You are on page 1of 38

The most tragic word in the English language is potential

Arthur Lotti

6
Relativistic Particles and Fields in
External Electromagnetic Potential
Having found classical fields to describe relativistic particles, we may ask what corrections the relativistic motion causes to the Schrodinger description of atoms. We
shall therefore study Klein-Gordon and the Dirac equation in an external field. Let
A (x) be the associated four-vector potential, accounting for electric and magnetic
field strengths via (4.227) and (4.228). For classical relativistic point particles, an
interaction with these external fields is introduced via the so-called minimal substitution rule, which we are now going to discuss. Throughout this Chapter, the
presence of an external electromagnetic field will be assumed, without asking for its
origin.
An important property of the electromagnetic field is its description in terms of a
vector potential A (x) and the gauge invariance of this description. In Eqs. (4.230)
and (4.231) we have expressed electric and magnetic field strength as components
of a four-curl F = A A of a vector potential A (x). This four-curl is
invariant under gauge transformations
A (x) A (x) + (x).

(6.1)

The gauge invariance restricts strongly the possibilities of introducing electromagnetic interactions into particle dynamics and the Lagrange densities (6.94) and (6.95)
of charged scalar and Dirac fields.
The prescription for coupling electromagnetism has been known for a long time
in the classical electrodynamics of point particles.

6.1

Charged Point Particles

A free relativistic particle moving along an arbitrarily parametrized path x ( ) in


four-space is described by an action
A = Mc

d q ( )q ( ).
428

(6.2)

429

6.1 Charged Point Particles

The physical time along the path is given by q 0 ( ) = ct, and the physical velocity
by v(t) dq(t)/dt. In terms of these, the action reads:
A=

6.1.1

dt L(t) Mc

v2 (t)
dt 1 2
c
"

#1/2

(6.3)

Coupling to Electromagnetism

If the particle has a charge e (in our convention, e has a negative value for electrons)
and lies at rest at position x, its electric potential energy is
V (x, t) = e(x, t)

(6.4)

(x, t) = A0 (x, t).

(6.5)

where

In our convention, the charge of the electron e has a negative value to have agreement
with the sign in the historic form of the Maxwell equations
E(x) = 2 (x) = (x),

B(x) E(x)
= A(x) E(x)

h
i
1

= 2 A(x) A(x) E(x)


= j(x).
c

(6.6)

If the electron moves along a trajectory q(t), its potential energy is


V (t) = e (q(t), t) .

(6.7)

In the Lagrangian L = T V , this contributes with the opposite sign


Lint (t) = eA0 (q(t), t)

(6.8)

giving a potential part of the interaction


Aint
pot = e

dt A0 (q(t), t) .

(6.9)

Since the time t coincides with q 0 ( )/c of the trajectory, this can be expressed as
Aint
pot =

eZ
dq 0 A0 .
c

(6.10)

In this form it is now quite simple to write down the complete electromagnetic interaction purely on the basis of relativistic invariance. The direct invariant extension
of (6.11) is obviously
Aint =

e
c

dq A (q).

(6.11)

430

6 Relativistic Particles and Fields in External Electromagnetic Potential

Thus, the full action of a point particle can be written in covariant form as
A = Mc

d q ( )q ( )

eZ
dq A (q),
c

(6.12)

or more explicitly as
A=

dt L(t) = Mc

v2
dt 1 2
c
"

#1/2

1
dt A v A .
c


(6.13)

The canonical formalism supplies us with the canonically conjugate momenta


P=

v
e
e
L
= Mq
+ A p + A.
v
c
1 v2 /c2 c

(6.14)

The Euler-Lagrange equation obtained by extremizing this equation is


L
d L
=
,
dt v(t)
q(t)

(6.15)

or

ed
e
d
p(t) =
A(q(t), t) eA0 (q(t), t) + v i Ai (q(t), t).
dt
c dt
c
We now split
d

A(q(t), t) = (v(t) )A(q(t), t) + A(q(t), t),


dt
t

(6.16)

(6.17)

and obtain
e
e
e
d
p(t) = (v(t) )A(q(t), t)
A(q(t), t) eA0 (q(t), t) + v i Ai (q(t), t).
dt
c
c t
c
(6.18)
The right-hand side contains the electric and magnetic fields (4.232) and (4.233), in
terms of which it takes the well-known form
d
v
p=e E+ B .
dt
c


(6.19)

This can be rewritten in terms of the proper time t/ as



d
e  0
Ep +p B ,
p=
d
Mc

(6.20)

Recalling Eqs. (4.230) and (4.231), this is recognized as the spatial part of the
covariant equation
d
e
p =
F p .
(6.21)
d
Mc
The temporal component of this equation
e
d 0
p =
Ep
d
Mc

(6.22)

431

6.1 Charged Point Particles

gives the energy increase of a particle running through an electromagnetic field. In


real time this is
v
d 0
p = eE .
(6.23)
dt
c
Combining this with (6.19), we find the acceleration
d
v
d p
e
v v
E+ B
v(t) = c
=
E
0
dt
dt p
M
c
c c




(6.24)

The velocity is related to the canonical momenta and external field via
e
P A
v
c
.
= s
2
c
e
P A + m2 c2
c

(6.25)

This can be used to calculate the Hamiltonian via the Legendre transform
H=

L
vL= PvL
v

(6.26)

giving
H=c

s


e
P A
c

2

+ m2 c2 + eA0 .

(6.27)

In the non-relativistic limit this has the expansion


1
e
H = mc +
P A
2m
c
2

2

+ eA0 + . . .

(6.28)

Thus, the free theory goes over into the interacting theory by the minimal substitution rule
e
e
p p A,
H H A0 .
(6.29)
c
c
or, in relativistic notation:
e
p p A .
c

6.1.2

(6.30)

Spin Precession in Atom

In 1926, Uhlenbeck and Goudsmit noticed that the observed Zeeman splitting of
atomic levels could be explained by an electron of spin 21 . Its magnetic moment is
usually expressed in terms of the combination of fundamental constants which have
the dimension of a magnetic moment, the Bohr magneton B = eh/Mc as

 = gB Sh ,

eh
,
2Mc

(6.31)

432

6 Relativistic Particles and Fields in External Electromagnetic Potential

where S = /2 is the spin matrix which has the commutation rules


[Si , Sj ] = ihijk Sk ,

(6.32)

and g is a dimensionless number called the gyromagnetic ratio or Lande factor.


If an electron moves in an orbit under the influence of a torque-free central force,
as an electron does in the Coulomb field of an atomic nucleus, the total angular
momentum is conserved. The spin, however, shows a precession just like a spinning
top. This precession has two main contributions: one is due to the magnetic coupling
of the magnetic moment of the spin to the magnetic field of the electron orbit, called
spin-orbit coupling. The other part is purely kinematical, it is the Thomas precession
discussed in Section (4.16), caused by the slightly relativistic nature of the electron
orbit.
The spin-orbit splitting of the atomic energy levels (to be pictured and discussed
further in Fig. 6.1) is caused by a magnetic interaction energy
H LS (r) =

g
1 dV (r)
SL
,
2
2
2M c
r dr

(6.33)

where V (r) is the atomic potential depending only on r = |x|. To derive this H LS (r),
we note that the spin precession of the electron at rest in a given magnetic field B
is given by the Heisenberg equation
dS
=
dt

 B,

(6.34)

where is the magnetic moment of the electron. In an atom, the magnetic field in
the rest frame of the electron is entirely due to the electric field in the rest frame of
the atom. A Lorentz transformation (4.276) to the electron velocity produces the
magnetic field in the electron rest frame:
B = Bel =

v
E,
c

1
=q
.
1 v 2 /c2

(6.35)

Since an atomic electron has a small velocity ratio v/c which is of the order of the
fine-structure constant 1/137, we can write approximately
Bel

v
E.
c

(6.36)

The electric field gives the electron an acceleration


v =

e
E,
M

(6.37)

Mc 1

v v,
e c2

(6.38)

so that we may also write


Bel

433

6.1 Charged Point Particles

and the precession equation (6.34) as


g
dS
S.
2 (v v)
dt
2c
This can be written as

(6.39)

dS
= LS S,
(6.40)
dt
where LS is the angular velocity of spin precession caused by the orbital magnetic
field at in the rest frame of the electron:
g

v v.
(6.41)
LS
2c2

In the rest frame of the atom where the electron is accelerated towards the center
along its orbit, this result receives a relativistic correction. To lowest order in 1/c,
we must add to LS the angular velocity T of Thomas precession, such that the
total angular velocity of precession becomes

=
LS +
T g 2 1 v v.

