Sei sulla pagina 1di 346

2-7 DISPERSION RELATIONS AND CAUSALITY 57

question depends critically on how the lifetime of the metastable state compares
with the collision time, that is, the time interval during which the atom is exposed
to the incident beam. To illustrate this point, let us first consider an experimental
arrangement in which the temporal duration of the incident photon beam is short
compared to the lifetime of the resonance. The energy resolution of such a photon
beam is necessarily ill-defined, A(nar) ) l*, because of the uncertainty principle.
It is then possible to select the excited resonant state by observing the exponential
decay of the resonance long after the irradiation of the atom has stopped. In such
a case, it is legitimate to regard the formation of the metastable state and its sub-
sequent decay as two independent quantum-mechanical processes.
On the other hand, the situation is very different with scattering of a photon
beam of a very well-defined energy. If the energy resolution of the incident beam is
so good (or if the lifetime of the resonant state is so short) that the relation
A(har) ( l" is satisfied, it is not possible to select the excited resonant state by
a delayed lifetime measurement or any other measurement without disturbing the
system. This is because the uncertainty principle demands that such a monochro-
matic photon beam must last for a time interval much longer than the lifetime of
the resonance. We can still determine the lifetime of the resonance by measuring
the decay width l" through the variation of the scattering cross section as a func-
tion of ar, using monochromatic beams of variable energies. Under such a cir-
cumstance thq interference effect between the nonresonant amplitudes and the
resonant amplitude discussed earlier can become important. In such a case, it is
meaningless to ask whether or not a particular photon has come from the resonance
state. If we are considering an everlasting, monochromatic, incident beam, re-
sonance fluorescence must be regarded as a single quantum-mechanical process
so long as the atom is undisturbed, or so long as no attempt is made to determine
the nature of the intermediate states.
In nuclear physics it is sometimes possible to obtain a sufficiently monochro-
matic photon beam so that we can study the Lorentzian energy dependence of the
cross section in the vicinity of a resonance whose lifetime is 5 10-t0 sec. For a
resonance with lifetime l0-10 sec, the mean life may be more readily determined
=
by observing the exponential decay. For a nuclear level with lifetime
- 10-r0 sec
(corresponding to a decay width of about 10-? eV) we can determine the lifetime
both by a delayed lifetime measurement and by a study of the energy dependence of
the cross section. Whenever both procedures are available, the agreement between
the lifetimes determined by the two entirely different methods is quite satisfactory.

2-:1. DISPERSION RELATIONS AND CAUSALITY

Real and imaginary parts of the forward scattering amplitude. In this section we shall
discuss the analytic properties of the amplitude for the elastic scattering of
photons by atoms in the forward direction, with no change in polarization. We
defined the coherent forward scattering amplitude/(ar) by
(do I d O)u:q,.(a) :.(c') : f @)l'
I (2.ret)
58 THE QUANTUM THEORY OF RADIATION 2-7

The sign of/(or) is taken to be opposite to that of the amplitude inside the absolute
value sign of (2.188). This is because the amplitude computed in Born approxima-
tion is positive if the interaction is attractive and negative if it is repulsive.f
The explicit form of f(a) can be obtained directly from (2.188) if the scatterer is
assumed to be the ground state of a single-electron atom:

l@)- -',[' - +4(ffi. #?#)) (2 ts2)


We now evaluate the real and imaginary parts of f(a).In atomic physics the level
width I, is of the order of l0-? eV, while the level spacing is of the order of 1 eV.
Hence in computing the real part we may as well ignore the presence of lr. This
approximation breaks down only in very narrow intervals near ha : E, - Eo.
Using the trick already employed in obtainingQ.l65), we have

: t 2l.oa' l(p: e@)),l=f,


Re [/(ar)] ( 2.193)
i mharn(os\.E- i.o')'
where h^ro: Er
- EA.In computing the imaginary part we can make anarrow-
width approximation in which the very sharp peak at ha = Er - Ea is replaced
by a 8-function. Thus

Imt/(a,)l :#4m
: ? ffitfe
.e'o'),nl'E(r,,
- a), (2.re4)
where we have used

-1,t €' : zE(x).


[m
X'
e-o
(2.tes)

We note that/(ar) acquires an imaginary part only when there is an intermediate


state into which the initial system y + A can make a transition without violating
energy conservation.
Because of (2.193) and (2.194) the real and imaginary parts of f(a) satisfy the
relation
Retf(a,)l:2+s-tffis (2.1e6)

This relation is known as the dispersion relation for scattering of light. It is equiva-
lent to a similar relation for the complex index of refraction (to be discussed later)
written by H. A. Kramers and R. Kronig in 1926-1927; hence it is sometimes
referred to as the Kramers-Kronig relation.$ The relation (2.196) can be shown
to be a direct consequence of a very general principle, loosely referred to as the
causality principle which we shall discuss presently. Its validity is independent of
the various approximations we have made, such as the narrow width approxima-
tion for states d the dipole approximation for the emission and absorption matrix
elements, and the perturbation-theoretic expansion.

f Merzbacher (1961), pp. 491493.


$The history of the Kramers-Kronig relation was traced back by M. L. Goldberger
to an1871 paper by Sellmeier.
2-7 DISPERSION RELATIONS AND CAUSALITY 59

Causality and analyticity. What do we mean by the causality principle ? In the


classical wave theory it simply means that no signal can travel with a velocity faster
than that of light. When applied to a scattering problem, it amounts to the require-
ment that no outgoing disturbance shall start until the incoming disturbance hits
the scatterer. In quantum field theory "causality" is often equated with the vanish-
ing of the commutator of two field operators taken at two space-time points sepa-
rated by a spacelike distance (cf. Eq. 2.85). As mentioned in Section 2-3, this
requirement, technically referred to as local commutativity, is directly related
to the notion that two measurements cannot influence each other if they are made
at a spacelike separation. Historically, the connection between the classical cau-
sality principle and the Kramers-Kronig relation was discussed by R. Kronig,
N. G. van Kampen, and J. S. Toll. In 1953 M. Gell-Mann, M. L. Goldberger, and
W. Thirring proved the dispersion relation for the scattering of light within the
framework of quantum field theory.
The derivation of the dispersion relation from the requirement of local
commutativity in quantum field theory is beyond the scope of this book. Instead,
we shall now demonstrate how the dispersion relation can be obtained from the
causality principle in the classical wave theory.
In studying the properties of the coherent forward amplitude for scattering of
light, we find that the photon spin is an inessential complication. So we shall
consider a scattering problem within the framework of the classical field theory of
a scalar zero-mass field f. Let us suppose that there is an incident monochromatic
wave whose propagation vector k is oriented along the z-axis. The asymptotic form
of the wave at large r is

$ - stu.*-t,t * J'@;0)stlklr-i,,tl, : tiat(z/c) -,r + .f(a;0)etalrr/ct-t)f ,, (2.197)

where f(r;O) is the scattering amplitude. We also use the notation /(ar) for the
forward scattering amplitude as in (2.191),

1@) : f(r; o)l':0. (2.1e8)

If we are to take advantage of the causal connection between the outgoing


wave and the incident wave, it is not so convenient to use a plane wave for the
incident disturbance. This is because the plane wave is everlasting both in space
and time. It is much better to use an incident wave train with a very sharp front.
For this reason we consider a 8-function pulse moving in the positive z-direction
with the property that the disturbance vanishes everywhere except at z : ct.
Because of the well-known relation

u(; -') : + t* *,'^""d-il 6r,, (2.tee)

the 8-function pulse contains negative as well as positive frequency components.


To obtain the scattering amplitude for negative values of or we exploit the reality
of the wave equation which tells us that, if { is a solution to the scattering problem,
then {* is also a solution with the same scatterer. Clearly, the form (2.197) is
preserved if we set

f(-^;0) : f*(^;0). (2.200)


60 THE QUANTUM THEORY OF RADIATION 2-7

Scatterer Scatterer

6-function d-function
disturbance disturbance
(a) (b)

Fig. 2-5. Causal relation between the scattered wave and the incident pulse: (a) before
scattering, (b) after scattering.

Using the superposition principle, we can write an asymptotic solution to the


scattering problem when the incident wave is given by a E-function pulse instead
of a plane wave. All we have to do is integrate (2. 197) with respect to or. Denoting
by .F(x, r) the outgoing wave at x (characterized by r,0, and S), we have the asymp-
totic form

u(: - * F(x, t) -+ Il-*""""d-tt 6J^ 1 i I --ffa;o)st'Vr/c)-tl


6!r.
')
(2.20t)
Initially, we have only the first term, the incident E-function pulse whose "trajec-
tory" is given by z : cl. The scattered wave or the outgoing disturbance should
not start until the incident pulse hits the scatterer; otherwise the causality principle
would be violated (Fig. 2-5).This requirement takes a particularly simple form in
the strictly forward direction t : z, cos d : l. We have
F(xrt)l* on thez-axis
: 0 for z ) ct, (2.202)

which because of (2.201) implies the following very stringent requirement for the
Fourier transform of the forward scattering amplitude:

i(i : : for r(0.


[*_*f{^)r-i" da
lIl(2r)t/'z1 o (2.203)

Assuming that /(r) is square integrable, we can invert the Fourier transform as
follows:

-f(^) : ltlQr)','l I*_f


(")ei" dr : ltl(2r),r21 !* I{ir*, ar. Q.204)

In other words, /(ar) is the Fourier transform of a function which vanishes for
r(0.
So far.f(o) is defined only for real values of ar. To study its analytic properties
we now define f(a) in the entire upper half of a complex ar-plane by (2.20$, that is,

f(r,*iro):Ul(2r7'r'1 li if">"'@'i+L@1)r clr ^ it-oo ( @, 1**,


for
[ 0{-r,tn < *oo,
(2.20s)
2-7 DISPERSION RELATIONS AND CAUSALITY 61

where or, and ai &ra the real and imaginary parts, respectively, of the complex
angular frequency co. Clearly, "f(r) with ar; ) 0 is better behaved than /(ar) with
ar purely real because of the exponential damping factor e-''' with z ) 0. Thus
/(ar) is a function analytic in the entire upper half of the ar-plane.
We now consider a line integral

g$)!-e-, (2.206)
J a
-O),
whose contour is shown in Fig. 2-6. The
point c'.r, is an arbitrary point on the real
axis, and the line integral along the real
axis is to be understood as the principal
value integral
Fig.2-6. Contour of the line integral

J:-: g U--'* J;..] e2o7)


(2.206).

There is no contribution from the semicircular path if l./(r) I - 0 at infinity faster


than l/lar l. Using the Cauchy relation we obtainf

f(,,): Itti f*-*fP)!".


J a', -
(r),
(2.208)

This leads to a pair of relations

Re [/(a,)]
_ I i- Imlf (a')l da'
tt J-* a'-0)
(2.20e)
r f- Relf(a,)lda,
Im[/(a,)] :-;J__ffi

where, although we have dropped subscripts r,the frequencies ar and a' are to be
understood as real. Whenever a function /(ar) satisfies (2.209), the real and imagi-
nary parts of/(ar) are said to be the Hilbert transforms of each other.$
In the realistic case of the scattering of light by atoms the coherent forward
scattering amplitude/(co) does not vanish at infinity. But we can repeat the fore-
going analyticity argument for the function f(co)la. Provided thatl(0) : 0 [which
is satisfied for the scattering of light by a bound electron; cf. Rayleigh's law (2.167)l
we have

Rer/(o)r :+r_t+#l# (2.2t0)

{In (2.208),11il rather than llLriappears because the point ro" lies right on the contour;
see Morse and Feshbach (1953), p. 368.
gTitchmarsh (1937), pp. ll9-128, gives a mathematically rigorous discussion of the
connection among (a) the vanishing of the Fourier transform for r { 0, (b) the analytical
extension of the function of a real variable into the entire upper half plane, and (c) the
Hilbert transform relation.
62 THE QUANTUM THEoRY oF RADIATIoN 2-7

Because of the reality condition (2.200) this can be rewritten as

Re t/(o)r : +[ Igg92
tr+r - o,'
_ 2a' S* Imf (r.o') d ^' -* ^)
(2.21t)
- ,t J o aGit -A'
which is identical to (2.197).*

Index of refraction; the optical theorem. The above relation for the scattering ampli-
tude can be converted into an equivalent relation for the complex index of refractton.
To demonstrate this, we first derive a relation between the forward scattering
amplitude and the index of refraction. Let us consider a plane wave siaz/c lslmally
incident on a slab of an infinitesimal thickness 8. (We are omitting the time de-
pendence e-i't throughout.) The slab, assnmed to be in the xy-plane, is made up
of N scatterers per unit volume. It is not difficult to show that the transmitted
wave at z : / is given by
,,',rc11 *T*af(,)], (2.2r2)

for sufficiently large l) cla. The first term, of course, is due to the original inci-
dent wave. To prove that the waves scattered in the various parts of the slab
actually add up to the second term, we simply note that the expression
f* -i1Q/s1\/-P2+12
NE I +f@,0)2rpdp
JU A/P -TI
(2.2r3)

[where p' : x2 + !', 0: tan-t (pll)l can be integrated by parts to give the second
term of (2.212) plus a term that varies as (c/arl), which is negligible for large /.$
Consider now a slab of finite thickness D. The phase change of the transmitted
wave can be computed by compounding the infinitesimal phase change given by
(2.212). We have

4]!Y@l' :
;g [t * e(2?,icN/6)r(a\I) (2.2r4)

This should be compared to etan(a)o/',where r(ar) is recognized to be the complex


index of refraction. In this manner, we obtain an equation that relates the micro-
scopic quantity, the forward scattering amplitude, to the macroscopically measur-
able quantity, the complex index of refraction:
n(a): I i 2r(cla)'Nf(a) (2.21s)
(first derived by H. A. Lorentz).

IIf /(ar) does not vanish at a:0, simply replace Re[/(ar)] by Rel,f@) -/(0)1. Inthis
connection we mention that /(0) for the scattering of light by any free particle of mass
M and electric charge q can rigorously be shown to be given by the Thomson amplitude
-(q2l4rMcz).
gstrictly speaking, the integral (2.213) is ill defined because the integrand oscillates with
a finite amplitude even for large values of p. Such an integral can be made convergent
by setting a --+ a * ie, where e is an infinitesimal real positive parameter.
2-7 DISPERSION RELATIONS AND CAUSALITY 63

It is now evident that we can write an integral representation for n(ar) analogous
to (2.211):
c l-a(:t')daJ,
Re[n(ar)] :taZ 7t Irr m In(a')l da'
--1 2
a
-a2
:l+ 7t Jo @
Q.Zl6\
where a(ar) is the absorption coefficient defined by
d.(a) : (2alc) Im [n(or)]. (2.217)
The relati on (2.216) is precisely the original dispersion relation derived by Kramers
and Kronig.
As another useful application of (2.215) we can show that the imaginary part
of f(co) is related to the total cross section orur. First let us note that the intensity
of the transmitted wave attenuates as
,-o1o1Nz : lei(.nz/crl, : g-(2o1 (2.2r8)
It then follows from (2.215) that
Im [/(ar)] : (ll4rt)(af c)o,",(a). (2.2te)
If we write it as

o.r(ar):YH@, (2.220)

we see that it is identical in form to the optical theorem (the Bohr-Peierls-Placzek


relation) in wave mechanics which can be derived from the requirement of the
conservation law of probability.f Because of this theorem we can rewrite the
dispersion relation as

Ret/(a,)l :*l-W (2.221)

To avoid any possible misunderstanding we emphasize that the optical theorem


(2.219) is exact, whereas the expression for the imaginary part given by (2.194),
being based on perturbation theory, is approximate. According to (2.194) the
imaginary part of f(ar) vanishes when a * cola; yet (2.188) tells us that there must
be a finite amount of scattering even when a * ar.t. There is no inconsistency
here. The nonresonant cross section obtained by squaring the perturbation ampli-
tude is of the order of ea; hence within the framework of an approximation in which
only terms of order e2 are kept, c61e or equivalently, Im lf@)1, does vanish ex-
cept at a * o)r,r.In this connection we may recall that when we originally inferred
the validity of the dispersion relation using (2.193) and (2.194), we were merely
equating terms of order e2, ignoring the ea contributions. In any event, the approxi-
mation implied in (2.194) is a very good one, since the cross section at resonance
is greater, by many orders of magnitude, than the nonresonant cross section.
The utility of relations such as (2.209) and (2.211) is not limited to the scattering
of light. Similar relations can be written for a wide class of problems in physics
whenever there is a causal connection between the input and the output, or, more

{Merzbacher (l 961), p. 499.


CHAPTER 3

RELATIVISTIC QUANTUM MECHANICS


OF SPIN.+ PARTICLES

3_1. PROBABILITY CONSERVATION IN RELATIVISTIC


QUANTUM MECHANICS
It is almost traditional to start an exposition of the Dirac theory of spin-f par-
ticles by discussing some of the "difficulties" of the Klein-Gordon theory. As we
shall see later, there is actually nothing wrong in the Klein-Gordon equation if
it is properly interpreted. We shall sketch, however, the usual arguments against
the Klein-Gordon equation since they played an important historic role in the
formulation of relativistic quantum mechanics.
In Schrddinger's wave mechanics we associate a complex-valued wave function
rf with a single particle, such that l*l'd'x gives the probability of finding the
particle in a volume element d3x. This interpretation is possible because the proba-
bility density P and the flux density S given by
P: lgl'> o (3.1)
and
S : -(ih l2m)(** V+ - qvf*) (3.2)
satisfy the continuity equation
0Pl0t + V.S:0 (3.3)
by virtue of the Schrodinger equation. Using Gauss' theorem, we also see that the
integral over all space t f dt x is a constant of the motion which can be set to unity
by appropriately normalizing +.
If one is to construct a relativistic quantum mechanics in analogy with non-
relativistic quantum mechanics, it may appear natural to impose the following
requirements on the theory. First, with the relativistic wave function we must be
able to construct bilinear forms which can be interpreted as the probability density
and the flux density satisfying a continuity equation of type (3.3). The probability
density we form must, of course, be positive definite. In addition, the special theory
of relativity requires that P must be the fourth component of a four-vector density.
To see this last point, we recall that d3 x "--+ ff xrF@ under a Lorentz
transformation because of the well-known Lorentz contraction of the volume
element; if Pdsx is to remain invariant, it is essential that P transforngglike the
fourthcomponentofafour-vectorP.---.Pl\m.Thecontinuityi:quation
(3.3) takes the following covariant form,
(0l7xr)sr : 0, (3.4)
where
Jr, : (S, icP). (3.5)

75
76 RELATIVISTIC QUANTUM MECHANICS OF SPIN-+ PARTICLES 3-1

Let us now see whether the relativistic quantum mechanics based on the Klein-
Gordon equation satisfies the above requirements. Consider a four-vector density
given by
su: A(0"#-So)'
\' 0x, 0x* ' /
(3.6)

where { is a solution to the free particle Klein-Gordon equation, and A is a mul-


tiplicative constant. The four-divergence of (3.6) vanishes,

h: ^lH #-(rd.)d - d*rd - H Pr) : o, (3.7)

by virtue of the Klein-Gordon equation. For a Klein-Gordon particle moving


at nonrelativistic velocities (E = mc'),

6 - *'-imc2tih' (3.8)

where rfn is the corresponding Schrodinger solution (cf. Problem l-2). The com-
ponents of s, are then given by

ro : -fs n = (2imclh)Al^hY, s: A["h*V^h - ryf-),rh]. (3.9)

If we set A : -ihl2m, then s and rs a.re precisely the flux density and c times
the probability density in the Schrtidinger theory. Thus we obtain a four-vector
current density from a solution of the Klein-Gordon equation with the following
properties: (i) the current density satisfies the continuity equation, and (ii) the
components of the current density coincide with the flux density and c times the
probability density in the nonrelativistic limit.
So far everything appears satisfactory. There is, however, a difficulty in inter-
preting
P:#(+.u# -'#+) (3.10)

as the probability density. In the Schrodinger theory, in which the time derivative
appears only linearly in the wave equation, the sign of the frequency is determined
by the eigenvalue of the Hamiltonian operator. In contrast, because the Klein-
Gordon equation is of second order in the time derivative, both u(x)s-trtru and
u*1x)e*'ut/u are equally good solutions for a given physical situation (cf. Problem
1-3). This means that P given by (3.10) can be positive or negative. We may arbi-
trarily omit all solutions of the form u(x)s-tntru with E < O. But this would be
unjustified because solutions of the form u(x)e-tztru with E > 0 alone do not
form a complete set. It appears that we must either abandon the interpretation
of (3.10) as the probability density or abandon the Klein-Gordon equation al-
together.
Let us analyze the origin of this difficulty a little more closely. From the deriva-
tion of the continuity equation (3.7) we may infer that the appearance of the linear
time derivative in so is unavoidable so long as the wave function satisfies a partial
differential equation quadratic in the time derivative. Perhaps we could avoid
this difficulty if we wrote a relativistic wave equation linear in the time derivative.
3-1 PROBABILITY CONSERVATION

In1928, in what is undoubtedly one of the most significant papers in the physics
of the twentieth century, P.A.M. Dirac succeeded in devising a relativistic waye
equation starting with the requirement that the wave equation be linear in 010t.
Using his equation, known to us as the Dirac equation, he was able to construct
a conserved four-vector density whose zeroth component is positive-definite. For
this reason, from 1928 until 1934 the Dirac equation was considered to be the
only correct wave equation in relativistic quantum mechanics.
In 1934 the Klein-Gordon equation was revived by W. Pauli and V. F. Weiss-
kopf. Their proposal was that, up to a proportionality factor, s, given by (3.6)
be interpreted as the charge-current density rather than as the probability-current
density. As we saw in Section 1-3, an interpretation of this kind is reasonable
in the classical field theory of a complex scalar field. The fact that the sign of so
changes when u*7x)e-tt'lz is substituted for z(x)erEt/h makes good sense if the
negqtive-energy solution is interpreted as the wave function for a particle with
opposite electric charge (cf. Problem 1-3).
The interpretation of s, as the charge-current density is even more satisfactory
for a theory in which a solution to the Klein-Gordon equation is to be interpreted
as a quantized field operator. In analogy with what we did for the complex scalar
field in classical field theory we form a non-Hermitian field operator d(;* 0t)
such that

+:u+' ++ (3.r 1)

where {1 and Qz are Hermitian operators whose properties are given in Problem
2-3. Consider now a four-current operator
j,:'(o- #,- #r) (3.12)

It is easy to show (Problem 3-l) that the fourth component ofTu has the property

(lr[.) - N[-'),
[<i^ttrldsx: e>
(3.13)

where
/lvr(i) - af.raur, (3.r4)
with
ek* : #(of,+ io1!,), .a{,x:
#(a[') +
ia,!)). (3.r5)

Physically, Nt*, and N(-) are the number operators for the Klein-Gordon particle
of charge e and for its antiparticle with charge -e. So the eigenvalue of the opera-
tor expression(3.13) is the total charge of the field. Usually the four-vector7, is
regarded as the charge-current density operator.
As emphasized in Chapter 2, quantum field theory accommodates physical
situations in which particles are created or annihilated. When there are processes
like .y --,7t* * tt- which take place in the Coulomb field of a nucleus, what is
cohserved is not the probability of finding a given particle integrated over all
space but rather the total charge of the field given by the eigenvalue of (3.13).
RELATTVTSTTC QUANTUM MECHANICS OF SPIN-+ PARTICLES 3-2

Coming back to the original argument against the Klein-Gordon equation, we see
that if we are to reject the Klein-Gordon equation on the ground that we cannot
form a positive-definite probability density, we might as well give up the Maxwell
theory which, as the reader may verify, cannot accommodate any conserved
four-vector density bilinear in the electromagnetic field.

3-2. THE DIRAC EQUATION


Derivation of the Dirac equation. Even though the Klein-Gordon equation is
quite satisfactory when properly interpreted, there is reason for rejecting it for
the description of an electron. The Klein-Gordon equation cannot accommodate
the spin-.| nature of the electron as naturally as the Dirac equation can. In this
connection, let us first study how to incorporate the electron spin in nonrelativistic
quantum mechanics.
In nonrelativistic quantum mechpnics, in order to account for the interaction
of the electron spin magnetic moment with the magnetic field, it is customary to
add a term
lI(spin) : -(ehl2mc)o.B (3.16)

to the usual Hamiltonian, as done originally by W. Pauli. This procedure appears


somewhat artificial, especially if we subscribe to the philosophy that the only
"fundamental" electromagnetic interactions are those which can be generated
by the substitutiotr pt,-pt,- eArf c. There is, however, a slightly less ad hoc
way of introducing the spin magn'etic moment interaction. In the usual wave-
mechanical treatment of the electron, the kinetic energy operator in the absence
of the vector potential is taken to be
II(KE) : p2f2m. (3.17)

However, for a spin-f particle we may just as well start with the expression

H$E) : (o.p)(o.p)12m. (3. l 8)

This alternative form is indistinguishable from (3.17) for all practical purposes
when there is no vector potential.f There is, however, a difference when we make
the substitution p --) p -eAlc. The expression (3.18) then becomes

h" ('-+)" ('-+)


: *(' - +)' + M,"'[(' - +), (' -+)]
I (- eA\' €h .,.8^
-m\P-?) -Mo'B' (3.1e)

where we have used


pxA:-ih(VxA)-Axp. (3.20)

(The operator p is assumed to act on everything that stands to the r.ight; in contrast,

fWe recall that the formula (o'AXc'B) : A'B + ic'(A x B) holds even if A and B
are operators.
3-2 THE DIRAC EQUATION

the V operator in (3.20) acts only on A.) Note that the spin magnetic moment
: 2.I
generated in this way has the correct gyromagnetic ratio g
Our object is to derive a relativistic wave equation for a spin--| particle.
Just as we incorporated the electron spin into the nonrelativistic theory by using
the kinetic energy operator (3.18), we can incorporate the electron spin into the
general framework of relativistic quantum mechanics by taking the operator analog
of the classical expression
(E'lc') -P2:(mc)' (3.21)
to be
(+ -o't) (T ro'e) : (mc)" (3.22)

where
g(oo)
- ih L : ihc ;0 , (3.23)
o[ oxo

and p : -ihY as before. This enables us to write a second-order equation (due


to B. L. van der Waerden),
o.tfrv) (,u*,- o.ihv)6 : (mc)zL, (3.24)
(,u
&*
for a free electron, where { is now a two-component wave function.
We are interested in obtaining a wave equation of first order in the time deriv-
ative. Relativistic covariance suggests that the wave equation linear in 0l0t must
be linear in V also. An analogy with the Maxwell theory may now be helpful.
The free-held D'Alembertian equation ZAr: 0 is a second-order equation,
while the free-field Maxwell equation (0l0xr) Fr,:0 is a first-order equation.
Note that F, obtained by differentiating Ar has more components than lu. This
increase in the number of components is the price we have to pay when we work
with the first-order equation.
Motivated by this analogy, we can define two two-component wave functions
0(8) and 0(r).
O(E): *(,0*- in".v)6, 6@ :6. (3.2s)

The total number of components has now been increased to four. The superscripts
R and Z come from the fact that as m --- ,0, dto) and (ltzr, respectively, describe
a right-handed (spin parallel to the momentum direction) and a left-handed (spin
antiparallel to the momentum direction) state of the spin-! particle, as we shall
see later. The second-order equation (3.24) is now equivalent to two first-order
equations
fiho.v - ih(010x)l+(L) : -mc$(nt, (3.26)
f-iho.V - ih(0l0xr)l+(8) : -mc+@)

fHistorically, all this was first obtained by working out the nonrelativistic limit of the
Dirac theory, as we shall show in the next section. For this reason, most textbooks state
that the g : 2 relation is a consequence of the Dirac theory. We have s@n, however,
that the g :2 relation follows just as naturally from the nonrelativistic Schr<idinger-
Pauli theory if we start with the kinetic energy operator (3.18). This point was emphasized
particularly by R. P. Feynman.
RELATTVTSTTC QUANTUM MECHANTCS OF SPrN-+ PARTTCLES 3-2

Note that unless the particle is massless, these first-order equations couple S(E)
and 4rzr just as the Maxwell equations, also first-order equations, couple E and B.
Equation (3.26) is equivalent to the celebrated wave equation of Dirac (cf.
Problem 3-5). To bring it to the form originally written by Dirac, we take the
sum and the difference of (3.26). We then have
:
-ih(o.V)16tnt - d(z)) - ih(0l0xo)(d(') * {tntl -mc(612) * 0(o)), 8.27\
ih(o.Y)(d(') + d(",) * ih(0l0xo)(d,",- d(')) - -mc(+(E) - +,")),
\ /
or, denoting the sum and the difference of d(t) and d(') by *t and r/n.a, we have

l',oolj'.T' *'rli;,],)ffJ
: -mcffJ (3 28)

Defining afour-component wave function 4, bV

{,^: : 1[:]), (3.2e)


ffiJ ffi:;l
we can rewrite (3.28) more concisely as

(r.o + r^dil)+ n T+ : o, (3.30)

or
(',*,*TW:o' (3.31)

where ,yuwith p,: 1,2,304 are 4 x 4 matrices given by

,r:(,0"r -fl ' 'vn: (J jr)' (3.32)

which really mean

nls _l,3:;':\ ^_l,l


:|; '':l:
?::\
-; ' etc.f (3.33)

:, : :,l' : _?,l
Equation (3. 3l) is the famous Dirac equation.$

fWe use the standard form of the 2 x 2 Pauli matrices

"':(? ), ",:(:
The symbol .I stands for the 2x
;)' ",:(l -?)
2 identity matrix

r:(r\o 0\.
tl
gln deriving the Dirac equation we have not followed the path originally taken by Dirac.
The first published account of the derivation given here is found in B. L. van der Waerden's
1932 book on group theory. The original approach of Dirac can be found in many books,
for example, Rose (1961), pp.3944; Bjorken and Drell (1964), pp. 6-8.
31 THE DIRAC EQUATION 81

We wish to emphasize that (3.31) is actually four differential equations that


couple the four components of rfn represented by a single-column matrix

-:ffii) (3.34)

A four-component object of this kind is known as a bispinor or, more commonly,


as a Dirac spinor. If there is any confusion as to the real meaning of (3.31), the
reader may write the matrix indices explicitly as follows:

i, i, [t',r.u {r+ ff)u,u] +u : o. (3.3s)

The fact thatrln has four components has nothing to do with the four-dimensional
nature of space-time; r/rB does not transform like a four-vector under a Lorentz
transformation, as we shall see in Section 3-4.
The 4 x 4 matrices ry, we introduced are called the gamma matrices or the
Dirac matrices. They satisfy the following anticommutation relations as we may
readily verify:
{ry*r"Yr}:'f p"Y, * nlunl,r:2Epr. (3.36a)
For instance,
: (; : (; ?)'
'' -:)'
nt,ntz *o,o, : (,.1,
-;') (!",-';') * (,], -;')[, -f')
: (o'o'["'"' : o' (3'36b)
o,oro*o,o,)
Moreover, each ry, is seen to be
".T;,:;,, G.37)
and traceless.f
Multiplying (3.30) by rya, we see that the Dirac equation can also be written
in the Hamiltonian form
H* : ih(0"lrlat), (3.38)
where
H:-icha.YlBmc', (3.3e)
with
II : (:,
B: ryn:
lo _;), dn: irynryn "J) (3.40)

fln the literature some people define gamma matrices that do not quite satisfy (3.35)
and (3.37). This is deplorable. Our notation agrees with that of Dirac's original paper
and Pauli's Handbuch article. The alternative notation sometimes found in the literature
is summarized in Appendix B.
82 RELATIvIsTIc euANTUM MEcHANIcs oF SPIN-+ PARTIcLES 3-2

The matrices c and B satisfyf


lar, B\ : o, B' : l, lar, at]t : 26rt. (3.41)

In Section 3-6, we shall make extensive use of the Hamiltonian formalism based
on (3.39).

