CHP 10
CHP 10
In this Chapter we will address the issue that the laws of physics must be formulated in a form
which is Lorentz–invariant, i.e., the description should not allow one to differentiate between frames
of reference which are moving relative to each other with a constant uniform velocity ~v . The
transformations beween such frames according to the Theory of Special Relativity are described
by Lorentz transformations. In case that ~v is oriented along the x1 –axis, i.e., ~v = v1 x̂1 , these
transformations are
x1 − v1 t 0 t − vc21 x1 0 0
x1 0 = q , t = q , x 2 = x2 ; x 3 = x3 (10.1)
v1 2 v1 2
1 − c 1 − c
which connect space time coordinates (x1 , x2 , x3 , t) in one frame with space time coordinates
(x01 , x02 , x03 , t0 ) in another frame. Here c denotes the velocity of light. We will introduce below
Lorentz-invariant differential equations which take the place of the Schrödinger equation of a par-
ticle of mass m and charge q in an electromagnetic field [c.f. (refeq:ham2, 8.45)] described by an
electrical potential V (~r, t) and a vector potential A(~~ r, t)
2
∂ 1 ~ q~
i~ ψ(~r, t) = ∇ − A(~r, t) + qV (~r, t) ψ(~r, t) (10.2)
∂t 2m i c
The replacement of (10.2) by Lorentz–invariant equations will have two surprising and extremely
important consequences: some of the equations need to be formulated in a representation for which
the wave functions ψ(~r, t) are vectors of dimension larger one, the components representing the
spin attribute of particles and also representing together with a particle its anti-particle. We will
find that actually several Lorentz–invariant equations which replace (10.2) will result, any of these
equations being specific for certain classes of particles, e.g., spin–0 particles, spin– 12 particles, etc.
As mentioned, some of the equations describe a particle together with its anti-particle. It is not
possible to uncouple the equations to describe only a single type particle without affecting nega-
tively the Lorentz invariance of the equations. Furthermore, the equations need to be interpreted
as actually describing many–particle-systems: the equivalence of mass and energy in relativistic
formulations of physics allows that energy converts into particles such that any particle described
will have ‘companions’ which assume at least a virtual existence.
Obviously, it will be necessary to begin this Chapter with an investigation of the group of Lorentz
transformations and their representation in the space of position ~r and time t. The representation
287
288 Relativistic Quantum Mechanics
in Sect. 10.1 will be extended in Sect. 10.4 to cover fields, i.e., wave functions ψ(~r, t) and vectors
with functions ψ(~r, t) as components. This will provide us with a general set of Lorentz–invariant
equations which for various particles take the place of the Schrödinger equation. Before introduc-
ing these general Lorentz–invariant field equations we will provide in Sects. 10.5, 10.7 a heuristic
derivation of the two most widely used and best known Lorentz–invariant field equations, namely
the Klein–Gordon (Sect. 10.5) and the Dirac (Sect. 10.7) equation.
where x0 = ct describes the time coordinate and (x1 , x2 , x3 ) = ~r describes the space coordinates.
Note that the components of xµ all have the same dimension, namely that of length. We will,
henceforth, assume new units for time such that the velocity of light c becomes c = 1. This choice
implies dim(time) = dim(length).
Minkowski Space
Historically, the Lorentz transformations were formulated in a space in which the time component
of xµ was chosen as a purely imaginary number and the space components real. This space is
called the Minkowski space. The reason for this choice is that the transformations (10.1) leave the
quantity
s2 = (x0 )2 − (x1 )2 − (x2 )2 − (x3 )2 (10.4)
The Lorentz transformations L describe the relationship between space-time coordinates xµ of two
reference frames which move relative to each other with uniform fixed velocity ~v and which might
be reoriented relative to each other by a rotation around a common origin. Denoting by xµ the
10.1: Natural Representation of the Lorentz Group 289
coordinates in one reference frame and by x0 µ the coordinates in the other reference frame, the
Lorentz transformations constitute a linear transformation which we denote by
3
0µ
X
x = Lµ ν xν . (10.6)
ν=0
Here Lµ ν are the elements of a 4 × 4–matrix representing the Lorentz transformation. The upper
index closer to ‘L’ denotes the first index of the matrix and the lower index ν further away from
‘L’ denotes the second index. [ A more conventional notation would be Lµν , however, the latter
notation will be used for different quantities further below.] The following possibilities exist for the
positioning of the indices µ, ν = 0, 1, 2, 3:
The reason for the notation is two-fold. First, the notation in (10.6) allows us to introduce the
so-called summation conventon: any time the same index appears in an upper and a lower position,
summation over that index is assumed without explicitly noting it, i.e.,
X3 X 3 X 3
yµ x µ = yµ xµ ; Aµ ν xν = Aµ ν xν ; Aµ ν B ν ρ = Aµ ν B ν ρ . (10.8)
| {z } | {z } | {z }
µ=0
new | {z } new |ν=0 {z } new |ν=0 {z }
old old old
The summation convention allows us to write (10.6) x0 µ = Lµ ν xν . The second reason is that upper
and lower positions allow us to accomodate the expression (10.5) into scalar products. This will be
explained further below.
The Lorentz transformations are non-singular 4 × 4–matrices with real coefficients, i.e., L ∈ GL(4,
R), the latter set constituting a group. The Lorentz transformations form the subgroup of all
matrices which leave the expression (10.5) invariant. This condition can be written
µ ν
xµ gµν xν = x0 gµν x0 (10.9)
where
1 0 0
0 −1 0 0
( gµν ) =
0 0 −1 0 = g .
(10.10)
0 0 0 −1
Combining condition (10.9) and (10.6) yields
This condition specifies the key property of Lorentz transformations. We will exploit this property
below to determine the general form of the Lorentz transformations. The subset of GL(4, R), the
290 Relativistic Quantum Mechanics
elements of which satisfy this condition, is called O(3,1). This set is identical with the set of all
Lorentz transformations L. We want to show now L = O(3,1) ⊂ GL(4, R) is a group.
To simplify the following proof of the key group properties we like to adopt the conventional matrix
notation for Lµ ν 0
L 0 L0 1 L0 2 L0 3
L1 0 L1 1 L1 2 L1 3
L = ( Lµ ν ) = L2 0 L2 1 L2 2 L3 3 .
(10.13)
L3 0 L3 1 L3 2 L3 3
Using the definition (10.10) of g one can rewrite the invariance property (10.12)
LT gL = g . (10.14)
we note first that the identity matrix 11 is an element of O(3,1) since it satisfies (10.14). We consider
then L1 , L2 ∈ O(3,1). For L3 = L1 L2 holds
i.e., L3 ∈ O(3,1). One can also show that if L ∈ O(3,1), the associated inverse obeys (10.14), i.e.,
L−1 ∈ O(3,1). In fact, employing expressions (10.16, 10.17) one obtains
Multiplying (10.14) from the right by gLT and using (10.15) one can derive LT gLgLT = LT and
multiplying this from the left by by g(LT )−1 yields
L g LT = g (10.21)
Using this result to simplify the r.h.s. of (10.20) results in the desired property
i.e., property (10.14) holds for the inverse of L. This stipulates that O(3,1) is, in fact, a group.
10.1: Natural Representation of the Lorentz Group 291
where L↑+ , L↓+ , L↑− , L↓− are disjunct sets defined as follows
2
3 3 3
X
0 j
X
0 2
X 2
A jB 0 ≤ A j Bj 0 . (10.34)
j=1 j=1 j=1
292 Relativistic Quantum Mechanics
conclude from (10.21) 3j=1 (A0 j )2 = (A0 0 )2 − 1. (10.34) provides then the estimate
P
2
3
X
A0 j B j 0 ≤ (A0 0 )2 − 1 (B 0 0 )2 − 1 < (A0 0 )2 (B 0 0 )2 .
(10.35)
j=1
A0 0 B 0 0 ≥ 1. Using the above expression for C 0 0 one can state C 0 0 > 0. In fact, since the group
property of O(3,1) ascertains CT gC = g it must hold C 0 0 ≥ 1.
The next group property of L↑+ to be demonstrated is the existence of the inverse. For the inverse of
any L ∈ L↑+ holds (10.16). This relationship shows (L−1 )0 0 = L0 0 , from which one can conclude
L−1 ∈ L↑+ . We also note that the identity operator 11 has elements
11µ ν = δ µ ν (10.36)
where we defined1
µ 1 for µ = ν
δ ν = (10.37)
0 for µ 6= ν
It holds, 110 0 = ≥ 1 and, hence, 11 ∈ L↑+ . Since the associative property holds for matrix multipli-
cation we have verified that L↑+ is indeed a subgroup of SO(3,1).
L↑+ is called the subgroup of proper, orthochronous Lorentz transformations. In the following we
will consider solely this subgroup of SO(3,1).
L µ ν = δ µ ν + µ ν ; µ ν small . (10.38)
For these transformations, obviously, holds L0 0 > 0 and the value of the determinant is close to
unity, i.e., if we enforce (10.14) actually L0 0 ≥ 1 and det L = 1 must hold. Property (10.14)
implies
11 + T g ( 11 + ) = g
(10.39)
where we have employed the matrix form defined as in (10.13). To order O(2 ) holds
T g + g = 0 . (10.40)
1 ρ
It should be noted that according to our present definition holds δµν = gµρ δ ν and, accordingly, δ00 = 1 and
δ11 = δ22 = δ33 = −1.
10.1: Natural Representation of the Lorentz Group 293
µ µ = 0
0 j = j 0 , j = 1, 2, 3
k
j k = − j , j, k = 1, 2, 3 (10.43)
Inspection shows that the matrix has 6 independent elements and can be written
0 −w1 −w2 −w3
−w1 0 −ϑ3 ϑ2
(ϑ1 , ϑ2 , ϑ3 , w1 , w2 , w3 ) =
−w2
. (10.44)
ϑ3 0 −ϑ1
−w3 −ϑ2 ϑ1 0
This result allows us now to define six generators for the Lorentz transformations(k = 1, 2, 3)
[ Jk , J` ] = k`m Jm (10.49)
[ Kk , K` ] = − k`m Jm
[ J k , K` ] = k`m Km .
Exercise 7.1:
Demonstrate the commutation relationships (10.49, 10.50).
The commutation relationships (10.49) define the Lie algebra associated with the Lie group L↑+ .
The commutation relationships imply that the algebra of the generators Jk , Kk , k = 1, 2, 3 is closed.
Following the treatment of the rotation group SO(3) one can express the elements of L↑+ through
the exponential operators
~ w)
L(ϑ, ~ = exp ϑ ~ · ~J + w ~
~ ·K ~ w
; ϑ, ~ ∈ R3 (10.51)
where the 3 × 3–matrices Lk are the generators of SO(3) defined in Chapter 5, that the transfor-
mations (10.51) for w
~ = 0 correspond to rotations of the spatial coordinates, i.e.,
~ 0 0
L(ϑ, w
~ = 0) = ~ . (10.53)
0 R(ϑ)
~ are the 3 × 3–rotation matrices constructed in Chapter 5. For the parameters ϑk of the
Here R(ϑ)
Lorentz transformations holds obviously
which, however, constitutes an overcomplete parametrization of the rotations (see Chapter 5).
We consider now the Lorentz transformations for ϑ~ = 0 which are referred to as ‘boosts’. A boost
in the x1 –direction is L = exp(w1 K1 ). To determine the explicit form of this transformation we
evaluate the exponential operator by Taylor expansion. In analogy to equation (5.35) it issufficient
to consider in the present case the 2 × 2–matrix
∞ n
w1n
0 0 −1 X 0 −1
L = exp w1 = (10.55)
−1 0 n! −1 0
n=0
since
0 0
L0
0 0
exp (w1 K1 ) = exp . (10.56)
0 0 1 0
0 0 0 1
Using the idempotence property
2
0 −1 1 0
= = 11 (10.57)
−1 0 0 1
10.1: Natural Representation of the Lorentz Group 295
The conventional form (10.1) of the Lorentz transformations is obtained through the parameter
change
sinh w1
v1 = = tanh w1 (10.59)
cosh w1
p p
Using cosh2 w1 − sinh2 w1 = 1 one can identify sinhw1 = cosh2 w1 − 1 and coshw1 = sinh2 w1 + 1.
