κ + ⎜α2 − ⎟G (r ) .
dr r ⎝ r ⎠
d d
ρ = α1α 2 r with = α1α 2 , (3.5.5)
dr dρ
and define G ( ρ ) ≡ G ( r ( ρ ) ) , F ( ρ ) ≡ F ( r ( ρ ) ) ,
54
dG ( ρ ) G(ρ) ⎛ Zα ⎞
α1α 2 = −κ α1α 2 + ⎜ α1 + α1α 2 F (ρ )
ρ ⎟⎠
i.e.
dρ ρ ⎝
dF ( ρ ) F (ρ) ⎛ Zα ⎞
α1α 2 = κ α1α 2 + ⎜ α 2 − α1α 2 G(ρ )
dρ ρ ⎝ ρ ⎟⎠
dG ( ρ ) G(ρ) ⎛ α1 Zα ⎞
or = −κ + ⎜⎜ + ⎟F (ρ )
dρ ρ ⎝ α2 ρ ⎟⎠
(3.5.6)
dF ( ρ ) F (ρ) ⎛ α 2 Zα ⎞
=κ + ⎜⎜ − ⎟G ( ρ ) .
dρ ρ ⎝ α1 ρ ⎟⎠
F ( ρ ) = F ( ρ ) e − ρ and G ( ρ ) = G ( ρ ) e − ρ (3.5.7)
we have
dG ( ρ ) G ( ρ ) ⎛ α1 Zα ⎞
= G(ρ) −κ + ⎜⎜ + ⎟F (ρ ) , (3.5.8)
dρ ρ ⎝ α2 ρ ⎟⎠
dF ( ρ ) F ( ρ ) ⎛ α 2 Zα ⎞
= F (ρ) + κ + ⎜⎜ − ⎟G ( ρ ) . (3.5.9)
dρ ρ ⎝ α1 ρ ⎟⎠
Inserting them in (3.5.8) and writing only the summands with the factor ρ s + μ −1 ,
where μ is an arbitrarily chosen m-value, we obtain
" + ( s + μ ) bμ ρ s + μ −1 + "
α1
= " + bμ −1ρ s + μ −1 " + " − κ bμ ρ s + μ −1 + " + a ρ s + μ −1 + " + Zα aμ ρ s + μ −1
α 2 μ −1
α1
−Zα aμ − a + ( s + μ + κ ) bμ − bμ −1 = 0 . (3.5.11)
α 2 μ −1
α2
( s + μ − κ ) aμ − aμ −1 + Zα bμ − b = 0. (3.5.12)
α1 μ −1
−Zα a0 + ( s + κ ) b0 = 0,
( s − κ ) a0 + Zα b0 = 0.
To obtain a non-trivial solution the determinant ( Zα ) + s 2 − κ 2 must vanish, i.e.
2
s = ± κ 2 − ( Zα ) .
2
(3.5.13)
K
⎛ ϕ (r ) ⎞
We investigate the sign of s. The wave function ⎜ K ⎟ must be normalizable i.e.
⎝ χ (r ) ⎠
(∫ ϕ ( rK )
V
2 K 2
)
+ χ ( r ) dV = finite, even in a volume near r = 0,
a ⎛ ⎛ G ( r ) ⎞2 ⎛ F ( r ) ⎞2 ⎞
∫0 ⎜⎜ − ⎜⎝ r ⎟⎠ + ⎜⎝ r ⎟⎠ ⎟⎟ r dr = finite, or
2
⎝ ⎠
a α1α 2
1
i.e. 2s + 1 ≥ 0 or s ≥ − .
2
1
It can be shown (Greiner, 1990, p.179) that the range of values 0 ≥ s ≥ − is also
2
excluded. Therefore,
s = + κ 2 − ( Zα )
2
(3.5.14)
⎡( s + μ + κ ) α 2 − Zα α1 ⎤ bμ = ⎡( s + μ − κ ) α1 + Zα α 2 ⎤ aμ . (3.5.15)
⎣ ⎦ ⎣ ⎦
α 2 bμ ≅ α1aμ . (3.5.16)
α1 α1 α1
μ aμ − Zα aμ ≅ aμ −1 + a ,
α2 α2 α 2 μ −1
α1
which yields for μ Zα the relations
α2
56
2 2
aμ ≅ aμ −1 and b μ ≅ bμ −1 . (3.5.17)
μ μ
Due to (3.5.10) and (3.5.17) the function F ( ρ ) can be written for large m's like this
⎛ 2 4 ⎞
F ( ρ ) ≅ " + ρ s ⎜⎜ am ρ m + am ρ m +1 + am ρ m + 2 + " ⎟⎟ ,
⎝ m +1 ( m + 1)( m + 2 ) ⎠
2ρ
⎛ 2 ρ ( 2 ρ )2 ( 2ρ )
n
⎞
ρ a0e
s
= ρ a0 ⎜ 1 +
s
+ +"+ + "⎟ .