Conserved current. We shall now derive the differential law of current conservation.
First let us define
$: *'ryn, (3.42)

where rp is called the adjoint spinor, as distinguished from the Hermitian conjugate
spinor rln+. Explicitly r| as well as +* can be represented by single-row matrices,
that is, if

d^ - /il;\, (3.43)

\il:/
then
gn : (^hf ,.hr*,
^ht,^ht), (3.44)
^7
: (./rf , n$t, -9f , -./"f ).
To obtain the wave equation for rp we start with the Hermitian conjugate of the
Dirac equation,
(3.4s)
*rrrr+ &*'ryn+T^hr:0.
Multiplying (3.45) by ry, from the right, we obtain the adioint equation

-t^fiv,rT+:0, (3.46)

where we have used


aaa (3.47)
aVf': Txry: - axn'
and ryp,ya : -?f anf*. We now multiply the original Dirac equation (3.31) from the
left by fr, the adjoint equation (3.46) from the right by {r, and subtract. Then
@lax)($ry,f) : o. (3.48)
Thus we see that
s, : ic$,yrt : @"l.nt a$, ir*r *) (3.4e)

tln addition to ryr, d.16,a;rld B, one also finds in the literature

: (: j)' : (,0, -f and o' : (;


o' o'
)' -?)
We shall, however, not use the rho matrices in this book.
THE DIRAC EQUATION

satisfies the continuity equation.f Using Green's theorem we see that

t+r^gosx - !+n^ho'*: const, (3.50)

which may be set to unity by appropriately normalizing +. Now, unlike (3.9),

$ryn^lr: qr./, : (3.s1)


F, {rfr*B
is necessarily positive-definite. So we may be tempted to identify Q"yn^h:,r/nn+
with the probability density as we did in the Schrtidinger theory. With this in-
terpretation,
51,: ic{la1rt : c*t Ao* (3.s2)
is identified with the flux density. For the covariance of the continuity equation,
we must, of course, prove rigorously that s" indeed transforms like a four-vector;
this will be done in Section 3-4.
For the time being, we assume that a solution r/n to the Dirac equation is a single-
particle wave function subject to the above interpretation of rfn+rln as the probability
density. Toward the end of this chapter, however, we shall point out some of the
difficulties associated with this interpretation. For a more satisfactory interpretation
of the Dirac theory, it is necessary to quantize the Dirac field. This we shall do in
Section 3-10.

Representation independence.In concluding this section we shall mention briefly


the notion of the representation independence of the Dirac equation. Suppose
somebody writes

(r,&+T)*' :0, (3.53)

where the only defining property of ry'* is that "y', with p : | . . .4 forms a set of
4 x 4 matrices satisfying
{'Y'*'Y|} - 2E,rr. (3.54)
Our assertion is that this equation is equivalent to the Dirac equation (3.31) where
the matrices y,, are explicitly given by (3.32). Note that we do not assert that for
a given physical situation rfn and $' are the same. Different sets of 4 x 4 matrices
satisfying (3.54) are referred to as sets of the gamma matrices in different repre-
sentations.
We can prove the representation independence of the Dirac equation by appeal-
ing to what is known as Pauli's fundamental theorem: Given two sets of 4 x 4
matrices satisfying
{ru,Y,}:28p, and {"fL,ry',} : 28p,

$A11 this can be readily proved even in the presence of the electromagnetic interaction
generated by the substitution

-ih!OXrt-> -ih!OXp-4r C

in (3.31).
RELATIVISTIC QUANTUM MECHANICS OF SPIN.+ PARTICLES 3-2

with /r, v : | . . . 4,there exists a nonsingular 4 x 4 matrix S such that


SryuS-t - rytr. (3.55)

Moreover, S is unique up to a multiplicative constant. The proof of the theorem


can be found in Appendix C. Assuming the validity of Pauli's theorem, we rewrite
(3.53) as

(tr,t -'&*ry),ss-'1 *' :0, (3.56)

where S is the matrix that relates the set {ryi,} and the set {"yr} via a relation of the
form (3.55). Multiplying by S-' from the left, we obtain

(r,#,*ry)s-'+'- o. (3.57)

This is the same as the original Dirac equation with S-lt' u, solution. In other
words, (3.53) is equivalent to the Dirac equation, (3.31), and the wave functions
r/r' and f are related by
f' : S./n. (3.58)

Let us consider the case where the.y', are also Hermitian. By taking the Hermitian
conjugate of (3.55) we see that S can be chosen to be unitary S+ : S-1. With a
unitary S we see that expressions like the probability density and the flux density
are the same:
$' Y'r*' : *'r'uL''rnl''
: S*Sryn S-t SryrS-t S^h
: ^trnt
fi'Yr*. (3.se)

Evidently all the physical consequences are the same regardless of whether we use
(3.31) or (3.53). Note, however, that the wave functions for the same physical
situation look ditrerent when different representations are used.f
In practice three representations of the gamma matrices are found in the
literature:
a) The standard (Dirac-Pauli) representation given explicitly by (3.32).
b) The Weyl representation in which not only ry7. but also rya are off-diagonal
(cf. Problem 3-5).

fActually we are already familiar with an analogous situation in the Pauli two-component
theory. In nonrelativistic quantum mechanics it is customary to use the standard repre-
sentation of the o matrices in which o, is diagonal. From the point of view of the com-
mutation relations, however, one may as well work with what we may call "a noncon-
formist's representation,"

",: (: ;t), ",: (l -)' ",: (? ;)


The spin-up spinor (the spin in the positive z-direction) is then given by

#(l) ratherthanby (;)


3-3 NONRELATIVISTIC APPROXIMATIONS; PLANE WAVES 85

c) The Majorana representation in which the ry7, are purely real and rya is purely
imaginary, hence ,yr(0l0xr) is purely real.
In this book, whenever explicit forms of Yp or "h are called for, only the standard
(Dirac-Pauli) representation is used.

3-3. STMPLE SOLUTIONS; NONRELATMSTIC APPROXIMATIONS;


PLANE WAVES

Large and small components. Before we stqdy the behavior of Dirac's wave func-
tion rfn under Lorentz transformations, let us examine the kind of physics buried
in the harmless-looking equation (3.31).
In the presence of electromagnetic couplings, the Dirac equation reads

(&- o*o,)"vr^l*T+: o' (3.60)

wheretheusualreplacement-ih@laxr)*-ih(7l7xr)-eArlcisassumedto
be valid. Assuming that Ar is time independent, we let the time dependence of
I b. given by
r/r : 9(*, t)11=rs-iEt/n (3.61)

(which, of course, means that rfn is an eigenfunction of ihAlAt with eigenvalue E).
We can then write the coupled equations for the upper and lower two components,
tdandfr, as follows (Eq. 3.28):

[".('- +)] *u: - eAo - mc')nlrn,


Ir' (3.62)

-1".(' - +)] i,n: -+(E - eAo '1 mc')^lu,


where A* : (A, iA) as before. Using the second equation, we can readily elimi-
nate "ln in the first equation to obtain

[".('-+)]|T=#-]["('-+)]*u=@-eA,-mc,)^!o.(3.63)
Up to now we have made no approximations. We now assume that
E = mcz, leArlKmc'. (3.64)

Defining the energy measured from mc'by


E(NR) : E mct, (3.6s)
-
we can make the following expansion:
c'| I 2mc" I [r --amAt-
EGR)-.eAo-,-...-l
E=AM: -- 2mlml: --l - rml' -
- I
.

(3.66)

This can be regarded as an expansion in powers of (ulc)' since E(NIi) - eA, is


roughly lp - @Alc)l'l2m = mu2f2. Keeping only the leading term in (3.66), we
86 RELATrvrsrrc eUANTUM MEcHANrcs oF sprN-+ pARTrcLEs 3-3

obtain

*".(, - +)".(r - +) *n: (E'<Nn1


- eA)*o, (3.67)

which, as we have already seen (cf. Eq. 3.19), becomes

l*G -+)' - *o.B * ,.t,f*^: E(NR)V^,. (3.68)

Thus to zeroth order in (ulc)z, rf"" it nothing more than the Schrtidinger-Pauli
two-component wave function in nonrelativistic quantum mechanics multiplied
by e-t^"otln. Using the second expression of (3.62), we see that $n is "smaller"
than $e by a factor of roughlv lp - e(Alc)ll2mc = uf2c, provided that (3.64) is
valid. For this reason with E - mc2, nht and nlto are respectively known as the
large and small componenfs of the Dirac wave function r/n.

Approximate Hamiltonian for an electrostatic problem. We shall now study the


consequences of keeping the second term in (3.66). For simplicity, let us treat
the case A : 0. The equation we must work with is
Hfs*": E(NR)V.{, (3.6e)
where

r11,,n; : (o.r)*(r - E-";#)t". p) * eAo. (3.70)

At first sight it might appear that we can regard (3.69) as the time-independent
Schrtidinger equation for r/rr. There are, however, three difficulties with this
interpretation. First, if we are working to order (ulr)', rfn" no longer satisfies the
normalization requirement because the probabilistic interpretation of the Dirac
theory requires that

I t+l+^ + ^hL^h)drx
: t, (3.7t)
where r|" is already of the order of ulc. Second, by expanding (3.70) it is easy to
see that ,F11Nn) contains a non-Hermitian term ihB.p. Third, (3.69) is not an
eigenvalue equation for the energy'since F1flo) contains E(NR) itself.
To overcome these difficulties, let us first note that the normalization require-
ment (3.71) can be written as

I+t(t * +*r,)ged3x--t (3.72)

to order ('ulc)'since, according to the.second expression of (3.62),

*ooH+" (3.73)

This suggests that we should work with a new two-component wave function V
defined by
v - ogt, (3.74)
where
O-l*(p'l8m'c'). (3.7s)
3-3 NONRELATMSTIC APPROXIMATIONS; PLANE WAVES 87

With this choice, V is normalized to order (ulc)'zsince

Jv'va'* = Igltl + (p'l4mzc2)fgndsx, (3.76)

where we have used (3.71). Multiplying (3.69) from the left by O-1 :
1
- (p'/8ffi'c'), we obtain
o-1 HAO-t v : frtNn)Q-z,En. (3.77)
Explicitly, to order (ulr)'we have

l#,* eAo
- {r#", ({*+ ee,)l -+(ry),".n)] v
_E(NR)(r _ (3.78)
#*)*,
or, writing E(NR) p' as f {ErNnr, Pr}, we have
It
lzm
+ r't, - Pn,1
W r g*72([P', (E(*o> - eA,)) - 2(o . p)(E(r{R)
- eAo)(o. n))]V
: f,(Nn)qr. Q.79)
ln general, for any pair of operators ,4 and B, we have
{Ar, B} - 2ABA : fA,lA, Bll. (3.80)
This very useful formula can be employed to simplify (3.79) where we set o.p : A
and E(NR) - eA, - B. Using
lo.p, (E tNn;
- eA)l: -ieho.E, (3.81)
and
: -ehzY 'E - 2eho'(E x p),
lo.p, -ieho'El (3.82)
both of which are valid since Y Ao : -E and V x E : 0, we finally obtain

lr** eA, - ,#" -'0";#i il - #,o."1 v - srNn)e. (3.83)


This equation is free of the difficulties mentioned earlier and can therefore be
regarded as the Schrodinger equation for the two-component wave function.
The physical significance of each term in (3.83) will now be given. The first two
need no explanation. The third term is due to the relativistic correction to the
kinetic energy as seen from the expansion

ffi-mcz: lpl,l2m -(lplnlSm'c,)+ ... (3.84)

The fourth term represents the spin interaction of the moving electron with the
electric field. Crudely speaking, we say this arises because the moving'electron
"sees" an apparent magnetic field given by E x (v/c). Naively, we expect in this
way -(ehl2mc) o.[E x (v/c)1, which is just twice the fourth term of (3.83). That
this argument is incomplete was shown within the framework of classical electro-
dynamics two years before the advent of the Dirac theory by L. H. Thomas, who
argued that a more careful treatment which would take into account the energy
associated with the precession of the electron spin would result in reduction by
88 RELATrvrsrrc eUANTUM MEcHANrcs oF sprN-+ pARTrcLEs 3-3

a factor of two, in agreement with the fourth term of (3.83).1 Hence the fourth
term of (3.83) is called the Thomas term. For a central potential
eAo: Y1r1, (3.85)
we obtain

-#o.(Ex p) #(-+#)".(x x p): #,f,#r." (3.86)

with S : hol2. Thus the well-known spin-orbit force in atomic physics, repre-
sented by (3.86), is an automatic consequence of the Dirac theory.
As for the last term of (3.83) we note that V.E is just the charge density. For
the hydrogen atom where V.E : -s$Gr(x), it gives rise to an energy shift

I #5rsr(x) | f(scrrro) l, d' x : #| "h[i;n.o) l, l*:0,


(3.87)

which is nonvanishing only for the s states [in contrast to (3.86) which affects
all but the s statesl. The last term of (3.86) was first studied in detail by C. G.
Darwin; hence it is called the Darwin term. We shall postpone the physical inter-
pretations of the Darwin term until Section 3-7.
Using the third, fourth, and fifth terms of (3.83) as the perturbation Hamiltonian
and the wave functions for the hydrogen atom in nonrelativistic quantum me-
chanics as the unperturbed wave functions, we can compute the lowest-order
relativistic correction to the energy levels of the hydrogen atom. Since this cal-
culation based on first-order time-independent perturbation theory is straight-
forward, we present only the results: $

AE: -(h)'(#)L,,(fr- *), (3.88)

which is to be added to the unperturbed energy levels


Ee) - (ezlStrannz). (3.8e)

The sum E(0' + AE correctly describes the fine structure of the hydrogen atom
to order (t+")'times the Rydberg energy (e'zl8ftas). Note that states with the
same n but different 7 (for example, 2prp - 2prp) are now split. On the other
hand, states with the same n and the same 7 (for example, 2sv, - 2prn) are still
degenerate. This degeneracy persists even in the exact treatment of the Coulomb
potential problem, as we shall see in Section 3-8.11

{For the derivation of the Thomas factor in classical electrodynamics consult Jackson
(1962), pp.364-368.
$See, for example, Bethe and Salpeter (1957), pp. 59-61.
llln practice it is of little interest to work out an expansion more accurate than (3.88),
using only the Dirac equation with V : -e2l(4rr), because other effects such as the
Lamb shift (discussed in the last chapter) and the hyperfine structure (to be discussed
in Section 3-8) become more significant.
3-3 NONRELATTVTSTTCAPPROXIMATIONS; PLANEWAVES 89

Free particles at rest. Let us now turn our attention to problems where exact
solutions to the Dirac equation are possible. The simplest solvable problem deals
with a free particle. We will first demonstrate that each component of the four-
component wave function satisfies the Klein-Gordon equation if the particle is
free. Although the validity of this statement is rather obvious if we go back to
d(') and d,*,, *. prove it by starting with the Dirac equation (3.31). Multiplying
(3.31) from the left by y,(010x,), we have

*,&,y,"t,,* - (T)'+ -- o. (3.e0)

Adding to (3.90) the same equation written in a form in which the summation
indices p, and v are interchanged, we obtain

*,t^(v,ryr l "vr"v,)t - r(H'+: o, (3.er)

which reduces to
It - (mclh)'z"1r :0 (3.e2)

by virtue of the anticommutation relation of the gamma matrices (3.35). Note


that (3.92) is to be understood as four separate uncoupled equations for each
component of r/r. Because of (3.92), the Dirac equation admits a free-particle
solution of the type
4, - u(p) exp [i(p.xlh) - i(hlQ] (3.e3)
with
P: an/fifVraWVf,, (3.e4)

where a(p) is a four-component spinor independent of x and l. Note that (3.93)


is a simultaneous eigenfunction of -ihY and ih@ftt) with eigenvalues p and d
respectively.
For a particle at rest (p : 0), the equation we must solve is
A , mc,
1[: - 7, Y'.
ryna1trt,.;, (3.es)

Inaccordance with (3.93) and (3.94), we first try the time-dependence r-tmczt/tt
Equation (3.95) then becomes
.#(I _ixt"[]) : -Tfi.[i)' (3e6)

which is satisfied only if the lower two-component spinor uo(O) vanishes. But,
using a similar argument, we see that (3.95) can be satisfied equally well by the
time-dependence e+inczt/h provided that the upper two-component spinor u"(0)
vanishes. As in the Pauli theory, the nonvanishing two-component spinors can be
taken as
/0\
/" and (,t/
\0/
RELATIVISTIC QUANTUM MECHANICS OF SPIN.+ PARTICLES 3-3

So there are four independent solutions to (3.e5):

-imc2t/h
(i) (;)'-,
(3.e7)

(f
fi)".'-', )'.'-"'
If we insist on the interpretation that ih(AlAt) is the Hamiltonian operator, the
first two are "positive-energy" solutions while the last two are "negative-energy"
solutions. Note that the eigenvalues of the Hamiltonian operator are Imc',
depending on whether the eigenvalues of rf r: B are * I ; this also follows directly
from the expression for the Hamiltonian (3.39). In Section 3-9 we shall show that
the existence of negative-energy solutions is intimately related to the fact that
the Dirac theory can accommodate a positron.
Earlier in this section we showed that in the nonrelativistic limit E = mc2,
the upper two-component spinor r/n" coincides with the Schrodinger wave function
apart from ,-imczt/tt. Therefore the first solution in (3.97) is the Dirac wave func-
tion fora particle at rest with spin up, since the eigenvalue of or ir + 1 when applied
," (;). rnit leads us to define a 4 x 4 matrix

s13----\o
nf tnlz
- ryzryt lat 0\
- - o,),
(3.e8)

whose eigenvalue is to be interpreted as the spin component in the positive z-


direction in units of hl2. This interpretation will be justified in later parts of this
chapter where we shall show that
a) the operator Ir/2 is the infinitesimal generator of a rotation about the z-axis
acting on the space-time independent part of the Dirac wave function; and that
b) for a central force problem the sum of x x p and h2l2 is indeed a constant of
the motion, to be identified with the total angular momentum, where I* is
defined as it was in (3.98), that is,

r,: ry: (;' J (iik)cycric.f (3.ee)

Evidently the solutions (3.97) are eigenfunctions of Ie with eigenvalues * I for


the first and third and - 1 for the second and fourth. Thus for each sign of the
energy there are two independent solutions corresponding to two spin states,
as expected from the spin-$ nature of the particle described by the Dirac equation.

fThe notation "(ryk) cyclic" stands for (i, j,k) : (1,2,3),(2,3,1), or (3, 1,2).
3-3 NONRELATTVISTICAPPROXIMATIONS; PLANE WAVES 91

It has sometimes been stated that we need a 2 x 2 : 4 component wave func-


tion in the relativistic electron theory to take into account the two energy states
and the two spin states of the electron for given p. This argument is incomplete
(if not incorrect). To see this we recall that although the wave function in the
Klein-Gordon theory is a single-component wave function, it can accommodate
the two energy states of a spin-zero particle. Furthermore, the two-component
second-order equation (3.24) of Waerden, which we started with, is a perfectly
valid equation for the electron; it can accommodate the two energy and the two
spin states of the electron just as the Dirac equation can.f In this connection
we note an important difference between a differential equation lineat in 010t,
such as the diffusion (heat) equation or the Schriidinger equation, and a differential
equation quadratic in the time derivative, such as the wave equation (for a vibrating
string) or the Klein-Gordon equation. When we solve the problem of the tem-
perature distribution of a potato placed in boiling water, it is sufficient to specify
initially only the temperature of the inside region of the potato; we do not need
to know how fast the potato is warming up. On the other hand, if we want to
predict the time development of a vibrating string for t > 0, it is not sufficient
to know only the displacement of the string at t:0; we must also know how
fast the various parts of the string are moving at t: 0. In general, then, when
solving a partial differential equation quadratic in the time derivative, we must
specify as initial conditions both the function and its time derivative. However,
when solving a partial differential equation linear in the time derivative, we need
specify only the function itself. Thus in the Dirac theory if we know I at t : 0,
we can predict the time development of {n for t > 0. By contrast, the two-com-
ponent equation (3.24) of Waerden is second order in the time derivative; so we
must know both { and A$lAt at t :0 to predict the future development of {.
For an energy eigenstate, specifyine AQIAt (in addition to {) amounts to specifVing
the sign of the energy. The number of independent components we must specify
is four whether we use the four-component Dirac equation or the two-component
Waerden equation.

Plane-wave solutions. Now let us return to the free-particle problem, this time
withp*0.Substituting
*: (il") : (;11]).*p(;n f -,+) (3.100)

into (3.62) with Ao : A : 0, we obtain

uo(P) : gJTp(a'p)',(p), uo(p):


Eh(o'p)',(p)' (3'101)

fOne of the reasons that the two-component formalism based on(3.24) is not used so
extensively as the Dirac theory is that a solution to (3.24) with the electromagnetic in-
teraction added has a rather complicated transformation property under parity (cf.
problem 3-5). Nevertheless, the two-component equation can be used for solving prac-
tical problems in quantum electrodynamics just as the Dirac equation can, as shown by
R. P. Feynman and L. M. Brown.
92 RELATrvrsrrc eUANTUM MEcHANrcs oF sprN-+ pARTrcLEs 3-3

ForE:\/W'mFc)0wemaytry,apartfromanormalizationconstant,
/o\
/tt and
\0/ lt/
for zr(p). We can easily find the lower two-component spinor *t] by using the
second equation of (3.101) if we recall that

o'P: ( n' Pt - iP'\


(3.102)
\Pt * iPz -Ps )
In this way we get two independent solutions for E > 0,

\ / 0

rz(,)(p):*( ; z(')(p) : ? \
p,cl@ + mc')
I\(p'*ip,)cl@*mc')l | "'u "l (p, - ip,)cl(E * mc,1l'
\ -p,cl(E*mc') I
(3.103)
N is to be determined later. For E -
where the normalization constant
-Jw<0,wemaystartwiththelowertwo.componentspinorz'
set to

(;) and (?)


so that in the case wher€ P : 0, r|n reduces to the third and fourth solutions of
(3.97). Using the first of (3.101) to obtain the upper spinor tta, wa have

| -p,cl(l Ell mc'z) \ / -(p,-ip,)cl(lEl*mc,)\


: p,cl(lEftmc'z)
Nl -(p,*ip,)cl(lEllmc'z) | uno rz(n)(p) : Nl
z(,)(p)
I'\oI\\?) I

(3'104)
Since
u@(p).,,n
[i'; -+]
satisfies the free-field Dirac equation, it is evident that each free-particle spinor
z(') with t:1,...,4 satisfies
p * mc)u{'t(p) : 0
(i,y. (3.10s)
with "y.p : nf ppp, p: (p,iElc) regardless of whether E > 0 or E { 0. This can,
of course, be checked by direct substitution.
As shown earlier, the four independent free-particle solutions with p : 0 written
in the form (3.97) are eigenspinors of the 4 x 4 matrix Ir. This is not true for the
free-particle solutions we have written for the case where p + 0, as we can directly
verify by applying I' to (3.103) and (3.104). But suppose we choose the z-axis
in the direction of momentum p so that Pt : pz :0. We then see that a(') with
r:1,...,4 are eigenspinors of I, with eigenvalues of +I, -I, +1, -1, respec-
tively. In general, although free-particle plane-wave solutions can always be chosen
3-3 NONRELATTVTSTTCAPPROXIMATIONS; PLANEWAVES 93

so that they are eigenfunctions of l.fl (where 0: p/lpl), it is not possible to


choose solutions in such a way that they are eigenfunctions of I.ff with an arbitrary
unit vector fi.f As we shall see in Section 3-5, this peculiarity of the plane-wave
solution in the Dirac theory stems from the fact that the operator I.fi does not
commute with the free-particle Hamiltonian unless fi : +0 or p : 0. The operator
1.fl which can be diagonalized simultaneously with the free-particle Hamiltonian
is called the helicity operator. The eigenstates of helicity with eigenvalues * I
and - 1 are referred to, respectively, as the right-handed state (spin parallel to
motion) and the left-handed state (spin opposite to motion).
It is easy to see that for a given fixed p, the free-particle spinors ,(')(p) given
by (3.103) and (3.104) with r - 1,...,4 are orthogonal to each other;
uo)+(p)uc')(p):0 for r*r,. (3.106)

For the normalization of uo), two conventions are found in the literature:
(a) (3.107)

which implies
(3.108)
hence
(3.10e)

(b) (3.1l0)
which means
(3.lll)
The second normalization convention which says that u+u transforms like the
zeroth component of a four-vector appears somewhat artificial at this stage. How-
ever, we shall see in the next section that this convention is quite natural from the
relativistic point of view. Throughout this book we shall use the normalization
condition given by the second form (3.110).
To summarize, the normalized plane-wave solutions for given p are:

*t : tl$ar("")(P) t"n [;P '+ - (3. l 12)


'+)'
for E : JWV{ +W7 } 0, and

^h-- fffirt:orn)(p).*n[if + +], (3. l r3)

fThis situation is in sharp contrast to the nonrelativistic Pauli theory in which any space-
iime independent two-component spinor can be regarded as an eigenspinor of o'ff, where
ft is a unit vector in some direction, that is,

(o.n)(;):(;)
Assuming that the spinor is normalized, all we have to do is set 4 : cos (0012)s-to'tz
and 6 : sin (00f2)s*ft,/2,.where do and fo characterize the orientation of the unit vector
along which the spin component is sharp.
RELATIVISTIC QUANTUM MECHANICS OF SPIN.+ PARTICLES 3-3

for E : -JEWI +WVa { 0, where

'r (,,'-ilr*,)] (3 114)

forE>0,and

(3.lrs)

for E < 0. The square root factors in (3.112) and (3.113) merely compensate for
lEl'zlmc'zin (3.110), so that

I,^h'^lr|3x
: l. (3.1l6)

As V ----+ oo, the allowed values of E form a continuous spectrum. For positive-
energy free-particle solutions, mcz < E { €, whereas for negative-energy solu-
tions, -oo ( E I -mc', aS shown in Fig. 3-1.
Allowed E>0

E:0--- -Forbidden

N
I

E:-mc2 |

Allowed E<0

Fig. 3-1. The allowed values of E for free particles.