(6.42)

Since g is very close to 2, the Thomas precession explains why the spin-orbit splitting
was initially found to be in agreement with a normal gyromagnetic ratio g = 1, the
characteristic value for a rotating charged sphere.
If there is also an external magnetic field, this is transformed to the electron
rest frame by a Lorentz transformation (4.273), where it leads to an approximate
equation of motion for the spin
dS
=
dt

 B 

v
E .
c


(6.43)

 via Eq. (6.31), this becomes




v
eg
dS
S
em
S B E .
(6.44)
dt
2Mc
c
This equation defines the frequency
em of precession due to the magnetic end elec-

Expressing

tric fields in the rest frame of the electron. Expressing E in terms of the acceleration
via Eq. (6.37), this becomes
g
M
dS
.

S eB
(v v)
dt
2Mc
c


(6.45)

The acceleration can be expressed in terms of the central Coulomb potential V (r)
as
x 1 dV
v =
.
(6.46)
r M dr
the spin precession rate in the electrons rest frame is
dS
g
x 1 dV
=
S eB + v
dt
2Mc
r c dr

g
1 dV
=
S eB
L
2Mc
Mc dr

. (6.47)

434

6 Relativistic Particles and Fields in External Electromagnetic Potential

There exists a simple Hamiltonian operator for spin-orbit interaction H LS (t)], from
which this equation can be derived via Heisenbergs equation (1.279)
i

S(t)
= [S(t), H LS (t)].
h

(6.48)

The operator is

1
dV
L
H LS (r) = B
Mc e dr
ge
g
1 dV
=
SB+
SL
.
2
2
2Mc
2M c
r dr

(6.49)

Indeed, using the commutation rules (6.32), we find immediately (6.46). Historically,
the interaction energy (6.49) was used to explain the experimental level splittings
assuming a gyromagnetic ratio g 1 for the electron.
Without the external magnetic field, the angular velocity of precession caused
by spin-orbit coupling is is
g
1 V
L
.
(6.50)
LS =
2
2
2M c
r r
It was realized by Thomas in 1927 that the relativistic motion of the electron change
the factor g as in (6.42) to g 1, so that the true precession frequency is

g 1 1 V
L
.

=
LS +
T = 2M
2 c2
r r

(6.51)

This implied that the experimental data gave g 1 1, so that g is really twice
as large as expected for a rotating charged sphere. Indeed, the value g 2. was
indeed predicted by the Dirac theory of the electron.
In Section 12.15 we shall find that the magnetic moment of the electron has a g
factor slightly larger than the Dirac value 2, the relative deviations a (g 2)/2
being defined as the anomalous magnetic moments. From measurements of the
above precession rate, experimentalists have deduced the values
a(e ) = (115 965.77 0.35) 108 ,
a(e+ ) = (116 030 120) 108 ,
a( ) = (116 616 31) 108 .

(6.52)
(6.53)
(6.54)

In quantum electrodynamics, the gyromagnetic ratio will receive further small


corrections, as will be discussed in detail in Chapter 12.

6.1.3

Relativistic Equation of Motion for Spin Vector and


Thomas Precession

If an electron moves in an orbit under the influence of a torque-free central force,


such as an electron in the Coulomb field of an atomic nucleus, the total angular

435

6.1 Charged Point Particles

momentum is conserved. The spin, however, performs a Thomas precession as


discussed in the previous section. There exists a covariant equation of motion for the
spin four-vector introduced in Eq. (4.761) which describes this precession. Along a
particle orbit parametrized by a parameter , for instance the proper time, we form
the derivative with respect to , assuming that the motion proceeds at fixed total
angular momentum:
dS
dp
= J
.
d
d

(6.55)

The right-hand side can be simplified by multiplying it with the trivial expression
1
g p p = 1.
M 2 c2

(6.56)

Now we use the identity for the -tensor


g = g + g + g + g ,

(6.57)

which can easily be verified using the antisymmetry of the -tensor and considering
= 0123. Then the right-hand side becomes a sum of the four terms

1 

, (6.58)

J
p
p
p
+
J
p
p
p
+
J
p
p
p
+
J
p
p
p

M 2 c2

where p dp /d . The first term vanishes, since p p = (1/2)dp2 /d =


(1/2)dM 2 c2 /d = 0. The last term is equal to S p p /M 2 c2 . Inserting the identity (6.57) into the second and third terms, we obtain twice the left-hand side of
(6.55). Taking this to the left-hand side, we find the equation of motion
dS
1
dp
= 2 2 S
p .
d
M c
d

(6.59)

Note that on account of this equation, the time derivative dS /d points in the
direction of p .
Let us verify that this equation yields indeed the Thomas precession. Denoting
the derivatives with respect to the time t = by a dot, we can rewrite (6.59) as
dS
S
=
dt
dS0
=
S 0
dt


1 dS
1 
2
v,
= 2 2 S 0 p0 + S p p = 2 (S v)
d
M c
c
1d
2
.
(S v) = 2 (S v)
c dt
c

(6.60)
(6.61)

We now differentiate Eq. (4.774) with respect to the time using the relation =
2

3 vv/c
, and find
S R = S

1 0
3
1 0
1
S 0 v.

S
v

S
(v v)
v

+ 1 c2
+ 1 c2
( + 1)2 c4

(6.62)

436

6 Relativistic Particles and Fields in External Electromagnetic Potential

Inserting here Eqs. (6.60) and (6.61), we obtain


S R =

1 0
3
2 1

S 0 v.

(S

v)v

S
(v v)
v

+ 1 c2
+ 1 c2
( + 1)2

(6.63)

On the right-hand side we return to the spin vector SR using Eqs. (4.773) and
(4.776), and find
S R =

2 1
(SR v)v]
=
[(SR v)v
+ 1 c2

T SR,

(6.64)

with the Thomas precession frequency

T = ( + 1) c12 v v,
2

(6.65)

which agrees with the result (4B.26) derived from purely group-theoretic considerations.
In an external electromagnetic field, there is an additional precession. For slow
particles, it is given by Eq. (6.45). If the electron moves fast, we transform the
electromagnetic field to the electron rest frame by a Lorentz transformation (4.273),
and obtain an equation of motion for the spin becomes

v
2 v
S R = B = B E
c
+1c
Expressing

" 

v
B
c

#

(6.66)

 via Eq. (6.31), this becomes

S R SR

em

eg
=
SR
2Mc

"

v
v
B E
c
+1c


v
B
c

#

(6.67)

which is the relativistic generalization of Eq. (6.44). It is easy to see that the
associated fully covariant equation is
1 d
eg
1
g
dS
eF S +
F S + 2 2 p S F p .
=
p S p =
d
2Mc
Mc
d
2Mc
M c
"

(6.68)

On the right-hand side we have inserted the relativistic equation of motion (6.21)
of a point particle in an external electromagnetic field.
If we add to this the relativistic Thomas precession rate (6.59), we obtain the
covariant Bargmann-Michel-Telegdi equation1
dS
g2 d
e
g2
1
egF S +
gF S + 2 2 p S F p .
=
p S p =
d
2Mc
Mc
d
2Mc
M c
(6.69)
"

V. Bargmann, L. Michel, and V.L. Telegdi, Phys. Rev. Lett. 2 , 435 (1959).

437

6.2 Charged Particle in Schr


odinger Theory

For the spin vector SR in the electron rest frame this implies a change in the
electromagnetic precession rate in Eq. (6.67) to2
dS
=
dt

em T S (
em +
T ) S

(6.70)

with a frequency given by the Thomas equation

e
=
Mc

"

g
g
v
1

v
g
B
1 +
1
B

2
+1 c
c
2 +1

v
E .
em T
c
(6.71)
The contribution of the Thomas precession is the part without the gyromagnetic
retio g:

"

e
1
1
1
=
1
B+
(v

B)
v
+
vE .
Mc

+1 c2
+1 c

(6.72)

This is agrees with the Thomas frequency (6.65) after inserting the acceleration
(6.24).
The Thomas equation (6.71) can be used to calculate the time dependence of
of an electron, i.e., its component of the spin in the direction
the helicity h SR v
of motion. Using the chain rule of differentiation,
1
d
d
) = S R v
+ [SR (
(SR v
v SR )
v] v
dt
v
dt

(6.73)

and inserting (6.70) as well as the equation for the acceleration (6.24), we obtain
dh
e
=
SR
dt
Mc



c
g
gv
B+
E .
1 v

2
2c v


(6.74)

where SR is the component of the spin vector orthogonal to v. This equation shows
that for a Dirac electron which has g = 2 the helicity remains constant in a purely
magnetic field. Moreover, if the electron moves ultrarelativistically (v c), the
value g = 2 makes the last term extremely small, (e/Mc) 2 SR E, so that the
helicity is almost unaffected by an electric field. The anomalous magnetic moment
of the electron, however, changes this a finite value (e/Mc)aSR E. This drastic
effect was used to measure the experimental values listed in Eqs. (6.52)(6.54).