⎜ 1 1⋅ 2 n! ⎟
⎝ ⎠
F (r ) F ( ρ ) F ( ρ ) − ρ
f (r ) = = α1α 2 = e α1α 2 (3.5.18)
r ρ ρ
is like ρ s −1e ρ , which is not normalizable. Therefore, the series (3.5.10) must stop
for a certain value m = ν , i.e.
aν +1 = bν +1 = 0 . (3.5.19)
α1
−Zα aν +1 − a + ( s + ν + 1 + κ ) bν +1 − bν = 0 .
α2 ν
α1
− a = bν for ν = 0,1, 2," , (3.5.20)
α2 ν
α1
− ⎡⎣( s + ν + κ ) α 2 − Zα α1 ⎤⎦ a = ⎡( s + ν − κ ) α1 + Zα α 2 ⎤⎦ aν , (3.5.21)
α2 ν ⎣
2E
i.e. 2 α1α 2 ( s + ν ) = Zα (α1 − α 2 ) = Zα , (3.5.22)
=c
where (3.5.3) is applied. The equation (3.5.22) is squared and reads
(m c 2
0
4
− E 2 ) ( s + ν ) = E 2Z 2α 2 or
2
1 1 ⎛ Z 2α 2 ⎞ (3.5.23)
= ⎜ 1 + ⎟.
E2 m02c 4 ⎜⎝ ( s + ν )2 ⎟⎠
57
n ≡ ν + κ ≡ ν + j + 1 / 2, n = 1, 2," . (3.5.24)
⎢ ( Zα ) ⎥
2
E = m0c ⎢1 +
2
2⎥
. (3.5.25)
⎢
⎢⎣ (
n − κ + κ 2 − ( Zα )
2
) ⎥
⎥⎦
e 2 1
Because of α ≡ ≅ (cf.(3.5.3)) for small Z's we have
=c 137
( Zα )
2
1 (3.5.26)
x 3x 2 x
(1 + x ) , (1 − x ) ≅ 1 + 2 x and (1 − x )
−1 / 2 −2 1/ 2
≅ 1− + ≅ 1−
2 8 2 (3.5.27)
for x 1
we write (3.5.25) this way
58
1
−
⎡ ⎤ 2
⎢ ⎥
⎢ ( Zα )
2 ⎥
E ≅ m0c 2 ⎢1 + ⎥
⎢ ⎛ Zα ⎞ ⎥
2
⎞
⎢ ⎜ n − κ + κ ⎜1− ( ) ⎟ ⎟ ⎥
⎛ 2
⎢ ⎜ ⎜ 2κ 2 ⎟⎠ ⎟ ⎥
⎣ ⎝ ⎝ ⎠ ⎦
⎡ ⎤
⎢ ⎥
⎢ ( α ) ⎥
2
α
4
Z 3 ⎛ Z ⎞
≅ m0c 2 ⎢1 − + ⎜ ⎟ ⎥
κ ⎛ Zα ⎞2 ⎞ 8 ⎝ n ⎠ ⎥
2
⎢ ⎛
⎢ 2 ⎜⎜ n − 2 ⎜ κ ⎟ ⎟⎟ ⎥
⎣⎢ ⎝ ⎝ ⎠ ⎠ ⎦⎥
⎡ ( Zα )2 ⎛ ( Zα )2 ⎞ 3 ⎛ Zα ⎞ 4 ⎤
≅ m0c ⎢1 − 2
⎜1+ ⎟+ ⎥
⎢⎣ 2n 2 ⎜⎝ κ n ⎟⎠ 8 ⎜⎝ n ⎟⎠ ⎥
⎦
⎡ ( Zα )2 ( Zα )4 ⎛ n 3 ⎞⎤
≅ m0c ⎢1 − 2
− ⎜ − ⎟⎥ ,
⎢⎣ 2n 2 2n 4 ⎝ j + 1 / 2 4 ⎠ ⎥⎦ (3.5.28)
⎡ ( Zα )2 ( Zα )4 ⎛ n 3 ⎞⎤
Eion. = m0c − E ≅ m0c ⎢
2
+ 2
⎜ − ⎟⎥ . (3.5.29)
⎢⎣ 2n
2
2n 4 ⎝ j + 1 / 2 4 ⎠ ⎥⎦
( Zα )
2
Z 2e 4 Z 2e 4
m0c 2
= m0 = m
2=2n 2 ( 4πε 0 )
0
2n 2 2= 2 n 2 2
(3.5.30)
(using (3.5.3) and (3.5.1))
1 1/ 2 13.60583 13.60601
2 1/ 2 3.40146 3.40151
2 3/ 2 " 3.40147
3 1/ 2 1.51176 1.551178
3 3/ 2 " 1.551177
3 5/ 2 " 1.551176
The total energy splitting of fine structure levels, ΔEfine , for a given n is determined
by the extreme values of j +1/2. The minimal value is ½ + ½ = 1 and the maximum
is given by n = ν + j + 1 / 2 (see (3.5.24)). I.e. n = j + 1 / 2 holds for ν = 0 .