When the particles are not free, problems for which the Dirac equation can be
solved exactly are rather scarce. In Section 3-8 we shall treat the classical problem
of the electron in the Coulomb potential for which exact solutions can be found.
Another problem that can be solved exactly is that of an electron in a uniform
magnetic field, which is left as an exercise (Problem 3-2).
RELATIVISTIC COVARIANCE

3-4. RELATIVISTIC COVARIANCE


Lorentz transformations and rotations. Before we establish the relativistic covariance
of the Dirac equation, we shall review briefly the properties of Lorentz transfor-
mations. Since a Lorentz transformation is essentially a rotation in Minkowski
space, let us first examine ordinary rotations in the usual three-dimensional space.
Consider a rotation through an angle co, of the coordinate system in the 1-2
plane about the third axis. The sense of rotation is that of a right-handed screw
advancing in the positive xr-direction if or ) 0. A point described by (xt, x2, xr) in
the old coordinate system is described in the new (primed) coordinate system by
(x't, xL, xl) where
x'r: Jrlcos a * xzsinar, xL: -xtsino * xzcos @. (3.117)

Meanwhile, when we perform a Lorentz transformation along the x1-direction


such that the primed system moves with velocity a: Fc, (x, lxo) and (x', ix's) ate
related by
--,:-
-'r x1 Bxo
^/l - P' n/l - P'
x'r : xr cosh y * ixr sinh 1, xL: -ix, sinh y I x+ cosh 1, (3.119)

where we have set


tanh 1 : B. (3.120)

Since the cosine and sine of an imaginary angle iyare coshl and i sinh 1, the above
Lorentz transformation may be viewed as a rotation in Minkowski space by an
angle iy "in" the l-4 plane, just as the transformation (3.1l7) represents a rotation
in the 1-2 plane by a real angle o. For both three-dimensional rotations and Lorentz
transformations we have
oruQ)rv : Epl , (3.tzt)
where a, is defined by
x'r: apvXv. (3.122)

From now on, the term Lorentz transformation will be used both for a three-
dimensional rotation of the form (3.117) and for a "pure" Lorentz transformation
of the form (3.119). At this stage we shall consider only those coordinate trans-
formations which can be obtained by successive applications of a transformation
that differs infinitesimally from the identity transformation. For example, (3.117)
can be obtained by successively compounding an infinitesimal rotation
xl: xr f 8arxr, Jcl : -8o xr * xz, (3.123)

where Dar is the infinitesimal angle of rotation. This means that we restrict our
considerations to the cases where
det (ar,) : l, a44 ) 0. (3.124)

Covariance of the Dirac equation. What is meant by the covariance of the Dirac
equation under Lorentztransformations ? First of all, if someone, working with the
primed system, were to formulate a relativistic electron theory, his first-order
wave equation would look hke the Dirac equation. Second, there exists an explicit
96 RELATTVTSTTC QUANTUM MECHANTCS OF SprN-+ PARTTCLES

prescription that relates f(") and "$'(x'), where f(") and r/n'(x') are the wave
functions corresponding to a given physical situation viewed in the unprimed
and the primed system, respectively. Third, using the foregoing prescription, we
must be able to show that the "Dirac-like" equation in the primed system not
only looks like the Dirac equation in the unprimed system but actually is equivalent
to it.
We take the point of view that the gamma matrices are introduced merely as
useful short-hand devices that enable us to keep track of how the various com-
ponents of ./, are coupled to each other. Hence the explicit forms of the gamma
matrices are assumed to be unchanged under Lorentz transformations.f Note,
in particular, that the gamma matrices themselves are not to be regarded as com-
ponents of a four-vector even though, as we shall show in a moment, r|ryrrfn does
transform like a four-vector. On the other hand, we expect that for a given physical
situation the wave functions in different Lorentz frames are no longer the same.
With this point of view, if the Dirac equation is to look the same in the primed
system, we must have
,v
-&*'@) + T.ih'(x') : o. (3.r2s)

Note that the gamma matrices themselves are not primed. The question we must
now ask is: How are rfn and rfn' related? An analogy with electrodynamics may be
helpful here. When we perform a Lorentz transformation without a simultaneous
change in the gauge, the components of Ar(x) and A'r(x') are related by
A'r(''): aruA,(x), (3'126)
where a* given by (3.122) depends only on the nature of the transformation and
is independent of the space-time coordinates. Similarly, we may assume that the
prescription that relates f(") and !r'(x') is a linear one that can be written as

"h'@'): st(x), (3.t27)


where S is a 4 x 4 matrix which depends only on the nature of the Lorentz trans-
formation and is completely independent of x and r. We rewrite (3.125), using
0f0x'u: ar,(010x) [which was proved in Chapter 1 (cf. Eq. l.2l)] as follows:
nrt,ot,htO +Tsf : o (3. r28)

or, multiplying ,S-r from the left, we have

S-'ry, Sor,*,,h + r.
T+: (3.12e)

We now ask: Is (3.129) equivalent to the Dirac equation ? The answer to this ques-
tion is affirmative provided that we can find an S that satisfies
,S-tryrS op,,:.f, (3.130)

fThis means that we are considering a Lorentz transformation not accompanied by


a simultaneous change in the representation of the gamma matrices.
RELATIVISTIC COVARIANCE

or, multiplying by o^, and summing over v, we obtain


S-try^S:ry,o^,, (3.131)

where we have used (3.121).f The problem of demonstrating the relativistic covari-
ance of the Dirac equation is now reduced to that of finding an S that satisfies
(3.1 3 l).
We shall first treat the case of an ordinary rotation in three dimensions which
we refer to as a "pure rotation." In looking for an S appropriate for the transfor-
mation (3.117), we recall that for a particle at rest the upper two-component
spinor r,la is identical with the corresponding two-component spinor in the non-
relativistic Pauli theory. So in this particular case we know from nonrelativistic
quantum mechanics how za transforms under three-dimensional rotations: if
a, is the Pauli spinor in the unprimed system, then

('"'f *io,sin+) (3,132)

times rz, is the Pauli spinor corresponding to the same physical situation seen in
the primed system defined by (3.117).$ Since the form of S is independent of
whether it acts on the "at-rest" spinor or a more general wave function, it is natural
to try for S the 4 x 4 analog of (3.132) given by
s"ut: cosf + t >, sinf
(3.133)
:coS|+ryr,yrsin|,
which is now assumed to act on a four-component wave function. Since
(ry, ryr), : nlrnfz.lrf z : -"y1ry7: - 1, ^S;1 is given by

S;i:cosf -?fffzsinf . (3.134)

fNote that the matrices au tedrta.nge the coordinates x and xo, whereas the gamma
matrices and the 4 x 4 matrix S rearrange the components of +. Although a* and 'f ,
are both 4 x 4 matrices, they are in entirely different spaces. The relation (3.131), for
instance, really means
) (S-t)"B(ryr)Br(S)ra : E aLv(yv)d6.

$See, for example, Dicke and Wittke


(1960), p. 255. As an example, let us consider an
6lectron whose spin is in the positive x1-direction. Its wave function is represented by
I /l\
E\r)'
which is evidently an eigenspinor of o, with eigenvalue *1. In a primed system obtained
by a rotation of 90' about the xr-axis, the electron spin is in the negative x!-direction.
Accordingly, the wave function in the primed system is
( zr ,'4)L( 1\:1/1 +t\
\cosz-+to3sr +/\/2\r/ =z\t -i)'
This is indeed seen to be an eigenspinor of o, with eigenvalue -1.
98 RELATrvrsrrc eUANTUM MEcHANrcs oF sprN-+ pARTrcLEs 34

So the relation to be checked is (cf. Eq. 3.131)

("or i - ntrtzsin :
+)v^ (cos i + ry,.y, sin +) r,ott,. (3.135)
For ),:3,4 we have o),,:01,; hence (3.135) is trivially satisfied by virtue of
f ffznft,E: ?f t,+nfff, and "f rf znft,tnlt|yz : -'fz,+. For ), : 1,

("or; - nrrzsin
+)v, (cos $ + v,v,sin|)
: tr cos'
| * 2"yrsin f cos
f - ,y, sin' | '

: lr cos al * ry, sin ar. (3.136)


But according to (3.117) cosc,l : att, sinar : at2.So (3.135) is indeed satisfied
for ), : 1; similarly for ), : 2. Thus we have proved that the S"o, that satisfies
(3.131) is given correctly by (3.133).
According to (3.119) a "pure" Lorentz transformation is nothing more than
a rotation in the 1-4 plane by an imaginary angle i1. Hence for corresponding
^S
to (3.119) we try a matrix analogous to (3.133) with &) -+ iX, nlr - nt+l

SLo. : cosh (yl2) * i,yr,yn sinh (1/2). (3.1 37)


The relation we must check this time is

(cosrr +- iny,ntnsinh
+)v^ (costr ! + int,ntnsinh +) : r,e,,,. (3.138)
Using similar techniques, we can verify this. For instance,

("oth +- int,ntnsinh
! + int,,yn sinh +)
+)v, (cosh

-- ,rt cosh, + - 2iry, cosh f sinh ! + ,yn sinh, !


:f+cosh1+,y{-isinhl)
: +ota * rytau.
?f
(3.13e)
Since we have found S that satisfy (3.131) for the transformation (3.117) and
(3.119), we have established the covariance of the Dirac equation under (3.117)
and (3.119). In general, the s that relates r/r and ,l' via (i.lz7) is given by

:
Srnt cos
f * iori sin
!- (3.140)

for a rotation about the kth-axis (ijk cyclic) by an angle ar, and

SLo": cosh!-otcrsinhf (3.141)

for a Lorentz transformation along the kth-axis with B : tan 1, where we have
introduced new 4 x 4 matrices
ct,: (llzi)Wp,nl,l: -i,f ,nf,, tt + u. (3.r42)
34 RELATrvrsrrc covARIANcE 99

More explicitly,

Orj: -6jt: Xt: Qik) cYctic,


6- ), (3.143)
ck;: dt:
l0 o*\
-cl,n- \o* o)'
It is easily verified that
Slor: cosf - iotisin
!: S-'r, (3.r44)

but
Sio, : coshf - cx+sinh : S"n" + S;.t". (3.145)
f
Thus, unlike S"ut, Sr,o" is not unitary. This is not catastrophic, however; S".. should
not be unitary if 9*^h is to transform like the fourth component of a four-vector
under Lorentz transformations. It is very important to note that for both pure
rotations and pure Lorentz transformations we have
S-1 : rynSrryo, ,St : ynS-tryn, (3.146)

ry4 commutes with oq brtt anticommutes with o1a. We shall


since make extensive
use of (3.146) in the next section.

Space inversion. So far we have considered only those transformations which


can be obtained by compounding transformations that differ infinitesimally from
the identity transformation. Under such transformations the right-handed coor-
dinate system remains right-handed. In contrast, space inversion (often called
the parity operation) represented by
X'--X, tt:t (3.r47)

changes a right-handed coordinate system into a left-handed one; hence it is outside


the class of transformations considered so far. We shall now demonstrate that
the Dirac theory is covariant under space inversion even in the presence of Ar.
According to the Maxwell theory the four-vector potential transforms under
space inversion asf
A'(x', t') : -A(x, f), A'n(x', t') : An(x, t). (3.148)

{To prove this we first note that if the equation for the Lorentz fotce
dl v \ :'ln
*fr\6) *(v/c) xBI
is to take the same form in the space-inverted system, the electromagnetic fields must
transform as E': -E,B': B sinCeY/: -Y. Meanwhile, E andBarerelatedtoAand
Ao: iAtbY
E:-YA,-+K, B:vxA.
Hence we must have (3.148).
100 RELATIVISTIC QUANTUM MECHANICS OF SPIN.+ PARTICLES 34

If the Dirac equation is to look similar in the space-inverted system, we have

(&- # n;)r**' +ry*' :0, (3.14e)

or, equivalently,

l_(*r_ #Ar)or + (*^- #n')r^f+' + Tg' : o. (3.150)

We try as before
t'(*', t') : S"r/r(x, r), (3.rs1)
where S' is a 4 x 4 matrix independent of the space-time coordinates. Multiplying
(3.150) by St'from the left, we see that (3.149) is equivalent to the Dirac equation
if there exists ,S" with the property
S;try*S" : -rlk, SFtrynS.p : yr. (3.152)
Since ry4 commutes with rya but anticommutes with "yr,
Se : T'f +, S r' : 'YnlT
(3.1 s3)

will do the job, where 2 is some multiplicative constant. The invariance of the
probability density ,,/r*^/, further requires lql' : l, or rJ : eif with @ real. It is
customary to set this phase factor 7 equal to I even though no experiment in the
world can uniquely determine what this phase factor is.f We shall take
S.p : Yr (3.1 54)

for an electron wave function.


Closely related to the parity operation is an operation known as mirror reflection,
for example,
(xi, xL, x'r) : (xr, xr,
-xr), xt : x'0. (3. l ss)
Since (3.155) is nothing more than the parity operation followed by a 180" rotation
about the third axis, the covariance of the Dirac equation under mirror reflection
has already been demonstrated implicitly. Similar statements hold for more general
transformations with
'
det (ar,) - - l, aaa ) 0, (3.156)

which are called improper orthochronous Lorentz transformations in contrast to


proper orthochronous Lorentz transformations that satisfy
det (au,) : 1, ano ) 0. (3. l 57)

Simple examples. To appreciate the real physical significance of S".1, ,S1u., S", it is
instructive to work out some examples at this stage. As a first example, let us con-
sider an infinitesimal form of (3.1l7) in which cos o and sin al are set, respectively,
to I and 8ar. The wave functions in the two systems are related by
*'(*'): [1 + f ),(8ar/2)l^lr@), (3. rs8)

lsome authors argue that one can narrow down the choice for ? to +1, +i by requiring
that four successive inversion operations return the wave function to itself. However,
there does not appear to be any deep physical significance attached to such a requirement.
34 RELATIVISTIC COVARIANCE 101

wherex'-x*Exwith
8x : (xrDco,
-xr8ar,0). (3. r se)
But

*'(*'):9'(x) * Ex,
ffr+ U*rH (3.160)

Consequently,

f'(") : [r +
t >,? - (",0, *,- x,Ear
&))g(x)
: f(") * i Do
t-\
L? * *(-,u.,#,r ihx,3-)] +t"t (3.16r)

We see that the change in the functional form of * induced by the infinitesimal
rotation consists of two parts: the space-time independent operator i )e 8ol2 acting
on the "internal" part of ./r(x) and the familiar iLr}alh operator affecting just
the spatial part of the wave function. The sum
(t lh)[(h > ,12) * L,] (3.162)
is to be identified with the third component of the total angular-momentum opera-
tor in units of h since it generates an infinitesimal rotation around the third axis.f

Fig. 3-2. A positive-helicity electron moving with momentum p along the xr-axis. The
electron is at rest in the primed system. The gray arrow indicates the spin direction.

As a second example, let us consider a free positive-energy electron of helicity


* I and momentum p along the positive xr-direction. We choose a primed system
in such a way that it will coincide with the rest system of the electron (Fig. 3-2). In
the primed system the electron wave function can be written

- imczt,tn (3.163)
"lr'(x'):

fUsually an operator that rotates the physical system around the third axis by an angle
Daris I - i}rar(Jslh). But in our case we are rotating the coordinate system rather than the
physical system. This explains why we have I * i Dor"fr/ft instead of the above operator.
r02 RELATIVISTIC QUANTUM MECHANICS OF SPIN-+ PARTICLES 34

The question is: What is the wave function for the same physical situation in the
unprimed system? According to (3.141) and (3.137)
9(x) : ,S;ot"./r'(x'), (3.164)
where
,Stl" : cosh (yl2) - iry rry n sinh (yl2), (3.16s)
with 1 given by
coshl : Elmc', sinhl : pslmc. (3. l 66)
Since
\rc +Zos'-in:
cosh(yl2): @, (3.167)
sinh(1/2):@:prcf@,
we obtain

I IEW 't, psccs \ /l\

"-(i)
IlTiWll'l
|\@i;:;;, w- ll
^,1
'
/\o/
I

(3.1 68)

This result is in complete agreement with u,t,(p), with p, : p2 :0 obtained


earlier by solving directly the Dirac equation (cf. Eq. 3.114). It is amusing that
the normalization constant which we pick up automatically is precisely the one
that appears when z(p) is normalized accordingto (3.110), which says that u+u is
the fourth component of a four-vector. As for the space-time dependence of the
wave function, we merely note that
tt:tcoshl -@rlc)sinh1
: (Efmc2)t - (prlmc2)xr. (3. l 6e)
So we find that
9(x) : S;,1..h'(x')
:hz(')(P) -+l
'*nlintf,
: ,r[-*Lz(')(P)t*n [;'+ -'#]' (3.170)

where we have used y: (mc2lE)V' that follows from the Lorentz contraction
of the normalization volume along the direction of motion. Thus we see that
once we know the form of the wave function for a particle at rest, the correct
wave function for a moving particle of definite momentum can be constructed
just by applying S;1.. This operation is sometimes known as the Lorentz boost.
34 RELATIVISTIC COVARIANCE 103

To work out an example that involves S", let us look at a Dirac wave function
with a definite parity:
fI O(t, t) : _[r/^(*, l). (3.171)

According to (3.151) and (3.154) the functional form of the wave function in the
space-inverted system is

^h'(t,
t): rynnlr(-x, /), (3.r72)
which is to be identified with fh/n(*, t).Thus

(; -?)ffi-:, 1): +ff.[:,] Qr73)

If, in addition, *^ and 9r can be assumed to be eigenstates of orbital angular


momentum, then
9r(-*, t) : (-l)''fr(x, t): I*n(x, t), (3'174)
-gr(-t , t) : -(- l)"r/nr(x, t) : *qo(x, /),
where lo and I, are the orbital angular momenta of the two-component wave
functions. Thus
(-l)" : -(-l)". (3.175)

At first sight thisappears to be a peculiar result, since it implies that if rfn" is a


two-component wave function with an even (odd) orbital angular momentum,
then rfu is a two-component wave function with an odd (even) orbital angular
'momentum. Actually, this is not too surprising in view of the second part of (3.62),
which, for a central force problem with A : 0, takes the form

*u: #T (3'176)
^P(o'P)f''
Let us suppose that rlrn is an s,r, state wave function with spin up so that
/l\
*o: nf')(o)t-"""n' (3.t77)

Then we find that

ihc
1aa*' *'-'u*)(^y)'
^t^
Yn: -E=T + m7l I -i(Et/h)
a a
\di-'6i .Tx, I
,,__.:,)(L),_,,",,,,
_ ____,!q_ +#(.,| ,*,
: z=hw#l- ^lT"'(l). (3,78)
^lY"*(?)),-,*',,,
part
which is recognized as a wave function whose angular is that of 7 Ptn wave
function with 7, : ;. Thus r/n, and t, have opposite parities in agreement with
(3.175). We could have guessed that this would be so, since the operator that
multiplies 9" ir (3.176) is a pseudoscalar operator which does not change 7 and
104 RELATIVISTIC QUANTUM MECHANICS OF SPIN.+ PARTICLES 3-5

7' but does change the parity. We shall make use of this property in Section 3-8
when we discuss central-force problems in detail.
An even more striking consequence of Sr : yr may now be discussed. Consider
a positive-energy free-particle wave function and a negative-energy free-particle
wave function both with p : 0:

(1"') e-imc2t/n and (r1,,) ,imc2t/h, (3.17e)

where X(*) may stand for

(;) .r (?)
Since they are eigenstates of ryn with eigenvalues *l and
-1, respectively, we
have the following far-reaching result: a positive energy electron at rest and a
negative energy electron at rest have opposite parities. This will be shown to imply
that an electron and a positron have opposite "intrinsic" parities when a negative
energy state is properly interpreted in Section 3-9. For instance, the parity of an
e+e--system in a relative s state must be odd despite its even orbital parity. This
remarkable prediction of the Dirac theory has been checked experimentally in
the decay of a positronium and will be discussed in Chapter 4.

3-5. BILINEAR COVARIANTS


Transformation properties of bilinear densities. We are in a position to discuss
bilinear densities of the forrn .}f^h, where I is a product of gamma matrices.
Such densities are called bilinear covariants since they ha.ve definite transforma-
tion properties under Lorentz transformations, as will be shown in a moment.
Let us first note that because of (3.146) the relation*'(x'): Sf(") implies that
fi'(*'): ft(x)Sr"yn : *r(*)Vn"ynS*yn : r/(x)S-r, (3. I 80)
whether ,S stands for S.o, or S1o.. Clearly this relation holds also for ,Sp. Using
(3.180) we immediately see that fr$ ir invariant:

fi'(*') f'(x') : O(") f(") (3. 1 81)

under pure rotations, pure Lorentz iransformations, and space inversion; hence
rp{n (not is a scalar density. Tc investigate the transformation properties
^/r*^/r)
of firyrt, it is sufficient to recall (3.131). We have
:
fi'(*')"y r^lr'(x') :
fr(")S-'ry"Sf(x) a*,fi(x)"y,^!r(x) (3.182)
under pure rotations and pure Lorentz transformations. For the behavior under
space inversion, we obtain

vl,,Ir': rs,.'1fi}**: {-$T^[] (3,83)

Hence fi,yr{ is a four-vector density whose space components change under


parity. Consequently the flux density and the probability density defined earlier
(cf. Eq. 3.49) do indeed form a four-vector. Using similar techniques we find
3-5 BILINEAR COVARIANTS 105

Table 3-1
BEHAVIOR OF BILINEAR COVARIANTS UNDER LORENTZ
TRANSFORMATIONS

Proper orthochronous . Space


Lor entz transformat i ons lnverslon

Scalar 'V^h {rh ^Ig


Vector $vr9 ar,$v,"1, I -try-91
t 9ryn,./'l
Tensor (antisymmetric,
second rank)
$orr^1, apxaro{ro:^o"lr Il-*"t"*'lh\
rn*I
Axial vector t{r,Yst r,"h a*i$v5v,.r!, I i{t'Y s'Y k+l
(pseudovector) I -;^Iryuryn
_gvu*
+l
Pseudoscalar $vu\, gvu*

that $or,* : -i@"yr,y,rl, with p # a (which is necessarily antisymmetric in p, and


z) is a second-rank tensor density. At this point it is advantageous to define a
Hermitian 4 x 4 matrix
(3.184)
This ryu matrix has the remarkab;';:il,1", ,, anticommutes with every one
of yrwithp:1,...,4,
{,yu, nlu} : 0, p + 5, (3.185)

as seen, for instance, from ,yzryfyzrfsnla: (-l)tryrryrryrnltryz, where we have used


the fact that'y2 commutes with ryr but anticommutes with ,f b,ys, ry+. Note also that

"y?: I (3.186)
just as ry* with p: | 4. As for its explicit form it is easy to show that
t0-1\
: (-r (3'187)
v' o/
in the standard (Dirac-Pauli) representation. Using (3.185), we see that
Sij,ryuSr,o" : ys, ,S-tyuS"ot : Ysr (3. r 88)

since ry' commutes with or, but


SFtryuSr : r8e) (3.
-nf s,
since ryu anticommutes with ryn. Hence fi"yrt transforms exactly like the scalar
density rlr/n under proper orthochronous Lorentz transformations but changes
its sign under space inversion. This transformation is characteristic of a pseudo-
scalar density. Finally, using similar arguments, we can easily see that i.luuf r^1,
transforms in the same way as F,yr* under proper orthochronous Lorentz trans-
formations but in exactly the opposite way under space inversion. This is expected
of an axial vector (pseudovector) density. Table 3-l summarizes the results.
The question naturally arises: Have we listed all possible bilinear covariants
of the form .}f.h? To answer this question let us start multiplying the ryr. If we
106 RELATTVTSTTC QUANTUM MECHANTCS OF SprN-+ PARTTCLES 3-5

multiply any pair of gamma matrices, we get either a12r: 1 when the two matrices
are the Same Of f *rl,:
-rf,rfp: iop, When the twO matriCes are different. When
we multiply three gamma matrices, we get back only one of the y, up to sign unless
all three are different (for example, 1yfyztf r: -?f ff flz: -nlz); when the three
matrices are different, we do get a new matrix I pnl,nf tt. But ,y rry,ryr with y, * v # ),,
can always be written in the formyryoup to sign,where o * pt,,z,),(for example,
ry{lzfs?y+f t : nls,f r). Finally, when we multiply four gamma matrices, we get
only one new matrix, ryu : ,f rf z"f sf, (which is, of course, equal to -rf zrf sry{fb
"f z,y{f r"f z,
etc.). Needless to say, when we multiply five or more gamma matrices,
we obtain nothing new. So

lz : l, nlp, ctr: -i"yrrl, 0n * u), iryuy, and f s (3.190)


represents all we can get. This means that there are in all sixteen independent
4 x 4 matrices (as we might have guessed): the identity matrix, the four ry, matrices,
the six o* matrices (antisymmetric in p, and u), the four inyu"y, matrices, and
the.y5 matrix. The factors *i in (3.190) are inserted so that

D:1, (3.lel)
for A - l, . . . ,16. The l" are all traceless with the obvious exception of the
identity matrix, as the reader may easily verify by using the explicit forms of l,
in the standard (Dirac-Pauli) representationl (cf. Appendix B). Moreover, they
are all linearly independent. Consequently any 4 x 4 matrix can be written as
a unique linear combination of the sixteen L. We can find the coefficient ),, in the
expansion of an arbitrary 4 x 4 matrix A
16

A: X
A
trrlr, (3.1e2)
by simply evaluating
Zr(Alr) : Tr(2 ).B ls|") : 4\o, (3.1e3)
B

where we have used the fact that lrf, is traceless when B + A and is equal to the
identity matrix when A : B. The algebra generated by l, is called Clffird algebra
after W. K. Clifford, who studied generalized quaternions half a century before
the advent of the Dirac theory.
Let us return now to our discussion of the bilinear covariants. It is worth keep-
ing in mind that Table 3-1 exhausts all possible bilinear covariants of the form
.7f9, as first shown by J. von Neumann in 1928. For instance, note that we have
no way to write a symmetric second-rank tensor of the form .71.h. This does not
mean, however, that we cannot form a symmetric second-rank tensor in the Dirac

fTo prove this without appealing to any particular representation, first verify that for
every f,4 there exists at least one f a (different from f,J such that fzfr :
-l.ala. Then
-Tr(l ) : Tr(l nhfo) : Tr(l%l t) : Tr(1,)
which would be impossible unless fr were traceless.
3-5 BILINEAR COVARIANTS lo7

theory. If we start introducing derivatives of rfn and r|, we can develop an expres-
sion like

ru, : -'+ (+r,# - *ro,+) + T (3.re4)


"n,,*(firy,ry,"v,*),
which can be shown to be the energy-momen,rr- Lnror of the Dirac wave func-
tion.
For not too relativistic electrons of positive energies some bilinear covariants
are "large" while others are "small." To see this we first recall that if E - mcz
and VKmc', then r/nr is of order (alc) compared to $". We can see thenthat
rlrf and $"y n* given by
.7.h :
^lrr "yn"l,
: *ln*o - rhL*u,
(3.1es)
$"yn*: ^h*^/,
: *ln''|," + ^hL\,"
are both "large" and in fact equal if terms of order (ulc)' or higher are ignored.
Similarly, since
.
iry,"yr:
lor o\
(,d : 4 _rlou
^ o\I
_or), o,,
\ o ool'
(3.196)

it follows that i$,yryrrfn and {or,* (iik cyclic) are "large" and indistinguishable
up to order uf c:

r{";,ir}g x g\o,,\,^ (3.1e7)

In contrast to I , Iu i,ys,f o, and cri1, which connect r/r| with rfnr, the matrices ryp,
iryunyn, cpa, artd ry5 connect./rl(91) with rf'"(r/rr). Hence the corresponding bilinear
covariants are "small" or, more precisely, of order uf c. For instance,

$Vr,l, : (,.!rt,-./rl) (-: -ilffi;)


: -^ht^hu -l ^lrL*". (3.1e8)

Gordon decomposition of the vector current. The remaining part of this section is
devoted to a detailed discussion of the vector covariant $"yr^!r, which occurs most
frequently. We argued earlier that within the framework of the single-particle
Dirac theory, s* : ic{rryr.h is to be regarded as the four-vector probability current.
We therefore define
i* -- €sr": iec$,yr"lr, (3.1ee)
which is to be interpreted as the charge-current density. Using steps analogous
to (3.45) through (3.48) we can show that j, satisfies the continuity equation even
in the presence of the electromagnetic interaction. With (3.199) as the charge-
current density, the Hamiltonian density for the electrornagnetic interaction of the
charged Dirac particle is given by
ff ,n, : -j rArfc : -ie$"yr*A,
: -e"h+4.A + ,{r*Ao. (3.200)
108 RELATrvrsrrc eUANTUM MEcHANrcs oF sprN-+ rARTTcLES 3-5

This relationship can also be inferred from the Hamiltonian form of the Dirac
equation (cf. Eqs. 3.38 and 3.39):
ihQlAt)$: l(ichY - eA).ct f B*r' -f eA,l*. (3.201)
To appreciate the physical significance of7, we rewrite (3.199) as

jr: (iecl2)($,yr+ + fi,y,*)


iehl
- 2mL -+r*r.(*- #n.)+ * {(#,+ #u.)+\,y,,y."t), (3.202)

where we have used (3.60) and its analog for the adjoint wave function r|. This
encourages us to split 7, into two parts:
jr: jt) + jff), (3.203)
according to dependency on whether or not the summation index z in (3.202)
coincides or does not coincide with p. We have

it]' :t+,(#,^h- {,H,) - #or{,y (3.204)

and

j t:,
: i+l-+r,r, f;,+ . (#.) %,Y,*

+ #A'fivrry'"1' * #A'r"r,v**f u,,


eh A /T (3.205)

This decomposition is known as the Gordon decomposition, named for W. Gordon.