6.2

Charged Particle in Schr


odinger Theory

When going over from quantum mechanics to second quantized field theories we
found the rule that a non-relativistic Hamiltonian
H=
2

L.T. Thomas, Phil. Mag. 3 , 1 (1927).

p2
+ V (x)
2m

(6.75)

438

6 Relativistic Particles and Fields in External Electromagnetic Potential

became an operator
H=

2
+ V (x) (x, t),
d x (x, t)
2m
3

"

(6.76)

where we have omitted the operator hats, for brevity. With the same rules we see
that the second quantized form of the interacting nonrelativistic Hamiltonian in a
static A(x) field,
H=

(p eA)2 e 0
+ A ,
2m
c

(6.77)

is given by
H=

"

e
1
i A
d x (x, t)
2m
c
3

2

+ eA (x) (x, t).

(6.78)

When going to the action of this theory we find


A=

dtL =

dt

d x (x, t) it + eA0 (x, t)


+

e
1
(x, t) i A
2m
c


2

(x, t) .

(6.79)

It is easy to verify that (6.78) reemerges from the Legendre transform


H=

L
(x, t) L.

(x, t)

(6.80)

The action (6.79) holds also for time-dependent A (x) fields.


We can now deduce the second quantized form of the minimal substitution rule
(6.29) which is
e
i A(x, t),
c
t t + ieA0 (x, t),

(6.81)

or covariantly:

e
+ i A (x).
(6.82)
c
This substitution rule has the important property that the gauge invariance of
the free photon action is preserved by the interacting theory: If we perform the
gauge transformation
A (x) A (x) + (x),

(6.83)

A0 (x, t) A0 (x, t) + t (x, t)


A(x, t) A(x, t) (x, t),

(6.84)

i.e.,

439

6.2 Charged Particle in Schr


odinger Theory

the action remains invariant if we simultaneously change the fields (x, t) of the
charged particles by a spacetime-dependent phase
e

(x, t) ei c (x,t) (x, t).

(6.85)

Under this transformation, the derivatives of the field change like


e
i (x, t) ,
(x, t) e
c
i ec (x,t)
(t iet ) (x, t).
t e
i ec (x,t)

(6.86)

The modified derivatives appearing in the action have therefore the following simple
transformation law:
e
e
e
i A (x, t) ei c (x,t) i A (x, t),
c 
c



e 0
e
t + i A (x, t) ei c (x,t) t + ieA0 (x, t).
c

(6.87)

These combinations of derivatives and gauge fields are called covariant derivatives
and written as
e
i A (x, t)
c 

Dt (x, t) t + ieA0 (x, t)
D(x, t)

(6.88)

or, in four-vector notation,


D (x) =

e
+ i A (x).
c

(6.89)

Here the adjective covariant does not refer to the Lorentz group but to the gauge
group. It records the fact that D transforms under local gauge changes (6.81) of
in the same way as itself in (6.85):
e

D (x) ei c (x) D (x).

(6.90)

With the help of this covariant derivative any action, which is invariant under global
phase changes by a constant phase angle
(x) ei (x),

(6.91)

can easily be made invariant under local gauge transformations (6.83). We merely
have to replace all derivatives by covariant derivatives (6.89) and add the gauge
invariant photon action (4.234).

440

6.3
6.3.1

6 Relativistic Particles and Fields in External Electromagnetic Potential

Charged Relativistic Fields


Scalar Field

The Lagrangian density of a free charged scalar field is from Eqs (4.162)
L = (x) (x) M 2 (x)(x)

(6.92)

For a relativistic scalar field of charge e, the interacting Lagrange density is a


straightforward generalization of the Schrodinger case (6.79):
L = [D (x)] D (x) M 2 (x)(x)




e
e
= i A (x) (x) + i A (x) (x) M 2 (x)(x)
c
c

6.3.2

(6.93)

Dirac Field

The Lagrangian densities of a free charged spin-1/2 field is from Eq. (4.498):
L = (x) (x) M 2 (x)(x)

(6.94)

(i/
m) (x).
L(x) = (x)

(6.95)

and

If the particle carries a charge e, we replace in this Lagrangian:

/ =

e
e
+ i A = / + i A
/
c
c

D
/.

(6.96)

In this way we arrive at the Lagrangian of quantum electrodynamics (QED)


1 2
(i/
L(x) = (x)
D m) (x) F
.
4

(6.97)

The classical field equations can easily be found by extremizing the action and
variation with respect to all fields which gives
A
= (i/
D M) (x) = 0

(x)
1
A
= F (x) j (x) = 0.
A (x)
c
with the current density

(6.98)
(6.99)

j (x) e c (x)
(x).

(6.100)

1
F (x) = j (x).
c

(6.101)

Equation (6.99) is the Maxwell equation for the electromagnetic field around a classical four-dimensional vector current j (x):

441

6.4 Pauli Equation from Dirac Theory

In the Lorentz gauge A (x) = 0, this equation reads simply


1
2 A (x) = j (x).
c

(6.102)

The current j combines the charge density (x) and the current density j of
particles of charge e in a four-vector
j = (c, j) .

(6.103)

In terms of electric and magnetic fields E i = F i0 , B i = F jk , the field equations


(6.101) turn into the Maxwell equations
0 = e
E = = e
= 1 j = e .
BE
c
c

(6.104)

The first is Coulombs law, the second Amperes law in the presence of charges and
currents.
Note that the physical units employed here differ from those used in many books
of classical electrodynamics3 by the absence of a factor 1/4 on the right-hand side.
The Lagrangian used in those books is
1
1 2
F (x) j (x)A (x)
8
c


i
1 h 2
1
2
=
E B (x) j A (x),
4
c

L(x) =

(6.105)

which leads to Maxwells field equations


E = 4
4
j.
B =
c

(6.106)

The form employed


conventionally

in quantum field theory arises from this by replacing A 4A and e 4e. The charge of the electron in our units has
therefore the numerical value
q

(6.107)
e = 4 4/137

rather than e = .

6.4

Pauli Equation from Dirac Theory

It is instructive to take the Dirac equation (6.98) to a two-component form corresponding to (4.561) and (4.573), and further to (4.579). Due to the fundamental
3

See for example J.D. Jackson, Classical Electrodynamics, Wiley and Sons, New York, 1967.

442

6 Relativistic Particles and Fields in External Electromagnetic Potential

nature of the equations to be derived we shall not work with natural units in this section but carry all fundamental constants along explicitly. As in (4.580), we multiply
(6.98) by (ihD
/ Mc) and work out the product
e
+ i A + Mc
(ihD
/ Mc) (ihD
/ + Mc) = i h
c




e
+ i A Mc
i h
c
(6.108)


We now use the relation


1
1
= ( + ) + ( ) = g i ,
2
2
with from (4.514), and find

(6.109)

e
+ i A
c



e
e
i
e
e

h
+ i A + i A [h + i A , h
=g
+ i A ]
c
c
2
c
c

e
h
+ i A
c



e
= h
+ i A
c


2

1 eh
F .
2 c

(6.110)

Thus we obtain as a generalization of Eqs. (4.579) the Pauli equation


"

e
h
+ i A
c


2

1 eh
F M 2 c2 (x) = 0,

2 c

(6.111)

and the same equation once more for the other two-component spinor field (x).
Note that in this equation, electromagnetism is not coupled minimally. In fact,
there is a non-minimal coupling of the spin via a term

1
F = H + i E.
2
where in the chiral and Dirac representations

D = 
0

(6.112)


0

(6.113)

respectively. Thus, in the chiral representation, Eq. (6.111) decomposes into two
separate two-component equations for the upper and lower spinor components (x)
and (x) in (x):
"

e
h
+ i A
c


2

 (H iE) M

2 2

#(

(x)
(x)

= 0.

(6.114)

In the nonrelativistic limit where c , we remove the fast oscillations from


2
(x), setting (x) eiM c t/h (x, t)/ 2M as in (4.153), and find for (x, t) the
nonrelativistic equation
h
2
e
it +
i A
2M
ch

"

2

e
H eA0 (x) (x, t) = 0.
+
2Mc

(6.115)

443

6.5 Relativistic Wave Equations in Coulomb Potential

This corresponds to a magnetic interaction energy


(6.116)
For a small magnet with a magnetic moment , the magnetic interaction energy is
Hmag =  H.
(6.117)
Hmag =

eh
H.
2Mc

A Dirac particle with has therefore a magnetic moment


e

e h
=2
S,
 = Mc
2
2Mc

(6.118)

where S = /2 is the spin matrix. Experimentally, one parametrizes a magnetic


moment of a fundamental particles as
e
S
 = g 2Mc

(6.119)

with the gyromagnetic ratio g, which is equal to unity for a uniformly charged sphere.
According to the Dirac theory, an electron has a gyromagnetic ratio
ge |Dirac = 2.