Therefore, we have
( Zα )
4
ΔEfine ( Zα )
2
≅ ( n − 1) .
Eion. n2
ΔEfine
With n = 2 and Z = 1 we have ≅ 1.33 ⋅ 10−5 in agreement with table 3.5.1.
Eion.
⎛ ϕ0 ⎞
⎜ 3 ⎟ KK
K N ⎜ ⎟ ei( pr −Et ) / = .
Ψ ( r ,t ) = ∑ σip i
(3.6.2)
⎜
V i =1 ⎟
⎜⎜ ϕ 0 ⎟⎟
⎝ p ⎠
⎛ 1⎞
⎜ ⎟
Ψ E ,p,m =
N ⎜ 0 ⎟ ei( pz −Et ) / = ,
0 ,+
V ⎜ 1⎟
⎜⎜ ⎟⎟
⎝0⎠
(3.6.3)
⎛0⎞
⎜ ⎟
Ψ E ,p,m =
N ⎜ 1 ⎟ ei( pz −Et ) / =
0 ,−
V ⎜0⎟
⎜⎜ ⎟⎟
⎝ −1⎠
⎡ ⎛ 3 ⎞ ⎤
⎜ ∑σ i p
i
⎢ 0 ⎟ ⎥
=
⎡⎣Sp , P ⎤⎦ = ⎢ ⎜ i =1 ⎟ , β P0 ⎥
⎢2 p ⎜ 3
⎟ ⎥
⎢ ⎜ 0
⎝
∑ σ i pi ⎟
⎠
⎥
⎣ i =1 ⎦
= ⎛ 3 ⎛σ i 0 ⎞⎛ 1 0 ⎞ i 3 ⎛1 0 ⎞ ⎛σ i 0⎞ ⎞
= ⎜ ∑⎜ ⎟⎜ ⎟ p P0 − ∑ ⎜ ⎟⎜ ⎟ P0 p ⎟
i
(3.6.4)
2 p ⎜⎝ i =1 ⎜⎝ 0 σ i ⎟⎠ ⎝ 0 −1⎠ i =1 ⎝ 0 −1⎠ ⎜ 0
⎝ σ i ⎟⎠ ⎟
⎠
= ⎛σ i 0 ⎞ i = ⎛σ i 0 ⎞ i
⎟ ( p P0 − P0 p ) =
3 3
=
2p
∑ ⎜
⎜
i =1 ⎝ 0
⎟
−σ i ⎠
i
2p
∑ ⎜⎜ 0 ⎟ p P ⋅2 ≠ 0
−σ i ⎟⎠
i =1 ⎝
n → p + e− + ν ,
Detailed experimental analyses have shown that the spin of the neutrino is
antiparallel to the momentum direction, i.e. it looks backward. Therefore, we can
describe it by means of Ψ E ,p,m0 ,− , (3.6.3), which is a helicity eigenfunction with
eigenvalue − = / 2 . In accordance with our predictions the neutrino has no definite
parity. Namely, measurements reveal that in weak interactions parity violation
occurs.
Analogously, experimentally, the spin of the antineutrio has been found to be
parallel to the momentum, i.e. the helicity is well defined, and one has detected
also, that the parity is not definite, as we expect.
The ability to describe the relation between definite helicity and undefined parity of
neutrinos is an additional highlight of the Dirac theory.
⎛ 1 ⎞
⎜ ⎟
⎜ 0 ⎟
K N
Ψ + ( r ,t,p ) = ⎜ cp ⎟ ei( pz −Et ) / =
V ⎜ ⎟
⎜ m0c + E ⎟
2
⎜ ⎟
⎝ 0 ⎠
(3.7.1)
⎛ 0 ⎞
⎜ ⎟
⎜ 1 ⎟
K N
Ψ − ( r ,t,p ) = ⎜ 0 ⎟ ei( pz −Et ) / = .