Let us look at each component of (3.20a) and (3.205) a little more closely. The
four-vector (3.20\ does not contain any gamma matrix. In fact 7[t) would be
formally identical to the expression for the three-vector current density in the
Schrodinger theory if we could replace the Dirac wave function by the Schrodinger
wave function. This is rather gratifying since we know that in the nonrelativistic
limit, @$lAxr)!r, etc., can be legitimately replaced by @^h\lAd*.t, etc. As for
,rlt) we can easily show that

itD:#+v-#$,hA, (3.206)

when the time dependenee e-iEt/h is assumed. This reduces to ic times the charge
current density e*L\ru in the Schrodinger theory, provided E - mc2 and I eArlK
mc2. As for (3.205) we recall that in the nonrelativistic limit

$oun*: -$onu*
can be ignored while Fo,n* can be interpreted as the spin density $\o,"1r,, (j kl
cyclic). In other words, the kth component of (3.205) is -ehl2m times

*,r+1",A) - f;,{"l,ro,^h), ukt) cyclic, (3.207)


3-5 BILINEAR COVARIANTS 109

which is just the kth component of the curl of the spin density.f Thus for slowly
moving electrons the Gordon decomposilion of j*can be regarded as the separation
ofTu into the convection current due to the moving charge and the current associated
with the intrinsic magnetization (magnetic dipole density) of the electron.
When jp interacts with A, via (3.200), we have the interaction Hamiltonian
density
ehla, I
-i*4-
c 2mc lox,:^ho'r+))A'
: - mdi(fi",,*)
eh 0A

#l+'u+ r+ ", r^h) ++


*tf ",, rh)]
(3.208)
*I+r,rfio,r"lrl,
where we have dropped @lAfi($a,rtAr) which is irrelevant when the interaction
density is integrated. Noting that
*F,r$r,rr/r = B.(rhIo*^) (3.209)
in the nonrelativistic limit, we see that (3.208) can indeed account for the spin
magnetic moment interaction with the gyromagnetic ratio g :2, in agreement
with our earlier discussion based on (3.67).
Experimentally, as first shown by P. Kusch in 1947, the observed gyromagnetic
ratio of the electron is not exactly 2 but rather

s:2[t* (h){"+ ],
(3.210)

which holds also for the muon. The origin of the extra magnetic moment was
satisfactorily explained in 1947 by J. Schwinger who took into account the fact
that the physical electron can emit or absorb a virtual photon, as we noted in
Section 2-8. (We shall come back to this point in Chapter 4.) When the magnetic
moment is not correctly given by g : 2 we may add a phenomenological term to
the interaction Hamiltonian of the form

ff,n, ffi8r,,-{o,**7, (3.2n)

called an anomalous moment (Pauli moment) interaction.$ The total magnetic

fln this connection recall that according to classical electrodynamics a magnetic dipole
density 7/l gives rise to an effective current density
i1"rrl : cY x '
"0(
See Panofsky and Phillips (1955), p. 120; Jackson (.1962),p. 152.
gThe interaction (3.211) is to be interpreted as an effective Hamiltonian density. To the
extent that Schwinger's correction is computable on the basis of the interaction (3.200),
there is no need to postulate an additional "fundamental" interaction of the type (3.21l),
at least for the electron and the muon.
ll0 RELATIVISTIC QUANTUM MECHANICS OF SPIN.+ PARTICLES 3-5

moment computed from the sum of (3.208) and (3.21l) is


p: (ehl2mc)(t * rc). (3.212)
For the proton, replacing e and m in (3.208) and (3.211) by le I and the proton
mass mt, we have
pr,: (lelhl2m,,c)(l -l- r,,), (3.213)
where experimentally rc,, :
1.79, as O. Stern first determined. We thus see that
roughly 60/, of the observed proton magnetic moment is "anomalous." According
to (3.208), spin-f particles should have zero magnetic moments as e-----) 0, yet
experimentally neutral particles, such as the neutron and the A-hyperon, are
known to possess sizable magnetic moments. They can again be represented
phenomenologically by interactions of the type (3.211). In contrast to the anom-
alous moments of the electron and the muon, the anomalous moments of the
profon, the neutron, the A-hyperon, and so forth, cannot be accounted for by
Schwinger's mechanism. Therefore their existence appears to indicate a failure
of the simple prescription pp pp - qArlc. However, if we consider that these
- surrounded by virtual meson clouds, the failure
particles are complicated objects
of the simple prescription does not seem surprising. Spin-f particles whose
electromagnetic properties can be understood on the basis of (3.200) alone are
sometimes referred to as pure Dirac particles.

Vector covariant for free particles. To investigate further the physical meaning
of the gamma matrices, let us consider this time fiflu.lrr, where rfni and ^17 are
E > 0 plane-wave solutions:

*, : ur)(p)."n (T -'+),
^tr (3.2r4)
,lrr :
ffi"r')(p') exp (,e;" - +)
This is of some practical interest in connection with the scattering of a Dirac
particle by an external vector potential A, since the transition matrix element
which we compute in the Born approximation is essentially
-ie I fifyollrrArd'x.
Assuming for simplicity that the vector potential is time independent so that
4 : E' (elastic scattering), we can evaluate iry* taken between uo''(p) and rz(')(p)
as follows:
i uo') (p, ) "y ouo )
(p) : uc' )+
(p, ) a ru(,) (p)

: ('rW)' (r"'*, x'''r ;?#p)(o r ?)("4; r,,,)


Pr * p^' io-{(p'-p)><iiu}
-- x l--T*,
^.(.c')+ f
- r

2mc I,"', (3.21s)

initial and final Pauli two-component spinors. The first


where X(') and XG') are the
term, of course, corresponds to the convection current.l[t) of (3.201). To see the
meaning of the second term, we note that it appears in the transition matrix
3-5 BILINEAR COVARIANTS 111

element as

-, Qlry| + I,o, **u't+ lio'{(p' -!) x r,', *, [;(e;p,)*]


a}]

: Hffi) I no' **u,-". [(v.^n t%ili]) . r] x,',


H*,*xG)+oxG''l,o'*(v x A) *n [i+], Q.216)

which is recognized as the perturbation matrix element expected from the spin
magnetic moment interaction.
Finally, let us consider the case 9r : r/n1. Using (3.215) with P : p', we havef

i
I,{,,tr^hds
x : t,^hror^hd3 x : (ffiH : P#. (3 .2r7)

But this is nothing more than the classical particle velocity divided by c. It is
interesting to recall in this connection that the analo g of - ie I F,y rlrA, d3 x in the
clossical electrodynamics of a point particle is
Hcrassicar : (-ev.Alc) a eAo. (3.218)
Since plane-wave solutions of the form (3.214) are orthonormal, the result
(3.217) can be generalized to any wave function that can be expressed as a super-
position of E > 0 free-particle plane-wave solutions:

S',,: ? ,r:,,r^/Hcr,,u@(p).^n (;# - +), (3.21e)

where co," is a Fourier coefficient whose modulus squared directly gives the pro-
ability for finding the electron in state (p, r). With the help of (3.215), we obtain

(ao)* = x
J" *Lrodu"lrurod,
:xI
: (prclE), ( 3.220)
;
,o)'rr'r'o',r'''r'or,'1n):
where * stands for positive energy. Using similar techniques with negative energy
spinors, we can readily obtain
(ar)- : -(prcflEl> Q.221)
for a wave function made up exclusively of negative-energy plane-wave solutions.
We shall come back later to these very important relations.
Although we have treated only i$,yr{: *ra,,r/n it detail, the reader may
work out the analogs of (3.216) for other matrices ?f u inls1yp, ou,, 1rs, and so forth.
For instance, the interpretation of Qorn"h is of interest in connection with electron-
neutron scattering (Problem 3-6).

fNote that, although we have used the two-component language, we have not made
a single nonrelativistic approximation in obtainine Q.215) and (3.217).
tt2 RELATTVISTTC QUANTUM MECHANTCS OF SprN-+ PARTTCLES 3-6

3_6. DIRAC OPERATORS IN THE HEISENBERG REPRESENTATION


Heisenberg equation of motion. Up to now we have regarded the Dirac matrices
as short-hand devices that rearrange the various components of t. In discussing
the time development of a matrix element such as I+'*(*, t)ao$(x, t)ds x,we shall
find it sometimes more convenient to introduce a time-dependent operator af+t(t)
(where ,F1 stands for Heisenberg) with the property

t)arg(x, t) d3 x: ,o)aln(r)f(*, 0) dt x. (3.222)


J +'n{*, J +'*{*
The time development of J 9"(t, t)ar$(x, t)d3 x can then be inferred directly
from the differential equation that governs the behavior of the operator aL*). AII
this amounts to working in the Heisenberg representation, where the state vector
is time independent and the dynamical operator is time dependent. For this reason
let us briefly review the connection between the Heisenberg representation and the
Schrodinger representation.
By virtue of the Schrodinger equation [or the Dirac equation written in the
Hamiltonian form (3.38) and (3.39)1, a Schrodinger (or Dirac) wave function, which
is not necessarily assumed to be an energy eigenfunction, can be written as
l) - e-iHtln*\(x,0), (3.223)
"/r(x,
where f1 is the Hamiltonian operator that acts on the wave function. Associated
with a time-independent operator O(s) where ^S stands for Schriidinger, the cor-
responding operator in the Heisenberg representation can be defined as
gltrl(l)
- eiut/hA,s) e-iHtih (3.224)
Clearly, at t :0, dltttr coincides with O(s):
gltzr(0) o(s).
-
With (3.223) and (3.224), the matrix elements of a given operator in the two re-
presentations taken between any initial and final state are seen to be the same
provided that in evaluating the matrix element in the Heisenberg representation
we use the wave function at t :0. that is,

J +'*(*, r)otsrr/n(x , t) ds x: J +'*(*,0) orrrg(x, 0) d3x. , (3.225)

By considering an infinitesimal displacement in time, we are able to readily deduce


the Heisenberg equation of motion:

ry::+[H,o(")] *ry:' (3.226)

where the last term is nonvanishing only when O depends explicitly on time.
Note that this equation is equivalent to

# I +'rt*, /)o<^sr9 (x, t) d3 x


: +J +'*1*, t)lH,o(")1.1,(x, t) d3 x+ J',/r'*{*, D(ilo"D lilt)\h!, t) d3x. (3.227)
DIRAC OPERATORS IN THE HEISENBERG REPRESENTATION 113

With a Dirac matrix, ar(F) we may associate a Heisenberg dynamical operator


df:')(P(")) with the property that its explicit matrix representation at t :0 is
dr(p).Since the operalot ,-itrt/h that connects dL') with dr, FL') with B, and
so forth, is unitary, the anticommutation relations among ar and B hold also for
the corresponding dynamical operators. In this section only, we shall omit the
superscript (H) so that a* will stand for the dynamical operator that corresponds
to the matrix ar discussed in the previous sections. Similarly, the symbols x,
p, and L will stand for the velocity operator, the momentum operator, and the
orbital angular momentum operator in the Heisenberg representation.

Constants of the motion. With the aid of the Heisenberg equation of motion we
can immediately determine whether or not a given observable is a constant of the
motion. For instance, for a free particle whose Hamiltonian isf
H-cdipi+B*r, (3.228)

we have the Heisenberg equation for the momentum

#:f,w,nd :0, (3.22e)

since pe commutes with both cdipi and Bmcz. Equation (3.229) tells us that we
can find a solution to the Dirac equation which is a simultaneous eigenfunction
of the Hamiltonian and the momentum. But we know this already, since the plane
wave solutions obtained in Section 3-3 are simultaneous eigenfunctions of H
and p.
As a less trivial example, let us consider L for a free particle. For the x-com-
ponent of L we have

lH, Lrl : fcd*pt, (xrp, - xrp)l : -ihc(d.zps - drpr), (3.230)

and similarly for L, and I'. So

dLldt:c(a x p). (3.23t)


This means that for a free Dirac particle, L is not a constant of the motion, in
sharp contrast with the corresponding operator in the Schrcidinger theory.
,
Next let us consider 2. First it is useful to note that
dn: -2x1s: -yslir, (3.232)

which is obvious from the explicit forms of otr, ryu, and 2r.$ Therefore

1I1, )rl : lcdrpx, Ir] : -cl"yuZrp*, ItI


: 2ic(d.zp, - drpr), (3.233)

fsince it is customary to use dp and B to write H, we shall, in this section, make exclusive
use of ap and B rather than opn and "ya.
$To prove this without recourse to any particular representation, note, for instance,
ds: -irys1y +:iryfyz"f t"fz"yznl+: -Irys,, where we have used (yryr)' -1. :
tt4 RELATTVTSTTC QUANTUM MECHANTCS OF SprN-+ PARTTCLES 3-6

where we have used ryrlrlr : -i,yu2r: id* and so forth. Consequently,


d> 2c.
-At:-h\oxP). (3.234)

Thus the spin angular momentum of a free electron is not a constant of the motion,
either. But taking the dot product of (2.234) and p and remembering (3.229), we
obtain
d(z.p)ldt:0, (3.23s)
which means that the helicity 2.p/lp I is a constant of the motion, as shown in
Section 3-3.
Let us now consider
J:L+h>12. (3.236)
Because of (3.231) and (3.234) we have
dJldt :0. (3.237)
Thus, although L and h>12 taken separately are not constants of the motion, the
sum (3.236) which, according to (3.161), should be identified with the total angular
momentum is a constant of the motion. As is well known, the constancy of J is
a consequence of invariance under rotation. Hence we may argue that J must be
a constant of the motion even if we add to the free-particle Hamiltonian a central
(spherically symmetric) potential V(r). Indeed, the relation (3.237) still holds in
the presence of Z(r) since both Lr and lr commute with Z(r).
Next we shall consider the time derivative of the mechanical momentum
7t:p-eAlc (3.238)
(as opposed to the canonical momentum p) of an electron in the presence of Ar.
Using the Hamiltonian
H : cd.7l * eAo I Bmc', (3.239)
we obtain

i r, : itr, rt nl - +* : + difn i, ?t,l + f, |,to, o rl. (3.240)

But we know that


lAo, nrl: ,uW,
and

ltr,, r,l :'+#, -'+H: Y t,, etc. (3.241)

Hence
tt:e(E* cxB). (3.242)
Since (ao)* has been shown to correspond to the classical particle velocity in
units of c (cf. Eq. 3.220), we may be tempted to identify (3.242) with the operator
equation for the Lorentz force. However, as we shall see later, we have to be
somewhat careful in regarding cc as the operator that corresponds to the particle
velocity in the usual sense.
3-6 DIRAC OPERATORS IN THE HEISENBERG REPRESENTATION ll5

It is also of interest to study the time dependence of 2.t for an electron in the
electromagnetic field. Assuming that A, is time-independent, we obtain

ry: +I(-'r"y,2'x + B*r'|- eA),2.r1: +lAo,2.7rl:er.E, (3.243)

since both ,y5 and B commute with2-rt. Let us suppose that there is no electric
field. We know that the magnitude of the mechanical momentum of a charged
particle is unchanged in a time-independent magnetic field. The constancy of
2.rr then amounts to the constancy of the helicity. It follows that a longitudinally
polarized electron (one whose helicity is * I or 1) entering a region with a mag-
-
netic field will remain longitudinally polarized no matter how complicated B may
be. When an electron is injected into a region with a uniform magnetic field B
whose direction is perpendicular to the initial electron velocity, the electron follows
a circular path with an angular frequency, known as the cyclotron frequency:
a)L: @lBllmc) JfF". (3.244)
In this particularly simple case, the constancy of helicity implies that the electron
spin precesses in such a way that its precession angular frequency or" is equal to
a)7,, ds pictorially represented in Fig. 3-3.

Fig. 3-3. Spin precession of a moving electron in a uniform magnetic field. The gray
arrows indicate the spin direction.

In deriving the equality of a"and c,rs we implicitly assumed that the gyromagnetic
ratio of the electron is strictly two. In reality, because of the anomalous moment
of the electron, a" and c,ls &r0 not quite equal. It can be shown that the correct
relation is given by

(3.24s)

This means that as the initially longitudinally polarized electron makes one orbital
turn, its spin direction departs from the direction of motion by a very small amount,
lll37 rad if the electron is nonrelativistic (cf. Eq. 3.210). This principle has been
used experimentally to make precise measurements on the anomalous moments
of the electron and the muon.

"Velocity" in the Dirac theory. Let us now return to the free-particle case. Consider
(ilh)lH, xf, -- (iclh)la,p,, xf, : cd6,
*n: (3.246)
which says that a is the velocity in units of c. (Actually this relation holds even
I 16 RELATrvrsrrc eUANTUM MEcHANrcs oF sprN-+ IARTTcLES 3-6

in tlre presence of A *.) At first sight this appears quite reasonable in view of (3.220),
which says that (au)* is the same as the expectation value of pycH-t. Note, how-
ever, that the eigenvalue of au is *l or -1. Hence the eigenvalue of the velocity
operator is *c, as first pointed out by G. Breit in 1928. This is a truly remarkable
result, since for a particle of finite mass the classical velocity cannot be equal to
:Lc. We may also note that because ar and a1 do not commute when k + l, a
measurement of the x-component of the velocity is incompatible with a measure-
ment of the/-component of the velocity; this may appear strange since we know
that pl and p, commute
In spite of these peculiarities, there is actually no contradiction with the results
derived earlier. The plane-wave solutions (3.114) and (3.115) which are eigen-
functions of p are not eigenfunctions of ar (unless the particle is massless), as
the reader may verify by directly applyin g d* In fact, since an fails to commute
with the Hamiltonian (see Eq.3.248), no energy eigenfunctions are expected to be
simultaneous eigenfunctions of ar.
Let us now look at the time derivative of Ar:
dx : Qlh)lH, af,
: (ilh)(-2arH * {H, au})
: (ilh)(-2auH * 2c p), (3.247)

where we have used the fact that Ar anticommutes with every term in f1 except
the term that involves cuft itself. This equation can also be written as

2iar?mc2
dr: -+() x p),,
- - (3.248)

We see that the velocity operator *r: capis not aconstant of themotiondespite
the fact that the particle is free. Contrast this with Eq. (3.229), which says that
the momentum of a free particle ls a constant of the motion.
The relation (3.247) can be regarded as a differential equation for au|).
Keeping in mind that pr and H are constants of the motion, we easily see by direct
substitution that the solution of this differential equation is
arQ) : cprH-t * (ao(0) - cpuH-t) e-zillt/h (3.24e)

The first term of Q.2a\ is reasonable since, for an eigenstate of momentum and
energy, it gives cprlE in agreement with (3.220) and(3.221). But what is the physical
significance of the second term? Taken literally, it seems to say that the velocity
of the electron has an additional term that fluctuates rapidly about its average
value even in the absence of any potential.
As for the coordinate operator, the relation (3.249) can be easily integrated
to yield
xu(/) : x*(0) * c'p,,H-1t + (ichl2)(a-(0) cprH-')H-t e-2iIIt/h. (3.250)
-
The first and the second terms are understandable because their expectation values
are seen to give the trajectory of the wave packet according to the classical law:
x[,,1**')1/; : x[.r,,'*)(O) I t(puc2/6:;t"t"'), (3.2s1)
3-7 "zITTERBEwEGUNG"ANDNEGATIVE-ENERGysoLUTIoNs ll7

just as they do in nonrelativistic quantum mechanics. The presence of the third


term in (3.250) (which, of course, is a consequence of the second term in Eq. 3.249)
appears to imply that the free electron executes very rapid oscillations in addition
to the uniform rectilinear motion (3.251). This oscillatory motion, first discussed
by E. Schrodinger in 1930, is called Zitterbewegung (literally "quivering motion").
We shall say more about this in the next section.

3_7. ZITTERBEWEGUNG AND NEGATIVE.ENERGY SOLUTIONS


Expectation values of c and x. The algebraic techniques extensively employed in
the previous section are quite powerful when we want to obtain the constants
of the motion or establish a correspondence with the classical theory. However, in
order to analyze the peculiarities we encountered at the end of the previous section,
it is instructive to go back to the Schrodinger representation and reinterpret the
operator relations (3.249) and (3.250), using specific wave functions.
Since the positive- and negative-energy plane-wave solutions (3.1l4) and (3.115)
with all possible p form a complete orthonormal set, the most general free-particle
wave function can be written as

r) : exP lip'x rlElr\


^1,(x, ? ,Z,r'rmco,'rz(')(P) \ft - h)

^'tm
| ? ,4, c'''u(il(P)'.P (+r *'+)' Q'2s2)

where co,, is to be determined from the Fourier expansion of at t :0. By ap- *


propriately choosin E cp.,, we can write the wave function for an arbitrarily localized
free-particle wave packet in the form (3.252). Let us now evaluate (au). A
straightlorward calculation using (3.220) and (3.221) gives

(a r) _,:,:^,:;u,
: | ;: lo).
u,t,., t,
ffi
+>p r'-t,2r'-3,4 lLl
* cp,,,c;,,uo)t(p)aru?')(p) e2itzttinl. (3.253)
The first (second) term which is time independent represents the group velocity
of the wave packet made up exclusively of positive- (negative-) energy plane-wave
components. The last two terms which are time dependent are more interesting.
First note that a* taken between ,(3'4)t(p) and u(t't)(p) is "large" when lplK*r,
in sharp contrast with a, taken between ,r(t'2)+(p) and u(t''zr(p) which is of order
o/c. Specifically,
uir or a)r(p;
d,,tt(r.. r)(p) - XG-)+ okx('.) + Q(llpllmclr), (3.2s4)
where X(t) and X('-) are the two-component Pauli spinors corresponding to
,(t or 2) and a(3 u" a). Therefore the last two 'terms of (3.253) represent a superposi-
tion of violent and rapid oscillations, each with an angular frequency -2mc2lh
: 1.5 X l02t Sec-r and an amplitude -lc(r'z).6(r'+)1.
ll8 RELATIVISTIC QUANTUM MECHANICS OF SPIN-+ PARTICLES 3-7

As for (x) we first observe that the operator relation Q.ZaQ implies

(dldt)
J +*(r, t)xr"l(x, t)d3 x : , !9*(x, t)ar"l(x, t) d3 x. (3.2ss)

We can therefore obtain (x*) by integrating c times (3.253) with respect to time:

(x*):("r),:o*h A,,l rn,rf t a@f - >,1,nl ro,,r #,


t IE h
+ >,?",,4, nh ffi)' k {,,, r r, 1. u@ + (p) d ru@ (p) e - 2 tt /

- cor,c[ruo)+(p)d.ku(r')(p) e2itEtt/rtl. Q.256)

Thus on top of the rectilinear motion of the wave packet we again have a super-
position of violent oscillations, each with an angular frequency -2mc'lh.If both
cp, r or 2 and cp, 3 or a a;ta appreciable, the fluctuation of the electron coordinate due

to these oscillations is of the order of hlmc :3.9 x 10-11 cm. It is very important
to note that the peculiar oscillatory behavior, Zitterbewegung, of both (ar) and
(x*) is due solely to an interference between the positive- and negative-energy
components in the wave packet. The Zitterbewegung is completely absent for
a wave packet made up exclusively of positive- (negative-) energy plane-wave
solutions.

Presence of negative-energy components. At this stage we may naturally ask why


we cannot simply forget all about the trouble-making negative-energy components
in constructing a wave packet. This can certainly be done for a free-particle wave
packet since in a potential-free region a wave packet made up exclusively of posi-
tive-energy plane waves at a given time does not develop negative-energy com-
ponents at later times. On the other hand, we can show that the wave function
for a well-localized particle contains, in general, plane-wave components of nega-
tive energies. As an illustration, let us consider a harmlesslooking four-component
wave function at t :0,

f(x, o) :[f) (3.2s7)

where ld(x)l' is assumed to be appreciable only in a region whose linear dimension


is Ax*. From our earlier discussion in Section 3-3, we see that r/n corresponds
to the four-component wave function for a nonrelativistic particle with spin-up
localized to - A,xx. We now propose to expand this wave function in various
plane-wave components so that r/r(x, 0) takes the form (3.252), evaluated at t :0.
The appropriate Fourier coefficients can be readily found by multiplying (3.257)
from the left by yc)+(p)s-to'x/n and integrating over the space coordinates.
Although the coefficients themselves depend on the detailed form of f(x), we
3-7 "ZITTERBEwEGUNG" AND NEGATIVE-ENERGy soLUTIoNs 119

easily see that


cp,s Ptc _ _(pt - ipr)c
cp,+ . (3.258)
r-: - - @lprz' cp,r- lEllmc'z
This means that a negative-energy component is comparable in importance to
the corresponding positive-energy component whenever f has a Fourier com-
ponent with momentum comparable to mc. On the other hand, we know that
the Fourier transform of f is appreciable in a region in momentum-space whose
linear dimension is
LPr - hlA'xv. (3.2se)

Suppose the state in question is localized to Lxr S hlmc. The uncertainty relation
(3.259) tells us that we need plane-wave components of momenta lplZ mc.
From (3.258), we then infer that there must be appreciable amounts of negative-
energy components.
We have seen that a well-localized state contains, in general, plane-wave com-
ponents of negative energies. Conversely, we may ask how well localized a state
will be which we can form using only plane-wave components of positive energies.
A rather careful analysis by T. D. Newton and E. P. Wigner indicates that the
best-localized state we can construct in this way is one in which the characteristic
linear dimension of the wave packet is - hlmc but no smaller. This statement can be
shown to be valid also for a Klein-Gordon particle.
It is interesting to note that a positive-energy bound-state wave function when
expanded in free-particle plane waves usually contains some negotive-energy com-
ponents. For instance, let us take the case of the ground-state wave function of the
hydrogen atom whose energy is evidently positive, E : mcz - (e'zl8ztao) ) 0. When
this bound-state wave function (whose explicit form will be given in the next
section) is expanded in free-particle plane waves, we do obtain nonvanishing
coefficients for negative-energy plane waves. An immediate consequence of this
is that the electron in the hydrogen atom exhibits Zitterbewegung. As a result,
the effective potential that the electron at x feels is no longer just Z(x) but rather
V(x + 8x), where Ex characterizes the fluctuation of the electron coordinate.
Note now that V(x + 8x) can be expanded as follows:

v(x +Ex) : v(x) +8x.V, + +8xi8x1 .. . .260)


#,+ (3

Assuming that x fluctuates with magnitude l8x | = hlmc without any preferred
direction; we obtain for the time average of the difference between
V(x* E*) -V(x):
:
(AZ),,,"u averase =+ +(*)' v2 v
+(*)' 5ttr(x). (3.261)

Apart from a numerical factor (f instead of +), this is just the effective potential
needed to explain the Darwin term discussed earlier in connection with the ap-
proximate treatment of the hydrogen atom (cf. Eqs. 3.83 and 3.87).
120 RELATIvIsTIc euANTUM MEcHANIcs oF sPIN-+ PARTIcLES 3:7

Region I Region II

Oscillatory

Fig. 3-4. One-dimensional potential with mc2 >E- ht) -mcz.

Klein's paradox. As a final illustration of the peculiarities attending the negative-


energy solutions, let us consider a simple one-dimensional potential (Fig. 3-4).
In Region I the particle is free; the height of the potential in Region II is assumed
to be Zo. When considering a region in which the potential is not varying rapidly,
we can proceed directly to obtain the functional form of the wave function. For our
purpose it is actually sufficient to look at f1 only:

(o'p) cT+W@'p)c*u: *rt - V- mc')nlo. Q.262)

Wherever Z is locally independent of x, we obtain

*nn " p(ry -+)x, exP


?ry -,ui)*, (3.263)

where :
p'cn (E - V I mc')(E - V - mcz). (3.264)
With p' > 0 we have an oscillatory sohttion, while with p2 ( 0 we have an ex-
ponentially damped solution. Let us suppose that
mc2>E-Vo)-mcz. (3.26s)
The Region II is a classically forbidden region (p' < 0), and the free-particle
wave function in Region I dies out exponentially as it enters Region II.
So far everything ha\ been straightforward. Let us now consider the potential
given by Fig. 3-5. Our experience with nonrelativistic quantum mechanics tells
us that the wave function in Region III is eyen more strongly damped. However,
when the potential becomes so strongly repulsive that
Vn
- E > mc2, (3.266)
(3.264) tells us that just the opposite is true. Since both E-V*mc' and E-Y-mcz
are now negative, we have p' > 0; hence the solution in Region III is oscil-
latory just as is the free-particle solution in Region I. This result is exactly the
opposite of the one we set out to find. Semiclassically speaking, a particle initially
confined in Region I can tunnel through Region II (just as the a-particle inside
an a-emitting nucleus), and behaves in Region III as though it were in an attractive
potential instead of the very strong repulsive potential implied by (3.266). This
theory is named Kletn's paradox for O. Klein, who worried about this interesting
point in 1930.
3-7 "zITTERBEwEcuNc" AND NEcATIVE-ENERGy soLUTIoNs 121

V1

!ig. 3_5. Potential to illustrate Klein's paradox. Oscillatory solutions are expected in
shaded regions.