(6.120)

The experimental value is very close to this. The small deviation called anomalous
magnetic moment will be explained in Chapter 12.
The nonrelativistic Pauli equation (6.115) could also have been obtained by introducing a minimal coupling directly into the nonrelativistic two-component equation
(4.576). This changes it it eA0 and ( )2 [ ( ieA)]2 . The latter
is worked out as in (6.121), yielding

[ ( ieA)]2 = ij + iijk k (i ieAi )(j ieAj )

= ( ieA)2 + iijk k (i ieAi )(j ieAj ) + e H, (6.121)

so that Eq. (4.578) goes over into (6.115), after reinserting all fundamental constants,

6.5

Relativistic Wave Equations in Coulomb Potential

It is now easy to write down field equations for a Klein-Gordon and a Dirac field
in the presence of an external Coulomb potential of charge Ze. In natural units we
have

Z
VC (x) =
,
r = x2 ,
(6.122)
r
corresponding to a four-vector potential
eA (x) = (VC (x, 0), 0).

(6.123)

444

6 Relativistic Particles and Fields in External Electromagnetic Potential

Since this does not depend on time, we can consider the wave equations for wave
functions (x) = eiEt E (x) and (x) = eiEt E (x), and find the time-independent
equations
(E 2 + 2 M)E (x) = 0

(6.124)

and

( 0 E + i M)E (x) = 0.

(6.125)

In these equations we simply perform the minimal substitution


EE+

Z
.
r

(6.126)

The energy-eigenvalues obtained from the resulting equations can be compared with
those of hydrogen-like atoms. The velocity of an electron in the ground state is of
the order Zc. Thus for rather high Z, the electron has a relativistic velocity and
there must be corrections significant deviations from Schrodinger theory. We shall
see that the Dirac equation in an external field reproduces quite well a number of
features resulting from the relativistic motion.

6.5.1

Reminder of Schr
odinger Equation with Coulomb
Potential

The time-independent Schrodinger equation reads




1
Z
2
E E (x) = 0.
2M
r


(6.127)

The Laplacian may be decomposed into radial and angular parts by writing
2 =

2
2
2
L
+

,
r 2 r r
r2

(6.128)

= xp
are the differential operators for the generators of angular momenwhere L
tum [the spatial part of Li = L23 of (4.95)]. Then (6.127) reads
2 2ZM
2
L
2
2
+ 2
2ME E (x) = 0.
r
r r
r
r
!

(6.129)

2 are the spherical harmonics Ylm (, ), which diagonalize also


The eigenstates of L
with the eigenvalues
the third component of L,
2 Ylm (, ) = l(l + 1)Ylm (, ),
L
3 Ylm (, ) = mYlm (, ).
L

(6.130)

6.5 Relativistic Wave Equations in Coulomb Potential

445

The wave functions may be factorized into a radial wave function Rnl (r) and a
spherical harmonic:
nlm (x) = Rnl (r)Ylm(, ).
(6.131)
Explicitly,
Rnl (r) =

1
1/2
aB n (2l

1
+ 1)!

v
u
u
t

(n + l)!
(n l 1)!

(6.132)

(2r/naB )l+1 er/naB M(n + l + 1, 2l + 2, 2r/naB )


v
u

1 u (n l 1)! r/naB
e
(2r/naB )l+1 L2l+1
= 1/2 t
nl1 (2r/naB ),
(n
+
l)!
aB n
where aB is the Bohr radius, which in natural units h
= c = 1 is equal to
aB =

1
,
ZM

(6.133)

For a hydrogen atom with Z = 1, this is about 1/137 times the Compton wavelength
of the electron e h
/Me c. The classical velocity of the electron on the lowest Bohr
orbit is vB = c. Thus it is almost nonrelativistic, which is the reason what the
Schrodinger equation explains the hydrogen spectrum quite well. The functions
M(a, b, z) are confluent hypergeometric functions or Kummer functions, defined by
the power series
a
a(a + 1) z
M(a, b, z) F1,1 (a, b, z) = 1 + z +
+ ... .
b
b(b + 1) z!

(6.134)

For b = n they are polynomial related to the Laguerre polynomials4 Ln (z) by


Ln (z)

(n + )!
M(n, + 1, z),
n!!

(6.135)

The radial wave functions are normalized to


Z

drRnr l (r)Rnr l (r) = nr nr .

(6.136)

They have an asymptotic behavior Pnl (r/n)er/naB , where Pnl (r/n) is a polynomial
of degree nr = n l 1, which defines the radial quantum number. The energy
eigenvalues depend on are n in the well-known way:
En = Z 2
4

M2
.
2n2

(6.137)

I.S. Gradshteyn and I.M. Ryzhik, op. cit., Formula 8.970 (our definition differs from that in
L.D. Landau and E.M. Lifshitz, Quantum Mechanics, Pergamon Press, New York, 1965, Eq. (d.13):
Our Ln = () /(n + )!Ln+ |L.L. ).

446

6 Relativistic Particles and Fields in External Electromagnetic Potential

The number 2 M/2 is the Rydberg-constant:


Ry =

2 M
27.21

eV 3.288 1015 Hz.


2
2

(6.138)

Later in Section 12.21) we shall need the value of the wave function at the origin.
It is non-zero only for s-waves where it is equal to
1
|n00 (0)| =

6.5.2

ZM
1
=
n

1
.
naB

(6.139)

Klein-Gordon Field in Coulomb Potential

After the substitution (6.126) into (6.124), we find the Klein-Gordon equation in
the Coulomb potential (6.122):
"

Z
E+
r

2

+ M

E (x) = 0.

(6.140)

With the angular decomposition (6.128), this becomes


2 Z 2 2 2ZE
2
L
2
2
+

(E 2 M 2 ) E (x) = 0.
r
r r
r2
r

"

(6.141)

The solutions of this equation can be obtained from those of the nonrelativistic
Schrodinger equation (6.129) by replacing
2 L
2 Z 2 2 ,
L
E
,
M
E2 M 2
E
.
2M

(6.142)
(6.143)
(6.144)

The replacement (6.142) is done most efficiently by defining the eigenvalues l(l +
2 Z 2 2 by analogy with those of L
2 as ( + 1):
1) Z 2 2 of the operator L
l(l + 1) Z 2 2 ( + 1).

(6.145)

Then the quantum number l of the Schrodinger wave functions is simply replaced
by l = l l , where
l

"

2

1
1
= l+ l+
2
2
2 2
Z
+ O(4 ).
=
2l + 1

#1/2

(6.146)

The other solution of relation (6.145) with the opposite sign in front of the square
root is unphysical since the associated wave functions are too singular at the origin

447

6.5 Relativistic Wave Equations in Coulomb Potential

to be normalizable. As before, the radial quantum number nr which determines the


degree of the polynomial Pnl (r/n) in the wave functions must be an integer. This
is no longer true for the combination of quantum numbers which determines the
energy, which is now given by
nr + + 1 = nr + l + 1 l = n l .

(6.147)

This leads to the equation for the energy eigenvalues


Enl 2 M 2
Z 2 M2 Enl 2
1
=
,
2M
2
M 2 (n l )2

(6.148)

with the solution


M
Enl = q
2
1 + Z 2 /(n l )2

(Z)2 3 (Z)4
(Z)4
= M 1
+

+ O(Z 6 6 ) .
2n2
8 n4
n3 (2l + 1)
"

(6.149)

The first two terms correspond to the Schrodinger energies (6.137) (plus rest energy),
the next two are relativistic corrections. The first of these breaks the degeneracy
between the levels of the same n and different l, generated in the Schrodinger theory
by the Lentz-Runge vector [O(4)-invariance].
The correction terms become large for large central charge Z. In particular, the
lowest energy and successively the higher ones become complex for central charges
Z > 137/2. The physical reason for this is that the large potential gradient near the
origin can create pairs of particles from the vacuum. This phenomenon can only be
properly understood after quantizing the field theory.
As for the free Klein-Gordon field, the energy appears with both signs.

6.5.3

Dirac Field in Coulomb Potential

After the substitution (6.126) into (6.125), we find the Dirac equation in the
Coulomb potential (6.122):

Z 0
E+
+ i M E (x) = 0.
r



(6.150)

In order to find the energy spectrum it is useful to establish contact with the
Klein-Gordon case. Multiplying (6.150) by the operator


we obtain

"

Z
E+
r

2

Z 0
+ i + M,
E+
r


+ i

M 2 E (x) = 0.,
r

(6.151)

(6.152)

448

6 Relativistic Particles and Fields in External Electromagnetic Potential

In the chiral representation, 44 -matrix 0 = has a diagonal block form (4.557).