V ⎜ ⎟
⎜ −cp ⎟
⎜ m c2 + E ⎟
⎝ 0 ⎠
As shown at the end of section 3.1, they are eigenfunctions of the spin operators
Sz and S 2 with the eigenvalues ±=s and = 2s ( s + 1) respectively with s = ½. We
concentrate on the four-component spinors
⎛ 1 ⎞ ⎛ 0 ⎞
⎜ ⎟ ⎜ ⎟
⎜ 0 ⎟ ⎜ 1 ⎟
ω + ( p ) = ⎜ cp ⎟ , ω − ( p ) = ⎜ 0 ⎟ (3.7.2)
⎜ ⎟ ⎜ ⎟
⎜ m0c + E ⎟ ⎜ −cp ⎟
2
⎜ ⎟ ⎜ m c2 + E ⎟
⎝ 0 ⎠ ⎝ 0 ⎠
62
⎛ 1 ⎞ ⎛ 0 ⎞ ⎛ 1 ⎞
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ 0 ⎟ ⎜ 1 ⎟ ⎜ 0 ⎟
ω + ,α ( p ) ω − ,β ( p ) ω + ,γ ( p ) = ⎜ cp ⎟ ⎜ 0 ⎟ ⎜ cp ⎟ . (3.7.3)
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ m0c + E ⎟ ⎜ −cp ⎟ ⎜ m0c + E ⎟
2 2
⎜ ⎟ ⎜ m c2 + E ⎟ ⎜ ⎟
⎝ 0 ⎠α ⎝ 0 ⎠β ⎝ 0 ⎠γ
ω αβγ ( p,i = 0 ) = ω + ,α ( p ) ω + ,β ( p ) ω + ,γ ( p ) ,
ω αβγ ( p,i = 1) = ω + ,α ( p ) ω + ,β ( p ) ω − ,γ ( p ) + ω + ,α ( p ) ω − ,β ( p ) ω + ,γ ( p )
+ ω − ,α ( p ) ω + ,β ( p ) ω + ,γ ( p ) ,
(3.7.4)
ω αβγ ( p,i = 2 ) = ω + ,α ( p ) ω − ,β ( p ) ω − ,γ ( p ) + ω − ,α ( p ) ω + ,β ( p ) ω − ,γ ( p )
+ ω − ,α ( p ) ω − ,β ( p ) ω + ,γ ( p ) ,
ω αβγ ( p,i = 3 ) = ω − ,α ( p ) ω − ,β ( p ) ω − ,γ ( p ) .
Obviously, the index i gives the number of ω − ( p ) - spinors in every triple. For an
arbitrary spin s we have 2s + 1 expressions ω ( p,i ) in analogy with (3.7.4)
containing "products" with 2s four-spinors.
= ⎛= ⎞
Sz,α ,β ,γ ω αβγ ( i = 0 ) = ω + ,α ω + ,β ω + ,γ + ω + ,α ⎜ ω + ,β ⎟ ω + ,γ
2 ⎝2 ⎠
⎛= ⎞ 3
+ ω + ,α ω + ,β ⎜ ω + ,γ ⎟ = =ω αβγ ( i = 0 ) ,
⎝2 ⎠ 2
=
Sz,α ,β ,γ ω αβγ ( i = 1) = (ω + ,α ω + ,β ω − ,γ + ω + ,α ω + ,β ω − ,γ − ω + ,α ω + ,β ω − ,γ )
2
=
+ (ω + ,α ω − ,β ω + ,γ − ω + ,α ω − ,β ω + ,γ + ω + ,α ω − ,β ω + ,γ ) (3.7.6)
2
=
+ ( −ω − ,α ω + ,β ω + ,γ + ω − ,α ω + ,β ω + ,γ + ω − ,α ω + ,β ω + ,γ )
2
=
= ω αβγ ( i = 1) ,
2
=
Sz,α ,β ,γ ω αβγ ( i = 2 ) = − ω αβγ ( i = 2 ) ,
2
3
Sz,αβγ ω αβγ ( i = 3 ) = − =ω αβγ ( i = 3 ) ,
2
where the parameter p was omitted. Generally (with arbitrary spin s) one obtains,
instead of (3.7.6),
Now, we will show that the expressions ω α ,β ,",τ ( p,i ) are also eigenfunctions of S 2
with the well-known eigenvalues s ( s + 1) = 2 . We again restrict ourselves on s =3/2
and write explicitly
Sx2,α ,β ,γ = ( Sx ,α + Sx ,β + Sx ,γ )( Sx ,α + Sx ,β + Sx ,γ )
(3.7.9)
= Sx2,α + Sx2,β + Sx2,γ + 2Sx ,α Sx ,β + 2Sx ,α Sx ,γ + 2Sx ,β Sx ,γ .