What is the origin of this peculiar behavior? Let us recall that the free-particle
solutions to the Dirac equation exhibit an energy spectrum ranging from
-mcz
to -oo as well as from *mc' to oo. Now suppose we apply a small positive poten-
tial V. The condition that we have negative-energy oscillatory solutions now
becomes
-oo { E 1-mc' + V. (3.267)

As V is increased adiabatically, we see that eventually ^E need not even be negative


for (3.267) to be satisfied. Coming back to Fig. 3-5, we see that the oscillatory
solution in Region III is essentially a negative-energy solution despite E > O,
since it can be obtained from the wave function whose space-time dependence is
expli(pxlh) + i(lEltlh)l by simply increasing V adiabatically. Klein's paradox
arises because when the potential V is sufficiently positive, an oscillatory negative-
energy solution in Region III can have the same positive energy as an oscillatory
positive-energy solution in Region I. The tunneling of the electron from Region
I into Region III must therefore be viewed as a transition from a positive-energy
to a negative-energy state. We shall say more about such a transition in Section
3-9, pp. l3l-l43.In any case, we find that our intuitive notion that a strong posi-
tive potential can repulse the particle breaks down completely when Z becomes
comparable to 2mc2.
Similar peculiarities are present also for strongly attractive potentials. With a
moderately attractive finite-ranged potential we can have bound-state solutions
(E < mcz) which fall off outside the range of the potential, just as they do in non-
relativistic quantum mechanics so long as the attraction does not exceed a certain
critical strength. But when the potential becomes too strong, the Dirac theory
starts accommodating solutions with -E /ess than mcz which are oscillatory and
undamped outside the range of the potential. The interested reader may verify this
point in detail by studying the behavior of the Dirac particle in a deep spherical
well (Problem 3-10c).
122 RELATrvrsrrc euANTUM MEcHANIcs oF spIN-+ PARTIcLES

3_8. CENTRAL FORCE PROBLEMS; THE HYDROGEN ATOM


General considerations. In this section we shall first study some general properties
of the wave function for an electron in a spherically symmetric potential. As we
have already seen in Section 3-6, the total angular momentum operator J is a
constant of the motion whenever the Hamiltonian is given by
H-ca.p+Bmc'*V(r). (3.268)

Let us now look for other constants of the motion. Intuitively, we expect that we
must be able to specify whether the electron spin is parallel or antiparallel to the
total angular momentum. In nonrelativistic quantum mechanics these two pos-
sibilities are distinguished by the eigenvalues of
o.J : o.(L * hol2): (l/n)(J2 -L2 + th'). (3.26e)

Alternatively, we may specify /, which can be eitherT * t or j - f. For a relativistic


electron we might try the 4 x 4 generalization of (3.269), namely I.J. However,
the commutator of 'H with 2.J turns out to be rather involved, as the reader may
verify. Instead, then, let us try B2.J which has the same nonrelativistic limit
as I.J:
lH, p >.Jl : lH, pl
>.J + plH, >l'J
: -2cF(cr.pX>.J) + 2icB(a x p).J, (3.270)
where we have used (3.233), and
lH, Pl: cd'PP
- Fre'P
: -2c9a.p. (3.27t)
If we take advantage of
(a.A)(2.8) : -ryu(2.AX>.8)
: -ryrA.B f ic.(A x B), (3.272)
we can simplify expression (3.270):
H, B >'J] : 2c F,.vu(p'J)
::':!:;i
t

: :,:i,l,ll, (3.273)
where we have used ^,
p.L : -ihV.[* x (-;nV;] : g (3.274)
and (3.271). Therefore an operator K defined by
K: F >.J - phl2: p(>.L + h) (3.27s)
does commute with I/:
lH, K1: g. (3.276)
Furthermore, using the fact that J commutes with B and >. L, we readily see that
[J,K1 :9 (3.277)
for an electron in a central potential, we can construct a simultaneous
as wel[. Thus,
eigenfunction of H, K, J2, and ,/r. The corresponding eigenvalues are denoted
by E, - K, j(j + t)h', and ish.
3-8 CENTRAL FORCE PROBLEMS; THE HYDROGEN ATOM t23

We shall now derive an important relation between rc and j. First let us consider
1s2-p(>.t+h)p(>.L+h)
: (>'L + h)'
:Lz + t>.(L x t) + zh>.L + h'
:L2+n>.L+h'. (3.278)
At the same time, since
J' : L' + hZ.L + 3h'z 14, (3.27e)
we obtain
Kz __ J, + (3.280)
+hr,
which means that the eigenvalues of J' and K2 are related to each other by
tc2
- j(j + l)h'+ *h': (j + +)'h' (3.281)
So we must have
rc:*(i++). (3.282)
Thus rc is a nonzero integer which can be positive or negative. Pictorially speaking,
the sign of r determines whether the spin is antiparallel (rc > 0) or parallel (r < 0)
to the total angular momentum in the nonrelativistic limit.
Explicitly, the operator K is given by

*:(o'L+h o \ (3.283)
0 -o.L - h)'
Thus, if the four-component wave function r/n (assumed to be an energy eigen-
function) is a simultaneous eigenfunction of K, J', and "Ir, then
(o.L * h)*o: -rch*d, @.L + h)"h": rch{n, (3.284)
and
J'*^,o : (L + nol2)'"h",": i(i { l)hz\ra,n,
(3.285)
Js$.r,n : (L, * hosl2).lre,u: jrhnlta,n.

The operator L2 is equal to J2 ho.L


- - *h'when it acts on the two-component
wave functions *;_and r/^". This means that any two-component eigenfunction of
o.L + h and J2 is automatically an eigenfunction of L2. Thus, although the four-
component wave function rfn is not an eigenfunction of L2 (since fl does not com-
mute with L2), *d and Sa separately are eigenfunctions of L2 whose eigenvalues
are denoted by l^(1" * l)h' and l"(lo * l)h2. From (3.284) and (3.285) we then
obtain
-tc: j(j +l)- lA(la + 1) ++, *: j(j + l) - l,,(lu + 1) ++. (3.286)

Using (3.282) and (3.386), we can determine lo and lu for a given rc. The results
are summarized in Table 3-2.
For a givenj, locan assume two possible values corresponding to the two possible
values of r. This fact is already familiar from our study of nonrelativistic quantum
mechanics. For example, forT : *, Io can be either 0 or I (s] or p$),depending
on whether rc is negative or positive. What is new is that for a fixed rc the orbital
124 RELATIVISTIC QUANTUM MECHANICS OF SPIN.+ PARTICLES 3-8

Table 3-2
RELATIONS AMONG r, j, It AND /B

parities of and qB are necessarily opposite. As we showed in Section 3-4,


"h"
this result can also be derived from the requirement that the four-component
wave function r/n have a definite parity (cf. Eqs. 3.172 through 3.175).
We can now write * as

o:(ilJ:f,fK;;,,), (3.287)

where Uli stands for a normalized spin-angular function (an r-independent


eigenfunction of J', J r, L2, and, of course, 52) formed by the combination of the
Pauli spinor with the spherical harmonics of order /. Explicitly,

uJi: Y1,-'rz(;) - (3.288)


"r'.t,'(? )
forT: l!!,and
(3.28e)
",".',(?)
forT - I - f.f The radial functions f and g depend, of course, on r. The factor
i multiplying/has been inserted to make/and g real for bound-state (or standing-
wave) solutions.
Before we substitute (3.287) in the Dirac equation written in the form

c(o.p)"ho: (E - V(r) - mc')rlra, c(o'p)*e: (E - V(r) a mc'),\u, Q.290)


let us note that
o.p :A?(o.x)(o.p)
:9(-n,fi + io.n). (3.2er)

Moreover, the pseudoscalar operator (o.x)/r acting on U'1i^ must give an eigen-
function of J2, Jr, and L2 with the sameT andTr but of opposite orbital parity.There-
fore [(o .x)lr]UJi^ is equal to UJi, itself up to a multiplicative phase factor. lNote:
(o.x)'lr2: 1.1 It is not difficult to show that this phase factor is minus one if we

[See, for example, Merzbacher (1961), p. 402. Throughout this book we shall follolv
the phase convention used in Condon and Shortley (1951), Rose (1957, 1961), Merzbacher
(1961), and Messiah (1962). The phase convention used by Bethe and Salpeter (1957) is
slightly different due to an unconventional definition of Yf.
3-8 CENTRAL FORCE PROBLEMS; THE HYDROGEN ATOM t25

conform to the phase convention used in writing (3.288) and (3.289). In fact,
we have already verified this for the special case j : jr':
t, lt:0,
as seen from
(3.178). similarly [(o.x)lr] acting on wJi" gives wJi^ apart from a minus sign.
Thus

p)*" :,@;f) (-
(o.
fi + io.t)fu ji"
in,

: ,(o;d) (-in,ff + i(* - ryn1)uii"


(3.2e2)

Similarly
(o.p)*u: ih#*rr" a ;U-D! sg1i". (3.2e3)

we now observe that the spin-angular functions completely drop out when we
rewrite (3.290), using (3.292) and (3.293):

-rh+drr-0:rc)hf :(E v-mc,)s,


(3.2e4)
,h#*U
dr '
+-rc)n g: (E
r - V'+ mc,)f.
Introducing
F(r): rf(r), G(r) : rg(r) (3.295)
as in nonrelativistic quantum mechanics, we finally get radial equations:

- ldF tt
-\ :
o, - ; F)
\a,- -(E - v - mc2)G,
Q'296)
n, (4+ 11 C):
-\ (E V + mcz)F.
maz\F
\d- -
Hydrogen atom. On the basis of the coupled equations (3.2g6)a variety of problems
can be attacked. We shall consider only one problem along this line; the remaining
part of this section will be devoted to a discussion of an electron bound to the
atomic nucleus by a Coulomb potential. This classical problem (first treated by
C. G. Darwin and W. Gordon in 1928) can be solved exactly. The reader who is
interested in other central-force problems-the anomalous Zeeman effect, free
spherical waves, exact solutions (as opposed to Born approximation solutions
.to be discussed in Chapter 4) to the Coulomb scattering problem, etc.-may
consult Rose's book.f
In order to simplify (3.296) when Z is given by
V : -(Zezf 4rr), (3.2e7)

fRose (1961), Chapter 5.


126 RELATrvrsrrc euANTUM MEcHANICS oF sPIN-+ PARTICLES 3-8

we introduce
d.1 : * E)lhc, dz:
(mc' (mc'
- E)lhc, (3.2e8)
,: (Ze'zl4zrhc) : Zd. - Zll37, p : n@rr.
Note that ht/a%: Jfrcr E/c is just the magnitude of the imaginary mo-
mentum of an electron of energy E. The coupled equations we must solve are

(fr- t), - (^l*,- G : o, (h. t), - Gfn #)F: o


#)
(3.zee)
As in the nonrelativistic treatment of the hydrogen atom, we seek solutions to
(3.299) of the form
F : e-P P'Ao^r*, G : P'Pob*P*' (3'3oo)
'-o
Substituting (3.300) in (3.299), and equating the coefficients of e-op'po-r, we obtain
the recursion relations
(s * 4- rc)ao
- aq-, * - JAJdrbq-r : 0,
rybo
(3.301)
(s * 4* rc)bo
- bn_,
-,f as - Jdrldroq_1 :0.
For q : 0, we have
(s
- rc)ao * ybo : 0, (s f r)bo - Iao: 0. (3.302)

Since a, and bo are not zero, the secular determinant of (3.302) must vanish; hence

t: tJP - "y'. (3.303)

We must require that J *r*dt, be finite. This requirement amounts to

J lFl'dp
( @, J tct' dp I a. (3.304)

Thus ,F and G must behave better than p-'tz at the origin, which means s>-f.
Since
Kz
- ,y'> min (rcz) - ,y' -l - (Zll31)'z, (3.305)

the above requirement cannot be satisfied if we take the negative root of (3.303).
So we are led to take the positive root.I
It is not difficult to show that F and G would increase exponentially as p ----- oo
(that is, F, G - e+p at infinity) if the power series (3.300) did not terminate.$
Assuming that the two series terminate with the same power, there must exist
n' with the property
on,+t: bn,*t: 0, on,* 0, bn,* 0. (3.306)

fFor lrcl : 1, f : Flr and g : Glr diverge at the origin (since s < 1); yet (3.304) is
satisfied.
gThe ep behavior for F and G at infinity is allowed if E > mc2, which means a purely
imaginary p. Indeed the oscillatory behavior of the radial functions at infinity is charac-
teristic of scattering-state solutions which exhibit a continuous energy spectrum.
3-8 CENTRAL FORCE PROBLEMS; THE HYDROGEN ATOM 127

Setting Q: n' * I in (3.301) we obtain


an, -- -"/aJarO", (3.307)
from both parts of (3.301) (which incidentally justifies our assumption that the
two series terminate with the same power). To write an equation that involves
only an, and bn, whose ratio is now known, we multiply the first of (3.301) by
ar,the second bv .;@r, setq: n'this time, and subtract:
[a'(s I n' -,c) + ,r@rntlan, - tr@(s * n'* ") - drf)b,, - 0; (3.308)
hence
2\/dtd., (s * n') : ny(dr
- dr), (3.30e)
or
J@ry-P (s * n') - E,y. (3.3r0)
Thus we obtain the energy eigenvaluesl
mc2
E: mc2
Zzn
(3.311)
_
'(n'+ffi72
Note that E depends only on n' and jI t:l*1.
In order to compare (3.311) with the corresponding expression obtained in the
Schrodinger theory, we define
n-nt+U++):n'*|rc|. (3.312)
Since the minimum value of n' is zero, we have
n>j*t:l*1, (3.3r3)
which is at least unity. Expanding (3.311) we get

E : mc2[r _ +ry _ +s#(** _ *) _ ] (3 3r4)

Since$
$d.2mc2 : e2 f (8tt a*o1,,), (3.315)
we see that n is indeed identical with the familiar principal quantum number in
nonrelativistic quantum mechanics. Note also that the leading correction to the
Balmer formula is precisely the fine-structure splitting (3.88) which tells us that,
for a given n,higher j-states are at higher levels.
In the Dirac theory each state of a hydrogen atom can be completely charac-
terized by n', rc, and "lr. V7r can translate this classification scheme into the more
familiar one based on spectroscopic notation. This is done in Table 3-3 which
can be obtained with the help of Table 3-2 and Eqs. (3.312) and (3.313). Note that
although L2 is not "good" in the relativistic theory, it is customary to use the

fFormula (3.311) was first obtained by A. Sommerfeld, using a relativistic version of


N. Bohr's old quantum theory.
$In this section we shall us€ 4s61p rather than ao for the Bohr radius to avoid a possible
confusion with the coefficient ao in (3.300).
128 RELATIVIsTIc euANTUM MECHANIcs oF SPIN-+ PARTICLES

Table 3-3
RELATIVISTIC QUANTUM NUMBERS AND SPECTROSCOPIC
NOTATION
A pair of states which have the same energy according to (3.311) are denoted by "deg."

n nt:n -lrl>0 rc:*(i*L\ Notation

I -1 lsi
) -l 2sLl
) +1 zpLl
2pZ
2 -2 3s*
3 -l )

3 +1 tpLI
3 -2 3pil
3 +2 3d:,l
3 -3 3dz

notation ptp, atc., which actually means ln: I with 7 : $, etc. In other words the
orbital angular momentum of the upper two-component wave function (which
becomes the wave function of the Schrodinger-Pauli theory in the nonrelativistic
limit) determines the orbital angular momentum in the spectroscopic language.
The reader may wonder why we have omitted in Table 3-2 the rc ) 0 states when-
ever n' : 0. The reason for this omission is evident if we go back to the second
expression of (3.302) and (3.307) which together imply

f/(s * rc) : -Jdrl-dt, fr' : 0 only. (3.316)

This can be satisfied only if rc is negative because s is a positive number smaller


than lrc | (cf. Eq. 3.303). The absence of the r ) 0 states for n' : 0 corresponds
to the familiar rule in nonrelativistic quantum mechanics: The maximum value
of/isn-l,notn.
For the ground state (n' :0, K: -l), the relation (3.311) simplifies to
tr :7n62"rc. (3.317)
So
Jdtdr: Zd'mc2lhc : Zlauon,, (3.318)

bo_
_: Za _ (1 -"re (3.31e)
as l+JI-(Za\'- Za
Up to an overall multiplicative constant denoted by N we can readily write the
ground-state wave function

- z a 7z -1
txG)\
{ a : h (*)"
(

s
r- z r / a soa,
(*)"' - r**rq! -,
r,
\,, ,)

(3.320)
CENTRAL FORCE PROBLEMS; THE HYDROGEN ATOM 129

where X(') is the Pauli spinor

(;) .r fi),
depending on whether jr: t or
-+. A straightforward calculation shows that
the normalization requirement for ,f gives

,u-)JT=tzatz-r @
i'i-' (3.321)
hlw,
where

r(") : [ ,-t p1' 4y,


(3.322)
l(m) : (m - l)! for la : positive integer.
Note that N approaches I as Zd. -,0. Furthermore
(Zrf a",y,)J@-r is essentially unity except at distances of order

', _ l37h 0_2(137)2/22 (3.323)


2mcZ"
As r .---- 0, r/r exhibits a mild singularity. This, however, is of academic interest
only, since the wave function at short distances must be modified because of the
finite charge distribution of the nucleus. Thus we see that for the ground states of
hydrogen-like atoms with low Z, the upper two-component wave function is
essentially identical to the Schrodinger wave function multiplied by a Pauli spinor.
As for the lower two-component wave function, we merely remark that, apart
from i(o.x)lr, the ratio of the lower to the upper components is given by (cf.
Eq. 3.307)

(3.324)

where a is the "velocity" of the electron in Bohr's circular orbit theory. This result
is in agreement with our earlier discussion in Section 3-3 on the ratio of r/nr to
*".
In 1947 W. E. Lamb and R. C. Retherford observed a splitting between the
2st- and 2pt-states of the hydrogen atom not given by (3.311). As already dis-
cussed in Section 2-8, the main part of this "Lamb shift" can be satisfactorily
accounted for when we consider the interaction of the electron with the quantized
radiation field.
Another important effect not contained in (3.311) arises from the interaction
between the magnetic moment of the nucleus and the magnetic moment of the
electron. In the case of the hydrogen atom, for instance, when we compound the
electron spin with the proton spin, the net result is ,F : I (triplet) or F: 0 (sin-
glet), where F is the quantum number corresponding to the total spin. Since the
magnetic interaction is dependent on the relative orientation of the two magnetic
dipole moments, each level of the hydrogen atom characterized by n, j, I(: /r) is
split further into two sublevels corresponding to the two possible values of f'
130 RELATIVISTIC QUANTUM MECHANICS OF SPIN.+ PARTICLES 3-8

even in the absence of any external magnetic field. This is known as a hyperfine
splitting. Let us estimate it for the s-states using nonrelativistic quantum mechanics.
Classically the magnetic field created by the proton magnetic momenf l\{(ctass) ir1

B:v*(*,"*.)xv#r) Gszs)
Quantum-mechanically we replace llf(crass) by the magnetic moment operator
M:Woo,
zmec
(3.326)

where o,, is the Pauli matrix for the proton spin and 2(L + r) is the g-factor of the
proton. (We assume that the magnetic moment distribution of the proton is point-
like.) Within the framework of the Schrddinger-Pauli theory we obtain the in-
teraction Hamiltonian operator

Irr(hf)- -p,.Y *(*


"V#)
: (p.M) r, (+,) - tc.r. v)(M .yl #,
: ?Qr.M)v' (#)- [,".vXM.v) - !*.*o')#,, G.327)

where trt, : (ehl2m"c)o". The quantity in the brackets transforms like a traceless
tensor of rank two; so when it is integrated with a function of x, f(x), it gives
a nonvanishing contribution only if the expansion of/(x) in spherical harmonics
contains YT. For the spherically symmetric s states, only the first term of (3.327)
is relevant. Using the nonrelativistic wave function 9,,, w€ obtain the energy shift

'i AEn- -Stt.VrJat'r1x)lflx)l'd'x


:l+#l t+(#)'l ou.op

: 3 a'(rr (fr)# (3.328)


{_::, i:":,
rc)

as first shown by E. Fermi in 1930. Note that the order of magnitude of this split-
ting is the fine-structure splitting multiplied by (m"lm). For the lsf-state, the
above energy difference corresponds to a radio microwave of 1420 Mc or 2lcm;
the radiative transition between these two hyperfine levels is of fundamental
importance in radio astronomy. We may parenthetically mention that this energy
difference is one of the most accurately measured quantities in modern physics;
S. B. Crampton, D. Kteppner, and N. F. Ramsey have determined that the corre-
sponding radiofrequency is (1420.405751800 + 0.000000028) Mc.
There are other corrections to formula (3.31l). First, we must take into account
the motion of the nucleus since the mass of the nucleus is not infinite; a major

fPanofsky and Phillips (1955), p. l2O; Jackson (1962) p. 146.


3-9 HOLE THEORY AND CHARGE CONJUGATION l3l

part of this correction can be taken care of if we use everywhere the reduced mass
mum,,f (m. * mn) in place of m". Second, there are other contributions to the Lamb
shift not discussed in Chapter 2; especially important is the vacuum polarization
effect to be discussed later. Third, the finite size of the nucleus also modifies for-
mula (3.311) especially for the s states which are sensitive to small deviations from
Coulomb's law at close distances; in the interesting case of the 2s state of the
hydrogen atom, however, we can estimate the energy shift due to this effect to be
only 0.1 Mc, using the observed proton charge radius - 0.7 x l0-t3 cm.
The utility of the Dirac theory in atomic physics is not limited to light hydrogen-
like atoms. For heavy atoms where (Za)'is not very small compared with unity
(0.45 for uranium), the relativistic effects must be taken into account even for
understanding the qualitative features of the energy levels. Although we cannot,
in practice, study one-electron ions of heavy atoms, it is actually possible to check
the quantitative predictions of the Dirac theory by looking at the energy levels
of the innermost (K-shell and Z-shell) electrons of high Z atoms which can be
inferred experimentally from X-ray spectra. Similar studies have been carried
out with muonic atoms (atoms in which one of the electrons is replaced by a nega-
tive muon).
Although we shall not discuss the emission and absorption of radiation using
the Dirac theory, the results of Section 2-4 are applicable mutatis mutandis.
All we need to do is make the following replacements:I
lr^,^\eh
- ,*(P'A -r A'P) - ffio'B-------+ -a'A
r;lr( schriidinser-Pauli ) -----> Dirac)
{n(
th(Schriidineer-Pauli)f (nOt.[roiru"l;. (3.32g)
--1,{n(Dirac)t

3_9. HOLE THEORY AND CHARGE CONJUGATION


Holes and positrons. Although we have shown that the Dirac theory accommodates
negative-energy solutions whose existence should not be ignored altogether,
we have as yet not examined their physical significance. This section is devoteij
to the physical interpretation of the negative-energy states within the framework
of a theory in which the electron field is not quantized.
As a simple example to illustrate some of the difficulties with the original Dirac
theory of 1928, let us consider an atomic electron. According to the quantum
theory of radiation developed in the previous chapter, an excited atomic state
can lose its energy by spontaneously emitting a photon even in the absence of
any external field. This is why all atomic states, with the exception of the ground
states, have finite lifetimes. In the Dirac theory, however, the so-called ground
state of an atom is not really the lowest state since there exists a continuum of
negative-energy states from
-mc2 to -co for any potential that vanishes at

fWe have to be a little more careful when we treat a process in which the quadratic A2
term in the nonrelativistic Hamiltonian is important. This point will be discussed in the
next section with reference to Thomson scattering.
t32 RELATIVISTIC QUANTUM MECHANICS OF SPIN.+ PARTICLES 3-9

infinity. We know that an excited atomic state makes a radiative transition to the
ground state; similarly, we expect that the atomic electron in the ground state
with energy mcz -lE""l can emit spontaneously a photon of energy Z2mc2
and fall into a negative-energy state. Furthermore, once it reaches a negative-
energy state, it will keep on lowering its energy indefinitely by emitting photons
since there is no lower bound to the negative-energy spectrum. Since we know
that the ground state of an atom is stable, we must somehow prevent such cata-
strophic transitions.
Faced with this difficulty, Dirac proposed, in 1930, that all the negative-energy
states are completely filled under normal conditions. The catastrophic transitions
mentioned above are then prevented because of the Pauli exclusion principle.
What we usually call the vacuum is actually an infinite sea of negative-energy
electrons. Occasionally one of the negative-energy electrons in the Dirac sea can
absorb a photon of energy ha ) 2mc' and become an E > 0 state. As a result,
a "hole" is created in the Dirac sea. The observable energy of the Dirac sea is
now the energy of the vacuum minus the negative energy of the vacated state,
hence a positive quantity. In this way we expect that the absence of a negative-
energy electron appears as the presence of a positive-energy particle. Similarly,
when a hole is created in the Dirac sea, the total charge of the Dirac sea becomes
Q: Quu.ru.r €: Quu.uu.r ?l,l): Quu"uo'', * lel; (3.330)
hence the observable charge of the hole is

Qurs: Q - Quu",rrr,,': lgl' (3.331)

This means that a hole in the sea of negative-energy states looks like a positive-
energy particle of charge le l. Thus once we accept (a) that the negative-energy
states are completely filled under normal conditions and (b) that a negative-energy
electron can absorb a photon of energy > 2mc2 (just as a positive-energy electron
can) to become a positive-energy electron, we are unambiguously led to predict
the existence of a particle of charge I e I with a positive energy.
When Dirac proposed this "hole theory," there was no good candidate for the
predicted positively charged particle. In the beginning Dirac even thought that
the hole in the negative-energy state should be identified with the proton. How-
ever, it was quickly pointed out by J. R. Oppenheimer that if this interpretation
were correct, the hydrogen atom would undergo a self-annihilation into two
photons with a lifetime - lQ-to sec.f Moreover, H. Weyl, who looked at the
symmetry properties of the Dirac equation, proved that the mass of the particle
associated with the hole must be the same as the electron mass. Prior to 1932,
because of the experimental absence of the conjectured particle, Dirac's hole
theory was not taken seriously. To recapture the prevailing atmosphere of the
time, we quote from W. Pauli's Handbucft article.$

fThis number assumes that the energy released is2m"c2.If we take the energy released
to be m.cz * m,,cz, the lifetime is even shorter.
gThe translation from the original German text is the work of J. Alexander, G. F. Chew,
W. Selove, and C. N. Yang.
3-9 HOLE THEORY AND CHARGE CONJUGATION 133

Recently Dirac attempted the explanation, already discussed by Oppenheimer, of


identifying the holes with antielectrons, particles of charge + I e I and the electron mass.
Likewise, in addition to protons, there must be antiprotons. The experimental absence
of such particles is then traced back to a special initial state in which only one of the
two kinds of particles is present. We see that this already appears to be unsatisfactory
because the laws of nature in this theory with respect to electrons and antielectrons
are exactly symmetrical. Thus .f-ray photons (at least two in order to satisfy the laws
of conservation of energy and momentum) must be able to transform, by themselves,
into an electron and an antielectron. We do not believe, therefore, that this explana-
tion can be seriously considered.
When the article appeared in print, however, C.D. Anderson had already de-
monstrated the existence of a positron. Many years later Pauli made the following
famous remark on Dirac:
. . . with his fine instinct for physical realities he started his argument without knowing
the end of it.