Correspondingly, one obtains from (10.59)
p
cosh2 w1 − 1 sinh w1
v1 = = p . (10.60)
cosh w1 sinh2 w1 + 1
√−v1
1
√ 0 0
1 − v12 1 − v12
√−v1 2 1
√ 0 0
exp (w1 K1 ) =
1 − v1 1 − v12
(10.62)
0 0 1 0
0 0 0 1
According to (10.3, 10.6, 10.51) the explicit transformation for space–time–coordinates is then
x1 − v1 t t − v1 x1
x01 = p 2
, t0 = p 2
, x02 = x2 , x03 = x3 (10.63)
1 − v1 1 − v1
Scalars The quantities with the simplest transformation behaviour are so-called scalars f ∈ R
which are invariant under transformations, i.e.,
f0 = f . (10.65)
An example is s2 defined in (10.4), another example is the rest mass m of a particle. However, not
any physical property f ∈ R is a scalar. Counterexamples are the energy, the charge density, the
~ r, t)|2 or the scalar product ~r1 · ~r2
z–component x3 of a particle, the square of the electric field |E(~
of two particle positions. We will see below how true scalars under Lorentz transformations can be
constructed.
4-Vectors The quantities with the transformation behaviour like that of the position–time vector
xµ defined in (10.3) are the so-called 4–vectors aµ . These quantites always come as four components
(a0 , a1 , a2 , a3 )T and transform according to
µ
a0 = Lµ ν aν . (10.66)
Examples of 4-vectors beside xµ are the momentum 4-vector
m m ~v
pµ = (E, p~) , E = √ , p~ = √ (10.67)
1 − ~v 2 1 − ~v 2
the transformation behaviour of which we will demonstrate further below. A third 4-vector is the
so-called current vector
~
J µ = (ρ, J) (10.68)
~ r, t) are the charge density and the current density, respectively, of a system of
where ρ(~r, t) and J(~
charges. Another example is the potential 4-vector
~
Aµ = (V, A) (10.69)
~ r, t) are the electrical and the vector potential of an electromagnetic field. The
where V (~r, t) and A(~
4-vector character of J µ and of Aµ will be demonstrated further below.
10.2: Scalars, 4–Vectors and Tensors 297
Scalar Product 4-vectors allow one to construct scalar quantities. If aµ and bµ are 4-vectors
then
aµ gµν bν (10.70)
is a scalar. This property follows from (10.66) together with (10.12)
µ ν
a0 gµν b0 = Lµ ρ gµν Lν σ aρ bσ = aρ gρσ bσ (10.71)
where aν is a vector with transformation behaviour as stated in (10.66). One calls 4-vectors aµ
covariant and 4-vectors aµ contravariant. Covariant 4-vectors transform like
a0 µ = gµν Lν ρ g ρσ aσ (10.73)
where we defined
g µν = gµν . (10.74)
We like to point out that from definition (10.72) of the covariant 4-vector follows =aµ g µν aν .
In
µν µ
fact, one can employ the tensors g and gµν to raise and lower indices of L ν as well. We do not
establish here the consistency of the ensuing notation. In any case one can express (10.73)
a0µ = Lµ σ aσ . (10.75)
Note that according to (10.17) Lµ σ is the transformation inverse to Lσ µ . In fact, one can express
[(L−1 )T ]µ ν = (L−1 )ν µ and, accordingly, (10.17) can be written
(L−1 )ν µ = Lµ ν . (10.76)
def ∂
∂µ0 = . (10.78)
∂x0µ
To prove the 4-vector property of ∂µ we will show that g µν ∂ν transforms like a contravariant 4-
vector, i.e., g µν ∂ν0 = Lµ ρ g ρσ ∂σ . We start from x0 µ = Lµ ν xν . Multiplication (and summation) of
x0 µ = Lµ ν xν by Lρ σ gρµ yields, using (10.12), gσν xν = Lρ σ gρµ x0 µ and g µσ gσν = δ µ ν ,
µ
xν = g νσ Lρ σ gρµ x0 . (10.79)
This is the inverse Lorentz transformation consistent with (10.16). We have duplicated the expres-
sion for the inverse of Lµ ν to obtain the correct notation in terms of covariant, i.e., lower, and
298 Relativistic Quantum Mechanics
contravariant, i.e., upper, indices. (10.79) allows us to determine the connection between ∂µ and
∂µ0 . Using the chain rule of differential calculus we obtain
3
X ∂xν ∂
∂µ0 = = g νσ Lρ σ gρµ ∂ν = Lµ ν ∂ν (10.80)
∂x0µ ∂xν
ν=0
g λµ ∂µ0 = Lλ σ g σν ∂ν , (10.81)
d’Alembert Operator We want to construct now a scalar differential operator. For this purpose
we define first the contravariant differential operator
µ µν ∂
∂ = g ∂ν = , −∇ . (10.82)
∂t
Proof that pµ is a 4-vector We will demonstrate now that the momentum 4-vector pµ defined
in (10.67) transforms like (10.66). For this purpose we consider the scalar differential
It holds 2
dτ
= 1 − (~v )2 (10.85)
dt
from which follows
d 1 d
= √ . (10.86)
dτ 1 − ~v 2 dt
One can write
m m dt
p0 = E = √ = √ . (10.87)
1 − ~v 2 1 − ~v dt
2
m v1 m dx1
p1 = √ = √ . (10.88)
1 − ~v 2 1 − ~v 2 dt
One can express then the momentum vector
m dxµ d
pµ = √ = m xµ . (10.89)
1 − ~v 2 dt dτ
10.3: Relativistic Electrodynamics 299
d
The operator m dτ transforms like a scalar. Since xµ transforms like a contravariant 4-vector, the
r.h.s. of (10.89) alltogether transforms like a contravariant 4-vector, and, hence, pµ on the l.h.s. of
(10.89) must be a 4-vector.
The momentum 4-vector allows us to construct a scalar quantity, namely
pµ pµ = pµ gµν pν = E 2 − p~ 2 (10.90)
Evaluation of the r.h.s. yields according to (10.67)
m2 m2~v 2
E 2 − p~ 2 = − = m2 (10.91)
1 − ~v 2 1 − ~v 2
or
p µ p µ = m2 (10.92)
which, in fact, is a scalar. We like to rewrite the last result
E 2 = p~ 2 + m2 (10.93)
or p
E = ± p~ 2 + m2 . (10.94)
1
In the non-relativistic limit the rest energy m is the dominant contribution to E. Expansion in m
should then be rapidly convergent. One obtains
p~ 2 p 2 )2
2 3
(~ (~
p )
E = ±m ± ∓ 3
+ O . (10.95)
2m 4m 4m5
p
~ 2
This obviously describes the energy of a free particle with rest energy ±m, kinetic energy ± 2m and
relativistic corrections.
Lorentz Gauge In our previous description of the electrodynamic field we had introduced the
~ r, t), respectively, and had chosen the so-called Coulomb
scalar and vector potential V (~r, t) and A(~
~
gauge (8.12), i.e., ∇ · A = 0, for these potentials. This gauge is not Lorentz-invariant and we will
adopt here another gauge, namely,
~ r, t) = 0 .
∂t V (~r, t) + ∇ · A(~ (10.96)
The Lorentz-invariance of this gauge, the so-called Lorentz gauge, can be demonstrated readily
using the 4-vector notation (10.69) for the electrodynamic potential and the 4-vector derivative
(10.77) which allow one to express (10.96) in the form
∂µ Aµ = 0 . (10.97)
We have proven already that ∂µ is a contravariant 4-vector. If we can show that Aµ defined in
(10.69) is, in fact, a contravariant 4-vector then the l.h.s. in (10.97) and, equivalently, in (10.96) is
a scalar and, hence, Lorentz-invariant. We will demonstrate now the 4-vector property of Aµ .
300 Relativistic Quantum Mechanics
which reflects the principle of charge conservation. This principle should hold in any frame of
reference. Equation (10.98) can be written, using (10.77) and (10.68),
∂µ J µ (xµ ) = 0 . (10.99)
Since this equation must be true in any frame of reference the right hand side must vanish in all
frames, i.e., must be a scalar. Consequently, also the l.h.s. of (10.99) must be a scalar. Since ∂µ
transforms like a covariant 4-vector, it follows that J µ , in fact, has to transform like a contravariant
4-vector.
We want to derive now the differential equations which determine the 4-potential Aµ in the Lorentz
gauge (10.97) and, thereby, prove that Aµ is, in fact, a 4-vector. The respective equation for A0 = V
can be obtained from Eq. (8.13). Using ∇ · ∂t A(~ ~ r, t) = ∂t ∇ · A(~~ r, t) together with (10.96), i.e.,
~
∇ · A(~r, t) = −∂t V (~r, t), one obtains
In this equation the r.h.s. transforms like a 4-vector. The l.h.s. must transform likewise. Since
∂µ ∂ µ transforms like a scalar one can conclude that Aν (xσ ) must transform like a 4-vector.
F k0 = −F 0k = E k , F mn = −mn` B ` , k, `, m, n = 1, 2, 3 (10.104)
where mn` = mn` is the totally anti-symmetric three-dimensional tensor defined in (5.32).
One can readily verify, using (8.6) and (8.9), that F µν can be expressed through the potential Aµ
in (10.69) and ∂ µ in (10.82) as follows
F µν = ∂ µ Aν − ∂ ν Aµ . (10.105)
10.3: Relativistic Electrodynamics 301
~ r, t) and B(~
The relationships (10.103, 10.104) establishe the transformation behaviour of E(~ ~ r, t).
In a new frame of reference holds
F 0µν = Lµ α Lν β F αβ (10.106)
In case that the Lorentz transformation Lµ ν is given by (10.62) or, equivalently, by (10.63), one
obtains
Ey −v1 Bz Ez +v1 By
− √ − √
0 −Ex 2
1−v1 1−v12
Bz −v1 Ey By +v1 Ez
√ √
E x 0 −
0µν
1−v12 1−v12
F = E −v B
Bz −v1 Ey
(10.107)
√ 2
y 1 z
√ 2 0 −Bx
1−v1 1−v1
Ez +v1 By By +v1 Ez
√ 2 − √ 2 Bx 0
1−v1 1−v1
Comparision with
0 −Ex0 −Ey0 −Ez0
Ex0 0 −Bz0 By0
F 0µν =
Ey0 0 (10.108)
−Bx0
Bz 0
Ez0 −By0 Bx0 0
~ 0 and B
yields then the expressions for the transformed fields E ~ 0 . The results can be put into the
more general form
~ + ~v × B ~
~0
E = ~k ,
E ~ 0 = E⊥
E √ (10.109)
k ⊥
1 − ~v 2
~ − ~v × E ~
~0
B = ~k ,
B ~ 0 = B⊥
B √ (10.110)
k ⊥
1 − ~v 2
where E~ k, B
~ k and E
~ ⊥, B
~ ⊥ are, respectively, the components of the fields parallel and perpendicular
to the velocity ~v which determines the Lorentz transformation. These equations show that under
Lorentz transformations electric and magnetic fields convert into one another.
∂µ F µν = ∂µ ∂ µ Aν − ∂µ ∂ ν Aµ = ∂µ ∂ µ Aν − ∂ ν ∂µ Aµ = ∂µ ∂ µ Aν , (10.111)
where we used (10.105) and (10.97), one can conclude from (10.102)
∂µ F µν = 4π J ν . (10.112)
One can readily prove that this equation is equivalent to the two inhomogeneous Maxwell equations
(8.1, 8.2). From the definition (10.105) of the tensor F µν one can conclude the property
∂ σ F µν + ∂ µ F νσ + ∂ ν F σµ = 0 (10.113)
which can be shown to be equivalent to the two homogeneous Maxwell equations (8.3, 8.4).
302 Relativistic Quantum Mechanics
Lorentz Force
One important property of the electromagnetic field is the Lorentz force acting on charged particles
moving through the field. We want to express this force through the tensor F µν . It holds for a
particle with 4-momentum pµ as given by (10.67) and charge q
dpµ q
= pν F µν (10.114)
dτ m
where d/dτ is given by (10.86). We want to √ demonstrate now that this equation is equivalent to
the equation of motion (8.5) where p~ = m~v / 1 − v 2 .
To avoid √confusion we will employ in the following for the energy of the particle the notation
E = m/ 1 − v 2 [see (10.87)] and retain the definition E ~ for the electric field. The µ = 0
component of (10.114) reads then, using (10.104),
dE q ~
= p~ · E (10.115)
dτ m
or with (10.86)
dE q ~ .
= p~ · E (10.116)
dt E
From this one can conclude, employing (10.93),
1 dE 2 1 d~p 2 ~
= = q p~ · E (10.117)
2 dt 2 dt
This equation follows, however, also from the equation of motion (8.5) taking the scalar product
with p~
d~p ~
p~ · p·E
= q~ (10.118)
dt
√
where we exploited the fact that according to p~ = m~v / 1 − v 2 holds p~ k ~v .