⎛0 1 0 0⎞
⎜ ⎟
= ⎜1 0 0 0⎟ =2
Sx = and Sx =
2
⋅ 1 = Sy 2 = Sz 2 . (3.7.10)
2 ⎜0 0 0 1⎟ 4
⎜⎜ 0 ⎟
⎟
⎝ 0 1 0⎠
Sx ,α Sx ,β ω α ,β ,γ ( p,i = 0 )
⎛ 1 ⎞ ⎛ 1 ⎞ ⎛ 1 ⎞
⎛0 1 0 0⎞ ⎜ ⎟ ⎛0 1 0 0⎞ ⎜ ⎟ ⎜ ⎟
2 ⎜ ⎟ ⎜ ⎟ ⎜1 ⎟
0 0 0
= 1 0 0 0⎟ 0 0 0⎟ ⎜ ⎟ ⎜ ⎟
= ⎜ ⎜ cp ⎟ ⎜ ⎜ cp ⎟ ⎜ cp ⎟
4 ⎜0 0 0 1⎟ ⎜ ⎟ ⎜0 0 0 1⎟ ⎜ ⎟ ⎜ ⎟
⎟ ⎜ m0c + E ⎟ ⎜⎜ ⎟ ⎜ m0c + E ⎟ ⎜ m0c + E ⎟
2 2 2
⎜⎜ ⎟ ⎟
⎝0 0 1 0 ⎠α ⎜ ⎟ ⎝0 0 1 0 ⎠β ⎜ ⎟ ⎜ ⎟
⎝ 0 ⎠α ⎝ 0 ⎠β ⎝ 0 ⎠γ
⎛ 0 ⎞ ⎛ 0 ⎞ ⎛ 1 ⎞
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
2 ⎜
1 ⎟ ⎜ 1 ⎟ ⎜ 0 ⎟
=
= ⎜ 0 ⎟ ⎜ 0 ⎟ ⎜ cp ⎟ ,
4⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎟ ⎜ m0c + E ⎟
2
⎜ cp ⎟ ⎜ cp
⎜ m c2 + E ⎟ ⎜ m c2 + E ⎟ ⎜ ⎟
⎝ 0 ⎠α ⎝ 0 ⎠β ⎝ 0 ⎠γ
Sx2,α ,β ,γ ω α ,β ,γ ( p,i = 0 )
⎛ 0 ⎞ ⎛ 0 ⎞ ⎛ 1 ⎞
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ 1 ⎟ ⎜ 1 ⎟ ⎜ 0 ⎟
3 2 2
= = ω α ,β ,γ ( p,i = 0 ) + =2 ⎜ 0 ⎟ ⎜ 0 ⎟ ⎜ cp ⎟
4 4 ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎟ ⎜ m0c + E ⎟
2
⎜ cp ⎟ ⎜ cp
⎜ m c2 + E ⎟ ⎜ m c2 + E ⎟ ⎜ ⎟
⎝ 0 ⎠α ⎝ 0 ⎠β ⎝ 0 ⎠γ
⎛ 0 ⎞ ⎛ 1 ⎞ ⎛ 0 ⎞
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ 1 ⎟ ⎜ 0 ⎟ ⎜ 1 ⎟
2
+ =2 ⎜ 0 ⎟ ⎜ cp ⎟ ⎜ 0 ⎟
4 ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎟ ⎜ m0c + E ⎟ ⎜
2
⎜ cp cp ⎟
⎜ m c2 + E ⎟ ⎜ ⎟ ⎜ ⎟
⎠ β ⎝ m0c + E ⎠γ
2
⎝ 0 ⎠α ⎝ 0
⎛ 1 ⎞ ⎛ 0 ⎞ ⎛ 0 ⎞
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ 0 ⎟ ⎜ 1 ⎟ ⎜ 1 ⎟
2
+ =2 ⎜ cp ⎟ ⎜ 0 ⎟ ⎜ 0 ⎟
4 ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ m0c + E ⎟ ⎜
2
cp ⎟ ⎜ cp ⎟
⎜ ⎟ ⎜ m c2 + E ⎟ ⎜ m c2 + E ⎟
⎝ 0 ⎠α ⎝ 0 ⎠β ⎝ 0 ⎠γ
and analogously
65
Sy2,α ,β ,γ ω α ,β ,γ ( p,i = 0 )
⎛ 0 ⎞ ⎛ 0 ⎞ ⎛ 1 ⎞
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ 1 ⎟ ⎜ 1 ⎟ ⎜ 0 ⎟
3 2 2
= = ω α ,β ,γ ( p,i = 0 ) − =2 ⎜ 0 ⎟ ⎜ 0 ⎟ ⎜ cp ⎟
4 4 ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎟ ⎜ m0c + E ⎟
2
⎜ cp ⎟ ⎜ cp
⎜ m c2 + E ⎟ ⎜ m c2 + E ⎟ ⎜ ⎟
⎝ 0 ⎠α ⎝ 0 ⎠β ⎝ 0 ⎠γ
⎛ 0 ⎞ ⎛ 1 ⎞ ⎛ 0 ⎞
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ 1 ⎟ ⎜ 0 ⎟ ⎜ 1 ⎟
2
− =2 ⎜ 0 ⎟ ⎜ cp ⎟ ⎜ 0 ⎟
4 ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎟ ⎜ m0c + E ⎟ ⎜