We shall now examine a little more closely the absorption of a photon by one
of the negative-energy electrons in the Dirac sea. As stated earlier, if the photon
energy is sufficiently large, an electron in a negative-energy state may be "esca-
lated" to a positive-energy state
ea.o*1+az>0. (3.332)
According to the hole-theoretic interpretation, this appears as

?y ------> OEro * eiro, (3.333)

since the vacated negative-energy state is observable as a positron state. Atthough


a photon cannot produce an e-e* pair in free space without violating energy and
momentum conservation, the process (3.333) can take place in the Coulomb field
of a nucleus. As is well known, the production of an electron-positron pair is
a very frequent phenomenon when high-energy ry-rays go through matter. We may
also consider a closely related process,
€E>o+e|.o*21. (3.334)
Since all the negative-energy states are supposed to be filled under normal con-
ditions, (3.334) is forbidden except when there is a hole in the normally filled
negative-energy states. This means that whenever (3.334) is allowed, we can inter-
pret it as
erro * eiro+ 2"y. (3.33s)
This process has also been observed frequently as positrons slow down in solids.
We shall present a quantitative treatment of this electron-positron annihilation
process in Chapter 4.
At this stage we emphasize again that the electron must obey the Pauli exclusion
principle if the hole theory is to make sense. Otherwise we cannot attach much
meaning to the notion that the negative-energy states are completely filled. If it
were not for the exclusion principle, we could keep on for millions of years piling
up electrons in the same negative-energy state. Even though the energy spectrum
t34 RELATTVTSTTC QUANTUM MECHANTCS OF SPrN-+ PARTTCLES 3-9

Table 3-4
DYNAMICAL QUANTITIES IN THE HOLE THEORY

Charge Energy Momentum Spin Helicity "Velocity"

,a<0 El p
fi(>>
>.0
Electron state -l 'l -l 2
Y

Positron state +l +l El _fr<>> >.t


"l -p 2
Y

of free Klein-Gordon particles is identical to that of free Dirac particles, it is not


possible to construct a sensible hole theory out of Klein-Gordon particles which
obey Bose-Einstein statistics.
Let us study the connection between the various dynamical quantities of the
positron and those of the negative-energy electron whose absence appears as the
presence of the positron in question. We have already seen that both the charge
and the energy of the physical positron must be positive. What is the momentum
of the positron? Just as in the case of energy, the absence of momentum p in the
Dirac sea appears as the presence of momentum -p. Hence the momentum of the
physical (E > 0) positron state is opposite to that of the corresponding negative-
energy electron state. Similarly the absence of a spin-up E < 0 electron is to be
interpreted as the presence of a spin-down E > 0 positron. Thus we can construct
Table 3-4 for a free particle. (We have listed (I) rather than the eigenvalue of
13 since, in general, the plane-wave solutions are not eigenstates of Ia).
The entry "velocity" in Table 3-4 requires some explanation. Suppose we con-
sider a wave packet made up of negative-energy solutions whose momenta center
around a certain mean value. We can then associate a certain group-velocity
with the wave packet. The absence of this E { 0 wave packet must.appear as
a wave packet made up of E > 0 positron states moving in the same direction,
that is, the velocity of the positron wave packet must be the same as that of the
corresponding E ( 0 electron wave packet. It is not hard to see that this is pos-
sible only if the "velocity" of the negative-energy electron is opposite in direction
to its momentum. This appears somewhat strange but is completely consistent
with (3.221), which says that the expectation value of the velocity operator ca
is the negative of the expectation value of pc'zllEl. The reader who is still not
convinced may amuse himself by working out steps analogous to (3.163) through
(3.170) for a negative-energy plane wave. If we apply S;j, to the wave function
9(3,4) for an E < 0 electron at rest in the primed system, we obtain the wave
function +,G'4) which corresponds to the negative-energy electron whose momen-
tum (defined as the eigenvalue of -ihV) in the unprimed system is opposite to
the direction of motion of the primed system.

Thomson scattering in the Dirac theory. As a simple calculation that dramatically


illustrates the importance of the negative-energy states in an unexpected domain,
we shall now compute the cross section for Thomson scattering, that is, the scat-
3-9 HOLE THEORY AND CHARGE CONJUGATION 135

tering of a low-energy photon (ho K mcz) by a free electron. As in Section 2-5


(cf. Eqs. 2.158 and 2.168) we expect that the differential cross section is given by
rf;ls<"t.e@')12, which is also the same as the classical result. In the Diractheory
there is no analog of the seagull graph Fig. 21(c); we must compute the analogs
of Fig. 2-2 (a) and (b) for the free electron. We characterize the initial,,final,'and
intermediate states of the electron by (p, r), (p', r'), and (p", r") respectively. We
then obtain for the transition matrix element,

-#h,>
'21/
aa' n/ -r- ,"':1,2 \ E" - E - ha

* ), (3.336)

where we have used the rule stated in (3.329). Since all the negative-energy states
are supposed to be filled, the summation is over positive-energy states only
(r" : 1,2).The electron is initially at rest. So, as p*0, k-0, a typical matrix
element in (3.336) becomes

(p" r" la.e ("r


l0r) : (llV)"@ ! e-io"'*nuc\+(p',)(a.e@))ul)(g)dr x
: 8p,,ottQ"lt1p/,)(Cr. €(d))U(r)(Q) : 0, (3.337)
since the matrix element of ar taken between two at-rest .E > 0 spinors vanishes.
Because the final electron is also at rest for the scattering of a very soft photon,
we easily see that (3.336) is identically zero. This means that the Thomson scat-
tering cross section should vanish, in contradiction to both observation and non-
relativistic quantum mechanics.
What went wrong? In the hole theory we must take into account an additional
process which has no analog in nonrelativistic quantum mechanics. Consider
a negative-energy electron in the Dirac sea. It can absorb the incident photon
(say, at t : tr) and become a positive-energy electron. Even though this virtual
transition does not conserve energy (unless ha ) 2mc2), there is a finite matrix
element for it. At a subsequent time (t : tr) the initial electron can fiIl up the
vacated negative-energy state by emitting the outgoing photon. Meanwhile the
escalated electron goes on as the positive-energy final-state electron. All this
may be visualized physically as follows. The incident photon creates an electron-
positron pair at t : tr; subsequently at t : t, the positron is annihilated by the
initial electron, emitting the outgoing photon, as shown in Fig. 3-6(a). Similarly
it is possible for the outgoing photon to be emitted first as one of the E < 0 elec-
trons in the Dirac sea is escalated; subsequently the initial electron fills up the
vacated negative-energy state by absorbing the incident photon. This is illustrated
in Fig. 3-6(b), which physically represents the creation of an electron-positron
pair plus the outgoing photon followed by the annihilation of the positron with
the initial electron and the incident photon.
We shall now calculate the matrix elements for the two diagrams. According to
the hole theory, for electrons, the initial state is made up of the incident electron
and the Dirac sea. In Fig. 3-6(a), one of the negative-energy electrons denoted
136 RELATTVTSTIC QUANTUM MECHANICS OF SPIN-+ PARTICLES 3-9

(k" o')

Absence of (p', r')


(p", r")

(k, o)

(a) (b)

Fig. 3_6. Thomson scattering in the Dirac theory.

by (p", r") makes a transition to a positive-energy state (p', r') by absorbing the
photon (k, a). The relevant matrix element is -ecnfrffr^(p'r'lot'6c) srk'*lp" r").
Since the absorption takes place first, the energy denominator according to the
rules of Chapter 2 is Er- Ea-ha, where Er: Eu^"-(-lE" D+f + E'
and Et : E"u.I E. For the transition at t2 we have the matrix element
Working out Fig. 3-6(b) in a similar way,
-ecJniTVa(p,,r,,la.e{"')g-tr.'*lpr).
for both diagrams combined:
we obtain for the transition matrix element

h sr sr I (p" r" la'ee't r-tk"x pt) (p't' a'e'"'4lp" r')


e2 c2 l l
-ffif,,":-r,nr Lll
E'+lE"l-h^
r- (p" ! lo.r't du*lpD(p r'l \. r? j3R)
l '"'"""'
As before' we set ,E: E' - fficz, P: P':0 as k--- 0' Furthermore' because
of the space integration that appears in the evaluation of each matrix element,
p,, :0; hence lE"l:mcz.lt is now simple to evaluate each of the four matrix
elements. For examPle,

(0 r" l a. et"r l 0 t) : (0, X(' "'-) (o.1,*, "';'"') (';')


6 .6(c) (3.33e)
- *(s") 1(s).

As hor, hio' K mcz, the two terms in (3.338) combine. Taking into account the two
spin states of the negative-energy electron at rest, we obtain

r" =3, L

:
s" =1,2

- XG')+[(a.et"r;1o.r{al) + (o.e@'))(o.€(4))]X(s) .

2e@).€(o')Err,, (3.340)
-
3-9 HOLE THEORY AND CHARGE CONJUGATION 137

where we have used (3.339) and the closure property of the pauli spinors

*Z,rx""'xG')+: (i ),t o) + (?). 1)

:(; ?)
(3.341)
Thus (3.338) is
e2c2h I " (3.342)
2v\ffi mc'"
where @ is the angle between e(o) and e(''). Apart from the minus sign in front
this is exactly the time-independent part of (2.158), which has been shown to be
responsible for Thomson scattering.l It is amusing to note that the seagull graph
(Fig. 2-2c) which is the sole contributor to Thomson scattering in the nonrelati-
vistic theory is replaced in the relativistic theory by the two diagrams of Fig. 3-6
in which the photons are emitted and absorbed one at a time.
The moral we can draw from this calculation is twofold. First, it illustrates
(perhaps more vividly than any other example can) that it is absolutely necessary
to take into account transitions involving negative-energy states if we are to obtain
the correct nonrelativistic results. It is truly remarkable that only by invoking the
concept of a negative-energy state (or a positron state), which is completely foreign
to nonrelativistic quantum mechanics, can we arrive at the correct Thomson
amplitude. Second, comparine (3.336) with (3.338) and noting that because of
energy conservation, the energy denominators in (3.338) can be written as
E' + lE" | - ha : -(8" - E* ha),
(3.343)
E' +lE"|* ha' : -(8" - E- ha),
we observe that the negative of (3.338) is formally identical with (3.336) as we
replace vt' - 3,4 with rtt - l, 2 despite the reversal in the time orderings of the
matrix elements.$ This kind of observation provides a natural justification of
R.P. Feynman's point of view according to which a negative-energy electron go-
ing " backward in time " is to be regarded as a positron going "forward in time."
We shall say more about this in the next chapter.
We might add that we can evaluate expressions (3.336) and (3.338) without
making the approximation hc'>Kmcz, E':77rg2. We then obtain the famous

Iff (3.336) is finite (as in the case of the scattering of a high-energy photon), the relative
sign of (3.336) and (3.338) is important. Actually the correct amplitude turns out to be
the difference of (3.336) and (3.338). This is because for Fig. 3-6 we must take into account
the minus sign arising from the fact that the initial E > 0 electron is "exchanged" with
one of the E ( 0 electrons in the Dirac sea. The reader need not worry about such subtle
sign changes; we shall show in the next chapter that the covariant prescription based
on the quantized Dirac theory automatically gives the correct signs for these matrix
elements. (See Problem 4-12.)
, $In writing (3.336) and (3.338) we have used the usual convention in which the matrix
t element standing to the right represents the perturbation acting earlier.
138 RELATrvrsrrc eUANTUM MEcHANICS oF sPIN-+ PARTIcLES 3-9

/\++
/\
/\
Fig. }-7. Pictoria I representation of vacuum polarization.

formula for Compton scattering

$il : +d(*)'(# * * - 2 + 4cos' @) (3.344)

derived by O. Klein and Y. Nishina in 1929.I

Virtual electron-positron pairs. As another example illustrating the importance


of electron-positron pairs we mention a phenomenon known as vacuum polariza-
tion. According to the hole theory, the completely filled sea of negative-energy
electrons has no observable effects; in particular, what we normally call the vacuum
is homogeneous and has no preferred direction. Let us, however, consider a nu-
cleus of charge Q : Zle I placed in the Dirac sea. There is now a departure from
complete homogeneity because the charge distribution of the negative-energy
electrons is different from that of the free-field case. In terms of the electron-
positron language this is due to the fact that a virtual electron-positron pair created
in the Coulomb field behaves in such a way that the electron tends to be attracted
to the nucleus while the positron tends to escape from the nucleus. This is pictori-
ally represented in Fig. 3-7. Note that the "induced" negative charges in the
vicinity of the nucleus are compensated for by positive charges that 'oescape to
infinity." As a result, the net charge observed at large but finite distances is smaller
than the bare charge of the nucleus. In fact what is usually called the observed
charge of the nucleus is the original bare charge of the nucleus partially canceled
by the charges of the virtual electrons surrounding the nucleus. This situation is
rather analogous to that of a charge placed in a dielectric material; the effective
charge in a polarized medium is the original charge divided by u, where e is the
dielectric constant. In other words, because of virtual electron-positron pairs the
vacuum behaves like a polarizable medium.

fFormula (3.344) can be obtained more readily using covariant perturbation theory
(cf. Section 4-4).
3-9 HOLE THEORY AND CHARGE CONJUGATION 139

On the other hand, at very close distances to the nucleus the "bare charge"
itself may be explored; the electron in a hydrogen-like atom should feel at very
short distances a stronger attraction than the attraction determined by the Cou-
lomb potential due to the usual observed charge. Since the s state electrons have
a greater probability of penetrating the nucleus, we expect that the energy levels
of the s states should be displaced to lower levels. Although the argument pre-
sented here is rather qualitative, the 2s*
- 2p! splitting of the hydrogen atom due
to this vacuum polarization effect turns out to be calculable. In 1935, E. A. Uehling
predicted that the 2sf-state should lie lower by 27 Mc than the 2p$-state.f The
experimentally observed Lamb shift, as we have seen in Section 2-8, has the oppo-
site sign and is about 40 times larger in magnitude. Although the major paii of
the Lamb shift is not due to this Uehling effect, for a precise comparison of the
experimental value with the theoretical value of the Lamb shift (measured to an
accuracy of 0.2 Mc) it has been proved essential to take vacuum polarization
seriously. The effect of vacuum polarization is also observable in r mesic and
muonic atoms.
The notion that the negative-energy states are completely filled becomes rather
treacherous when applied to a particle subject to an external potential. The results
of Section 3-7 show that even the (positive-energy) wave function of the hydrogen
atom when expanded in plane waves contains small negative-energy components.
At first we may be tempted to simply drop the negative-energy components by
saying that these states are completely filled. But this cannot be right because,
if we do so, we do not even obtain the correct energy levels; in fact, we may recall
that the Darwin term (needed for the 2slapg degeneracy) can be qualitatively
explained by invokingZitterbewegung, which arises from interference of the posi-
tive- and negative-energy plane-wave components of the bound-state wave func-
tion.
A crude physical argument for the Zitterbewegung of an atomic electron within
the framework of the hole theory goes as follows. We note that in the Coulomb
field of the nucleus a negative-energy electron can make a virtual transition to
a positive-energy state (which is equivalent to saying that the Coulomb field can
create a virtual electron-positron pair). Now the fact that the wave function in
the hydrogen atom contains negative-energy components implies that the atomic
electron in the orbit can fill up the hole in the negative-energy state (which means
that the atomic electron can annihilate with the positron of the virtual pair).
The escalated electron which is left over can now go around the nucleus as the
atomic electron. In short, the atomic electron and one of the E < 0 electrons in
the Dirac sea are visualized as undergoing "exchange scattering." What is the order
of magnitude of the distance over which this effect takes place? From the uncer-
tainty principle we expect that the energy violation by an amount 2mc2 involved
in the escalation of the negative energy electron is allowed only for a time interval
A't - hf 2mc2. (Note incidentally that this is of the order of the reciprocal of the

,
IUsing covariant perturbation theory, we shall briefly outline the calculation of the
Uehling effect in Chapter 5.
140 RELATrvrsrrc euANTUM MEcHANIcs oF spIN-+ rARTICLES 3-9

Zitterbewegung frequency.) At the time the original atomic electron fills up the
vacated negative-energy state, the escalated electron is at most c(Ar) - hl2mc
away from the original electron. This distance is precisely the order of magnitude
of the fluctuation of the electron coordinate due to the Zitterbewegung.

Charge.conjugate wave function. It is not entirely obvious from the form of the
Dirac equation that the space-time development of an electron state in a given
potential Aris identical to that of the corresponding positron state in the potential
-At. For this reason let us cast the Dirac theory into a form which makes the
symmetry between the electron and the positron self-evident. This can be best
done using a method originally exploited by H. A. Kramers, E. Majorana, and
W. Pauli.
We first ask whether the theory based on the Dirac equation with the sign of
eA, reversed,

(#,. #u,)ry,*' +Te" :0, (3.34s)

is equivalent to the one based on the original Dirac equation (3.60). We assume
as usual that there is a definite prescription that relates r/rd (called the charge-
conjugate wave function) and {r which are respectively solutions to (3.345) and
(3.60). Motivated by our experience with the Klein-Gordon theory (cf. Problem
1-3), we tryl
*t : S"9*, (3.346)

where Sa is a 4x 4 matrix. We must now show that

l&+ #Ar)rr * (*+ #n,)r,fs"f* +rys".1,* :0 (3.347)

is as good as the original Dirac equation (3.60). Taking the complex conjugate
of Q3al we have

l&- #dr)vi + (- ft+ |e,)vrfsi+ +rysfie:0. (3.348)

Multiplying (3.348) bV (Sd)-' from the left and comparing the result with the
original Dirac equation, we see that the equivalence of (3.346) and (3.60) can be
established if there exists ^56 such that
(S})-'ryf Sf :,tr,,
(3.34e)
(Sf)-'.yfSf -- -rf*
In the standard (Dirac-Pauli) representation,,y, and ?y4 are purely real while ry,
and ry, are purely imaginary. It is easy to see that in this representation,
Sc : cf z: Sf : (Si)-t (3.350)

fNote that \hc is to be represented by a single column matrix, since rf* (complex con-
jugate) rather than r/r+ (Hermitian conjugate) enters.
3-9 HoLE THEoRy AND cHARGE coNJUGATToN I4l

will do:{

',{:_l]},,: {li} (3.351)

since we have demonstrated the..t#: ;;J; that satisfies (3.34e), we have


proved the equivalence of (3.3a5) and (3.60).
It is very important to note that Sc : ,f z is true only in the standard (Dirac-
Pauli) representation. In fact the particular forms of S" depend on the particular
representations we happen to use. For instance, in the Majorana representation
in which ryn is purely imaginary and ryp is purely real, r/r" it simply f* itself, as
can readily be seen by complex-conjugating (3.60).$ This situation should be
contrasted with the parity case where S" : rya (up to a phase factor) holds in any
representation.
What does (3.346) mean for some of the familiar wave functions we have obtained
in the previous sections? Take, for instance, the first of the E ) 0 plane-wave
solutions (3.114). We have

J-4rnv, [z.r1n1*p (+r .'#)]


t o 0 0 -l\/ I
p4pf oor oll
:4--ffi[ \* o
oro;ll p,ct(E*mc,) l*n(-%.*'#)
\- t o o o/ \(p, * ip,)cl(E -Y mcz)l
l-(p, - ip,)cl(lEl + mc2)\
:Jryl o'ct(tEto mc')
l*n(-# *,+)
\-;)
: rffi-
- rqvu@(-P)exP(-# *'+)' (3'3s2)
^,1

Note that the eigenvalues of -ihV and ih(010t) are -p and -lE I respectively.
Similarly

^l# v,lu<'t1p1.*p %I - +)). : ffir,nt(')(-p) .*p ( - # *'+)


(3.353)

tMore generally we have Sc : 7ryr, where 7 is an undetermined phase factor: however,


this phase factor can be set to I by convention.
gln a more advanced treatment of the subject the relations (3.346) and (3.349) are often
written as +c : C$', C-r,f pC : -fTr, where Tstands for "transpose." In the standard
representation C:,Yz,la upto aphasefactorsincerfc :fzf +({r,y)r --"lz'l{t{*:ryzg*.
{
t42 RELATTVTSTTC QUANTUM MECHANTCS OF SprN+ PARTTCLES 3-9

Thus the charge-conjugate wave function r/nc obtained from the positive-energy
plane-wave solution t bV means of (3.3a6) is the wave function for a negative-
energy plane wave whose magnitude of the energy is the same and whose momen-
tum is opposite. Moreover, the spin direction (or if p is not along the z-axis, the
expectation value of X) is also reversed since the index 4(3) goes with the index
l(2). If we now invoke the hole theory, we see that the charge-conjugate wave
function describes the dynamical behavior of the negative-energy state whose
absence appears as the E > 0 positron of the same p and same (2) (cf. Table 3-4).
Likewise, when r/n represents a negative-energy electron state whose absence
appears as the positron state of p and (2), then rf^c represents the positive-energy
electron state of p and (X).
As another example, let us compare the probability distribution with the
^hr ^h
corresponding r/^d+'{rd where for the sake of definiteness, t may be taken to be
the wave function for the ground state of the hydrogen atom. In general, we have

^ht'^h'
: (yr"ln*)*("yr.h*) : +*t+* : "lrr+. (3.3s4)

For the electron in the hydrogen atom, the Ao(: -tA) that appears in (3.60)
islell4ttri *t, according to (3.3af, is a solution to the Coulomb energy problem
in the negative electrostatic potential -l ell4rr. Evidently the energy eigenvalue
of ,lrt is the negative of that of r/n because of the complex conjugation that appears
in (3.3a6). Thus the relation (3.35a) implies that the negative-energy electron going
around the negative electrostatic potential has the same probability distribution
as the corresponding positive-energy electron going around the positive electro-
static potential. This means that in an electrostatic potential that appears repulsive
to the positive-energy electron (for example, in the Coulomb field of the antipro-
ton), the negative-energy electron behaves dynamically as though it were in an
attractive force field. Invoking now the hole-theoretic interpretation, we see that
an antiatom in which a positron is bound to the center by Ao:
-l ell4rr looks
like the usual atom in which an electron is bound to the center by Ao : lell4ttr.
We define the charge-conjugation operation such that its application on the
electron (positron) state of momentum p and the spin-expectation value h<2>12
results in the positron (electron) state of momentum p and the spin-expectation
value h<2>12. The equivalence of (3.345) and (3.60) implies that if
^lr!, t) charac-
terizes the space-time behavior of an E > 0 electron state in a potential A, then
its charge-conjugate wave function 9"(x, t) characterizes the space-time behavior
ofthe negative-energy electron state whose obsence appears os the chorge-conjugote
(positron) state in the potential -Ap.
When we start computing the expectation values of the various dynamical
variables using t", we obtain results which may appear somewhat confusing at
first sight. For example, if we naively evaluate the expectation value of p with
respect to rfn and $t, we obtain the result: (p) is opposite to (p)", as can be seen
directly from (3.352) and (3.353) for the free-particle case, and similarly for (2).
But we know that the momentum and the spin direction are unchanged under
charge conjugation which transforms the electron state of momentum p and
3-10 QUANTTZATTON OF THE DrRAC FIELD 143

(I) into the positron state of momentum p and (X) (not -p and -(I)). This
peculiarity is due to the fact that in the unquantized Dirac theory the so-called
charge-conjugate wave function r/r" is not the wave function of the charge-con-
jugate state but rather that of the state (subject to the potential whose sign is
opposite to the original one) whose qbsence appears as the charge-conjugate
state.
What is even more striking, the space integral of the charge density

Q : e
I W^Vdsx
: e
t ^hn^hA'x,
(3.355)

which is the total charge, cannot possibly change its sign when we replace ,tfl by
its charge-conjugate wave function *o, io sharp contrast to the Klein-Gordon
case where the substitution { <* {* results in the reversal of the charge-current
density (cf. Eqs. 1.55 and 3.127).I Actually this is expected because when ,rfn it
a positive-energy wave function, rfn" ir the wave function for an E < 0 negotively
charged particle even though it "behaves dynamically" like a positively charged
particle in an external electromagnetic field. For a more satisfactory formulation
of charge conjugation, it is essential to quantize the electron field according to
Fermi-Dirac statistics.

3-r0. QUANTTZATION OF THE DIRAC FrELD


Difficulties of the unquantized Dirac theory. One of the great triumphs of rela-
tivistic quantum theory is that it has succeeded in providing a theoretical frame-
work within which we can discuss quantitatively a variety of physical phenomena
involving the creation and annihilation of various particles. We learned in the
last chapter that the "natural language" used to describe the creation and an-
nihilation of photons is that of quantum field theory. In the previous section
we did discuss phenomena such as pair creation and pair annihilation. The lan-
guage used there, however, is very different from that of quantum field theory;
instead of saying that the number of electrons is not conserved in pair production,
we have argued that the number of electrons actually is conserved and that all that
happens is just the escalation of a negative-energy electron. In other words, we
have tried to describe phenomena such as pair production without abandoning
the single-particle interpretation of the Dirac wave function according to which
the space integral of .ln*+ is a constant of the motion even in the presence of the
electromagnetic interaction. In doing so, however, we were forced to depart very
radically from the single-particle theory itself; in fact, we had to introduce a sea
of an infinite number of negative-energy particles.
There are essentially two reasons why the hole-theoretic description works.
First, as we have already mentioned, crucial to the success of the hole theory is
the assumption that the electron obeys the Pauli exclusion principle. Second,

fln fact rlnrrln cannot have its sign changed under any transformation that preserves its
positive-defi nite form.
144 RELATrvrsrrc eUANTUM MEcHANrcs oF sprN+ pARTrcLEs 3-10

although electrons and positrons can be created or annihilated, the basic interac-
tion in electrodynamics is such that the difference between the number of (positive-
energy) electrons and the number of (positive-energy) positrons,
N: i(e-) - N(e*), (3.356)
is conserved. In the hole-theoretic description what we do is just set
:
trf(e;'o) N(e-),
(3.357)
N(e".') : -N(e*) * constant background,
so that the newly defined electron number, given by the sum
N':N(e;ro)*N(eE.o), (3.358)
is necessarily conserved whenever (3.356) is conserved.
In the "real world" there are nonelectromagnetic phenomena which do not
conserve (3.356). Take, for instance, a beta (plus) decay
p'--+n*e+*v. (3.3se)
Although the free proton cannot undergo this disintegration process because
of energy conservation, a proton bound in a nucleus can emit a positron and
neutrino and turns itself into a neutron. In the hole-theoretic interpretation we
may try to attribute the presence of the e* in the final state to the absence of a
negative-energy electron in the Dirac sea. But where is the electron which used
to occupy the now vacated negative-energy state? It is apparent that the proba-
bility of finding the electron is no longer conserved.t

Second quantization. Our beta-decay example reveals that it is actually much


more sensible to construct a formalism in which we allow electrons and positrons
to be destroyed or created more freely. Guided by the success of the quantum
theory of radiation, we are tempted to follow, as much as possible, the quantiza-
tion procedure we used in the photon case. We shall first construct a "classicalo'
theory of the Dirac field using the standard Lagrangian formalism of Chapter I
and then quantize the dynamical excitations of the Dirac field by replacing the
Fourier coefficients by creation and annihilation operators. At this stage it is not
completely clear whether this method is a legitimate one. As we emphasized in
Section 2-3, the classical field theory is a limit of the quantum field theory where
the occupation number goes to infinity, but we know that the occupation number
of a particular electron state is at most one. However, let us go ahead with the
Lagrangian formulation of the classical Dirac field.$

fWe might argue that in this B+ process a negative-energy electron gives up its charge
to the proton and gets escalated to a (positive-energy) neutrino state. But note that the
electron and the neutrino are different particles which must be described by different
wave equations.
$The reader who is unhappy with our procedure may study an alternative, more axio-
matic, approach based on J. Schwinger's action principle, discussed, for instance, in
Chapter I of Jauch and Rohrlich (1955). In Schwinger's formalism the field variables
are treated as operators from the very beginning.
3-10 QUANTIZATTON OF THE DIRAC FIELD I45

The basic free-field Lagrangian density from which the field equation may be
derived is taken to be
g) : -ch{,yr(010x)$ - mc, V+
- -ch+"(,y)"8(0l0xr)*e - mcz 8"8fr"9p, (3.360)

which is a Lorentz invariant scalar density. In the Lagrangian formulation each


of the four components of + and r| is to be regarded as an independent field
variable. Varying fr" [*hi.h actually stands for (r/n+)"(.yr)r"] we obtain four Euler-
Lagrange equations of the form AglA{r":0 which can be summarized as the
single Dirac equation (3.31). To obtain the field equation for r|, we first make
the replacement

- cn g,(v ),u
hfiB
----+, o (hF). {v,), u*u, (3.361)

which is justified since the difference is just a four-divergence. Varying fu *t


then get the adjoint equation (3.46). The "canonical momentum" z conjugate
to {n isf
7tp: ag ih$,(,y)"o: ih^hh. (3.362)
A,-?MAD:
The Hamiltonian density is then obtainable by the standard prescription (1.4):

./f - citpaff-s
: ch(t+- #- i$,nffi+ $rrhr) * mc'V+
:,,h+(-ihca.Y + Bmcr)*. (3.363)

Thus the total Hamiltonian of the free Dirac field is

H : I g+(-ihca.v + Bmcz).,h dsx. (3.364)

Since the plane-wave solutions (3.114) and (3.115) takenat t:0 form a com-
plete orthonormal set, an arbitrary four-component field at t : 0 can be expanded
in free-particle plane waves. The Dirac field r/n becomes a quantized field if we
replace the Fourier coefficients in the plane-wave expansion by operators of the
type considered at the end of Section 2-2. We have

f(x, r) : h? i, (3.365)
^lffibpe)u<,>(D)ern',n,
where rf is now an operator assumed to act on state vectors in occupation number
space. We interprct bf) and D5')+ as respectively the annihilation and the creation
operators for state (p, r). A single electron state characterized by (p, r) is represented
by D[rtt (0) l0>. As we have already seen in Chapter 2, the Pauli exclusion prin-

fAs we have written the Lagrangian density, the canonical momentum conjugate to
r| vanishes.
146 RELATrvrsrrc eUANTUM MEcHANrcs oF sprN-+ pARTrcLEs 3-10

ciple is guaranteed if we use the Jordan-Wigner anticommutation relations (cf.