We therefore decompose
!
E (x)
E (x) =
,
(6.153)
E (x)
and find the equation for the upper two-component spinors
"

Z
E+
r

2

Z
M 2 E (x) = 0,
+ +i
r
2

(6.154)

and the same equation for E (x) with i replaced by i. Expressing 2 via (6.128)
and writing 1/r =
x/r 2 , we obtain the differential equation

2 Z 2 2 + iZ x
2ZE
L
+

(E 2 M 2 ) E (x) = 0,
r2
r
(6.155)
and a corresponding equation for E (x).
, the total angular momentum
Due to the rotation invariance of x
"

2
2

+
r 2 r r

=L
+S=L
+
J

(6.156)

2 and
commutes with the differential operator in (6.155). Thus we can diagonalize J
1
3
J3 with eigenvalues j(j + 1) and m. For a fixed value of j = 2 , 1, 2 , . . . , the orbital
angular momentum can have the value l+ = j + 21 and l = j 1/2. The two states
is a pseudoscalar and must necessarily
have opposite parities. The operator x
is the unit matrix, its eigenvalues must
change the parities. Since the square of x
be 1. Moreover, the unit vector x changes necessarily the parity and thus l. Thus,
in the two-component Hilbert space of fixed quantum numbers j and m with orbital
angular momenta l = l = j 1/2, the diagonal matrix elements vanish




(6.157)

|jm, +i = 1.
hjm, | x

(6.158)

|jm, +i = 0,
hjm, +| x

|jm, i = 0.
hjm, | x

For the off-diagonal elements we easily calculate

|jm, i = 1,
hjm, +| x

The central parentheses in (6.155) have therefore the matrix elements




L Z iZ
x =

(j + 21 )(j + 23 ) Z 2 2
iZ
iZ
(j 12 )(j + 12 ) Z 2 2

(6.159)
In analogy with the Klein-Gordon case, we denote the eigenvalues of this matrix by
( + 1). The corresponding values of are found to be
j + =

"

1
j+
2

2

#1/2

j = j+ 1.

(6.160)

449

6.5 Relativistic Wave Equations in Coulomb Potential

These may be written as


j = j

1
j l j ,
2

(6.161)

where l 21 is the orbital angular momentum, and


1
j j +
2

"

1
j+
2

2

#1/2

Z 2 2
+ O(Z 4 4 ).
2j + 1

(6.162)

When solving Eq. (6.155), the solutions consist, as in the nonrelativistic hydrogen
atom, of an exponential factor multiplied by a polynomial of degree nr which is the
radial quantum number. It is related to the quantum numbers of spin and orbital
angular momentum, and to the principal quantum number n, by
nr + j + 1 = nr + l + 1 l = n j .

(6.163)

In terms of j , the energies obey the same equation as in (6.149), so that we obtain
Enj = q

M
1 + Z 2 2 /(n j )2

(Z)4
(Z)2 3 (Z)4
+

+ O(Z 6 6 ) .
= M 1
2n2
8 n4
n3 (2j + 1)
#

"

(6.164)

The condition nr 0 implies that


j n

j n

3
2
1
2

for

j+ = j + 21 j ,

j = j 12 j .

(6.165)

For n = 1, 2, 3, . . . , the total angular momentum runs through j = 12 , 23 , . . . , n 21 .


The spectrum of the hydrogen atom according to the Dirac theory is shown in
Fig. 6.1. As a remnant of the O(4)-degeneracy of the levels with l = 0, 1, 2, . . . , n 1
and fixed n in the Schrodinger spectrum, there is now a twofold degeneracy of levels
of equal n and j, with adjacent l-values, which are levels of opposite parity. An
exception is the highest total angular momentum j = n 1/2 at each n which
occurs only once. The lowest degenerate pair consists of the levels 2S1/2 and 2P1/2 .5
It was an important experimental discovery to find that this prediction is wrong.
There is a splitting of about 10% of the fine-structure splitting. This is called the
Lamb shift. Its explanation is one of the early triumphs of quantum electrodynamics,
which will be discussed in detail in Chapter .
As in the Klein-Gordon case, there are complex energies, here for Z > 137, with
S1/2 being the first level to become complex.
5

Recall the notation in atomic physics for an electronic state: n2S+1 LJ , where n is the principal
quantum number, L the orbital angular momentum, J the total angular momentum, and S the
total spin. In a one-electron system such as the hydrogen atom, the trivial superscript 2S + 1 = 2
may be omitted.

450

6 Relativistic Particles and Fields in External Electromagnetic Potential

Figure 6.1 Hydrogen spectrum according to Diracs theory. The splittings are shown
only schematically. The fine-structure splitting of the 2P -levels is about 10 times as big
as the hyperfine splitting and Lamb shift.

An important correct prediction of the Dirac theory is the presence of fine structure. States with the same n and l but with different j are split apart by the forth
term in Eq. (6.164) MZ 4 4 n3 /(2j+1). For the states 2P1/2 and 2P3/2 , the splitting
is
Z 4 2 2
fine E2P =
M.
(6.166)
32
In a hydrogen atom, this is equal to
fine E2P = 3.10.95 GHz.

(6.167)

Thus it is roughly of the order of the splitting caused by the interaction of the magnetic moment of the electron with that of the proton, the so-called hyperfine-splitting.
This is for 2S 1/2 , 2P 1/2 , and 2P 3/2 levels approximately equal to 1, 1/8, 1/24, 1/60
times 1 420 MHz.6
In a hydrogen atom, the electronic motion is only slightly relativistic, the velocities being of the order c, i.e., only about 1% of the light velocity. If one is not
only interested in the spectrum but also in the wave functions it is advantageous
to solve directly the Dirac equation (6.150) with the gamma matrices in the Dirac
6

See H.A. Bethe and E.E. Salpeter in Encyclopedia of Physics (Handbuch der Physik) 335 ,
Springer, Berlin, 1957, p. 196.

451

6.5 Relativistic Wave Equations in Coulomb Potential

representation (4.544). Multiplying (6.150) by 0 and inserting for 0


Dirac matrix (4.556), we obtain

Z
i

EM +

Z
E+M +
i
r

E (x)
E (x)

= 0.

= the
(6.168)

This is of course just the time-independent version of (4.573) extended by the


Coulomb potential according to the minimal substitution rule (6.126). To lowest
order in , the lower spinor is related to the upper by
E (x) i

 E (x).

(6.169)

2M

We may take care of rotational symmetry of the system by splitting the spinor wave
functions into radial and angular parts
Gjl (r) l
yj,m(, )
i

r
,
E (x) =
F (r)

jl
l
yj,m(, )
x
r

(6.170)

l
where yj,m
(, ) denote the spinor spherical harmonics. They are composed from
the ordinary spherical harmonics Ylm (, ) and the basis spinors (s3 ) of (4.443) via
Clebsch-Gordan coefficients (see Appendix 4C):
l
yj,m
(, ) = hj, m|l, m ; 12 , s3 iYlm (, )(s3 ).

(6.171)

The derivation is given in Appendix 6A.


The explicit form of the spinor spherical harmonics (6.171) is for l = l :

!
l+ m + 12 Yl+ ,m 1 (, )
1
l+
2

yj,m (, ) =
,
(6.172)
2l+ + 1 l+ + m + 21 Yl+ ,m+ 1 (, )
2

l
yj,m
(, )

1
2l + 1

!
l + m + 21 Yl ,m 1 (, )
2

.
l m + 12 Yl ,m+ 1 (, )
2

On these eigenfunctions, the operator L


L

(6.173)

 has the eigenvalues

l
l
 yj,m
(, ) = (1 + )yj,m(, ),

(6.174)

with

1
1
= (j + ), j = l .
(6.175)
2
2
We can now go from Eqs. (6.168) to radial differential equations by using the
trivial identity,

f (r) l+
y

r l,m

 x ( x) i ( ) f (r) yl

l,m ,
+

r2

(6.176)

452

6 Relativistic Particles and Fields in External Electromagnetic Potential

and the algebraic relation Eq. (4.461) in the form

( a)( b) = i (a b) + i(a b),

(6.177)

to bring the right-hand side to

 x (irr i L) f (r) yl

r2

l,m

"

f (r)
f (r)
= ir
i (1 + ) 2
r
r

l
 x yl,m
.
+

(6.178)

In this way we find the radial differential equations for the functions Fjl (r) and
Gjl (r):
Z
d
1
EM +
Gjl (r) = Fjl (r) (j + 1/2) Fjl (r),
(6.179)
r
dr
r


d
1
Z
Fjl (r) =
Gjl (r) (j + 1/2) Gjl (r).
(6.180)
E+M +
r
dr
r
Tosolve these, dimensionless variables 2r/ are introduced, with =
1/ M 2 E 2 , writing


F (r) =

1 E/Me/2 (F1 F2 )(), G(r) =

1 + E/Me/2 (F1 + F2 )(). (6.181)

The functions F1,2 () satisfy a degenerate hypergeometric differential equation of


the form
#
"
d
d2
(6.182)
2 + (b ) a F (a, b; ) = 0,
d
d
and the solutions are
F2 () = l F ( ZE, 2 + 1; ),
ZE
F1 () = l
F ( + 1 ZE, 2 + 1; ).
1/ + ZE

(6.183)

The constant is Einsteins gamma parameter = 1 v 2 /c2 for the atomic unit
velocity v = Zc. It has the expansion = 1 Z 2 2 /2,
As an example, we write down explicitly the ground state wave functions of the
1/2
1S state:

1
0
v
u

u (2MZ)3

0
1
1+
emZr
t

1S 1/2 , 1 =
1
i .
i 1 cos
1
i
sin
e
2
4
2(1 + 2) (2MZ)

Z
Z
1
1
i
i Z sin e i Z cos
(6.184)
The first column is for m = 1/2, the second for m = 1/2. For small , Einsteins
gamma parameter has the expansion = 1 Z 2 2 /2, and we see that in the
0, the upper componentsqof the spinor wave functions tend to the nonrelativistic
Schrodinger wave function 2 (ZM)3 /4e multiplied by Pauli spinors (4.443).
In general,
l
(6.185)
j,m
(x) = hj, m|l, m; 12 , s3 inlm (x)(s3 ).
The lower, small components vanish.