2
⎜ cp cp ⎟
⎜ m c2 + E ⎟ ⎜ ⎟ ⎜ ⎟
⎠ β ⎝ m0c + E ⎠γ
2
⎝ 0 ⎠α ⎝ 0
⎛ 1 ⎞ ⎛ 0 ⎞ ⎛ 0 ⎞
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ 0 ⎟ ⎜ 1 ⎟ ⎜ 1 ⎟
2
− =2 ⎜ cp ⎟ ⎜ 0 ⎟ ⎜ 0 ⎟
4 ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ m0c + E ⎟ ⎜
2
cp ⎟ ⎜ cp ⎟
⎜ ⎟ ⎜ m c2 + E ⎟ ⎜ m c2 + E ⎟
⎝ 0 ⎠α ⎝ 0 ⎠β ⎝ 0 ⎠γ
and
Sz,2 α ,β ,γ ω α ,β ,γ ( p,i = 0 )
⎛ 1 ⎞ ⎛ 1 ⎞ ⎛ 1 ⎞
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ 0 ⎟ ⎜ 0 ⎟ ⎜ 0 ⎟
3 2 2 .
= = ω α ,β ,γ ( p,i = 0 ) + 3 =2 ⎜ cp ⎟ ⎜ cp ⎟ ⎜ cp ⎟
4 4 ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ m0c + E ⎟ ⎜ m0c + E ⎟ ⎜ m0c + E ⎟
2 2 2
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎝ 0 ⎠α ⎝ 0 ⎠β ⎝ 0 ⎠γ
Finally, we have
S 2α ,β ,γ ω α ,β ,γ ( p,i = 0 )
⎛3 2⎞ 3 5 (3.7.11)
= 3 ⎜ + ⎟ = 2 ω α ,β ,γ ( p,i = 0 ) = ⋅ =2 ω α ,β ,γ ( p,i = 0 ) .
⎝4 4⎠ 2 2
Equations (3.7.7) and (3.7.12) show that the symmetric multispinors ω α ,β ,",τ ( p,i )
belonging to a given spin s may be interpreted as spinor part of the wave functions
of a particle with spin s, where the z-component can assume 2s + 1 different
values.
The whole wave function of such a particle reads in analogy with (3.7.1)
K N
Ψ α ,β ,",τ ( r ,t,p,i ) = ω α ,β ,",τ ( p,i ) ei( pz-Et ) / = . (3.7.13)
V
We define a Dirac equation which can only act on the α -part of (3.7.13) according
to (3.1.1)
⎛ 3 ⎞ K
⎜ c ∑ α i p + β m0c − E ⎟ Ψ α ( r ,t,p ) = 0 ,
i 2
(3.7.14)
⎝ i =1 ⎠α
where Ψ α has the form of the function (3.7.1). No matter if Ψ α is of the type
Ψ + or Ψ − (cf. (3.7.1)) it is a solution of the Dirac equation (3.7.14). Therefore the
multispinor function (3.7.13) is also a solution of (3.7.14), i.e.
⎛ 3 ⎞ K
⎜ c ∑ α j p + β m0 c − E ⎟ Ψ α ,β ,",τ ( r ,t,p,i ) = 0 .
j 2
⎝ j =1 ⎠α
⎛ 3 ⎞ K
⎜ c ∑ α j p + β m0 c − E ⎟ Ψ α ,β ,",τ ( r ,t,p,i ) = 0 with ς = α , β ," ,τ . (3.7.15)
j 2
⎝ j =1 ⎠ς
These are named the Bargmann-Wigner equations. Of course, they hold also with
K
linear combinations of functions Ψ α ,β ,",τ ( r ,t,p,i ) with various i-values.