For the spatial components, e.g., for µ = 1, (10.114) reads using (10.103)
dpx q
= ( EEx + py Bz − pz By ) . (10.119)
dτ m
√
Employing again (10.86) and (10.67), i.e., E = m/ 1 − v 2 , yields
dpx h
~ x
i
= q Ex + (~v × B) (10.120)
dt
which is the x-component of the equation of motion (8.5). We have, hence, demonstrated that
(10.114) is, in fact, equivalent to (8.5). The term on the r.h.s. of (10.120) is referred to as the
Lorentz force. Equation (10.114), hence, provides an alternative description of the action of the
Lorentz force.
10.4: Function Space Representation of Lorentz Group 303
which states that the function values ψ 0 (x0µ ) at each point x0µ in the new frame are identical to the
function values ψ(xµ ) in the old frame taken at the same space–time point xµ , i.e., taken at the pairs
of points (x0µ = Lµ ν xν , xµ ). We need to emphasize that (10.121) covers solely the transformation
behaviour of scalar functions. Functions which represent 4-vectorsor other non-scalar entities, e.g.,
the charge-current density in case of Sect. 10.3 or the bi-spinor wave function of electron-positron
pairs in Sect. 10.7, obey a different transformation law.
We like to express now ψ 0 (x0µ ) in terms of the old coordinates xµ . For this purpose one replaces
xµ in (10.121) by (L−1 )µ ν xν and obtains
This result gives rise to the definition of the function space representation ρ(Lµ ν ) of the Lorentz
group
def
(ρ(Lµ ν )ψ)(xµ ) = ψ((L−1 )µ ν xν ) . (10.123)
This definition corresponds closely to the function space representation (5.42) of SO(3). In analogy
to the situation for SO(3) we seek an expression for ρ(Lµ ν ) in terms of an exponential operator and
transformation parameters ϑ, ~ w,
~ i.e., we seek an expression which corresponds to (10.51) for the
natural representation of the Lorentz group. The resulting expression should be a generalization of
the function space representation (5.48) of SO(3), in as far as SO(3,1) is a generalization (rotation
+ boosts) of the group SO(3). We will denote the intended representation by
~ w) def ~ w)) ~ ~ ~ K
~
L(ϑ, ~ = ρ(Lµ ν (ϑ, ~ = ρ eϑ·J + w· (10.124)
1 h ϑk J k i
Jk = lim ρ e − 11 (10.126)
ϑk →0 ϑ1
and Kk
1
ρ ewk Kk − 11 .
Kk = lim (10.127)
wk →0 w1
304 Relativistic Quantum Mechanics
One obtains
J1 = x3 ∂2 − x2 ∂3 ; K1 = x0 ∂1 + x1 ∂0
J2 = x1 ∂3 − x3 ∂1 ; K2 = x0 ∂2 + x2 ∂0
J3 = x2 ∂1 − x1 ∂2 ; K3 = x0 ∂3 + x3 ∂0 (10.128)
[ Jk , J` ] = k`m Jm
[ Kk , K` ] = − k`m Jm
[ Jk , K` ] = k`m Km . (10.133)
We demonstrate this for three cases, namely [J1 , J2 ] = J3 , [K1 , K2 ] = −J3 , and [J1 , K2 ] = K3 :
[ J1 , J2 ] = [x3 ∂2 − x2 ∂3 , x1 ∂3 − x3 ∂1 ]
= [x3 ∂2 , x1 ∂3 ] − [x2 ∂3 , x3 ∂1 ]
= −x1 ∂2 + x2 ∂1 = J3 , (10.134)
10.4: Function Space Representation of Lorentz Group 305
[ K1 , K2 ] = [x0 ∂1 + x1 ∂0 , x0 ∂2 + x2 ∂0 ]
= [x0 ∂1 , x2 ∂0 ] − [x1 ∂0 , x0 ∂2 ]
= −x2 ∂1 + x1 ∂2 = − J3 , (10.135)
[ J1 , K2 ] = [x3 ∂2 − x2 ∂3 , x0 ∂2 + x2 ∂0 ]
= [x3 ∂2 , x2 ∂0 ] − [x2 ∂3 , x0 ∂2 ]
= x3 ∂0 + x0 ∂3 = K3 . (10.136)
L(w3 ) = exp w3 K3 ,
(10.137)
where K3 is given in (10.128), can be simplified considerably. For this purpose one expresses K3 in
terms of hyperbolic coordinates R, Ω which are connected with x0 , x3 as follows
∂R x0 ∂R x0
= , = − (10.141)
∂x0 R ∂x3 R
and
∂Ω ∂Ω ∂tanhΩ 1
= = cosh2 Ω 0
∂x3 ∂tanhΩ ∂x 3 x
∂Ω ∂Ω ∂cothΩ 1
= = − sinh2 Ω 3 . (10.142)
∂x0 ∂cothΩ ∂x 0 x
The chain rule yields then
∂R ∂ ∂Ω ∂ x0 ∂ 1 ∂
∂0 = 0
+ 0
= − sinh2 Ω 3
∂x ∂R ∂x ∂Ω R ∂R x ∂Ω
∂R ∂ ∂Ω ∂ x3 ∂ 1 ∂
∂3 = 3
+ 3
= − + cosh2 Ω 0 . (10.143)
∂x ∂R ∂x ∂Ω R ∂R x ∂Ω
306 Relativistic Quantum Mechanics
∂
K3 = x0 ∂3 + x3 ∂0 = . (10.144)
∂Ω
The action of the exponential operator (10.137) on a function f (Ω) ∈ C∞ (1) is then that of a shift
operator
3 3 ∂
L(w ) f (Ω) = exp w f (Ω) = f (Ω + w3 ) . (10.145)
∂Ω
Applying the correspondence principle to (10.92) one obtains the wave equation
or
~2 ∂ µ ∂µ + m2 ψ(xν ) = 0 .
(10.149)
where ψ(xµ ) is a scalar, complex-valued function. The latter property implies that upon change of
reference frame ψ(xµ ) transforms according to (10.121, 10.122). The partial differential equation
(10.151) is called the Klein-Gordon equation.
In the following we will employ so-called natural units ~ = c = 1. In these units the quantities
energy, momentum, mass, (length)−1 , and (time)−1 all have the same dimension. In natural units
the Klein–Gordon equation (10.151) reads
∂µ ∂ µ + m2 ψ(xµ ) = 0
(10.150)
10.5: Klein–Gordon Equation 307
or
∂t2 − ∇2 + m2 ψ(xµ ) = 0 .
(10.151)
One can notice immeadiately that (10.150) is invariant under Lorentz transformations. This follows
from the fact that ∂µ ∂ µ and m2 are scalars, and that (as postulated) ψ(xµ ) is a scalar. Under
Lorentz transformations the free particle Klein–Gordon equation (10.150) becomes
∂µ0 ∂ 0µ + m2 ψ 0 (x0µ ) = 0
(10.152)
which has the same form as the Klein–Gordon equation in the original frame.
This conservation law differs in one important aspect from that of the Schrödinger equation (10.154),
namely, in that the expression for ρKG is not positive definite. When the Klein-Gordon equation
had been initially suggested this lack of positive definiteness worried physicists to a degree that the
Klein–Gordon equation was rejected and the search for a Lorentz–invariant quantum mechanical
wave equation continued. Today, the Klein-Gordon equation is considered as a suitable equation
to describe spin–0 particles, for example pions. The proper interpretation of ρKG (~r, t), it had been
realized later, is actually that of a charge density, not of particle probability.
which results in the expected [see (10.93] dispersion relationship connecting E0 , p~0 , m
∂µ ∂ µ + m2 ψo (~p, λ|xµ )
= 0 (10.166)
p
p, λ|xµ ) = Nλ,p ei(~p·~r − λEo (~p)t) Eo (~
ψo (~ p) = m2 + p~o2 , λ = ±
We want to demonstrate now that the wave functions for p~ 6= 0 in (??) can be obtained through
appropriate Lorentz transformation of (10.168). For this purpose we consider the wave function for
a particle moving with momentum velocity v in the direction of the x3 –axis. Such wave function
should be generated by applying the Lorentz transformation in the function space representation
p
(10.145) choosing m = sinhw3 . This yields, in fact, for the wave function (10.168), using (10.138)
to replace t = x0 by hyperbolic coordinates R, Ω,
3 ∂
3
L(w )ψo (~ µ
p = 0, λ|x ) = exp w N e−iλmRcoshΩ
∂Ω
3)
= N e−iλmRcosh(Ω+w . (10.169)
The addition theorem of hyperbolic functions cosh(Ω + w3 ) = coshΩ coshw3 + sinhΩ sinhw3 allows
us to rewrite the exponent on the r.h.s. of (10.169)
The coordinate transformation (10.138) and the relationships (10.61) yield for this expression
m mv
−iλ √ x0 − iλ √ x3 . (10.171)
1−v 2 1−v 2
One can interpret then for λ = +, i.e., for positive energy solutions,
p
p = − mv/ 1 − v 2 (10.172)
which agrees with the expression given in (10.166). In case of λ = −, i.e., for negative energy
solutions, one has to interprete p
p = mv/ 1 − v 2 (10.175)
as the momentum of the particle and one obtains
3 + Eo (p)x0
L(w3 )ψo (~
p = 0, λ = −|xµ ) = N ei(px . (10.176)
momentum p~ ~
p~ − q A ˆ~ = −i∇
p ˆ~ − q A
p ˆ
~ = ~π
Table 10.1:
Coupling of a particle of charge q to an electromagnetic field described by the 4-vector potential
Aµ = (V, A) ~ or Aµ = (V, −A).~ According to the so-called minimum coupling principle the presence
of the field is accounted for by altering energy, momenta for classical particles and the respective
operators for quantum mechanical particles in the manner shown. See also Eq. (10.147).
To obtain the appropriate wave equation we follow the derivation of the free particle Klein–Gordon
equation above and apply again the correspondence principle to (10.93), albeit in a form, which
couples a particle of charge q to an electromagnetic field described through the potential Aµ (xν ).
According to the principle of minimal coupling [see (10.69)] one replaces the quantum mechanical
operators, i.e., i∂t and −i∇ in (10.150), according to the rules shown in Table 10.1. For this purpose
one writes the Klein-Gordon equation (10.150)
and seek an equation for Ψ(~r, t). We will also assume, in keeping withnthe non-relativistic limit,
that the mass m of the particle, i.e., it’s rest energy, is much larger than all other energy terms, in
10.6: Klein–Gordon Equation with Electromagnetic Field 311
where we neglected all terms which did not contain factors m. The approximation is justified on
the ground of the inequalities (10.183). The Klein-Gordon equation (10.181) reads then
" #
ˆ~ − q A(~
[p ~ r, t)]2
i ∂t Ψ(~r, t) = + qV (~r, t) Ψ(~r, t) (10.185)
2m
This is, however, identical to the Schrödinger equation (10.2) of a non-relativistic spin-0 particle
moving in an electromagnetic field.
Pionic Atoms
To apply the Klein–Gordon equation (10.181) to a physical system we consider pionic atoms, i.e.,
atoms in which one or more electrons are replaced by π − mesons. This application demonstrates
that the Klein–Gordon equation describes spin zero particles, e.g., spin-0 mesons.
To ‘manufacture’ pionic atoms, π − mesons are generated through inelastic proton–proton scattering
p + p −→ p + p + π − + π + , (10.186)
then are slowed down, filtered out of the beam and finally fall as slow pions onto elements for which
a pionic variant is to be studied. The process of π − meson capture involves the so-called Auger
effect, the binding of a negative charge (typically an electron) while at the same time a lower shell
electron is being emitted
We want to investigate in the following a description of a stationary state of a pionic atom involving
a nucleus with charge +Ze and a π − meson. A stationary state of the Klein–Gordon equation is
described by a wave function
ψ(xµ ) = ϕ(~r ) e−it . (10.188)
Inserting this into (10.181) yields (we assume now that the Klein–Gordon equation describes a
2
~ r, t) ≡ 0
particle with mass mπ and charge −e) for qV (~r, t) = − Zer and A(~
" 2 #
Ze2 2 2
+ + ∇ − mπ ϕ(~r ) = 0 . (10.189)
r
312 Relativistic Quantum Mechanics
Because of the radial symmetry of the Coulomb potential we express this equation in terms of
spherical coordinates r, θ, φ. The Laplacian is
1 2 1 1 1 L̂2
∇2 = ∂r r + 2 2 ∂θ sinθ∂θ + 2 2 ∂φ2 = ∂r2 r − 2 . (10.190)
r r sin θ r sin θ r r
Ze2 2
With this expression and after expanding ( + r ) one obtains
!
d2 L̂2 − Z 2 e4 2Ze2
− + + 2 − m2π r φ(~r) = 0 . (10.191)
dr2 r2 r
The operator L̂2 in this equation suggests to choose a solution of the type
R` (r)
ϕ(~r ) = Y`m (θ, φ) (10.192)
r
where the functions Y`m (θ, φ) are spherical harmonics, i.e., the eigenfunctions of the operator L̂2
in (10.191)
L̂2 Y`m (θ, φ) = ` (` + 1) Y`m (θ, φ) . (10.193)
(10.192) leads then to the ordinary differential equation
d2 `(` + 1) − Z 2 e4 2Ze2
− + + 2 − m2π R` (r) = 0 . (10.194)
dr2 r2 r
Bound state solutions can be obtained readily noticing that this equation is essentially identical to
2
that posed by the Coulomb problem (potential − Zer ) for the Schrödinger equation
d2 2mπ Ze2
`(` + 1)
− + + 2mπ E R` (r) = 0 (10.195)
dr2 r2 r
mπ (Ze2 )2
En = − ; n = 1, 2, . . . ; ` = 0, 1, . . . n − 1 . (10.196)
2n2
In this expression the number n0 defined through
n0 = n − ` − 1 (10.197)
counts the number of nodes of the wave function, i.e., this quantity definitely must be an integer.