453

6.6 Green Function in External Electromagnetic Field

6.6

Green Function in External Electromagnetic Field

An important physical quantity of a propagating field is the Green function, defined


as the solution of the equation of motion with a -function source term [recall (1.314)
and (2.403)]. For external electromagnetic fields which are constant or plane waves,
this Green function can be calculated exactly.

6.6.1

Scalar Field in Constant Electromagnetic Field

For a scalar field, the Green function G(x, x ) is defined by the inhomogeneous
differential equation
( 2 M 2 )G(x, x ) = i (4) (x x ),

(6.186)

whose solution can immediately be expressed as a Fourier integral:


Z Z
d4 p
d4 p ip(xx )+i (p2 M 2 +i)
i
ip(xx )
d
e
=
e
.
(2)4 p2 M 2 + i
(2)4
0
(6.187)
A detailed discussion of this function will be given in Subsection 7.2.2.
Here we shall address the problem of calculating the corresponding Green function in the presence of a static electromagnetic field, which obeys the more complicated differential equation

G(xx ) =

[i eA(x)]2 M 2 G(x, x ) = i (4) (x x ),

(6.188)

for which a Fourier decomposition is no longer helpful. For a constant and an


oscillating electromagnetic field, however, this equation can be solved by an elegant
method due to Fock and Schwinger [1].
Generalizing the right-hand side of (6.187), we find the representation
G(x x ) =

d hx|ei [(ieA)

M 2 +i]

|x i.

(6.189)

The integrand contains the time-evolution operator associated with the Hamiltonian
operator

H(x,
i) (i eA)2 + M 2 .
(6.190)
This is the Schroedinger representation of the operator
= H(
H
x, p) = P 2 + M 2 ,

(6.191)

where P p eA (
x) is the canonical momentum in the presence of electromagnetism.
We shall calculate the evolution operator in (6.189) by introducing timedependent Heisenberg position and momentum operators obeying the HeisenbergEhrenfest equations of motion [recall (1.276)]:
h

d
x ( )
x )] = 2P ( )
= i H,
d
h
i

dP ( )
P ( ) = 2eF (
= i H,
x( ))P ( ) + ie F (
x( )).
d

(6.192)
(6.193)

454

6 Relativistic Particles and Fields in External Electromagnetic Potential

In a constant field where F (


x( )) is a constant matrix F , the last term in
the second equation is absent and we find directly the solution


P ( ) = e2eF


where the matrix e2eF




e2eF

P (0).

(6.194)

is defined by its formal power series expansion

= + 2eF + 4e2 F F

2
+ ... .
2

(6.195)

Inserting (6.194) into Eq. (6.192), we find the time-dependent operator x ( ):

x ( ) x (0) =

e2eF 1
eF

P (0).

(6.196)

where the matrix on the right-hand side is again defined by its formal power series
e2eF 1
eF

(2 )3
= 2 + e F F
+ ... .
3!
2

(6.197)

Note that division by eF is not a matrix multiplication by the inverse of the matrix
eF but indicates the reduction of the expansion powers of eF by one unit. This is
defined also if eF does not have an inverse.
We can invert Eq. (6.196) to find
eeF
1
eF
P (0) =
2
sinh eF
"

[
x( ) x(0)] ,

(6.198)

and, using (6.194),


P ( ) = L (eF ) [
x( ) x(0)] ,
with the matrix

1
eeF
L (eF )
eF
2
sinh eF
"

(6.199)

(6.200)

By squaring (6.199) we obtain


P 2 ( ) = [
x( ) x(0)] K (eF ) [
x( ) x(0)] ,

(6.201)

where
K (eF ) = L (eF )L (eF ).

(6.202)

Using the antisymmetry of the matrix F , we can rewrite this as


1
e2 F 2
K (eF ) = L (eF )L (eF ) =
4 sinh2 eF

"

(6.203)

455

6.6 Green Function in External Electromagnetic Field

The commutator between two operators x( ) at different times is


e2eF 1
[
x ( ), x (0)] = i
eF

(6.204)

and
!

e2eF 1
x ( ), x (0) + x ( ), x (0) = i
eF
!
#
"
e2eF e2eF
sinh 2eF
= i
= 2i
,

eF
eF

e2eF 1
+i
eF T

(6.205)

With the help of this commutator, we can expand (6.201) in such a way in powers
of operators x( ) and x(0), that the later operators x( ) come to lie to the left of
the earlier operators x(0) as follows:
H(
x( ), x(0); ) =
x ( )K (eF )
x ( ) x (0)K (eF )
x (0)
i
+ 2
x ( )K (eF )
x (0) tr [eF coth eF ] + M 2 .
2

(6.206)

Given this form of the Hamiltonian operator it is easy to calculate the time evolution
amplitude in Eq. (6.189):

hx, |x 0i hx|eiH |x i.
(6.207)
It satisfies the differential equation

eiH |x i
eiH |x i = hx|eiH eiH H
i hx, |x 0i hx|H
x( ), P ( ))|x , 0i.
= hx, |H(

(6.208)

Replacing the operator H(


x( ), P ( )) by H(
x( ), x(0); ) of Eq. (6.206), the matrix
elements on the right-hand side can immediately be evaluated using the property
hx, |
x( ) = xhx, |,

x(0)|x , 0i = x |x , 0i,

(6.209)

and the differential equation (6.210) becomes


i hx, |x 0i H(x, x ; )hx, |x 0i,
or
hx, |x 0i = C(x, x )E(x, x ; ) C(x, x )ei

d H(x,x ; )

(6.210)
.

(6.211)

The prefactor C(x, x ) contains a possible constant of integration in the exponent


which may have an arbitrary dependence on x and x . The following integrals are
needed:
Z

1
d K(eF ) =
4

1
e2 F 2
= eF coth eF ,
2
4
sinh eF

(6.212)

456

6 Relativistic Particles and Fields in External Electromagnetic Potential

and
Z

sinh eF
sinh eF
= tr log
+ 4 log .
eF
eF

d tr [eF coth eF ] = tr log

(6.213)

these results following again from a Taylor expansion of both sides. The exponential
factor E(x, x ; ) in (6.211) becomes, therefore,
)

1
i
1
sinh eF
E(x, x ; ) = 2 exp (xx ) [eF coth eF ] (xx ) iM 2 tr log
.

4
2
eF
(6.214)
The last term produces a prefactor

det

sinh eF
eF

1/2

(6.215)

The time-independent integration constant is fixed by the differential equation


with respect to x:

[i eA (x)] hx, |x 0i = hx|P eiH |x i = hx|eiH eiH P eiH |x i

= hx, |P ( )|x 0i,

(6.216)

which becomes, after inserting (6.199):


[i eA (x)] hx, |x 0i = L (eF )(x x ) hx, |x 0i,

(6.217)

Calculating the partial derivative we find


i hx, |x 0i = [i C(x, x )]E(x, x ; ) + C(x, x )[i E(x, x ; )]
1
= [i C(x, x )]E(x, x ; ) + C(x, x ) [eF coth eF ] (x x ) E(x, x ; ).
2
Subtracting from this eA (x)hx, |x 0i, and inserting (6.211), the right-hand side of
(6.217) is equal to [i C(x, x )]E(x, x ; ) plus


1
L (eF )(x x ) [eF coth eF ] (x x ) C(x, x )E(x, x ; ). (6.218)
2

Inserting Eq. (6.200), this simplifies to


e
F (x x ) C(x, x )E(x, x ; ),
2

(6.219)

so that C(x, x ) satisfies the time-independent differential equation


e
i eA (x) F (x x ) C(x, x ) = 0.
2

(6.220)

This is solved by

C(x, x ) = C exp ie

x
x

1
A () + F ( x )
2



(6.221)

457

6.6 Green Function in External Electromagnetic Field

The contour of integration is arbitrary since A () A () + 12 F ( x ) has a


vanishing curl:
A (x) A (x) = 0.

(6.222)

We can therefore choose the contour to be a straight line connecting x and x, in


which case the F -term does not contribute in (6.221), since d points in the same
direction of x x as x and F is antisymmetric. Hence we may write for
a straight-line connection


C(x, x ) = C exp ie

x
x

d A () .