Eq.2.a9)
{b{"bg',)n} : 8,,,800.,
{b{>, bS't} :
g, (3.366)

{b{'l.r, b$')+} :
g,

from which it follows that the eigenvalue of the number operator defined by
Nf;, - b{>+ 6c> (3.367)
15 ZerO Or One.
We assume that the Hamiltonian operator of the quantized Dirac field has the
same form as the classical expression (3.363). We then havef

H : +J ; ; i, i, 6ffi,bg)+uo+(p)e-to.*rn)(-ihca.y I Bmc,)
" (,rffirbfi't
u<'t1n'1 eto"*n) dex

(3.368)

because of the orthogonality and normalization relations (3.106) and (3.110),


and the energy momentum relation

E, : +JWTT +W for ,,' : {l-'2.' (3.369)


l.3, +.
Recall that by our definition, the creation and annihilation operators Df;)+ and
b{) are time-dependentoperators. Their time dependence can be inferred from the
Heisenberg equation of motion,

b{,: *tr,bg)l: T + bg,l El for r - b.1r,2,q.


'.' ' (3.370)
bgtn - :
itu,Dg')+l ++b?lEl for r -
lr,2,
4, t3,
where we have taken advantage of the very useful relation

lAB, Cl : A{8, C} - lA, C}8. (3.37r)


Thus we have
b{)(t) : Df;)(Q) e+rtEtt/h for , : (3.372)
{!r:,'o',

|Previously we used the symbol H for the Hamiltonian operator (-ihca.V + Fmc')
acting on the Dirac wave function. In this section, f/ stands for the total Hamiltonian
operator of the free Dirac field acting on state vectors in occupation-number space.
3-10 QUANTTZATTON OF THE DrRAC FrELD 147

and a similar relation for Df;)+(l). The expansion (3.365) now becomes

./,(*, r) : +E ,m(,:;,,rf,,0)z(')(p)*p [+I -'+]


* ,4,n Df)(o) uo(p)."p [qI . '#]) e373)
The quantized free field is now seen to satisfy the same field equation (viz. the
Dirac equation) as the one derived from the classical variational principle. Note
that this is not a prtori self-evident; recall, in particular, that b and Di can no
longer be written as linear combinations of P and Q satisfyingIe, pI: ih. What is
even more striking, the form of the field equation obtainable from the Heisenberg
equation of motion is the same whether the creation and annihilation operators
satisfy anticommutation relations or commutation relations, as the reader may
readily verify.f Mathematically this remarkable feature is a consequence of the
fact that the commutator IAB,C) that appears in (3.371) can also be written as

lAB, C7 : AlB, Cl + lA, ClB. (3.374)


From the fact that we can get the same field equation whether the electron satisfies
Fermi-Dirac statistics or Bose-Einstein statistics, we may be tempted to infer that
quantum field theory does not "know" which statistics the electron is supposed
to satisfy. This inference, however, is not correct, as we shall see shortly.
At this point it turns out to be more convenient to redefine the b and D+ so that
they now become time-independent operators. We do this because we would like
to exhibit the time dependence of the field operator rf more explicitly. We set
Df;)tn"*) : rf.X.td)(g),
(3.375)
bf;l<orol(l) - !(r)(new) e+rlEft/h fOr , : [:, ?,
13,4.
The form of the anticommutation relations is unchanged under this replacement.
In the previous sections the symbol r/n stood for a single-particle wave function
while in the present section the same symbol rln is used to denote the quantized
Dirac field which is an operator that can act on state vectors in occupation-number
space. To distinguish between the two possibilities the wave function r/n is often
called a c-number field; the field operator *, a q-number field. We naturally ask:
what is the connection between the c-number {r and the q-number r/r? For a single-
particle plane-wave state the desired connection is easily established:

!h(c-nun,ber)
: (0 l r/n(a-"o"'1u., l Df)*Oo), (3.376)
where f(q-nunrber) is given by (3.373) with D5'r replacing Df)(O). To see this, just note
<olb{,) b{)* I o> : g",,gno,. (3.377)

{Note that (3.368) would still be valid even if the electron satisfied Bose-Einstein
statistics.
148 RELATTVTSTTC QUANTUM MECHANICS OF SPIN-+ PARTICLES 3-r0

Therefore, for r : lr2, we have


(0 1
gro-""mbe,) I DS)*<oo)
^ll@"(')(p)
: (; .'p t# -'+), (3.378)

which is indeed the wave function for a positive-energy plane wave characterized
by (p, r). The transition from f(c-number) to 'tfn(c-number) is sometimes called second
quantization f Although in the above example we considered just a single-particle
state, it is important to keep in mind that, in general, f(o-number) can actually
operate on the state vector for an assembly of electrons (and positrons). The Dirac
equation in the quantized theory should be regarded as a differential equation
that determines the dynamical behavior of the entire aggregate of electrons (and
positrons).
Let us now go back to the expression for the Hamiltonian operator (3.368),
which is unchanged under the replacement (3.375). This expression for the total
Hamiltonian makes good sense since we showed in Chapter 2 that ba)+ bo) (whose
eigenvalue is zero or one) is to be interpreted as the number operator. The energy
of a one-electron state characterized by (p,t) is just lEl or -lEl depending on
whether r : l, 2 or r : 3, 4. The total energy of an ensemble of E > 0 electrons
is just the sum of the energies of the individual electrons.
We can also compute the total charge operator. Following steps analogous to
(1.51) through (1.54), we can readily show that i{nyrrf satisfies the continuity
equation. Assuming that the charge density operator is given by ,*r* even in
the 4-number theory, we get for the total charge operator,

Q:tl *r* d'*


:eZ p'rr'

: e 2 Ab{'r og'' (3'379)

This is again expected.$ As for the total momentum of the Dirac field we may
start with (cf. Problem 1-l)
Pk -- -t J ,v ayd3x, (3.380)
c
where
_ _ as M_a& as
vq.4te (3.381)
O@\r"l7xn) 7xr Axr A(0il"10xn)
We then obtain
P : -ihJ +-v+ d'x: 4 l,r pb;'t+ 6<'t (3.382)

*By first quantizatioz one simply means p(classical)


-, -ihY, etc., for the dynamical
variables of a single Particle.
$Recall that the negative-energy electron has electric charge e - -leleven though it
"behaves dynamically" like a positively charged particle.
3-10 QUANTTZATTON OF THE DrRAC FrELD 149

Positron operators and positron spinors. Although (3.368), (3.379), and (3.382)
are satisfactory from the hole-theoretic point of view, the persistent appearance
of negative energies seems somewhat distasteful. It is much better to have a for-
malism in which the free-particle energy is always positive while the total charge
is positive or negative (depending on whether there are more positrons or electrons).
With this aim in mind let us define b$'), df;), a(')(p), and n(')(p) with r 1,2 such :
that
bf' : b{), (r : s) for r : 1,2,
for r :4,
dgrr - +byl {t: t forr:3; (3.383)
[s:2
and
a(')(p) - u?)(p), (r : s) for.r :1,2,
o(')(p) : szar(_p) Js:l forr:4, (3.384)
[s:2 forr:3.
The basic motivation for all this stems from the fact that the annihilation of a
negative-energy electron of momentum -p and spin-down appears as the creation
of a positron with momentum *p and spin-up.We later see that dr (d) can indeed
be interpreted as the creation (annihilation) operator of a positron. We have
reshuffied the order of the r- and s-indices and inserted minus signs in such a way
that
,S"ar(')*(p) : nlzrz(')*(p)
- u(')(p),
(3.385)
Sco{'l*1n; : ryro(s)*(p) : z(')(p),

with the same r (: 1, 2) (ct.3.352). Note that the d and dt satisfy the same anticom-
mutation relations as the b and the bt:
{df,>, /a't+} : 8,,., 8nn,,
(3.386)
{df>, ag't1: {d5.-)*, d$e'r+}
: g.
We also have
{b['), d[r] : {b5'), dyr}: {b[')*, df:t+1
: {rf;)*, df;r} :0. (3.387)

For later purposes it turns out to be useful to collect formulas for u and u. First,
(3.105) now becomes
(i,y.p ! mc)uF)(p):0, (3.388)
(-i"y. p + mc)u@(p) -- 0,
where p : (p,iElc) with E positive, even in the equation for o(')(p). The orthogo-
nality and normalization relations (3.106) and (3.1l0) become
u(s)r(p)a(s)(p) : 8,,,(Elmcz), 2ts')+(p)ots)(p) : 6,",(Elmcz),
(3.38e)
pts'r+(-n;z(')(p)
- u(")+(-p)o(')(p) : 0,
150 RELATTVTSTTC QUANTUM MECHANTCS OF SprN-+ PARTTCLES 3-10

where E is again understood to be positive. In terms of t(')(p) and t(')(p) we obtain


from the Hermitian conjugate of (3.388),
A(')(pXfry-p1'mc):9,
(3.3e0)
O(')(p)(-iV.p+mc):9,
where we have used (f i,y. p + mc)rnyn: ryt(tiny. p + mc). rt is also straight-
forward to prove with the aid of (3.389)
ua't(p)uF)(p) : 8,,,, ,ts')(p)pts)(p) : -0,,,,
(3.3e1)
;ts')(p)2tsr(p) : rts'r1p)4ts)(p) : 0,

for example, by multiplying the first expression of (3.383) from the left by z('')(p)ryn,
multiplying the first expression of (3.390) (with s replaced by s') from the right by
rynz("(p), and adding the two. The expansions for r/n and r| now take the form

.itn(*, r) : h?,I,,tr(ag, r,',1n) exp


t# -'+)
* t-F *'+l),
df;r+rt')(P) exp
(3.3e2)
fr(*, r) : ,rL?,I, .t$Qf)ou)(p).*p [qr - +]
* bgr+;r-)(p).*o [-fP'* LZI\.
"'vL h +, h)1,
where from now on it is understood that E shall always standfor the positive square
root JffiF +frF. Note that in obtaining (3.392) from (3.373) we used the
fact that the sum over p runs over all directions (-p as well as p).
Going back to the Hamiltonian operator and the total charge operator, we
can now rewrite (3.368) and (3.379) as follows:

H- ps
E(b$'r+6t'r
- dgldql+)

:: E(D[s)tbG) * dgttdf) - 1), (3.3e3)


ps
and
Q: e ? } (b$s)tb(s) + dllpdg})
: e2 > (05')*Df) _ dG)+d[.-) + 1). (3.3e4)

We recall that the anticommutation relations for the d and d+ are completely
identical in from with those of the b and b+. This means, among other things, that
the eigenvalue of d['r+4ts) is one or zero. From (3.393) and (3.394) we see that
if we interpret df>+ tat as the number operator for a positive-energy positron,
then we have the following satisfactory result: A state with an extra positron has
the expected extra positive energy and positive (-e:le) charge. Thus we take
N[u-,'l : b$s)f b(s), if[".'s) : d(s)f d(s) (3.3e5)
3-10 QUANTIZATION OF THE DIRAC FIELD 15I

to be the occupation-number operators for the electron and the positron state
characterized by (p,s). In hole-theory language, where r runs from I to 4, the ap-
plication of an E < 0 electron creation operator bf;'tt+ to the "physical vacuum"
must result in a null state since all the negative-energy states are already filled.
In our new notation this means that the application of d{'2) to the vacuum must
result in a null state. This is reasonable if dt'D is to be understood as the annihila-
tion operator for a positron with (p, s); in the vacuum there is no positron to be
annihilated. So for the vacuum state we require
b[') l0> : 0, d[') l0> : 0. (3.3e6)

From the anticommutation relation between N$".''1, d.f), and df'r, it follows
that the eigenvalue of N$"1') is zero for the vacuum state and one for a single
positron state d["+ 10) (cf. Eqs. 2.55 through 2.57). In other words, d$')n is the
creation operator for a positron.
The expressions (3.393) and (3.394) are still not completely satisfactory. It is
true that according to (3.393) the vacuum is the state with the lowest possible
energy; however, if we apply H to the vacuum state, we get -IpI,E which is - oo.
Physically this means that the infinite negative energy of the Dirac sea has not
yet been properly subtracted. We can redefine the energy scale so that 11 applied
to l0) gives a zero eigenvalue. We then have
H : 2 ftlfl",s) + N[e1s)) (3.3e7)
?
whose eigenvalue is necessarily positive semidefinite. Similarly subtracting the
infinite negative charge of the Dirac sea, we obtain
Q: e ? ? (N["-''r - N[".''r;
: -l"l? ? (N$"-'') - N["'''r;. (3.3e8)

This subtraction procedure amounts to starting with the charge densityf


p:e+++-"(fnf)'. (3.3ee)

Note that, unlike the total charge in the c-number theory (3.355) which is neces-
sarily negative, the eigenvalue of (3.398) can be negative or positive. Now at last
we can forget completely about negative energy electrons, the picturesque Dirac
sea, the negatively charged particles with E ( 0 that behave like positively charged
particles, the absence of p appearing as the presence of
-p, and all that. From now
on we can work with electrons and positrons of positive energies only.
We have seen that once we define both the energy and the charge of the vacuum
state to be zeio, then the total energy of the Dirac field is necessarily positive

lThe expression (3.399) can be shown to be equal to


(el2)(*+{ - *,r*'+r)
which is not zero in the q-number theory. This method of eliminating the undesirable
vacuum expectation value is due to W. Heisenberg.
152 RELATrvrsrrc eUANTUM MEcHANrcs oF sprN-+ pARTrcLEs 3-10

semidefinite, while the total charge can be negative or positive. We emphasize


that in obtaining this satisfactory result the anticommutation relations for the
creation and annihilation operators have played a crucial role. Had we used com-
mutation relations instead, we would have ended up with an expression for the
Hamiltonian operator whose eigenvalue has no lower bound [since b[tlt6ar *1,n
| :3,4 in (3.368) can take on an arbitrarily large positive numberl. Thus the
Dirac field must be quantized according to Fermi-Dirac statistics if we demand
that there be a state with the lowest energy. We have actually illustrated a special
case of a very general theorem which states that half-integral spin fields must be
quantized according to Fermi-Dirac statistics while integer spin fields must be
quantized according to Bose-Einstein statistics. This spin-statistics theorem,
first proved by W. Pauli in 1940, is one of the crowning achievements of relativistic
quantum theory.
Let us now look at the total momentum operator (3.382). We get
P : )p ) p(Df)*Df, + as;ag*;
s=1,2

p 8=1,2 p s=1,2

: F,4rp(N5"-''r
* N5"''"1 (3.400)

(since )pp:0). This justifies our earlier assertion that the physical momentum
of the positron state df)r l0) is p, not -p (cf. Eq. 3.38a).
In order to convince ourselves that D$')+ l0> and d[')r l0> with the same
(p, s) really are an electron and a positron state with the some spin direction, it is
instructive to work out the effect of applying the spin operator to these states.
Taking the spin density to be (hlz),lr+Z\h we obtain

sr3: (3.401)
@12) J+->sgd,x
for the z-component of the spin operator.f For an electron,
s'Ds"' o'
: !'i; l:;, Dr,+] c,3x I o)

rz(')(p)D['rt o). (3.402)


=
+ #z(')+(p)X, I

tWithin the framework of the Lagrangian formalism the ultimate justification for inter-
preting (hl2)rh+ I^/n as the spin density rests on the fact that the constancy of

!+.I-,ofx x v)s @12)2,1^hd,*


-r

is guaranteed by the invariance of the Lagrangian density under an infinitesimal rotation


around the z-axis [see, for example, Bjorken and Drell (1965), pp. 17-19,55]. Note that
no additive constant is needed because, with ,S, given bV (3.a01), S, l0) : 0 is auto-
matically satisfied by rotational invariance.
3-10 QUANTIZATION OF THE DIRAC FIELD 153

In contrast,
&agt+ lo> : [s,, df)l lo>
: -+ J {+*, df)'} x,./" ds xlo)

yg?,,(')+(P) x, o(')(P) df'* I o>' (3.403)

If the free-particle spinor corresponding to an electron state has an eigenvalue


Xs: *1, then (3.402) tells us that Sr:h12, as expected. On the other hand,
if we want to describe a positron with spin-up, (3.403) demands that we use a free-
particle spinor a for which the eigenvalue of 2, is minus one. As an example, take
the electron at rest with spin-up. The corresponding free particle spinor is r,r(t)(0)
with 2r - +1. For the positron state at rest with the same spin direction we are
supposed to use rttr(O): -v{a)(Q) which indeed has 2s - -1. This, of course,
is expected on the basis of the hole theory.
Under the charge conjugation operation defined in the previous section, a single
electron state b$')+ l0> goes into the corresponding positron state df'* l0) with
the same (p, s). Thus
bf> <- df>, D[srt r--+ 4rst+ (3.404)
under charge conjugation. We see that the total charge operator (3.398) indeed
changes its sign under charge conjugation in contrast to its c-number analog (3.355),
which always has the same sign. The Hamiltonian operator (3.397) does remain
invariant under charge conjugation, as it should.
In the previ'ous section we also examined the invariance property of the Dirac
equation under eA, --, -eAr. In the quantized theory the steps (3.345) through
(3.351) go through just as before if we replace rfn* by rln+r (where the transpose
operation brings the free-particle spinor back to the column matrix form without
affecting the creation and annihilation operators); ,{rd, given by

*'c : rr*+', (3.40s)


is now called the charge-conjugate field (rather than the charge-conjugate wave
function). Using (3.385), we find that "lrc in the free-field case is given by
: [-#s *'+) [T -+])
nl,'
#4 4 \re(rf'*rf' ..n + dPrz(')(p).^n

(3'406)
Comparing this with (3.39 2), wesee that the transformation
*' ,ln (3.407)
-
is achieved if we make the substitution (3.404) which is precisely the charge con-
jugation operationl r/nc annihilates positrons and creates electrons just as rf an-
nihilates electrons and creates positrons.
We shall now briefly mention the anticommutation relations among ,,lrn,
^lr,
and tp. First, it is evident that

[./r"(x), fu(x')] : {.ll(x),91(x')} : {^7"(x), frB(x')} : 0, (3.408)


154 RELATrvrsrrc eUANTUM MEcHANrcs oF sprN-+ pARTrcLEs 3-10

where x now stands for the four-vector (x, ict). For r/r(x) and rfr(x'), we have the
equal-time anticommutation relation first written by W. Heisenberg and W. Pauli:

{f"(", t),*b(x',t')}r=,: SdpD(3)(*


- *'), (3.40e)
the proof of which is left as an exercise (Problem 3-13). This also implies

[r/r"(*, t),fip(x',t')]r=r: (ryn),p8tsr(;


- xt). (3.410)
As for the anticommutation relation between r/' and $ ut different times, for our
purpose it suffices to remark that [r/r"(x), $u@')] is a function of the four-vector
x - x' such that it vanishes when x and x' are separated by a spaceJike distance:$
{^lr"(x),fru(x')} : 0
if (3.4r 1)
(x - x')' : (x - x')' - c,(t - t'), > 0.

Because of the anticommutation relation, it is clear that f(x) and r|(x') do not
commute when x
- x' is spacelike. This is not disturbing since rlr and r|, having
no classical analog, are not "measurable" in the same sense as E and B are measur-
able. On the other hand, for the charge-current density
j *(x) : iefi,y r* -,e(0 I .fr"yr.l I 0>, (3.412)
which ls "measurable," we obtain fiom (3.408) and (3.410),
Ur@),.1,(")l : o if (x - x')'> o, (3.4r3)
where we have used

lAB, CDI - -AC[D, B] + A[C,B]D - C{D, A}B + {C, A}DB. (3.414)


Thus measurements of charge-current densities performed at two different points
separated by a spacelike distance cannot influence each other, in agreement with
the causality principle.

Electromagnetic and Yukawa couplings. Let us now talk about the interaction of
electrons and positrons with the electromagnetic field. The Hamiltonian density
for the basic interaction is taken to be
ff6 - -it{^lr*A, (3.415)
where th is now the quantized electron field; Arcan be either classical or quan-
tized. This interaction can be derived from the Lagrangian density
gint : iefuYr"lrAu, (3.416)
since ff rn" is just the negative of g ;n whenevet I in does not contain time deriva-
tives of field operators. Strictly speaking, it{lr* should be replaced by (3.412),
but in practice the form (3.415) is sufficient since a constant (c-number) interaction
cannot cause a transition between different states.

fThe explicit form of [r/r,(x), FB(:)] known as -iS,B(x - x') may be found in more ad-
vanced textbooks, for example, Mandl (1959), pp.3G-35, pp.54-55; Schweber (1961),
pp. 180-182, pp. 225-227.
3-10 QUANTIZATION OF THE DIRAC FIELD 155

(a) ie9<-t7r{ft)A, (b) iegft)tu,lre)Au


"A"'v (c) iegft)t*{6'Au (d) te,1rt-t7u,l,e,Au
Fig. 3-8. Four possibilities realized by the interaction (3.415).

In lowest-order perturbation theory (and also in covariant perturbation based


on the interaction representation to be discussed in the next chapter), we may
regard r/r in (3.415) as the free field {n given by (3.392) and look at the effect of
.ffin, on a given initial state. It is convenient to decompose each of + and r| into
two parts as follows:
il :9(*) + f(-),
(3.417)
r|:r7;r*r*.frt-r,
where

f(*r: #+ 4 rreDf;rzr'r6; ."n (# -'+),


f(-r: W+ 4 ^lrydgsr+,u<'r1n; ..n (-#-: * +),
(3'418)
V,(+): #+ 4 (+r -'+),
^Eat(s)r(s)(p)."n

-#- *'#),
1
4ln(-)- ' SS
Y ,trbg'r+;tsr1n;
.*p (
"/Ti"
Note that 9<+r (which is called the positive frequency part of ./r) is linear in the
electron annihilation operators; it can therefore annihilate electrons but cannot
do anything else. Likewise .h(-), frt*', and rlr-r respectively create positrons,
annihilate positrons, and create electrons. So we see that the interaction (3.415)
can do four different types of things, as shown in Fig. 3-8 where the symbol x may
stand for an interaction with an external classical potential, the emission of a
photon, or the absorption of a photon. The question of which of the four pos-
sibilities given in Fig. 3-8 is actually realized in a particular physical process
depends entirely on what kind of initial and final state are present. For instance,
if initially there is only a positron, the action of (3.415) may result in the scattering
of the posi-tron, represented by Fig. 3-8(b). On the other hand, if initially there is
an electron-positron pair, and if we also know that there is neither an electron nor
a positron in the final state, the matrix element of ,ff intis that of a pair annihilation
process represented by Fig. 3-8(c). Note that in all cases the difference between the
number of electrons and the number of positrons (3.356) is conserved. In Chapter
4 we shall work out a number of problems in quantum electrodynamics based on
(3.41s).
156 RELATrvrsrrc eUANTUM MEcHANIcS oF spIN-+ rARTIcLES 3-11

The meson-nucleon interaction proposed by H. Yukawa in 1935 is patterned


after (3.415).t We have
ffrn:GW+, (3.41e)

where r/r and $ are now the nucleon and a neutral spin-zero meson field. Note
that r/r can annihilate nucleons and create antinucleons while r| can create nu-
cleons and annihilate antinucleons. The constant, G, characterizes the strength
of the coupling of the meson field to the nucleon. Alternatively we may have$
ffrn : iG{t"yu{r$. (3.420)

Both(3.419) and(3.420) are invariant under proper orthochronous Lorentz trans-


formations, as is evident from Table 3-1. The form of the interaction (3.a19)
can be made invariant under space inversion if f in the space-inverted system
is given by
6'@') - 6@), v,t - (-x, ict). (3.421)

On the other hand (3.420) can also be made invariant under space inversion if {
in the space-inverted system is given by
' 0'(x') : -d(x), y,t
- (-x, ict), (3.422)

since F"yr* changes its sign. If we require that parity be conserved, we may have
either (3.419) only or (3.420) only but not both simultaneously. In the next section
we shall point out that a Yukawa interaction of the type (3.419) leads to an s-wave
emission (absorption) of a meson by the nucleon while an interaction of the type
(3.420) leads lo a p-wave emission (absorption). The meson fields that transform
like (3.421) and (3.422) under space inversion are called respectively a scalar
and a pseudoscalar field; the couplings (3.419) and (3.420) are respectively called
a scalar coupling and a pseudoscalar (pseudoscalar) coupling.ll The meson fields
corresponding to the observed neutral spin-zero mesons, zr0 and ?, turn out to be
both pseudoscalar. We shall say more about the meson-nucleon interaction when
we discuss the one-pion exchange potential in the next chapter.

3-11. WEAK INTERACTIONS AND PARITY NONCONSERVATION;


THE TWO.COMPOI\ENT hTEUTRINO

Classification of interactions. To illustrate the power of the field-theoretic formalism


we have developed in the last section, we shall discuss in this section some simple
examples taken from the physics of so-called "weak interactions." As is well

fWe emphasize that the field-theoretic description of the meson-nucleon interaction


is not as firmly founded as that of the electromagnetic interaction of electrons and posi-
trons. Quantum field theory does, however, provide a convenient language to describe
some phenomena involving mesons and nucleons.
(fryu.ln)t :
$The iactor i is necessary to make the interaction density Hermitian. Note
^l+vsvr*: -^h./s.:1,'.
liso called to distinguish it from a pseudoscalar (pseudovector) coupling of the type
(iF hI m"c)(0 $ | 0x r) {r,y u'Y r*.
3-11 WEAK INTERACTIONS AND PARITY NONCONSERVATION I57

known, the basic interactions of elementary particles can be classified into three
categories:
a) strong interactions, for example, n*P--)n*p, r-*p----'A+Ko,
po-7t*+?t-;
b) electromagnetic interactions, for example, e* * e- '----,2ry, ./ + p ----+ p * ro,
Io * A *.y;
c) weak interactions, for example, n -+ p * e- * i, v'+p----, lL*+,r1',
K* ----- tt* * tt'.
To this list we should add a fourth (and the oldest) class of interactions, the grav-
itational interactions. But gravity turns out to be of little interest in our present-
day understanding of elementary particles. What is remarkable is that all four
classes of interactions are characterized by dimensionless coupling constants that
differ by many orders of magnitude; the dimensionless coupling constant that
characterizes electromagnetic interaction is (e'zl4rhc): d= r|". As we shall
show in Section 4-6, for the analogous coupling constant in pion-nucleon physics,
defined as in (3.420), we have (G'l4tthc) - 14, which is a typical strong-interaction
coupling constant. In contrast, the constant that characterizes the strength of a
typical weak-interaction process, when defined in a similar manner, turns out to
be as small as 10-12 to 10-1a; this will be illustrated in a moment when we discuss
A decay and pion decay. Gravitational interactions are even weaker; at the same
separation distance the gravitational attraction between two protons is about
10-37 times weaker than the electrostatic repulsion between them.
Another striking feature of the elementary particle interactions is that some
conservation laws which are obeyed to high degrees of accuracy by the strong
and electromagnetic interactions are known to be violated by the weak interac-
tions. In particular, weak-interaction processes in general do not conserve parity.
In this section we shall illustrate this point using the language of quantum field
theory.