The similarity of (10.194) and (10.195) can be made complete if one determines λ such that
d2 2Ze2
λ(`) ( λ(`) + 1 )
− + + 2 − m2π R` (r) = 0 . (10.200)
dr2 r2 r
The bound state solutions of this equation should correspond to values which can be obtained
from (10.196) if one makes the replacement
2 − m2π
E −→ , ` −→ λ(`) , e2 −→ e2 . (10.201)
2mπ mπ
One obtains
2
2 − m2π mπ Z 2 e4 m 2
π
= − . (10.202)
2mπ 2 (n0 + λ(`) + 1)2
Solving this for (choosing the root which renders ≤ mπ , i.e., which corresponds to a bound
state) yields
mπ
= q ; n0 = 0, 1, . . . ; ` = 0, 1, . . . . (10.203)
Z 2 e4
1 + ( n0 + λ(`)+1 )2
EKG (n, `, m) =
Using (10.197, 10.199) and defining EKG = results in the spectrum mπ
r
Z 2 e4
1 + 1
q
(n−`− 2
+ (`+ 12 )2 − Z 2 e4 )2
n = 1, 2, . . .
` = 0, 1, . . . , n − 1
m = −`, −` + 1, . . . , +`
(10.204)
In order to compare this result with the spectrum of the non-relativistic hydrogen-like atom we
expand in terms of the fine structure constant e2 to order O(8 ). Introducing α = Z 2 e4 and
β = ` + 12 (10.204) reads
1
q α
(10.205)
1+ √ 2 2(n − β + β −α)
1
α
q
1+ √
(n − β + β 2 −α)2
1
≈ q α
1+ (n − α
+ O(α2 ))2
2β
1
≈ q α
1+ n2 − α
n + O(α2 )
β
314 Relativistic Quantum Mechanics
1
≈ q
α α2
1+ n2
+ βn3
+ O(α3 )
1
≈ α α2 α2
1+ 2n2
+ 2βn3
− 8n4
+ O(α3 )
α α2
α2 α2
≈ 1− − + + + O(α3 ) . (10.206)
2n2 2βn3 8n4 4n4
Here the first term represents the rest energy, the second term the non-relativistic energy, and the
third term gives the leading relativistic correction. The latter term agrees with observations of
pionic atoms, however, it does not agree with observations of the hydrogen spectrum. The latter
spectrum shows, for example, a splitting of the six n = 2, ` = 1 states into groups of two and four
degenerate states. In order to describe electron spectra one must employ the Lorentz-invariant
wave equation for spin- 21 particles, i.e., the Dirac equation introduced below.
It must be pointed out here that does not denote energy, but in the present case rather the negative
of the energy. Also, the π − meson is a pseudoscalar particle, i.e., the wave function changes sign
under reflection.
equation p
i∂t Ψ(~r, t) = ± m2 − ∇2 Ψ(~r, t) . (10.208)
These two equation can be combined
p p
i∂t + m2 − ∇2 i∂t − m2 − ∇2 Ψ(~r, t) (10.209)
10.7: Dirac Equation 315
which, in fact, is identical to the two equations (10.208). Equations (10.208, 10.209), however,
are unsatisfactory since expansion of the square root operator involves all powers of the Laplace
operator, but not an operator i~γ · ∇ as suggested by the principle of relativity (equivalence of
space and time). Many attempts were made by theoretical physicists to ‘linearize’ the square root
operator in (10.208, 10.209), but for a long time to no avail. Finally, Dirac succeeded. His solution
to the problem involved an ingenious step, namely, the realization that the linearization can be
carried out only if one assumes a 4-dimensional representation of the coefficients γ µ .
Initially, it was assumed that the 4-dimensional space introduced by Dirac could be linked to 4-
vectors, i.e., quantities with the transformation law (10.66). However, this was not so. Instead,
the 4-dimensionsional representation discovered by Dirac involved new physical properties, spin- 12
and anti-particles. The discovery by Dirac, achieved through a beautiful mathematical theory,
strengthens the believe of many theoretical physicists today that the properties of physical matter
ultimately derive from a, yet to be discovered, beautiful mathematical theory and that, therefore,
one route to important discoveries in physics is the creation of new mathematical descriptions of
nature, these descriptions ultimately merging with the true theory of matter.
From this follows that γ 0 has real eigenvalues ±1 and γ j , j = 1, 2, 3 has imaginary eigenvalues ±i.
Accordingly, one can impose the condition
Obviously, as long as det(γ µ ) 6= 0 the dimension d of the square matrices γ µ must be even so that
(−1)d = 1.
For d = 2 there exist only three anti-commuting matrices, namely the Pauli matrices σ 1 , σ 2 , σ 3 for
which, in fact, holds
2
σj = 11 ; σ j σ k = − σ k σ j for j 6= k . (10.218)
The Pauli matrices allow one, however, to construct four matrices γ µ for the next possible dimension
d = 4. A proper choice is
σj
0 11 0 j 0
γ = ; γ = , (10.219)
0 −11 −σ j 0
Using property (10.218) of the Pauli matrices one can readily prove that condition (10.213) is
satisfied. We will argue further below that the choice f γ µ , except for similarity trasnformations, is
unique.
( iγ µ ∂µ + m ) ( iγ µ ∂µ − m ) Ψ(xµ ) = 0 (10.220)
where Ψ(xµ ) represents a 4-dimensional wave function, rather than a scalar wave function. From
this equation one can conclude that also the following should hold
( iγ µ ∂µ − m ) Ψ(xµ ) = 0 (10.221)
←
where we have defined Ψ† = (ψ1∗ , ψ2∗ , ψ3∗ , ψ4∗ ) and where ∂µ denotes the differential operator ∂ µ
operating to the left side, rather than to the right side. One can readily show using the hermitian
10.7: Dirac Equation 317
(γ µ )† = γ 0 γ µ γ 0 . (10.223)
Inserting this into (10.222) and multiplication from the right by γ 0 yields the adjoint Dirac equation
←
† µ 0 µ
Ψ (x ) γ iγ ∂µ + m = 0 . (10.224)
where
γ̃ µ = S γ µ S −1 (10.227)
A representation often adopted beside the one given by (10.222, 10.219) is the socalled chiral
representation defined through
µ µ 1 11 11
Ψ̃(x ) = S Ψ(x ) ; S = √ (10.228)
2 11 −11
and
−σ j
0 0 11 j 0
γ̃ = ; γ̃ = , j = 1, 2, 3 . (10.229)
11 0 σj 0
The similarity transformation (10.227) leaves the algebra of the Dirac matrices unaffected and
commutation property (10.213) still holds, i.e.,
[γ̃ µ , γ̃ ν ]+ = 2 g µν . (10.230)
These elements form a group since obviously any product of two Γr ’s can be expressed in terms
of a third Γr . The representations of this group are given by a set of 32 4 × 4–matrices which
are equivalent with respect to similarity transformations. Since the Γj are hermitian the similarity
transformations are actually given in terms of unitary transformations. One can conclude then that
also any set of 4 × 4–matrices obeying (10.235) can differ only with respect to unitary similarity
transformations. This property extends then to 4 × 4–matrices which obey (10.213), i.e., to Dirac
matrices.
To complete the proof in this section the reader may consult Miller ‘Symmetry Groups and their
Application’ Chapter 9.6 and R.H.Good, Rev.Mod.Phys. 27, (1955), page 187. The reader may
also want to establish the unitary transformation which connects the Dirac matrices in the form
(10.236) with the Dirac representation (10.219).
Exercise 7.2:
Demonstrate the anti-commutation relationships (10.218) of the Pauli matrices σ j .
Exercise 7.3:
Demonstrate the anti-commutation relationships (10.218) of the Dirac matrices γ µ .
Exercise 7.4:
Show that from (10.214) follows that γ 0 has real eigenvalues ±1 and can be represented by a
hermitian matrix, and γ j , j = 1, 2, 3 has imaginary eigenvalues ±i and can be represented by an
anti-hermitian matrix.
where S(Lµ ν ) is a non-singular 4 × 4–matrix, the coefficients of which depend on the matrix Lµ ν
defining the Lorentz transformation in such a way that S(Lµ ν ) = 11 for Lµ ν = δ µ ν holds. Ob-
viously, the transformation (10.241) implies a similarity transformation Sγ µ S −1 . One can, hence,
state that the Dirac equation (10.221) upon Lorentz transformation yields
µ
iS(Lη ξ )γ µ S −1 (Lη ξ ) Lν µ ∂ν0 − m Ψ0 (x0 ) = 0 .
(10.242)
The form invariance of the Dirac equation under this transformation implies then the condition
We want to determine now the 4 × 4–matrix S(Lη ξ ) which satisfies this condition.
The proper starting point for a constructiuon of S(Lη ξ ) is actually (10.243) in a form in which the
Lorentz transformation in the form Lµν is on the r.h.s. of the equation. For this purpose we exploit
(10.12) in the form Lν µ gνσ Lσ ρ = gµρ = gρµ . Multiplication of (10.243) from the left by Lσ ρ gσν
yields
S(Lη ξ )γ µ S −1 (Lη ξ ) gρµ = Lσ ρ gσν γσ . (10.244)
from which, using gρµ γ µ = γρ , follows
The construction of S(Lη ξ ) will proceed using the avenue of infinitesimal transformations. We had
introduced in (10.38) the infinitesimal Lorentz transformations in the form Lµ ν = δ µ ν + µ ν where
the infinitesimal operator µ ν obeyed T = − gg. Multiplication of this property by g from the
right yields (g)T = − g, i.e., g is an anti-symmetric matrix. The elements of g are, however,
µ ρ g ρν = µν and, hence, in the expression of the infinitesimal transformation
from which we can conclude σµν µν = −σµν νµ = −σνµ µν , i.e., it must hold
One can readily show expanding SS −1 = 11 to first order in µν that for the inverse infinitesimal
transformation holds
i
S −1 (µν ) = 11 + σµν µν (10.251)
4
Inserting (10.248, 10.251) into (10.246) results then in a condition for the generators σµν
i
− ( σαβ γ µ − γ µ σαβ ) αβ = νµ γν . (10.252)
4
Since six of the coefficients αβ can be chosen independently, this condition can actually be expressed
through six independent conditions. For this purpose one needs to express formally the r.h.s. of
(10.252) also as a sum over both indices of αβ . Furthermore, the expression on the r.h.s., like the
expression on the l.h.s., must be symmetric with respect to interchange of the indices α and β. For
this purpose we express
1 αµ 1 1 1
νµ γν = γα + βµ γβ = αβ δ µ β γα + αβ δ µ α γβ
2 2 2 2
1 αβ µ
= ( δ β γα − δ µ α γβ ) . (10.253)
2
Comparing this with the l.h.s. of (10.252) results in the condition for each α, β
[ σαβ , γ µ ]− = 2i ( δ µ β γα − δ µ α γβ ) . (10.254)
The proper σαβ must be anti-symmetric in the indices α, β and operate in the same space as the
Dirac matrices. In fact, a solution of condition (10.254) is
i
σαβ = [ γα , γβ ]− (10.255)
2
which can be demonstrated using the properties (10.213, 10.216) of the Dirac matrices.
Exercise 7.5:
Show that the σαβ defined through (10.255) satisfy condition (10.254).
σj
0 11 0
γ̃0 = ; γ̃j = , j = 1, 2, 3 . (10.256)
11 0 −σ j 0
322 Relativistic Quantum Mechanics
One can readily verify that the non-vanishing generators σ̃µν are given by (note σ̃µν = −σ̃νµ , i.e.
only six generators need to be determined)
−iσ j 0
`
i σ 0
σ̃0j = [γ̃0 , γ̃j ] = ; σ̃jk = [γ̃j , γ̃k ] = jk` . (10.257)
2 0 iσ j 0 σ`
Obviously, the algebra of these generators is closed under addition and multiplication, since both
operations convert block-diagonal operators
A 0
(10.258)
0 B
again into block-diagonal operators, and since the algebra of the Pauli matrices is closed.