(6.223)

The normalization constant C is finally fixed by the initial condition


lim hx, |x 0i = (4) (x x ),

(6.224)

which requires
C=

i
.
(4)2

(6.225)

Collecting all terms we obtain


x
i
1/2 sinh eF

hx, |x 0i =
d
A
()
det
exp
ie

(4 )2
eF
x


i
exp (xx ) [eF coth eF ] (xx ) iM 2 .
4

(6.226)

For zero field, this reduces to the relativistic free-particle amplitude


i (x x )2
i
exp

iM 2 .
hx, |x 0i =
2
(4 )
2
2
"

(6.227)

According to relation (6.189), the Green function of the scalar field is given by
the integral
Z

d hx, |x 0i.
(6.228)
G(x, x ) =
0

The functional trace of (6.226)


Trhx, |x 0i = V t

i
eE
2
(4 ) sinh eE

(6.229)

will be needed below. Due to tranlation invariance in spacetime, it carries a factor


total spatial volume V total time t of the universe.
The result (6.229) can be checked by a more elementary derivation [96]. We
let the constant electric field point in the z-direction, and represent it by a vector
potential to have only a zeroth component
A3 (x) = Ex0 .

(6.230)

458

6 Relativistic Particles and Fields in External Electromagnetic Potential

Then the Hamiltonian (6.191) becomes


=
2 + (
H
p20 + p
p3 + eEx0 )2 + M 2 ,

(6.231)

where p are the two-dimensional momenta in the xy-plane. Using the commutation
rule [p0 , x0 ] = i, this can be rewritten as
= eip0 p3 /eE H
eip0 p3 /eE
H

(6.232)

is the sum of two commuting Hamiltonians:


where H
= (
+ H
.
H
p20 e2 E 2 x20 ) + p2 + M 2 H
E

(6.233)

The first is a harmonic Hamiltonian with imaginary frequency E = ieE and an


energy spectrum 2(n + 1/2)ieE. The second describes a free particle in the xyplane. This makes it easy to calculate the functional trace. We insert a comlete set
of momentum states on either side of (6.207), so that the functional trace becomes
Trhx, |x 0i =

d x

d4 p
(2)4

d4 p i(pp )x

e
hp|ei (HE +H ) |p i
4
(2)

(6.234)

The matrix elements are

hp|ei H |p i = eip0 (x0 +p /eE) hp0 |eisHE |p0 iei (p +M


(2)2 (2) (p p )(2)(p3 p3 ).

2 i)

eip0 (x0 +p

3 /eE)

(6.235)

Inserting this into (6.234) and performing the integrals over the spatial parts of p
appearing in the -functions of (6.235) yields
d2 p i (p2 +M 2 i)

e
(2)2
Z
dp0 dp3 dp0 i(p0 p0 )(x0 +p3 /eE)

e
hp0 |eisHE |p0 i,

3
(2)

Trhx, |x 0i = V

dx0

(6.236)

which can be reduced to


i i (M 2 i) eE
e
Trhx, |x 0i = V t
4
2

"Z

dp0

hp0 |ei HE |p0 i .


2

(6.237)

The expression in brackets is the trace of ei HE , which is conveniently calculated


in the eigenstates |ni of the harmonic oscillator with eigenvalues 2(n + 1/2)E :

i H
E

Tre

ei 2(n+1/2)eE =

n=0

i
1
=
.
2 sin E
2 sinh eE

(6.238)

Thus we obtain
Trhx, |x 0i = V t

i
eE
,
4(2)2 2 sinh eE

(6.239)

459

6.6 Green Function in External Electromagnetic Field

6.6.2

Dirac Field in Constant Electromagnetic Field

For a Dirac field we have to solve the inhomogeneous differential equation


{i [ eA (x)] M} S(x, x ) = i (4) (x x ),

(6.240)

rather than (6.188). The solution can formally be written as


x ) = i (4) (x x ),
S(x, x ) = {i [ eA (x)] + M} G(x,

(6.241)

x ) solves a slight generalization of Eq. (6.188):


where G(x,

e
x ) = i (4) (x x ),
[i eA(x)]2 F M 2 G(x,
2

(6.242)

which is the Green function of the Pauli equation (6.111), in natural units. For
a constant field, the extra term enters the final result (6.241) in a trivial way if
we recall the relations to the Green function (6.189) and (6.228), which imply that
x ) contains the fields as follows:
G(x,
x ) =
G(x,

e
d exp i F hx, |x 0i.
2


(6.243)

Constant Electric Background Field


For a constant electric field in the z-direction, we choose the vector potential to have
only a zeroth component
A3 (x) = Ex0 .
(6.244)
Then, since F 30 = E, we have F3 0 = E and F0 3 = E, the field tensor F is
given by the matrix

F = E

0
0
0
1

0
0
0
0

0
0
0
0

1
0
0
0

= iE M3 ,

(6.245)

where M3 is the generator (4.60) of pure Lorentz transformations in the z-direction.


The exponential eeF is therefore equal to the boost transformation (4.59) B3 () =
eiM3 with a rapidity = E . Then we find from (4.14) the explicit matrices

cosh eE
0
0
sinh eE

0
1
0
0

0 sinh eE
0
0
1
0
0 cosh eE

(6.246)

0
0
0
sinh eE

0
0
0
0

0 sinh eE
0
0
0
0
0
0

(6.247)

eeF =
and hence

sinh eF =

460

6 Relativistic Particles and Fields in External Electromagnetic Potential

sinh eF
=
eF

and

sinh eE
eE
0
0

1
0

0
1

0
0
sinh eE
eE

0
1
0
0

0
0
0
0
1
0
0 coth eE

eF coth eF = eE
Thus we obtain

coth eE
0
0
0

(6.248)

(6.249)

x
i
eE
hx, |x 0i =
d A ()
(6.250)
exp
ie
2

(4 ) sinh eE
x
i
2
0
0
T 1
T
3
3
e 4 [(xx ) eE coth eE (xx ) +(xx ) (xx ) +(xx ) eE coth eE (xx ) ]iM ,

where the
superscript T indicates transverse directions to E. The prefactor
Rx
exp [ie x d A ()] is found by inserting (6.244) and integrating along the straight
line
= x + s(x x ), s [0, 1],
(6.251)
to be

exp ie

x
x

d A () = eieE(x0 x0 )

R1
0

ds[z +s(zz )]

= eieE(x0 x0 )(z+z ) . (6.252)

The exponential prefactor in the fermionic Green function (6.243) is calculated


in the chiral representation of the Dirac algebra where, due to (6.112) and (6.113)
e
exp i F
2


0
eeE
0 eeE

= exp (e E ) =

(6.253)

which is equal to
e
exp i F =
2


cosh eE sinh eE E
0

0
.
cosh eE +sinh eE E
(6.254)

Comparison with (4.503) shows that this is the Dirac representation of a Lorentz
boost with in the direction of E with rapidity = 2e|E| . The Dirac trace of the
evolution amplitude for Dirac fields is then simply
trhx, |x 0i =

i
eE
4 cosh eE,
2
(4 ) sinh eE

(6.255)

and the functional trace of this carries simply a total spacetime volume factor V t
as in the scalar expression (6.229).

6.6 Green Function in External Electromagnetic Field

461

Note that the Lorentz-transformation (6.254) has twice the rapidity of the transformation (6.246) in the defining representation, this being a manifestation of the gyromagnetic ratio of the electron in Diracs theory being equal to two [recall (6.120)].
The process of pair creation in a space- and time-dependent electromagetic field
is discussed in Ref. [3].
The above discussion becomes especially simple in 1+1 spacetime dimensions,
the so-called massive Schwinger model [4].

6.6.3

Dirac Field in Electromagnetic Plane-Wave Field

The constant-background field results in the last subsection simplify drastically if


electric and magnetic fields have the same size and are orthogonal to each other.
This is the case for a traveling plane wave of arbitrary shape [5] running along some
direction n with n2 = 0. If denotes the spatial coordinate along n, we may write
the vector potential as
A (x) = f (), nx.
(6.256)

where is some polarization vector with the normalization 2 = 1 in the gauge


n = 0. The field tensor is
F = f (),

n n ,

(6.257)

where the constant tensor satisfies


n = 0,

= 0,

= n n .

(6.258)

The Heisenberg equations of motion (6.192), (6.194) take the form


h

d
x ( )
x )] = 2P ( )
= i H,
(6.259)
d
h
i
dP ( )
)). (6.260)
)) + e n f ((
P ( ) = 2e P ( )f ((
= i H,
d
2

Note that the last term in (6.194) vanishes for a sourceless plane wave: F = 0.
Multiplying these equations by n we see that
n

d ( )
= 2n P ( ),
d

dP ( )
= 0.
d

(6.261)

Hence
nP ( ) = nP (0) = const,

) (0)
= n
(
x( ) n
x(0) = 2 nP ( ).