⎡ ∂ 3
⎛ ∂ i ⎞ 2⎤ K
⎢i= ∂t + e Φ + c ∑ ⎜ i=α i ∂x i − e A ⎟ − β m0c ⎥Ψ el ( r ,t ) . (3.8.1)
⎣ i =1 ⎝ ⎠ ⎦
K
The energy dependence of Ψ el ( r ,t ) is given by (3.1.2):
67
K K
Ψ el ( r ,t ) = ψ ( r ) e −iEt / = . (3.8.2)
By way of exception we choose the negative solution for the energy in (3.1.15)
E = − m0 2c 4 + c 2 p 2 = − E . (3.8.3)
⎡ ⎛ ∂ ⎞ 3
⎛ ∂ i ⎞ 2⎤ K
β
⎢ ⎜ −i= + e Φ ⎟ ∑ ⎜ −i= βα i * i − β e A ⎟ − m0c ⎥Ψ * el ,− ( r ,t ) = 0 (3.8.5)
+ c
⎣ ⎝ ∂t ⎠ i =1 ⎝ ∂x ⎠ ⎦
taking into account that Φ and Ai are real. According to (3.8.4) and (3.8.2) the
four-spinor Ψ * el ,− has the exponential energy dependence e −i E t / = . We multiply
−1
(3.8.5) from the left by a non-singular 4 × 4 -matrix C and insert C C = 1 in front of
Ψ * el ,− , which yields
⎡ −1 ⎛ ∂ ⎞ 3
⎛ −1 ∂ ⎞ ⎤
⎟ ∑ ⎜ −i=C βα i * C
−1
⎢C β C ⎜ − i= + e Φ + c − C β C e Ai ⎟ − m0c 2 ⎥
⎣ ⎝ δ∂t ⎠ i =1 ⎝ ∂x i
⎠ ⎦
K
⋅CΨ * el ,− ( r ,t ) = 0 .
(3.8.6)
On the other hand, analogously to (3.8.1), we write the Dirac equation for a
positron, which we multiply also with β from the left
⎡ ⎛ ∂ ⎞ 3
⎛ ∂ i ⎞ 2⎤ K
⎢ β ⎜ i= ∂t − e Φ ⎟ + c ∑ ⎜ i= βα i ∂x i + β e A ⎟ − m0c ⎥Ψ pos ,+ ( r ,t ) = 0 . (3.8.7)
⎣ ⎝ ⎠ i =1 ⎝ ⎠ ⎦
The wave function Ψ pos ,+ of the positron is taken for positive energy.
−1
Obviously C C = 1 is fulfilled. We check the relations in (3.8.8)
68
= i βα 2 β ( −i ) α 2 β = − β βα 2 α 2 β = − β ,
−1
Cβ C
= i βα 2 βα1 ( −i ) α 2 β = − βα 2 βα 2 α 1 β = α 1 β = − βα 1 ,
−1
C βα 1 * C
C βα 2 * C
−1
( )
= i βα 2 β −α 2 ( −i ) α 2 β = − βα 2 .
In this way the square brackets of (3.8.6) and (3.8.7) are identical and the
solutions must agree:
K K K
Ψ pos ,+ ( r ,t ) = CΨ * el ,− ( r ,t ) ≡ iβ α 2Ψ * el ,− ( r ,t ) . (3.8.10)
Hence, by means of the negative energy wave function of a particle the positive
energy function of the corresponding antiparticle can be formed in a simple way.
The transformed function iβ α 2Ψ * el ,− is named charge-conjugate function of the
electron.
69
∑(p )
3
E −c p ≡ E −c
2 2 2 2 2 i
= m02c 4 , (4.1.1)
i =1
⎛ ∂2 ∂2 ∂2 ∂ 2 m0 2c 2 ⎞ K
⎜ 2 2 − − − + ⎟Ψ ( r ,t ) = 0 . (4.1.2)
⎝ c ∂t ∂x ∂y ∂z = ⎠
2 2 2 2
Inserting (4.1.3) in (4.1.2) yields the relativistic relation (1.2.10) or (3.1.15), i.e.