We can finally note that the closedness of the algebra of the generators σµν is not affected by
similarity transformations and that, therefore, any representation of the generators, in particular,
the representation (10.255) yields a closed algebra.
S −1 = γ 0 S † γ 0 . (10.268)
We will prove this property in the chiral representation. Obviously, the property applies then in
any representation of S.
For our proof we note first
!
− 21 (w ~ σ
~ − iϑ)·~
e 0
S̃ −1 (w, ~ = S̃(−w,
~ ϑ) ~ −ϑ)~ = 1 ~ (10.269)
0 e 2 (w~ + iϑ)·~σ
One can readily show that the same operator is obtained evaluating
!
1
(w ~ σ
~ + iϑ)·~
0 † ~ 0 0 1
1 e 2 0 0 11
γ̃ S̃ (w,
~ ϑ)γ̃ = 1 ~ . (10.270)
11 0 0 e− 2 (w~ − iϑ)·~σ 11 0
in a moving frame of reference and the flux j µ in a frame at rest. Note that we have assumed
in (10.271) that the Dirac matrices are independent of the frame of reference. One obtains using
(10.268)
µ
j 0 = Ψ† (xµ )S † γ 0 γ µ SΨ(xµ ) = Ψ† (xµ )γ 0 S −1 γ µ SΨ(xµ ) . (10.272)
With S −1 (Lη ξ ) = S((L−1 )η ξ ) one can restate (10.246)
µ
S −1 (Lη ξ )γ µ S ( Lη ξ ) = (L−1 )νµ γν = (L−1 )ν γ ν = Lµ ν γν . (10.273)
where we have employed (10.76). Combining this with (10.272) results in the expected transfor-
mation behaviour
µ
j 0 = Lµ ν j ν . (10.274)
The free particle wave function is an eigenfunction of Ho , a property which leads to the energy–
momentum (dispersion) relationship of the Dirac particle. The additional degrees of freedom de-
scribed by the four components of the bispinor wave function require, as just mentioned, additional
characterizations, i.e., the identification of observables and their quantum mechanical operators, of
which the wave functions are eigenfunctions as well. As it turns out, only two degrees of freedom
of the bispinor four degrees of freedom are independent [c.f. (10.282, 10.283)]. The independent
degrees of freedom allow one to choose the states of the free Dirac particle as eigenstates of the
4-momentum operator p̂µ and of the helicity operator Γ ∼ ~σ · p ˆ~/|p
ˆ~| introduced below. These opera-
tors, as is required for the mentioned property, commute with each other. The operators commute
also with Ho in (10.233).
Like for the free particle wave functions of the non-relativistic Schrödinger and the Klein–Gordon
equations one expects that the space–time dependence is governed by a factor exp[i(~ p · ~r − t)].
As pointed out, the Dirac particles are described by 4-dimensional, bispinor wave functions and we
need to determine corresponding components of the wave function. For this purpose we consider
10.9: Solutions of the Free Particle Dirac Equation 325
φ(xµ )
φo
µ
Ψ(x ) = = ei(~p·~r − t) (10.276)
χ(xµ ) χo
where p~ and together represent four real constants, later to be identified with momentum and
energy, and φo , χo each represent a constant, two-dimensional spinor state. Inserting (10.276) into
(10.231, 10.232) leads to the 4-dimensional eigenvalue problem
m ~σ · p~ φo φo
= . (10.277)
~σ · p~ −m χo χo
( − m) 11 φo − ~σ · p~ χo = 0
− ~σ · p~ φo + ( + m) 11 χo = 0. (10.278)
Multiplication of the 1st equation by ( + m)11 and of the second equation by −~σ · p~ and subtraction
of the results yields the 2-dimensional equation
2
( − m2 ) 11 − (~σ · p~)2 φo = 0 .
(10.279)
According to the property (5.234) of Pauli matrices holds (~σ · p~)2 = p~ 2 11. One can, hence, conclude
from (10.279) the well-known relativistic dispersion relationship
2 = m2 + p~ 2 (10.280)
Obviously, the Dirac equation, like the Klein–Gordon equation, reproduce the classical relativistic
energy–momentum relationships (10.93, 10.94)
Equation (10.278) provides us with information about the components of the bispinor wave function
(10.276), namely φo and χo are related as follows
~σ · p~
φ0 = χ0 (10.282)
−m
~σ · p~
χo = φ0 , (10.283)
+m
where is defined in (10.281). These two relationships are consistent with each other. In fact, one
finds using (5.234) and (10.280)
~σ · p~ (~σ · p~)2 p~ 2
φ0 = χ0 = 2 χ0 = χ0 . (10.284)
+m ( + m) ( + m) − m2
The relationships (10.282, 10.283) imply that the bispinor part of the wave function allows only
two degrees of freedom to be chosen independently. We want to show now that these degrees of
freedom correspond to a spin-like property, the socalled helicity of the particle.
326 Relativistic Quantum Mechanics
For our further characterization we will deal with the positive and negative energy solutions [cf.
(10.281)] separately. For the positive energy solution, i.e., the solution for = +E(~
p), we present
φo through the normalized vector
u1
φo = = u ∈ C2 , u† u = |u1 |2 + |u2 |2 = 1 . (10.285)
u2
The corresponding free Dirac particle is then described through the wave function
!
u
Ψ(~p, +|xµ ) = N+ (~p) σ ·~
~ p ei(~p·~r − t) , = +E(~
p) . (10.286)
p) + m u
E(~
Here N+ (~
p) is a constant which will be chosen to satisfy the normalization condition
Ψ† (~
p, +) γ 0 Ψ† (~
p, +) = 1 , (10.287)
the form of which will be justified further below. Similarly, we present the negative energy solution,
i.e., the solution for = −E(~p), through χo given by
u1
χo = = u ∈ C2 , u† u = |u1 |2 + |u2 |2 = 1 . (10.288)
u2
corresponding to the wave function
!
−~ σ ·~
p
u
p, −|xµ ) = N− (~
Ψ(~ p) E(~
p) + m ei(~p·~r − t) , = −E(~
p) . (10.289)
u
Here N− (~
p) is a constant which will be chosen to satisfy the normalization condition
Ψ† (~
p, +) γ 0 Ψ† (~
p, +) = −1 , (10.290)
which differs from the normalization condition (10.287) in the minus sign on the r.h.s. The form
of this condition and of (10.287) will be justified now.
First, we demonstrate that the product Ψ† (~ p, ±)γ 0 Ψ(~p, ±), i.e., the l.h.s. of (10.287, 10.290), is
invariant under Lorentz transformations. One can see this as follows: Let Ψ(~ p, ±) denote the
solution of a free particle moving with momentum p~ in the laboratory frame, and let Ψ(0, ±)
denote the corresponding solution of a particle in its rest frame. The connection between the
solutions, according to (10.241), is Ψ(~p, ±) = S Ψ(0, ±) , where S is given by (10.259). Hence,
Ψ† (~
p, ±) γ 0 Ψ(~
p, ±) = Ψ† (0, ±) S † γ 0 S Ψ(0, ±)
= Ψ† (0, ±) γ 0 S −1 γ 0 γ 0 S Ψ(0, ±)
The same calculation for the negative energy wave function as given in (10.289) yields
2 0
† 0
Ψ (0, −) γ Ψ(0, −) = |N− (0)| (0, u ) γ† 0
= − |N− (0)|2 . (10.293)
u
Obviously, this requires the choice of a negative side on the r.h.s. of (10.290) to assign a positive
value to |N− (0)|2 . We can also conclude from our derivation
N± (0) = 1 . (10.294)
We want to determine now N± (~ p) for arbitrary p~. We consider first the positive energy solution.
Condition (10.287) written explicitly using (10.286) is
T ! !
u
~
σ · p
~
p) (u∗ )T ,
N+2 (~ u∗ γo σ ·~
~ p = 1 (10.295)
E(~
p) + m p) + m u
E(~
(~σ · p~)2
2 † †
N+ (~p) u u − u u = 1. (10.296)
(E(~p) + m)2
Replacing (~σ · p~)2 by p~ 2 [c.f. (5.234)] and using the normalization of u in (10.285) results in
p~ 2
2
N+ (~
p) 1 − = 1 (10.297)
(E(~p) + m)2
from which follows s
( m + E(~p) )2
N+ (~
p) = . (10.298)
p) )2 − p~ 2
( m + E(~
Noting
p) )2 − p~ 2 = m2 − p~ 2 + 2mE(~
( m + E(~ p) + E 2 (~
p) = 2(m + E(~
p)) m (10.299)
We consider now the negative energy solution. Condition (10.290) written explicitly using (10.289)
is T ! !
−~σ ·~
p
2 −~σ · p~ ∗ ∗ T o E(~
p) + m u
N− (~
p) u , (u ) γ = −1 (10.303)
E(~p) + m u
Evaluating the l.h.s. yields
(~σ · p~)2
† †
N−2 (~
p) u u − u u = −1 . (10.304)
p) + m)2
(E(~
This condition is, however, identical to the condition (10.296) for the normalization constant N+ (~
p)
of the positive energy solution. We can, hence, conclude
r
m + E(~ p)
N− (~
p) = (10.305)
2m
and, thereby, have completed the determination for the wave function (10.289).
The wave functions (10.286, 10.289, 10.300) have been constructed to satisfy the free particle Dirac
equation (10.275). Inserting (10.286) into (10.275) yields
p, λ|xµ ) = λ E(~
Ho Ψ(~ p, λ|xµ ) ,
p) Ψ(~ (10.306)
i.e., the wave functions constructed represent eigenstates of Ho . The wave functions are also
eigenstates of the momentum operator i∂µ , i.e.,
p, λ|xµ ) = pµ Ψ(~
i∂µ Ψ(~ p, λ|xµ ) (10.307)
Helicity
The free Dirac particle wave functions (10.286, 10.289) are not completely specified, the two com-
ponents of u indicate another degree of freedom which needs to be defined. This degree of freedom
describes a spin– 21 attribute. This attribute is the so-called helicity, defined as the component of
the particle spin along the direction of motion. The corresponding operator which measures this
observable is
1 ~p̂
Λ = σ· . (10.308)
2 |~p̂|
Note that ~p̂ represents here an operator, not a constant vector. Rather than considering the
observable (10.307) we investigate first the observable due to the simpler operator ~σ · ~p̂. We
want to show that this operator commutes with Ho and ~p̂ to ascertain that the free particle wave
function can be simultaneously an eigenvector of all three operators. The commutation property
[~σ · ~p̂, p̂j ] = 0 , j = 1, 2, 3 is fairly obvious. The property [~σ · ~p̂, Ho ] = 0 follows from (10.233) and
from the two identities
11 0 ~σ 0 ~ ~σ 0 ~ 11 0
· p̂ − · p̂ = 0 (10.309)
0 −11 0 ~σ 0 ~σ 0 −11
10.9: Solutions of the Free Particle Dirac Equation 329
and
0 ~σ ~σ 0 ~σ 0 0 ~
σ
· ~p̂ · ~p̂ − · ~p̂ · ~p̂
~σ 0 0 ~σ 0 ~σ ~σ 0
! !
0 (~σ · ~p̂)2 0 (~σ · ~p̂)2
= − = 0. (10.310)
(~σ · ~p̂)2 0 (~σ · ~p̂)2 0
We have shown altogether that the operators ~p̂, Ho and ~σ ·~p̂ commute with each other and, hence, can
be simultaneously diagonal. States which are simultaneously eigenvectors of these three operators
are also simulteneously eigenvectors of the three operators ~p̂, Ho and Λ defined in (10.308) above.
The condition that the wave functions (10.286) are eigenfunctions of Λ as well will specify now the
vectors u.
Since helicity is defined relative to the direction of motion of a particle the characterization of u
as an eigenvector of the helicity operator, in principle, is independent of the direction of motion
of the particle. We consider first the simplest case that particles move along the x3 –direction, i.e.,
p~ = (0, 0, p3 ). In this case Λ = 21 σ 3 .