(6.262)

Whereas the components of P ( ) parallel to n are time independent, those orthogonal to n have a nontrivial time dependence. To find it we multiply (6.260) by
and find

d
= en f ()(2n

d = en df () , (6.263)
P ( ) = 2e P f ()
P ) = en f ()
d
d
d

462

6 Relativistic Particles and Fields in External Electromagnetic Potential

which is integrated to

+ C ,
P ( ) = en f ()

(6.264)

with an operator integration constant C which commutes with the constant nP


and satisfies n C = 0 and and
) (0)

(
C = n (nP ) = n
.
2

(6.265)

Inserting this into (6.260) and integrating that equation yields


1
+ e2 n f 2 () + e n f ()
+D
,
2eC f ()
P ( ) =
2n
2


(6.266)

is again an interaction constant commuting with nP . With this, we can


where D
P and find
integrate the equation of motion (6.259) over d = d/2n
1
1
[
x( ) x(0)] =
2
(2nP )2

)
(

+D
+ e2 n f 2 () + e n f ()
.
d 2eC f ()

2
(0)
(6.267)
, which is reinserted into (6.266) to yield
This determines D
Z

1
P ( ) =
[
x ( ) x (0)]
2


Z (
)

e
2
2

h
d 2eC f () + e n f () + n f ()
i2

2
(0)
) (0)

)) . (6.268)
)) + e2 n f 2 ((
)) + e n f ((
+
2eC f ((

2
( ) (0)


Multiplying this by , and recalling (6.258) and (6.265), we find


1
[
x ( ) x (0)] +
2
Z (
)
en
+ en f ((
)).

d f ()

(0)
( ) (0)

P ( ) =

(6.269)

Inserting this into (6.264) determines the integration constant C :


Z (
)
en
1

d f ().
[
x ( ) x (0)]
C =

2
(0)
( ) (0)

(6.270)

It is useful to introduce the notation


1
hf i

( ) (0)
and

)
(

(0)

d f ().

h (f )2 i h (f hf i)2 i = h f 2 i h f i2 .

(6.271)

(6.272)

463

6.6 Green Function in External Electromagnetic Field

In order to calculate the matrix elements


0i = x P 2 + e F + M 2 x 0 ,
hx |H|x


2

(6.273)

we must time-order the operators x( ), x(0). For this we need the commutator
[
x ( ), x (0)] = 2i g .

(6.274)

This is deduced from Eq. (6.268) by commuting it with x( ) and using the trivially
), x ( )] = 0 as well as the nonequal-time
vanishing equal-time commutator [(

), x (0)] = 0, which
commutator [(0),
x ( )] = 2in . From the latter follows [(
is also needed for time-ordering. The result is



0i = 1 (x x )2 2 i + M 2 + e2 h()2 i 2 + e f () f ( ) .
hx |H|x
4 2


(6.275)

Integrating this in we obtain the exponential factor of the time-evolution amplitude


(6.211):
(


2
1
i
f () f ( )
E(x, x ; ) = 2 exp (xx )2 + M 2 +e2 h(f )2i i e
.

4

(6.276)

The time-independent prefactor C(x, x ) is again determined by the differential equation Eq. (6.216), which reduces here to
(x x )
h f i f () hx, |x 0i,
[i eA (x)] hx, |x 0i =

#

"

(6.277)

and is solved by
i
C(x, x ) =
exp ie
(4)2

dy

(xx )
A (y)

"Z

ny

#)!

f (y )
f (ny)
.
dy
ny
(6.278)

For a straight-line integration contour, the second term does not contribute, as
before.
Observe that in Eq. (6.276), the mass term M 2 is replaced by
2
Meff
= M 2 + e2 h(f )2 i,

(6.279)

implying that in an electromagnetic wave, a particle acquires a larger effective mass.


If the wave is periodic with frequency and wavelength = 2c/, the right-hand
side becomes M 2 + e2 h f 2 i. If the photon number density is , their energy density
is (in units with h
= 1), and we can calculate
e2 h f 2i = 4

h E 2i

= 4 .
2

(6.280)

464

6 Relativistic Particles and Fields in External Electromagnetic Potential

Hence we find a relative mass shift:


M 2
= 4
2e ,
M2

(6.281)

is the Compton wavelength of the


where e h
/Me c = 3.861592642(28) 103 A
electron. For visible light, the right-hand side is of the order of
A3 /100. Present
lasers achieve energy densities of 109 W/sec corresponding to a photon density
=

W
1 eV
1
109
2.082 107
.

sec

A3 h

(6.282)

which is too small to make M 2 /M 2 observable.

Appendix 6A

Spinor Spherical Harmonics

Equation (6.171) defines spinor spherical harmonics. In these, an orbital wave function of angular momentum l is coupled with spin 1/2 to a total angular momentum
j = l 1/2. For the configurations j = l + 1/2 with m2 = 1/2 the recursion
relation (4C.20) for the Clebsch-Gordan coefficients hs1 m1 ; s2 m2 |smi, simplifies by
having no second term. Inserting s1 = l , s2 = 1/2 and s = j = l + 1/2, we find
hl , m + 21 ; 21 , 12 |l + 12 , mi =

v
u
u l
t

m + 1/2
hl , m 12 ; 21 , 12 |l + 12 , m1i.
l m + 3/2

(6A.1)

This has to be iterated with the initial condition


hl , l ; 21 , 21 |l + 12 , l 21 i = 1,

(6A.2)

which follows from the fact that the state hl , l ; 21 , 12 i carries a unique magnetic
quantum number m = l 1/2 of the irreducible representation of total angular
momentum s = j = l + 1/2. The result of the iteration is
hl+ , m 21 ; 21 , 12 |l+ 12 , mi =

v
u
u l+
t

m + 1/2
.
2l+ + 1

(6A.3)

Similarly we may simplify the recursion relation (4C.21) for the configurations j =
l+ 1/2 with m2 = 1/2 to
hl , m 21 ; 21 , 21 |l + 12 , mi =

v
u
u l
t

+ m + 1/2
hl , m + 12 ; 21 , 21 |l + 12 , m+1i,
l + m + 3/2

(6A.4)

and iterating this with the initial condition


hl , l ; 12 n 21 |l + 12 , l + 21 i = 1,

(6A.5)

465

Notes and References

which expresses the fact that the state hl l ; 12 12 i is the state of maximal magnetic
quantum number m = l + 1/2 of the irreducible representation of total angular
momentum s = j = l + 1/2. The result of the iteration is
hl , m 12 ; 21 , 21 |l+ + 12 , mi =

v
u
u l+
t

+ m + 1/2
.
2l + 1

(6A.6)

Inserting (6A.3) and (6A.6) the expression (6.171) for the spinor spherical harmonic
of total angular momentum j = l + 1/2, which now reads
l

yj,m
(, ) = hl , m 12 ; 12 , 21 |l + 12 , mi Yl m1/2 (, )( 21 )
+ hl m + 21 ; 21 12 |l + 12 , mi Yl m+1/2 (, )( 12 ),

(6A.7)

and separating the spin-up and spin-down components, we obtain precisely (6.173).
In order to find corresponding result for j = l+ 1/2, we use the orthogonality
relation for states with the same l but different j = l 1/2:
hl + 21 , m|l 21 , mi = 0.

(6A.8)

Inserting a complete set of states in the direct product space yields


hl + 21 , m|l, m 12 ; 21 , 21 ihlm 12 ; 21 12 |l 12 , mi
+hl + 21 , m|l, m + 21 ; 21 , 12 ihl, m + 21 ; 12 , 12 |l 12 , mi = 0.

(6A.9)

Together with (6A.3) and (6A.6) we find


hl+ .m 12 ; 21 , 21 |l+ 12 , mi =
hl+ , m + 12 ; 21 , 12 |l+ 12 , mi =

v
u
u l+
t

+ m + 1/2
,
2l+ + 1

v
u
u l+
t

m + 1/2
,
2l+ + 1
(6A.10)

With this, the expression (6.171) for the spinor spherical harmonics written as
l

+
yj,m
(, ) = hl+ , m 12 ; 21 , 12 |l+ 12 , mi Yl,m1/2 (, )( 12 )
+ hl+ , m + 21 ; 21 , 12 |l+ 12 , mi Yl,m+1/2 (, )( 12 )

has the components given in (6.172).

Notes and References


[1] J. Schwinger, Phys. Rev. 82, 664 (1951); 93, 615 (1954); 94, 1362 (1954).
[2] C. Itzykson and J.B. Zuber, Quantum Field Theory, McGraw-Hill (1985).
[3] H. Kleinert, R. Ruffini, and X. SheShengPhys. Rev. D 78, 025011 (2008);
A. Chervyakov and H. Kleinert, Phys. Rev. D 80, 065010 (2009).
[4] M.P. Fry, Phys. Rev. D 45, 682 (1992).
[5] C. Itzykson and E. Brezin, Phys. Rev. D 2, 1191 (1970).

(6A.11)

You might also like