E2
2
= p 2 + m02c 2 .
c
Treating the equation (4.1.2) in the same way as in the Dirac equation in (2.3.1) up
to (2.3.3) one obtains by subtracting
1 ⎛ ∂2 ∂2 ⎞ 1 ∂⎛ ∂ ∂ ⎞
⎜Ψ * Ψ −Ψ Ψ * ⎟ = 2 ⎜Ψ * Ψ −Ψ Ψ * ⎟ . (4.1.4)
c2 ∂t ∂t ⎠ c ∂t ⎝ ∂t ∂t ⎠
2 2
⎝
With arguments used in section 2.3 one expects that the expression
∂Ψ ∂Ψ *
Ψ* −Ψ (4.1.5)
∂t ∂t
is proportional to the probability density ρ . However, this interpretation is doubtful
because the expression (4.1.5) can become negative, which is not permissible for
a probability. This is one of the shortcomings of the Klein-Gordon equation.
On the other hand, if ρ is defined like in (2.3.4) as
K
ρ = Ψ ( r ,t ) ,
2
it can be shown, that it isn't relativisticaly invariant (Yndurain, 1996, p. 24), which is
also unsatisfactory.
70
Yndurain also criticizes that, being of second order in the time derivative, the
K K
values of the wave function Ψ ( r ,t0 ) at time t0 do not uniquely determine Ψ ( r ,t ) :
K
knowledge of ∂Ψ ( r ,t ) / ∂t is also necessary.
There arise problems, when electrostatic fields, e.g. the field of a point charge,
have to be built in the Klein-Gordon equation. The often used form
=cα ⎞
2
⎛ K K
⎜E + ⎟ Ψ ( r ,t ) = ( m0 c + c p )Ψ ( r ,t ) ,
2 4 2 2
⎝ r ⎠
However, F. Gross (Gross, 1993, p.92) stresses that (4.2.1) is not manifestly
covariant. Dealing with this equation many problems are encountered. Anyhow, it
is an active area of current research.
71
References
Bjorken, J.D., Drell, S.D. (1964), Relativistic Quantum Mechanics, McGraw-Hill,
Inc.
Greiner, W. (1990), Relativistic Quantum mechanics, Springer-Verlag, Berlin.
Gross, F. (1993), Relativistic Quantum Mechanics and Field Theory, John Wiley &
Sons, Inc, New York.
Pfeifer, W. (1998), An Introduction to the Interacting Boson Model of the Atomic
Nucleus, www.walterpfeifer.ch.
Pfeifer, W. (2003), The Lie Algebras su(N), An Introduction. Birkhäuser Verlag,
Basel.
Yndurain, F.J. (1996), Relativistic Quantum Mechanics and Introduction to Field
Theory, Springer-Verlag, Berlin.
72
Index
angular momentum 25 Klein-Gordon Schrödinger equation
angular momentum operator 26, 44 70
antineutrino 60 Lamb shift 59
antiparticle 32, 68 Lorentz convention 8
arbitrary spin 61 Lorentz force 7
Bargman-Wigner equations 66 Lorentz group 4
Bohr-formula 58 Lorentz transformation 3, 18
boost 5 Lorentz-covariance 15
bound states 53 magnetic field strength 7
central potential 43, 46 magnetic moment 24
charge conjugation 66 anomalous 27
charge density 7 normal 27
circular path of the electron 43 massless Dirac particles 59
Clebsch Gordan coefficients 47 minimal substitution 15
conservation of flow of a fluid 20 momentum operator 6, 30
contravariant 4 multispinor 62
Coulomb potential 53 negative energy 66
covariant 70 negative energy state 32
current density 19 neutrino 60
cyclotron 38 nonrelativistic limit of the Dirac
density 16 equation 21
dielectricity 8 normalization factor 31
Dirac particles 34 parity operator 43, 60
Dirac sea 32 internal 43
electric current density 8 parity violation 61
electric field strength 7 Pauli equation 24
electromagnetic field 19, 66 Pauli realization 13
electromagnetic potentials 15 Pauli spin matrices 13
energy operator 6, 30 Pauli term 21, 24
energy splitting 59 Pauli's theorem 16
fine structure 57 permeability 8
flow conservation equation 20 plane wave 30
form invariant 19 point nucleus 53
free particle 7, 30 positron 66
Gauss system 7 principal quantum number 57
Greiner 29, 55 principle of minimal coupling 15
Gross 70 probability density 14, 19, 69
gyromagnetic ratio 24 probability density current 20
Hamiltonian 12 rest mass 7
helicity 30, 60 rotation of the coordinate 4
helicity operator 34 scalar potential 7
Hermite polynomial 40 Schrödinger equation 11, 26
Hermitian 11, 12, 29 Schrödinger-formula 58
homogeneous magnetic field 24, 38 secular equation 31
hyperfine structure 59 Sommerfeld structure formula 58
inertial system 15 special relativity 11
internal parity 47, 48 spherical harmonics 47, 50
Klein-Gordon equation 69 spherical potential 44
1