We assume first particles with positive energy, i.e., = +E(~ p). According to the definition (5.224)
of σ3 the two u vectors (1, 0)T and (0, 1)T are eigenstates of 12 σ 3 with eigenvalues ± 12 . Therefore,
the wave functions which are eigenstates of the helicity operator, are
1
0 ei(px − Ep t)
3
Ψ(pê3 , +, + 21 |~r, t) = Np
p 1
m + Ep 0
0
1 ei(px − Ep t)
3
Ψ(pê3 , +, − 12 |~r, t) = Np
(10.311)
−p 0
m + Ep 1
where ê3 denotes the unit vector in the x3 -direction and where
r
p
2 2
m + Ep
E p = m + p ; Np = . (10.312)
2m
We assume now particles with negative energy, i.e., = −E(~ p). The wave functions which are
eigenfunctions of the helicity operator are in this case
−p 1
m + Ep
0 ei(px + Ep t)
3
Ψ(pê3 , −, + 21 |~r, t) = Np
1
0
p 0
m + Ep
1 ei(px + Ep t)
3
Ψ(pê3 , −, − 12 |~r, t) = Np
(10.313)
0
1
To obtain free particle wave functions for arbitrary directions of p~ one can employ the wave functions
(10.311, 10.313) except that the states (1, 0)T and (0, 1)T have to be replaced by eigenstates u± (~ p)
of the spin operator along the direction of p~. These eigenstates are obtained through a rotational
transformation (5.222, 5.223) as follows
1 i ~ 1
u(~p, + ) = exp − ϑ(~ p) · ~σ , (10.314)
2 2 0
1 i ~ 0
u(~p, − ) = exp − ϑ(~ p) · ~σ (10.315)
2 2 1
where
~ p) = ê3 × p~ ∠(ê3 , p~)
ϑ(~ (10.316)
|~
p|
describes a rotation which aligns the x3 –axis with the direction of p~. [One can also express the
rotation through Euler angles α, β, γ , in which case the transformation is given by (5.220).] The
corresponding free particle wave functions are then
p, + 21 )
!
u(~
Ψ(~ 1
p, +, + 2 |~r, t) = Np p 1
ei(~p·~r − Ep t) (10.317)
m + Ep u(~
p , + 2 )
p, − 12 )
!
u(~
Ψ(~ 1
p, +, − 2 |~r, t) = Np −p 1
ei(~p·~r − Ep t) (10.318)
m + Ep u(~
p , − 2 )
−p 1
!
m + E u(~
p , + 2 )
Ψ(~p, −, + 12 |~r, t) = N√ p
ei(~p·~r + Ep t) (10.319)
p, + 12 )
u(~
p
p, − 12 )
!
m + Ep u(~
Ψ(~ 1
p, −, − 2 |~r, t) = Np 1
ei(~p·~r + Ep t) (10.320)
p, − 2 )
u(~
iγ̃ 0 ∂t − m Ψ̃(t) = 0 .
(10.321)
and are
1
0 −imt
Ψ̃(p = 0, +, 21 |t) = √12
1 e
,
0
10.9: Solutions of the Free Particle Dirac Equation 331
0
1 −imt
Ψ̃(p = 0, +, − 12 |t) = √12
0 e
,
1
1
0 +imt
Ψ̃(p = 0, −, 12 |t) = √12
−1 e
,
0
0
1 +imt
Ψ̃(p = 0, −, − 12 |t) = √12
0 e
. (10.322)
−1
The reader can readily verify that transformation of these solutions to the Dirac representationsas
defined in (10.228) yields the corresponding solutions (10.311, 10.313) in the p → 0 limit. This
correspondence justifies the characterization ±, ± 12 of the wave functions stated in (10.322).
The solutions (10.322) can be written in spinor form
1 φo ∓imt 1 0
√ e , φo , χo ∈ , (10.323)
2 χo 0 1
∓imt → i ( p3 x3 ∓ Et ) (10.324)
One should note that φo , χo are eigenstates of σ 3 with eigenvalues ±1. Applying (10.324, 10.325)
to (10.323) should yield the solutions for non-vanishing momentum p in the x3 –direction. For the
resulting wave functions in the chiral representation one can use then a notation corresponding to
that adopted in (10.311)
1
w 1
e 2 3
1 0 3
Ψ̃(p(w3 )ê3 , +, + 21 |~r, t) ei(px − Ep t)
= √
2 1 1
e− 2 w3
0
1 0
e− 2 w3
1 1 3
Ψ̃(p(w3 )ê3 , +, − 21 |~r, t) ei(px − Ep t)
= √
2 1 0
e 2 w3
1
332 Relativistic Quantum Mechanics
1
w 1
e2 3
1 0 3
Ψ̃(p(w3 )ê3 , −, + 12 |~r, t) ei(px + Ep t)
= √
2 1 1
−e− 2 w3
0
1 0
e− 2 w3
1 1 3
Ψ̃(p(w3 )ê3 , −, − 12 |~r, t) ei(px + Ep t)
= √ (10.326)
2 1 0
− e 2 w3
1
and similarly r
w3 Ep − m p
sinh = = p (10.330)
2 2m 2m (Ep + m)
Inserting expressions (10.329, 10.330) into (10.327), indeed, reproduces the positive energy wave
functions (10.311) as well as the negative energy solutions (10.313) for −p. The change of sign for
the latter solutions had to be expected as it was already noted for the negative energy solutions of
the Klein–Gordon equation (10.168–10.176).
where S(Lη ξ ) denotes again the transformation acting on the bispinor character of the wave function
Ψ(xµ ) and where ρ(Lη ξ ) denotes the transformation acting on the space-time character of the
wave function Ψ(xµ ). ρ(Lη ξ ) has been defined in (10.123) above and characterized there. Such
transformation had been applied by us, of course, when we generated the solutions Ψ(~ p, λ, Λ|xµ )
from the solutions describing particles at rest Ψ(~ p = 0, λ, Λ|t). We expect, in general, that if Ψ(xµ )
0 µ
is a solution of the Dirac equation that Ψ (x ) as given in (10.331) is a solution as well. Making
this expectation a postulate allows one to derive the condition (10.243) and, thereby, the proper
transformation S(Lη ξ ).
To show this we rewrite the Dirac equation (10.221) using (10.331)
Here we have made use of the fact that S(Lη ξ ) commutes with ∂µ and ρ(Lη ξ ) commutes with γ µ .
The fact that any such Ψ0 (xµ ) is a solution of the Dirac equation allows us to conclude
We will demonstrate now that the first condition is satisfied by ρ(Lη ξ ). The second condition is
identical to (10.243) and, of course, it is met by S(Lη ξ ) as given in the chiral representation by
(10.262).
As mentioned already we will show condition (10.334) for infinitesimal Lorentz transformations
Lη ξ . We will proceed by employing the generators (10.128) to express ρ(Lη ξ ) in its infinitesimal
form and evaluate the expression
~ · J~ + w
11 + ϑ ~ ∂µ 11 − ϑ
~ ·K ~ · J~ − w
~ ·K~
= ∂µ + M ν µ ∂ν + O(2 ) (10.335)
The result will show that the matrix M ν µ is identical to the generators of Lν µ for the six choices ϑ ~ =
(1, 0, 0), w ~
~ = (0, 0, 0), ϑ = (0, 1, 0), w
~ = (0, 0, 0), . . . , ϑ = (0, 0, 0), w
~ = (0, 0, 1) . Inspection of
(10.335) shows that we need to demonstrate
[J` , ∂µ ] = (J` )ν µ ∂ν ; [K` , ∂µ ] = (K` )ν µ ∂ν . (10.336)
We will proceed with this task considering all six cases:
0 µ = 0
0 µ = 1
[J1 , ∂µ ] = [x3 ∂2 − x2 ∂3 , ∂µ ] = (10.337)
∂3 µ = 2
−∂2 µ = 3
0 µ = 0
−∂3 µ = 1
[J2 , ∂µ ] = [x1 ∂3 − x3 ∂1 , ∂µ ] = (10.338)
0 µ = 2
∂1 µ = 3
0 µ = 0
∂2 µ = 1
[J3 , ∂µ ] = [x2 ∂1 − x1 ∂2 , ∂µ ] = (10.339)
−∂1 µ = 2
0 µ = 3
−∂1 µ = 0
−∂0 µ = 1
[K1 , ∂µ ] = [x0 ∂1 + x1 ∂0 , ∂µ ] = (10.340)
0 µ = 2
0 µ = 3
−∂2 µ = 0
0 µ = 1
[K2 , ∂µ ] = [x0 ∂2 + x2 ∂0 , ∂µ ] = (10.341)
−∂0
µ = 2
0 µ = 3
−∂3 µ = 0
0 µ = 1
[K3 , ∂µ ] = [x0 ∂3 + x3 ∂0 , ∂µ ] = (10.342)
0 µ = 2
−∂0 µ = 3
One can readily convince oneself that these results are consistent with (10.336). We have demon-
strated, therefore, that any solution Ψ(xµ ) transformed according to (10.331) is again a solution of
the Dirac equation, i.e., the Dirac equation is invariant under active Lorentz transformations.
10.10: Dirac Particles in Electromagnetic Field 335
One may also include the electromagnetic field in the Dirac equation given in the Schrödinger form
(10.233) by replacing i∂t by (see Table 10.1) i∂t − qV and pˆ~ by
ˆ = p
~π ˆ~ − q A
~. (10.344)
Non-Relativistic Limit
We want to consider now the Dirac equation (10.345) in the so-called non-relativistic limit in which
all energies are much smaller than m, e.g., for the scalar field V in (10.345) holds
φ(xµ )
µ
Ψ(x ) = . (10.347)
χ(xµ )
i∂t φ = ˆ χ + qV φ + mφ
~σ · ~π (10.348)
i∂t χ = ˆ φ + qV χ − mχ .
~σ · ~π (10.349)
We want to focus on the stationary positive energy solution. This solution exhibits a time-
dependence exp[−i(m + )t] where for holds in the non-relativistic limit || << m. Accordingly,
we define
i∂t Φ = ˆ X + qV Φ
~σ · ~π (10.353)
i∂t X = ˆ Φ + qV X − 2mX .
~σ · ~π (10.354)
ˆ Φ − 2m X .
0 ≈ ~σ · ~π (10.355)
ˆ
~σ · ~π
X ≈ Φ (10.356)
2m
to obtain a closed equation for Φ
2
ˆ
~σ · ~π
i∂t Φ ≈ Φ + qV Φ . (10.357)
2m
Equation (10.356), due to the m−1 factor, identifies X as the small component of the bi-spinor
wave function which, henceforth, does not need to be considered anymore.
Equation (10.357) for Φ can be reformulated by expansion of (~σ · ~πˆ )2 . For this purpose we employ
the identity (5.230), derived in Sect. 5.7, which in the present case states
(~σ · ~π ˆ 2 + i ~σ · (~π
ˆ )2 = ~π ˆ × ~π
ˆ) . (10.358)
ˆ × ~π
For the components of ~π ˆ holds
ˆ × ~π
~π ˆ = jk` ( πj πk − πk πj ) = jk` [πj , π` ] . (10.359)
`
Using
∂j Ak f = ((∂j Ak )) f + Ak ∂j f
∂k Aj f = ((∂k Aj )) f + Aj ∂k f
where ((∂j · · ·)) denotes confinement of the differential operator to within the brackets ((· · ·)), one
obtains
( [Aj , ∂k ] + [∂j , Ak ] ) f = [ ((∂j Ak )) − ((∂k Aj )) ] f (10.362)
~
or, using (10.360) and Aµ = (V, −A),
q q ~ jk` = − q B` jk`
[πj , πk ] = (( ∂j Ak − ∂k Aj )) = − ∇×A (10.363)
i i ` i
where we employed B(~ ~ r, t) = ∇ × A(~ ~ r, t) [see (8.6)]. Equations (10.344, 10.358, 10.359, 10.363)
allow us to write (10.357) in the final form
" #
ˆ~ − q A(~
[p ~ r, t)]2 q
i∂t Φ(~r, t) ≈ − ~ r, t) + q V (~r, t) Φ(~r, t)
~σ · B(~ (10.364)
2m 2m
Comparision of this expression with (5.222, 5.223) shows that the propagator in (10.366) can be
~ by an angle qtB, i.e., the interaction q~σ · B
interpreted as a rotation around the field B ~ induces a
1
precession of the spin- 2 around the magnetic field.
For the purpose of the solution we assume the chiral representation, i.e, we solve
where Ψ̃(xµ ) and γ̃ µ are defined in (10.228) and in (10.229), respectively. Employing πµ as defined
in Table 10.1 one can write (10.368)
( γ̃ µ πµ − m ) Ψ̃(xµ ) = 0 . (10.369)
For our solution we will adopt presently a strategy which follows closely that for the spectrum of
pionic atoms in Sect. 10.6. For this purpose we ‘square’ the Dirac equation, multiplying (10.369)
from the left by γ ν πν + m. This yields
Any solution of (10.368) is also a solution of (10.370), but the converse is not necessarily true.
However, once a solution Ψ̃(xµ ) of (10.370) is obtained then
3
X X
µ ν
γ̃ π̂µ γ̃ π̂ν = (γ̃ µ )2 π̂µ2 + γ̃ µ γ̃ ν π̂µ π̂ν . (10.374)
µ=0 µ,ν=1
µ 6=ν
The first term on the r.h.s. can be rewritten using, according to (10.230), (γ̃ 0 )2 = 11 and (γ̃ j )2 =
−11, j = 1, 2, 3,
3
ˆ2 .
X
(γ̃ µ )2 π̂µ2 = π̂02 − ~π (10.375)
µ=0
Following the algebra that connected Eqs. (5.231), (5.232) in Sect. 5.7 one can write the second
term in (10.374), noting from (10.230) γ̃ µ γ̃ ν = − γ̃ ν γ̃ µ , µ 6= ν and altering ‘dummy’ summation
10.10: Dirac Particles in Electromagnetic Field 339
indices,
X 1 X
γ̃ µ γ̃ ν π̂µ π̂ν = ( γ̃ µ γ̃ ν π̂µ π̂ν + γ̃ ν γ̃ µ π̂ν π̂µ )
2
µ,ν=1 µ,ν=1
µ 6=ν µ 6=ν
1 X
= ( γ̃ µ γ̃ ν π̂µ π̂ν − γ̃ ν γ̃ µ π̂µ π̂ν + γ̃ ν γ̃ µ π̂ν π̂µ − γ̃ µ γ̃ ν π̂ν π̂µ )
4
µ,ν=1
µ 6=ν
1 X µ ν
= [γ̃ , γ̃ ] [π̂µ , π̂ν ] (10.376)
4
µ,ν=1
µ 6=ν
~ = 0. Since
This expression can be simplified due to the special form (10.367) of Aµ , i.e., due to A
where f = f (~r, t) is a suitable test function and where ((· · ·)) denotes the range to which the
derivative is limited. Altogether, one can summarize (10.376–10.380)
X
µ ν ~σ 0
γ̃ γ̃ π̂µ π̂ν = i · ((∇qA0 )) (10.381)
0 −~σ
µ,ν=1
µ 6=ν
where r̂ = ~r/|~r| is a unit vector. Combining this result with (10.381), (10.374), (10.375) the
‘squared’ Dirac equation (10.368) reads
" 2 #
Ze2 Ze2
2 ~σ · r̂ 0 2
− ∂t − i + ∇ + i − m Ψ(xµ ) = 0 (10.383)
r 0 −~σ · r̂ r2
We seek stationary solutions of this equation. Such solutions are of the form
can be interpreted as the energy of the stationary state and, hence, it is this quantity that we
want to determine. Insertion of (10.384) into (10.383) yields the purely spatial four-dimensional
differential equation
" 2 #
Ze2 Ze2
2 ~σ · r̂ 0 2
+ + ∇ + i − m Φ(~r) = 0 . (10.385)
r 0 −~σ · r̂ r2
The expression (10.189) for the Laplacian and expansion of the term (· · ·)2 result in the two-
dimensional equation
" #
2 L̂2 − Z 2 e4 ∓ i~σ · r̂ Ze2 2Ze2 2 2
∂r − + + − m r φ± (~r) = 0 . (10.388)
r2 r
Except for the term i~σ · r̂ this equation is identical to that posed by the one-dimensional Klein-
Gordon equation for pionic atoms (10.191) solved in Sect. 10.6. In the latter case, a solution of
the form ∼ Y`m (r̂) can be obtained. The term i~σ · r̂, however, is genuinely two-dimensional and, in
fact, couples the orbital angular momentum of the electron to its spin- 12 . Accordingly, we express
the solution of (10.388) in terms of states introduced in Sect. 6.5 which describe the coupling of
orbital angular momentum and spin
According to the results in Sect. 6.5 the operator i~σ · r̂ is block-diagonal in this basis such that only
the states for identical j, m values are coupled, i.e., only the two states {Yjm (j − 12 , 12 |r̂), Yjm (j +
, |r̂)} as given in (6.147, 6.148). We note that these states are also eigenstates of the angular
1 1
2 2
10.10: Dirac Particles in Electromagnetic Field 341
momentum operator L̂2 [cf. (6.151]. We select, therefore, a specific pair of total spin-orbital angular
momentum quantum numbers j, m and expand
h± (r) g± (r)
φ± (~r) = Yjm (j − 12 , 12 |r̂) + Yjm (j + 12 , 12 |r̂) (10.390)
r r
Using
~σ · r̂ Yjm (j ± 12 , 12 |r̂) = − Yjm (j ∓ 12 , 12 |r̂) (10.391)
derived in Sect. 6.5 [c.f. (6.186)], property (6.151), which states that the states Yjm (j ± 21 , 12 |r̂) are
eigenfunctions of L̂2 , together with the orthonormality of these two states leads to the coupled
differential equation
" !
2 1 0
2Ze
∂r2 + + 2 − m (10.392)
r 0 1
!#
(j − 12 )(j + 21 ) − Z 2 e4 ±iZe2
1 h± (r)
− 2 = 0.
r ±i Ze2 (j + 21 )(j + 32 ) − Z 2 e4 g± (r)
We seek to bring (10.392) into diagonal form. Any similarity transformation leaves the first term in
(10.392), involving the 2 × 2 unit matrix, unaltered. However, such transformation can be chosen
as to diagonalize the second term. Since, in the present treatment, we want to determine solely the
spectrum, not the wave functions, we require only the eigenvalues of the matrices
!
(j − 12 )(j + 12 ) − Z 2 e4 ±iZe2
B± = , (10.393)
±i Ze2 (j + 12 )(j + 32 ) − Z 2 e4
but do not explicitly consider further the wavefunctions. Obviously, the eigenvalues are independent
of m. The two eigenvalues of both matrices are identical and can be written in the form
where
q
λ1 (j) = (j + 12 )2 − Z 2 e4 (10.395)
q
λ2 (j) = (j + 12 )2 − Z 2 e4 − 1 (10.396)
2Ze2
2 λ1,2 (j)[λ1,2 (j) + 1) 2 2
∂r − + + −m f1,2 (r) = 0 (10.397)
r2 r
This equation is identical to the Klein-Gordon equation for pionic atoms written in the form
(10.200), except for the slight difference in the expression of λ1,2 (j) as given by (10.395, 10.396)
and (10.199), namely, the missing additive term − 12 , the values of the argument of λ1,2 (j) being
j = 21 , 32 , . . . rather than ` = 0, 1, . . . as in the case of pionic atoms, and except for the fact that
we have two sets of values for λ1,2 (j), namely, λ1 (j) and λ2 (j).. We can, hence, conclude that the
342 Relativistic Quantum Mechanics
spectrum of (10.397) is again given by eq. (10.203), albeit with some modifications. Using (10.395,
10.396) we obtain, accordingly,
m
1 = r
2 4
; (10.398)
1 + qZ e
( n0 + 1 + (j+ 12 )2 − Z 2 e4 )2
m
2 = r
2 4
; (10.399)
1 + qZ e
( n0 + (j+ 12 )2 − Z 2 e4 )2
0
n = 0, 1, 2, . . . , j = 12 , 32 , . . . , m = −j, −j + 1, . . . , j
where 1 corresponds to λ1 (j) as given in (10.395) and 2 corresponds to λ2 (j) as given in (10.396).
For a given value of n0 the energies 1 and 2 for identical j-values correspond to mixtures of states
with orbital angular momentum ` = j − 21 and ` = j + 12 . The magnitude of the relativistic effect
is determined by Z 2 e4 . Expanding the energies in terms of this parameter allows one to identify
the relationship between the energies 1 and 2 and the non-relativistic spectrum. One obtains in
case of (10.398, 10.399)
mZ 2 e4
1 ≈ m − + O(Z 4 e8 ) (10.400)
2 (n0 + j + 32 )2
mZ 2 e4
2 ≈ m − + O(Z 4 e8 ) (10.401)
2 (n0 + j + 12 )2
n0 = 0, 1, 2, . . . , j = 12 , 32 , . . . , m = −j, −j + 1, . . . , j .
These expressions can be equated with the non-relativistic spectrum. Obviously, the second term on
the r.h.s. of these equations describe the binding energy. In case of non-relativistic hydrogen-type
atoms, including spin- 21 , the stationary states have binding energies
mZ 2 e4 n = 1, 2, . . . ` = 0, 1, . . . , n − 1
E = − , . (10.402)
2n2 m = −`, −` + 1, . . . , ` ms = ± 12
n = 1, 2, . . . ; ` = 0, 1, . . . , n − 1 ; m = −j, −j + 1, . . . , j
10.10: Dirac Particles in Electromagnetic Field 343
Table 10.2:
Binding energies for the hydrogen (Z = 1) atom. Degeneracies are denoted by ⇑. The energies
were evaluated with m = 511.0041 keV and e2 = 1/137.036 by means of Eqs. (10.402, 10.405).
ED (n, `, j, m) =
One can combine the expressions (10.403, 10.404) finally into the single formula m
r
Z 2 e4
1 + q
(n−j − 12 + (j+ 12 )2 −
n = 1, 2, . . .
` = 0, 1,1 . . . , n − 1
2
for ` = 0
j =
` ± 2 otherwise
1
m = −j, −j + 1, . . . , j
(10.405)
In order to demonstrate relativistic effects in the spectrum of the hydrogen atom we compare
in Table 10.2 the non-relativistic [cf. (10.402)] and the relativistic [cf. (10.405)] spectrum of the
hydrogen atom. The table entries demonstrate that the energies as given by the expression (10.405)
in terms of the non-relativistic quantum numbers n, ` relate closely to the corresponding non-
relativistic states, in fact, the non-relativistic and relativistic energies are hardly discernible. The
reason is that the mean kinetic energy of the electron in the hydrogen atom, is in the range of
10 eV, i.e., much less than the rest mass of the electron (511 keV). However, in case of heavier
nuclei the kinetic energy of bound electrons in the ground state scales with the nuclear charge Z
like Z 2 such that in case Z = 100 the kinetic energy is of the order of the rest mass and relativistic
effects become important. This is clearly demonstrated by the comparision of non-relativistic and
relativistic spectra of a hydrogen-type atom with Z = 100 in Table 10.3.
344 Relativistic Quantum Mechanics
Table 10.3:
Binding energies for the hydrogen-type (Z = 100) atom. Degeneracies are denoted by ⇑. The
energies were evaluated with m = 511.0041 keV and e2 = 1/137.036 by means of Eqs. (10.402,
10.405).
Of particular interest is the effect of spin-orbit coupling which removes, for example, the non-
relativistic degeneracy for the six 2p states of the hydrogen atom: in the present, i.e., relativistic,
case these six states are split into energetically different 2p 1 and 2p 3 states. The 2p 1 states with
2 2 2
1
j = 2 involve two degenerate states corresponding to Y 1 m (1, 12 |r̂) for m = ± 12 , the 2p 3 states with
2 2
j = 3
2
involve four degenerate states corresponding to Y 3 m (1, 12 |r̂) for m = ± 12 , ± 32 .
2
In order to investigate further the deviation between relativistic and non-relativistic spectra of
hydrogen-type atoms we expand the expression (10.405) to order O(Z 4 e8 ). Introducing α = Z 2 e4
and β = j + 21 (10.405) reads
1
q α
(10.406)
1+ √
(n − β + β 2 −α)2
This expression allows one, for example, to estimate the difference between the energies of the
states 2p 3 and 2p 1 (cf. Tables 10.2,10.3). It holds for n = 2 and j = 32 , 12
2 2
mZ 4 e8 mZ 4 e8
1
E (2p 3 ) − E (2p 1 ) ≈ − −1 = . (10.408)
2 2 2 · 23 2 32
10.10: Dirac Particles in Electromagnetic Field 345
We want to determine now the wave functions for the stationary states of a Dirac particle in a
4-vector potential
Aµ = (V (r), 0, 0, 0) (10.409)
where V (r) is spherically symmetric. An example for such potential is the Coulomb potential
V (r) = −Ze2 /r considered further below. We assume for the wave function the stationary state
form
µ −it Φ(~r)
Ψ(x ) = e , (10.410)
X (~r)
where Φ(~r), X (~r) ∈ C2 describe the spatial and spin- 21 degrees of freedom, but are time-independent.
The Dirac equation reads then, according to (10.232, 10.345),
In this equation a coupling between the wave functions Φ(~r) and X (~r) arises due to the term ~σ · p ˆ~.
This term has been discussed in detail in Sect. 6.5 [see, in particular, pp. 168]: the term is a scalar
(rank zero tensor) in the space of the spin-angular momentum states Yjm (j ± 12 , 12 |r̂) introduced in
Sect. 6.5, i.e., the term is block-diagonal in the space spanned by the states Yjm (j ± 12 , 12 |r̂) and does
not couple states with different j, m-values; ~σ · pˆ~ has odd parity and it holds [c.f. (6.197, 6.198)]
" #
3
ˆ~ f (r) Yjm (j + 1 , 1 |r̂) j+ 2
~σ · p 2 2
= i ∂r + f (r) Yjm (j − 12 , 12 |r̂)
r
(10.413)
" #
1
ˆ~ g(r) Yjm (j − 1 , 1 |r̂) 2 −j
~σ · p 2 2
= i ∂r + g(r) Yjm (j + 12 , 12 |r̂) .
r
(10.414)
1 1
∂r + = ∂r r
r r
The differential equations (10.411, 10.412) are four-dimensional with ~r-dependent wave functions.
The arguments above allow one to eliminate the angular dependence by expanding Φ(~r) and X (~r)
in terms of Yjm (j + 12 , 12 |r̂) and Yjm (j − 12 , 12 |r̂), i.e.,
a(r) b(r)
Φ(~r) Yjm (j + 2 , 2 |r̂) +
1 1
Yjm (j − 2 , 2 |r̂)
1 1
= r r
. (10.417)
c(r) d(r)
X (~r)
Yjm (j + 12 , 12 |r̂) + Yjm (j − 12 , 12 |r̂)
r r
In general, such expansion must include states with all possible j, m values. Presently, we consider
the case that only states for one specific j, m pair contribute. Inserting (10.417) into (10.411,
10.412), using (10.415, 10.416), the orthonormality property (6.157), and multiplying by r results
in the following two independent pairs of coupled differential equations
" #
j + 21
i ∂r − d(r) + [ m + V (r) − ] a(r) = 0
r
" #
j + 21
i ∂r + a(r) + [ −m + V (r) − ] d(r) = 0 (10.418)
r
and
" #
j + 12
i ∂r + c(r) + [ m + V (r) − ] b(r) = 0
r
" #
j + 12
i ∂r − b(r) + [ −m + V (r) − ] c(r) = 0. (10.419)
r
Obviously, only a(r), d(r) are coupled and b(r), c(r) are coupled. Accordingly, there exist two
independent solutions (10.417) of the form
f1 (r)
i r Yjm (j + 2 , 2 |r̂)
1 1
Φ(~r)
= (10.420)
X (~r)
g1 (r)
− Yjm (j − 2 , 2 |r̂)
1 1
r
f2 (r)
i r Yjm (j − 2 , 2 |r̂)
! 1 1
Φ(~r)
= (10.421)
X (~r) g2 (r)
− Yjm (j + 2 , 2 |r̂)
1 1
r
where the factors i and −1 have been introduced for convenience. According to (10.418) holds for
f1 (r), g1 (r)
" #
j + 21
∂r − g1 (r) + [ − m − V (r) ] f1 (r) = 0
r
" #
j + 12
∂r + f1 (r) − [ + m − V (r) ] g1 (r) = 0 (10.422)
r
10.10: Dirac Particles in Electromagnetic Field 347
" #
1
j+ 2
∂r + g2 (r) + [ − m − V (r) ] f2 (r) = 0
r
" #
1
j+ 2
∂r − f2 (r) − [ + m − V (r) ] g2 (r) = 0 (10.423)
r
Equations (10.422) and (10.423) are identical, except for the opposite sign of the term (j + 21 ); the
equations determine, together with the appropriate boundary conditions at r = 0 and r → ∞,
the radial wave functions for Dirac particles in the potential (10.409).
We want to determine now the wave functions of the stationary states of hydrogen-type atoms
which correspond to the energy levels (10.405). We assume the 4-vector potential of pure Coulomb
type (10.367) which is spherically symmetric such that equations (10.422, 10.423) apply for V (r) =
−Ze2 /r. Equation (10.422) determines solutions of the form (10.420). In the non-relativistic limit,
Φ in (10.420) is the large component and X is the small component. Hence, (10.422) corresponds
to states
f1 (r)
i Yjm (j + 2 , 2 |r̂)
1 1
Ψ(xµ ) ≈ r , (10.424)
0
i.e., to states with angular momentum ` = j + 12 . According to the discussion of the spectrum
(10.405) of the relativistic hydrogen atom the corresponding states have quantum numbers n =
1, 2, . . . , ` = 0, 1, . . . , n − 1. Hence, (10.422) describes the states 2p 1 , 3p 1 , 3d 3 , etc. Similarly,
2 2 2
(10.423), determining wave functions of the type (10.421), i.e., in the non-relativistic limit wave
functions
f2 (r)
i Yjm (j − 2 , 2 |r̂)
1 1
Ψ(xµ ) ≈ r , (10.425)
0
Behaviour at r → 0
We consider first the behaviour of the solutions f1 (r) and g1 (r) of (10.422) near r = 0. We note
that (10.422), for small r, can be written
" #
j + 21 Ze2
∂r − g1 (r) + f1 (r) = 0
r r
" #
j + 12 Ze2
∂r + f1 (r) − g1 (r) = 0 . (10.426)
r r
Setting
f1 (r) ∼ a rγ , g1 (r) ∼ b rγ (10.427)
→0 →0
yields
or
γ + (j + 12 ) −Ze2
a
= 0. (10.429)
Ze 2 γ − (j + 12 ) b
This equation
q poses an eigenvalue problem (eigenvalue −γ) for proper γ values. One obtains
γ = ± (j + 12 )2 − Z 2 e4 . The assumed r-dependence in (10.427) makes only the positive solution
possible. We have, hence, determined that the solutions f1 (r) and g1 (r), for small r, assume the
r-dependence in (10.427) with r
1
γ = (j + )2 − Z 2 e4 . (10.430)
2
Note that the exponent in (10.427), in case j + 12 )2 < Ze2 , becomes imaginary. Such r-dependence
would make the expectation value of the potential
Z
1
r2 dr ρ(~r) (10.431)
r
infinite since, according to (10.266, 10.267, 10.420), for the particle density holds then
1
ρ(~r) ∼ |rγ−1 |2 = . (10.432)
r2
Behaviour at r → ∞
For very large r values (10.422) becomes
∂r g1 (r) = − ( − m, ) f1 (r)
∂r f1 (r) = ( + m ) g1 (r) (10.433)
10.10: Dirac Particles in Electromagnetic Field 349
For bound states holds < m and, hence, µ is real. Let us consider then for the solution of (10.434)
√ √
f1 (r) = m + a e−µ r , g1 (r) = − m − a e−µr . (10.436)
The last two terms on the l.h.s. of both (10.440) and (10.441) correspond to (10.437) where they
cancelled in case f˜1 = g̃1 = a. In the present case the functions f˜1 and g̃1 cannot be chosen
identical due to the terms in the differential equations contributing for finite r. However, without
loss of generality we can choose
which leads to a partial cancellation of the asymptotically dominant terms. We also introduce the
new variable
ρ = 2µ r . (10.443)
350 Relativistic Quantum Mechanics
for γ given in (10.430) which conform to the proper r → 0 behaviour determined above [c.f.
(10.426–10.430)]. Inserting (10.448, 10.449) into (10.446, 10.447) leads to
X Ze2
(s + γ) αs + (j + 12 ) βs − √ αs
s m 2 − 2
mZe2
−√ βs ρs+γ−1 = 0 (10.450)
2
m − 2
mZe2
X
(s + γ) β2 + (j + 12 ) αs + √ αs
s m2 − 2
Ze2
+√ βs − βs−1 = 0 (10.451)
m2 − 2
From (10.450) follows
2
√mZe − (j + 12 )
αs m2 −2
= Ze2
. (10.452)
βs s + γ − √m 2 −2
Defining
Ze2
so = √ −γ (10.455)
m2 − 2
one obtains
s − so
βs = βs−1
s(s + 2γ)
(s − 1 − so )(s − so )
= βs−2
(s − 1)s (s − 1 + 2γ)(s + 2γ)
..
.
(1 − so )(2 − so ) . . . (s − so )
= β0 (10.456)
s! (2γ + 1)(2γ + 2) · · · (2γ + s)
One can relate the polynomials φ1 (ρ) and φ2 (ρ) defined through (10.448, 10.449) and (10.456,
10.457) with the confluent hypergeometric functions
a a(a + 1) x2
F (a, c; x) = 1 + x + + ... (10.458)
c c(c + 1) 2!
L(α)
n = F (−n, α + 1, x) . (10.459)
It holds
1 2
j+ − √mZe
2 m −2
2
φ1 (ρ) = β0 ργ F (−so , 2γ + 1; ρ) (10.460)
so
φ2 (ρ) = β0 ργ F (1 − so , 2γ + 1; ρ) . (10.461)
In order that the wave functions remain normalizable the power series (10.448, 10.449) must be of
finite order. This requires that all coefficients αs and βs must vanish for s ≥ n0 for some n0 ∈ N.
The expressions (10.456) and (10.457) for βs and αs imply that so must then be an integer, i.e.,
so = n0 . According to the definitions (10.430, 10.455) this confinement of so implies discrete values
for , namely,
m
(n0 ) = r 2 e4
, n0 = 0, 1, 2, . . . (10.462)
Z
1 + q
1
(n0 + (j+ 2 )2 − Z 2 e4 )2
352 Relativistic Quantum Mechanics
This expression agrees with the spectrum of relativistic hydrogen-type atoms derived above and
given by (10.405). Comparision with (10.405) allows one to identify n0 = n − j + 12 which, in fact,
is an integer. For example, for the states 2p 1 , 3p 1 , 3d 3 holds n0 = 1, 2, 1. We can, hence, conclude
2 2 2
that the polynomials in (10.461 for values given by (10.405) and the ensuing so values ( 10.455)
are finite.
Altogether we have determined the stationary states of the type (10.421) with radial wave functions
f1 (r), g1 (r) determined by (10.438, 10.439), (10.442), and (10.460, 10.461). The coefficients β0 in
(10.460, 10.461) are to be chosen to satisfy a normalization condition and to assign an overall
phase. Due to the form (10.410) of the stationary state wave function the density ρ(xµ ) of the
states under consideration, given by expression (10.267), is time-independent. The normalization
integral is then
Z ∞ Z π Z 2π
2
r dr sin θdθ dφ (|Φ(~r)|2 + |X (~r)|2 ) = 1 (10.463)
0 0 0
where Φ and X , as in (10.421), are two-dimensional vectors determined through the explicit form
of the spin-orbital angular momentum states Yjm (j ± 12 , 12 |r̂) in (6.147, 6.148). The orthonormality
properties (6.157, 6.158) of the latter states absorb the angular integral in (10.463) and yield [note
the 1/r factor in (10.421)]
Z ∞
dr (|f1 (r)|2 + |g1 (r)|2 ) = 1 (10.464)
0
The evaluation of the integrals, which involve the confluent hypergeometric functions in (10.460,
10.461), can follow the procedure adopted for the wave functions of the non-relativistic hydrogen
atom and will not be carried out here.
The wave functions (10.421) correspond to non-relativistic states with orbital angular momentum
` = j + 12 . They are described through quantum numbers n, j, ` = j + 21 , m. The complete wave
function is given by the following set of formulas
1
F1 (r) = F− (κ|r) , G1 (r) = F+ (κ|r) , κ = j+ 2 (10.466)
where2
(n0 + γ)m
γ−1 −µr
F± (κ|r) = ∓ N (2µr) e − κ F (−n0 , 2γ + 1; 2µr)
0 0
± n F (1 − n , 2γ + 1; 2µr) (10.467)
3
v
(2µ) 2 u m ∓ )Γ(2γ + n0 + 1)
N = (10.468)
u
Γ(2γ + 1) 4m (n0 +γ)m (n0 +γ)m − κ n0 !
t
2
This formula has been adapted from ”Relativistic Quantum Mechanics” by W. Greiner, (Springer, Berlin, 1990),
Sect. 9.6.
10.10: Dirac Particles in Electromagnetic Field 353
and
p
µ = (m − )(m + )
q
γ = (j + 12 )2 − Z 2 e4
n0 = n − j − 12
m
= q
2 e4
. (10.469)
1 + (nZ0 +γ) 2
We want to consider now the stationary states of the type (10.421) which, in the non-relativistic
limit, become
f2 (r)
i Yjm (j − 2 , 2 |r̂)
1 1
Ψ(xµ ) ≈ e−1t r . (10.470)
0
Obviously, this wavefunction has an orbital angular momentum quantum number ` = j − 12 and,
accordingly, describes the complementary set of states 1s 1 , 2s 1 , 2p 3 , 3s 1 , 3p 3 , 3d 5 , etc. not not
2 2 2 2 2 2
covered by the wave functions given by (10.465–10.469). The radial wave functions f2 (r) and g2 (r)
in (10.421) are governed by the radial Dirac equation (10.423) which differs from the radial Dirac
equation for f1 (r) and g1 (r) solely by the sign of the terms (j + 21 )/r. One can verify, tracing all
steps which lead from (10.422) to (10.469) that the following wave functions result