Parity. One learns in nonrelativistic quantum mechanics that a state with mo-
mentum p goes into a state with momentum -p under parity, as is evident from
the operator form of momentum, p : -ihV. For instance, in the Schriidinger
theory the plane-waYe solution f(x, t) : eXP (ip'xlh - iEtlh) goes into f(-x, r)
: eXp t(-p).xlh - iEtlhl, which we recognize as the plane-wave solution with
momentum opposite to the original one. The orbital angular momentum L is
unchanged under parity since it is given by x x p. Furthermore, we argue
that the spin angular momentum is also even under parity since space inversion
commutes with an infinitesimal rotation. As a result the magnetic moment inter-
action-(ehl2mc)o.8, for instance, is invariant under parity since B does not
change under parity either. From these considerations we expect that in field
theory an electron state D[')+ l0> goes into bg]+ l0> (with the same s), etc.
We shall see, in a moment, that this is essentially the case except that we
have to be careful with the relative transformation properties of the electron and
positron states.
158 RELATrvrsrrc eUANTUM MEcHANrcs oF sprN-+ pARTrcLEs 3-11

In Section 34 we proved that the Dirac wave function in the space-inverted


system x' - -x, t' : / is related to the wave function in the original system by
*'(x',t):,yn*(x,/) (up to a phase factor). The proof given there goes through
just as well even if r/n stands for the quantized Dirac field. If we now consider the
space inversion operation (commonly called the parity operation) that transforms
x into -x, we see that the functional form of the field operator changes as
f(x, t'(x, t) : ,yn*G-x, t),
r) (3.423)
since what used to be called ;1/
-
: -x is now x. What does this mean? Going back
to the plane-wave expansion of the filed operator, we have

ryn^/,(-x, r): ? 4,E[bf,ry,ru,(p)exp(-# -r+)


* dg+,yna(')(p) exp (t# .'#)). e.424)
But it is easy to show (cf. Problem 3-4) that
rynrz(')(p) : rzts)1-n;, nf+'uF)(p) : -o(')(-p). (3.42s)
Note the minus sign preceding the positron spinor, u("(-p); a special case of this
has already been worked out in Section 3-4, where we learned that a positive-
energy at-rest spinor and a negative-energy at-rest spinor behave oppositely under
parity. Thus

y,f(-x, r): -+)


? 4^Elbf;rz<,r6y.,,p(+I
- d[')+,t,r(p)exp (# .'+)). e.426)

Comparing this with (3.392) we deduce that the transformation (3.423)is accom-
if the creation and annihilation operators change in the following manner:
plished

bf;)* D9i, dP --> -dYtr,


(3.427)
6f;)n * bf}t, dff; ---+ -dg}*.
An immediate consequence of this is that a single-particle state with (p, s) goes
into a state with (-p, s), as expected from nonrelativistic quantum mechanics,
except that a positron state acquires a minus sign,
,f)* l0> ----+ fI b$')+ :
l0> bgl+ l0),
df)* l0) *
II d5'rt l0> :
-d9f l0>,
(3.428)

where fI is the parity operator that acts on state vectors.f


We shall now discuss the physical significance of the minus sign in the second
part of (3.428). Consider, for simplicity, an electron-positron system in which the
electron and the positron are both at rest, hence are in a relative s-state. According

lusing the operator fI, we can write (3.427) as rlrfs) fl-r : 69[, etc. we assume that
the vacuum state is even by convention: tI l0) : l0>.
3-1 l WEAK INTERACTIONS AND PARITY NONCONSERVATION I59

to (3.428), such a system transforms as

b52[ d$?J lo) lo> (3.42e)


- of-b5't-,d['J
under parity. This means that the parity an s-state e-e+ system is odd.I Actually
we could have anticipated this from the hole theory. According to the hole-theoretic
interpretation, what is physically observable is the relative parity between the
completely filled Dirac sea and the Dirac sea with one negative-energy electron
missing. Remembering that parity is a multiplicative concept in the sense that the
parity of a composite system is the product of the parities of the constituent systems,
we infer that the observable parity of the positron is the same as that of the missing
negative-energy electron. But, according to our earlier discussion following (3.179),
the parity of a negative-energy electron at rest is odd (when the phase factor 7 is
so chosen that the parity of a positive-energy electron at rest is even). Hence the
parity of a positron at rest is odd (relative to that of an electron at rest).
If we follow steps similar to (3.423) through (3.429) with a non-Hermitian
(charged) field {"6(x, l) which transforms as

d"r,(x, t) *d"n(-x, l) (3.430)


under parity,,it is not difficult - that a tt- state and a r+
to show state transform
in the some way under parity and that a ft* ft- system in a relative s-state is even
(in sharp contrast to the e*e- case). Quite generally, the "intrinsic parity" of the
"antiparticle" is opposite to that of the corresponding "particle" in the fermion
case and is the same in the boson case. This is one of the most important results
of relativistic quantum theory. In Section 4-4, we shall present experimental
evidence in favor of the odd parity of an s-state e+e- system.

Hyperon decay. So much for the transformation properties of the free-particle


states. Let us now examine how we may describe parity-nonconserving decay
processes using the language of quantum field theory. As a particularly simple
example, we shall consider the decay of a free A hyperon (known to be a spin-{
particle):
[ p * tt-. (3.431)
A
->
density (operator) that can account for this process is
simple interaction
,ff ,n, : 6lfir( + g',yr)*n * Hc, (3.432)
where 9,r can annihilate A hyperons and create anti-A hyperons,$ frn can annihi-
late antiprotons and create protons, and $! can create r- and annihilate z*.
We have added Hc which stands for the "Hermitian conjugate" because, without

fThe skeptical reader may demonstrate that this conclusion (which can be verified ex-
perimentally) is independent of our choice T : I in $'(xt) : qT r*@).
gThe anti-A hyperon is to be distinguished from the A hyperon. First, even though they
have the same charge (namely, zero), the same mass, and the same lifetime, their magnetic
moments are opposite. Second, annihilations of a .Lp system into mesons, for example,
A + p-----K+ * r* * zr-,havebeenobserved,whereasreactions of the type A +p-*
mesons are strictly forbidden.
160 RELATTVTSTTC QUANTUM MECHANTCS OF SprN-+ PARTTCLES 3-11

it, the interaction density would not be Hermitian.{ Explicitly stated,


Hc : ^1rl(s* * g'*,f u)/'n{ofl*
:.|o(B* - g'*,y)+pfl,. (3.433)
We see that the Hc can give rise to processes like
A->F*tt*, (3.434)

where L and B stand for an anti-A hyperon and an antiproton. It is a characteristic


feature of the quantum-field-theoreticdescription that if (3.431) takes place,
then (3.434) must also take place. At the end of this section, we shall show that
this feature is a consequence of what is known as CPT invariance. Experimentally
the decay process (3.434) has indeed been observed to be as follows:
p+B-----+A+L
I l-Fttt* (3.43s)
l*n * tt-.
We do not actually believe that the Hamiltonian density (3.432) is a "funda-
mental" interaction in nature in the same way as we believe that -iefi"ryr*"A*
is "fundamental.n' So many new particles and new decay processes are observed
nowadays in high-energy nuclear physics that if we were to introduce a new funda-
mental interaction every time a new decay process was discovered, a complete
list of the fundamental interactions would become ridiculously long. Anybody
in his right mind would then say that most (or perhaps all) of the interactions in
the list could not possibly be "fundamental." Unfortunately, as yet we do not
know what the basic interaction mechanisms are which give rise to a phenomeno-
logical interaction of the kind described by Q.a3D. In any case, for computational
purposes,let us go along with (3.432).
We shall now investigate the transformation properties of (3.432). First, we can
easily see that (3.432) is invariant under a proper orthochronous Lorentz trans-
formation,

(3.436)

since the form of S1,o"(,S1,"',) is independent of the type of spin-f field. Next, we
shall show that the interaction Q.a3\ is not invariant under parity unless g : 0
or g' : 0. Since 9o, f,r, and f,o transform under parity as

to* Tr'lt*o(-x,t), fn* Tt"lrtn(-x, t), Qn Tniln(-x, t),


- (3.437)

tThe use of a Hermitian Hamiltonian density is required since the fexpectation value o
the Hamiltonian operator must be real. Furthermore, in Section f2 we shall show that
the use of a non-Hermitian Hamiltonian violates probability conservation.
3-11 WEAK INTERACTIONS AND PARITY NONCONSERVATION 16I

the interaction density (3.432) changes under the parity operation as follows:
.ff in (x, t) qt qI q^61(-x, t)fr'(-x, t)(s -g'.yu)tn(-x, t) f Hc. (3.438)

Suppose g'- : 0; then # n transforms like a scalar density,


ff t) *
,ff in1(-x, t),
rn"(x, (3.43e)

provided we choose the phase factors inQ.a37) in such a way that qtqlrJr: l.
Similarly for g : 0, ff io, again transforms like a scalar density provided we set
,ttrtlrt^: -1. However, if both g and g' are nonvanishing, then ffin, cannot be
made to transform like a scalar density no matter how we choose Tpt T,r, and r1n.
This is what we mean by saying that the interaction is not invariant under parity.
All this appears somewhat formal. We now wish to exhibit some of the physical
manifestations of the parity-nonconserving interaction (3.432). For this purpose
let us obtain the transition matrix element for this process. First, recall that the
state vectors we have been using are time independent, while the interaction
Hamiltonian density is made up of time-dependent field operators. In the language
of time-dependent perturbation theory, discussed in Section 24, a state vector
in our occupation-number space corresponds to a time-independent wave func-
tion rz,(x) rather than to a time-dependent one un(x)s-tE"t/4. Recalling the connec-
tion between the operators in the Heisenberg and the Schrddinger representations,
we see that the matrix element of ! ff i^rdTx taken between the initial A state and
the final a-p state corresponds, in the language of Chapter 2, to the matrix ele-
ment of er*ot/hHtt-tvot/n taken between rz6(x) and u{x); the only new point is
that our Hamiltonian can now change the nature of the particles. Using this facc
and recalling the expression for c(r) given by (2.109) we see that c")(t) in our
case is given by

,ot(t) : (-ilh) (tlJ. dt'


! ff,,,"(*,t') o'*l), (3.440)

where the perturbation is assumed to be turned on at r 0.* :


In the language of quantum field theory the initial and final states in A decay
are given by
lt> : D[ '')t l0), (3.441)
lf> : ar(P)b9'"rt 1o),
where b$n''r+, bp'*'n, and a+(pt,) are the creation operators for the A hyperon,
the proton, and the r- meson. It is legitimate to replace bV its positive frequency
^/no
part'rfnl*) since the negative frequency part of fo acting on the initial state would
result in a (A + state which, because of orthogonality, gives zero when mul-
tiplied bV (/l 6I+, ^) from the left. Furthermore, the only part of rfnf*r which
contributes is just (U,JT) D[^'')ufr1p; exp (rp.x/h - iEtlh),since bS;"'t64"'* I 0):0,
unless p" : p, J" : s. All this amounts to saying that it is all right to replace the

lWe shall present a more formal discussion of the connection between the transition
matrix and the interaction Hamiltonian in Section 42,whenwe shall discuss the S-matrix
expansion in the interaction representation.
162 RELATrvrsrrc euANruM MEcHANrcs oF sprN-+ pARTrcLEs 3-11

field operator rfrr by the waye function of the initial A particle multiplied by the
corresponding annihilation operator. Likewise it is easy to see that {} can be
replaced by just

ar(p)cnfriT^J exp (-ip,.xlh I ia*t)


(with he)n: ffi), and .|o can be replaced by
exp (-fp' .xlh + iEptlh).
"MD[r'vr+r7t"r1n')
The result of all this gives

(tlJ ff,n,0'.1+
:(f lor(n)bf,,'ut+6a'llr)@J@)Wm,,,m^crc)rf ')(p'Xg+s,ry5)uf )(0)

" l+rr[,"n ?'#- T#) o, *f."n (i,", +'+ - i*k",),


(3.442)
where we have assumed that the initial A particle is at rest. Using

D[td)D[td)n l0) : (1
- D[t8)tr$td)) l0>, etc.,
it is not difficult to see that (/l ar(p")b9,'")+D(^'3) li) can eventually be reduced to
(010>: l. The exponential time dependence in (3.442) is precisely the kind that
appears in the derivation of the Golden Rule in time-dependent perturbation theory
(cf. Eq. 2.113); if we assume that the perturbation is switched on at I : 0 and
acts for a long time, the modulus squared of JB e"p (ico,t + iEpt'lh
- imnczt'lh)dt'
leads to 2rht times the usual I function that expresses energy conservation.
Note also that the space integral in(3.442) simply tells us that the transition matrix
element is zero unless momentum is conserved.
To sum up, the time-dependent matrix element that appears in the Golden
Rule can be obtained immediately from the Hamiltonian operator l,tinrdtx
just by replacing the quantized field operators in,ffim by the appropriate initial
and final wave functions with their time dependence omitted. In other words,
we get the correct results by pretending that the q-number density Q. 3\ made
up of the field operators is a c-number density made up of the initial- and final-state
wave functions.f
Let us now simplify the spinor product in Q.aa\. We have
tf,")(p') (g + g"/u) uf)(o)

l)(1"')
(3.443)

fln fact this kind of replacement is implicit in the discussions of beta decay that appear
in most textbooks on nuclear physics, for example, Segrd (1964), Chapter 9, and Preston
(1964), Chapter 15. We now understand why beta decay can be discussed at an elementary
level without using the language of quantum field theory.
3-11 wEAK rNTERAcrroNs AND pARrry NoNcoNSERVATToN 163

Assuming that the initial A is polarized with spin along the positive z-axis, we
obtain

(t * s' ;ffi)r"' : (o, * opcos d) (l) . a, sin u ," (or), Q.444)


where
: g, ap: g'(lp' lcl(E, I moc'))
os (3.44s)
and the angles 0 and { characterize the orientation of p' relative to the A spin
direction. The physical meaning of g and g' can now be seen as follows. lf g + 0,
g':0, the final state proton can be described by an st, j,: f wave function
(6). On the other hand, if g:0, g'* 0,thefinal proton is in a pure p*, j,:t
state described by cos d(6) + sin d etd(!). In other words, the scalar coupling
(the g-term) gives rise to an's{ a-p system, whereas the pseudoscalar (ryu) coupling
(the g'-term) gives rise to a p+ 7r-p system. If g and g' are both nonvanishing, both
sf and pl are allowed; in other words, the same initial state can go into final states
of opposite parities. Recalling that for g # 0, g' + 0 the interaction density (3.432)
is not invariant under parity, we see that an ff rn that does not transform like
(3.439) under parity indeed gives rise to final states of opposite parities.
We shall digress here and examine the pseudoscalar (pseudoscalar) coupling
of the pion to the nucleon (3.420) which gives rise to processes like p -----+ p I ro.
We may argue that when the proton dissociates itself into a tt, and a proton,
the final zrp system must be in p$ state. Unfortunately, such dissociation processes
are forbidden by energy momentum conservation if all the particles are free.
However, taking advantage of a reaction in which the nucleon is bound (specifi-
cally ft* + d -- p * p) it has been proved possible to show that the tt! is pseu-
doscalar (with the convention that the proton and the neutron are both even),
that is, d(*, l) --- -0(-x, t). Note that if there is just a ryu type coupling, parity
is conserved; the "intrinsic" odd parity of the pion is compensated for by the
odd orbital parity.f
Coming back to A decay, we can now compute the decay-angular distribution.
The relative probabilities of observing the proton with spin-up and spin-down
can be obtained immediately from (3.44$:
spin-up: la,* apcosdl2,
(3.446)
spin-down: I aolz sinz 0,
which results in the decay-angular distribution of the proton
l-acosd, (3.447)
where$
d.- _ 2Re(a,cf) . (3.448)
la,l' + laol'

fAn alternative way of understanding what is meant by the intrinsic odd parity of the
pion is to visualize the pion as a very tightly bound state of a nucleon (intrinsically even)
and an antinucleon (intrinsically odd) in a relative s-state.
$The minus sign in (3.447) arises from the fact that the experimentalists usually talk
about the decay-angular distribution of the pion relative to theAspin: I * acos?(")
Where COS d(") : _COS d.
164 RELATTVTSTTC QUANTUM MECHANTCS OF SprN-+ PARTTCLES 3-11

Recalling that 0 is measured from the A spin direction, we se.e that the angular
distribution (3.447) implies that whenever Re(a,a$) # 0 or equivalently
*
Re(gg'*) 0, there exists an observable effect that depends on (on) .p' , & pseudo-
scalar quantity that changes its sign under parity.
To really understand the meaning of parity nonconservation in this decay
process, it is instructive to work out the special decay configuration d : 0. This
is shown in Fig. 3-9(a). The transition probability for this process is, according
to (3.446),1a, * aolz apaft from kinematical factors. If we apply the parity opera-
tion to the decay configuration shown in Fig. 3-9(a), we obtain the decay configura-
tion shown in Fig. 3-9(b) since, under parity, momentum changes but spin does
not. However, according to (3.446), the transition probability for the physical
situation described by Fig. 3-9(b) is I a, - oplz since 0 : r. Thus the transition
probability for lr)*l,f) is not the same as that for fllr)*nlf> unless
Re(a,afi) : 0. Since the configuration in Fig. 3-9(b) is the mirror image of the
configuration in Fig. 3-9(a) apart from a 180' rotation about an axis perpendic-
ular to p', wo conclude that the mirror image of our world looks different from
our world if Re(gg'*) + O.

tpro,on
,#

Probability Probability
w^ lar*apl2 w^ lar- apl2

@
fProton
(a) (b)

Fig. 3-9. A decay. Parity conservation would require that the two decay configurations
(which go into each other under space inversion) be physically realizable with the same
transition probability. The gray arrows indicate the spin direction.

Although we cannot prepare a polarized sample of A hyperons using a magnet,


it turns out that the A hyperons produced in
rt-*p--->A+Ko (3.44e)
are strongly polarized in the direction pa incident X pn. Parity nonconservation
in A decay was unambiguously established, in 1957, by a Berkeley group and by
a Columbia-Michigan-Pisa-Bologna collaboration group who showed that there
ate more decay pions emitted with (pz incioent x p").p, o"".v ) 0 than with
(pzinciae,,t x pn).pad.ecay ( 0. Now, at last, it is possible to communicate even to
intelligent beings in outer space that the incident pion direction, the A direction,
3-11 WEAK INTERACTIONS AND PARITY NONCONSERVATION 165

and the preferential direction of the decay pion, taken in that order, form the
three axes of what we mean by a right-handed system.
Prior to 1956 practically everybody tacitly assumed that it was "illegal" to
write parity-nonconserving interactions like (3.432). There was a good reason for
this; the success of the parity selection rules in atomic and nuclear physics shows
that the principle of parity conservation holds to a high degree of accuracy in both
electromagnetic and strong interactions. In the years from 1954 to 1956, as various
experimental groups studied the properties of "strange" mesons called r* and 0*
which decay via weak interactions as
T+ ___) 2t+ | tt-, 0* _____>
E* * tto, (3.4s0)
it soon became evident that r+ and 0* have the same mass and the same lifetime;
it therefore appeared natural to assume that a r-like decay event and a d-like
decay event simply represent different decay modes of the same parent particle
(now called K*). However, using an ingenious argument based only on parity
and angular momentum conservation, R. H. Dalitz was able to show that the
experimental energy and angular distributions of the pions from r-decay strongly
suggest that the r and the 0 could not possibly have the same spin parity. Since,
at that time, people believed in parity conservation, this led to the famous r-6
puzzle.t Faced with this dilemma, in the spring of 1956 T. D. Lee and C. N. Yang
systematically investigated the validity of parity conservation in elementary particle
interactions. Their conclusion was that in the realm of weak interactions parity
conservation (which holds extremely well for the strong and electromagnetic
interactions) was "only an extrapolated hypothesis unsupported by experimental
evidence." Furthermore, they suggested a number of experiments that are really
sensitive to the question of whether or not parity is conserved. (Their list of sug-
gested experiments included the decay angular distribution of a polarized A
hyperon which we have been discussing.) As is well known, subsequent experi-
ments (beginning with the historic Co6o experiment of C. S. Wu and coworkers
and the ft-p-e experiments of J. I. Friedman and V. L. Telegdi and of R. L. Garwin,
L. M. Lederman, and M. Weinrich) have unequivocally supported the idea that
weak interactions in general do not conserve parity.
Coming back to A decay, let us work out the decay rate using the Golden Rule
and (3.442) through (3.446). The cos d term drops out as we integrate over all
angles. For the reciprocal of the partial lifetime, we get
l:l](A*pd)
Ts-'pr- h

:Ut!-Nmr92 *=E,(trp -p lg'l'lpl'c'\ arv lp'l'dlp'l


h 2con EoV Zmrc' \ro I ' E, mrct / (2tth)3 d(E" Eo)
* *
(3.4s1)

fFor detailed discussions of the r-0 puzzle see, for example, Nishijima (1964), pp.
315-323; Sakurai (1964), pp. 47 -51.
166 RELATrvrsrrc eUANTUM MEcHANrcs or srrN-| pARTrcLEs 3-11

where we have used


dlp'l _ dlp'l
d(Ep + E") - [( p' I c'l E) + (l p' lc'l E")] dlp'l
: haoE'
(3.4s2)
mnlp'lcn
If we insert the experimentally measured mean life of the A particle (2.6 x l0-ro
sec) and the branching ratio into the r- p decay mode (known to be about g),
we obtain

#-+ o.oo3
H=2 x lo-r4. (3.453)

Thus the dimensionless coupling constants


I
g l'z l@?thc) and I
g' l'zl@ithc)
for A decay are seen to be small compared to e'zl(4rhc) = T+T by many orders of
magnitude. The interaction responsible for A decay is indeed "weak."

Fermi theory of beta decay. Historically the theory of weak interactions started
when E. Fermi wrote, in 1932, a Hamiltonian density that involves the proton,
neutron, electron, and neutrino fields to account for nuclear beta decay:
n->P*e-*2. (3.454)
Fermi assumed for simplicity that the derivatives of the field operators do not
appear. With this hypothesis the most general interaction density invariant under
proper orthochronous Lorentz transformations has the form
ff ,n,: } (frol'r/.")[fr"I'(C, + Ci,'/)+,] * Hc, (3.4s5)
wheref
Ii : l, ,flr, ('xo, inf snf),t f s. (3.456)
We have subscribed to the usual convention according to which the light neutral
particle emitted together with the e- in (3.454) is an "antineutrinoo' (z), not a
"neutrino" (r), and the field operator nJt, annihilates neutrinos and creates antineu-
trinos. Evidently the explicitly written part of (3.a55) can account for the neutrino
induced reaction
v+n->e-+p (3.4s7)
as well as for B- decay (3.454). The Hc in (3.a55) can describe B* decay, K (electron)
capture, etc.:
p->n1'e+ lv,
(3'458)
e- * p --> rt * v,
since, when explicitly written, it contains the annihilation operators for protons
and electrons and the creation operators for neutrons, positrons, and neutrinos.

{We avoid the indices pc and z to prevent possible confusions with muon and neutrino.
3-1 I WEAK INTERACTIONS AND PARITY NONCONSERVATION t67

The constants C, and C't characterize the strength of the interactions of type i
(scalar, vector, tensor, axial vector, and pseudoscalar); they have the dimension
of energy times volume. From our earlier discussion on A decay, it is evident that
parity conservation requires either
C', : 0 for all i (rtt rl^T! T,: l),
or (3.45e)
C, : 0 for all i (rtt T"Tt q, : -l).
If neither of the two possibilities is satisfied, then the interaction density (3.455)
is not invariant under parity.
As it stands, (3.455) contains l0 arbitrary constants (which need not be purely
real). About a quarter-century after the appearance of Fermi's paper, it finally
became evident that the correct Hamiltonian density that phengrmenologically
described nuclear beta decay was

ff rn : Cr(fio,f^lh")[fr"ry^(l * ryrhhJ + CA("T]piryuryrf")lfi"i,y*y^(l + ryu]g,l * Hc


(3'460)
with
Cv : 6.2 x 10-44 MeV cm3 - (10-51\n) mucz(hf mrc)s,
(3.461)
colc" - -1.2.
The interaction (3.460) with C" -
-Cn is known as the V - A interaction; it was
written, on aesthetic grounds, by E. C. G. Sudarshan and R. E. Marshak, by R. p.
Feynman and M. Gell-Mann, and by J. J. Sakurai in advance of the confirming
experiments. Since the nucleon can be assumed to be nonrelativistic, only the time
component of the vector covariant and the space components of the axial vector
covariant contribute (ry1 and iyuryn are "small") unless the symmetry of the initial
and final nuclear states is such that the expectation values of I (the nonrelativistic
limit of frrrynr/l) and op (the nonrelativistic limit of tlrniryu,y7,fi) are both zero.
The vector interaction gives the Fermi selection rule A.,/: 0, no parity change,
while the axial-vector interaction gives the Gamow-Teller selection rule A,I: 0,
f l, no parity change, for the nuclear states.
We shall not discuss in detail the various aspects of nuclear beta decay: the
electron spectrum, the f-values, forbidden transitions, the electron-neutrino
angular correlation, the angular distributions of electrons from polarized nuclei,
etc. They are treated in standard textbooks on nuclear physics.t We concentrate
on just one aspect of (3.460), namely the physical meaning of (1 * ry)*,.

Two-component neutrino. The neutrino field n|n, can be expanded just as in (3.392).
The only difference is that its mass is consistent with zero, ftt, 1200eYlc2 ex-
perimentally. The positive frequency part of 9, is linear in the free-particle spinor
for an annihilated neutrino. So let us investigate the effect of (1 ryu) on a(')(p). *
fSee, for example, Preston (1962), Chapter 15; Kiill6n (1964), Chapter 13.
168 RELATrvrsrrc eUANTUM MEcHANIcs oF spIN-+ pARTIcLES 3-11

As m, ---' 0, we haYe

(3.462)

Meanwhile

->'o (r.r!,rrr,,,) : - f o,lu)[.#";,,,)


: -f;3"') (3.463)

This means that as mv ' 0, the ry5 operator (sometimes called the chirality oper
atorf) and the negative of the helicity operator have the same effect on a free-
particle spinor u(')(p). In particular, eigenspinors of 2.fl are also eigenspinors of
ryu with opposite eigenvalues, and (1 + ry5) acting on the free-particle spinor for
a right-handed particle gives zero as ffi,- 0. It then follows that (1 +.yu)9[.)
annihilates only left-handed (helicity : -1) neutrinos, denoted by zr. Likewise
it is easy to see that
(1 + ryu)ti-) creates trr,
.trt.)(l- yu) annihilates in,
^It-)(l -
yu) creates uL,
where 7" stands for a right-handed antineutrino. Now the Hc in (3.a60) contains

"7,(t - ryu) since


t-_
L*{,;;^}(1
+ vJt,]-: rr(t *'v,){-,T^r,} nta*.
: fr,(l - Yr) (3.464)
{*,T;;^}*",
where the upper signs are for the space components, the lower for the time com-
ponents. An immediate consequence of this is that the helicity of the neutrino
emitted in K (electron) capture [the second of (3.458)] is -1. This has indeed
been shown to be the case by a beautiful experiment of M. Goldhaber, L. Grodzins,
and A. W. Sunyar. The positive (negative) helicity of the antineutrino (neutrino)
emitted in F- (B*) decay has also been inferred from the electron (positron)
polarization and the e-l(e*z) angular correlation (cf. Problem 3-14). Parity con-
servation would require that in a physically realizable process the emission prob-
ability for a right-handed particle be the same as that for a left-handed particle
since the helicity changes sign under parity. It is therefore evident that the interac-
tion (3.460) which produces only a"(vo) in p. (P-) decay is incompatible with the
principle of parity conservation.

$This expression is derived from the Greek word XeLpmeaning "hand." The term "chira-
lity"was first used by Lord Kelvin in a somewhat different context.
3-t l WEAK INTERACTIONS AND PARITY NONCONSERVATION 169

There is another (simpler) way to see the connection between helicity and chira-
lity for a massless patticle. Let us go back to (3.26) which was obtained by lineariz-
ing the Waerden equation. When the fermion mass is zero, the two equations are
completely decoupled,f

(i".v -,h)4@) :0, (-,".v - ,*)d(o) : 0. (3.465)

Let us now postulate that only the first of (3.a65) has to do with physical reality.
Evidently a free-particle solution of the S(') equation in the c-number (wave-
function) language satisfies
o'D : - Elc, (3.466)

where E can be positive or negative. This means that the helicity of a positive-
energy neutrino is negative, while the helicity of a negative-energy neutrino is
positive. Using the hole theory, we infer that the helicity of a (positive-energy)
antineutrino is also positive (cf. Table 34). We should emphasize that the asser-
tion that the neutrino (antineutrino) is always left-handed (right-handed) makes
sense only if the mass is strictly zero. Otherwise we can easily perform a Lorentz
transformation that changes a left-handed particle into a right-handed one.
Meanwhile, according to Problem 3-5, in the Weyl representation of the Dirac
equation we have

+,:(la, (3.467)

and

r,: (-; ) (3.468)

It is then evident that (l + .y5)+, selects just f('), that is, only the lower two of
the four components of r/l in the Weyl representation:

(1 + ,yu)*:t(*or,) (3.46e)

tlt is amusing that the Maxwell equations applying to empty space can also be written in
a form similar to (3.465). The reader may show that (1.57) and (1.58) with p : 0, j : 0
can be combined to give

(-''.v-,*)o:0,
where

':'(l r -:)' .:'Li :il .:'(: l:)


and
.:(;i -"i),)
170 RELATrvrsrrc eUANTUM MEcHANIcs oF spIN-+ pARTIcLES 3-11

Suppose we demand that only


(l + "y)*, and fr,(t - ryu) : (l + "y)*,\+ryn
appear in the beta-decay interaction. Because of (3.469) this requirement amounts
to saying that only O(') and 0(')* appear, hence the neutrinos (antineutrinos)
emitted in F* (B-) decay are necessarily left- (right-) handed. We may argue that
the appearance of only d(') is due to either (a) an intrinsic property of the free
neutrino itself, or (b) a property of the parity-nonconserving beta-decay interoction
that just happens to select 6tzr (and also of other interactions in which the neutrino
participates). If the neutrino mass is strictly zero) the two points of view are,
in practice, completely indistinguishable.* A theory of neutrinos based on
4G) + 0, drnr : 0 (or vice versa) is called the two-component theory of the
neutrino.
Historically, the idea that the neutrino is described by {tzr only (or {(") only)
was advanced by H. Weyl in 1929; it was rejected by W. Pauli in his Handbuch
article on the grounds that the wave equation for 6tzr only (called the Weyl equa-
tion) is not manifestly covariant under space inversion (cf. Problem 3-5). With the
advent of parity nonconservation in 1957 the Weyl equation was revived by A.
Salam, by L. D. Landau, and by T. D. Lee and C. N. Yang.
Muon capture,
p- * P--->n*v', (3.470)

can also be described by an interaction of the form (3.455)i r/n" and .fn, are now
replaced by ** and particle v' whose mass is also consistent with zero
^lr,,.The
turns out again to be left-handed (its antiparticle l' is right-handed). For some
years it was generally believed that v and v' were identical; however, motivated
.y (which can be best understood if
by the experimental absence of p*
- e* +
v * v'), B. Pontecorvo and others proposed an experiment to test the assumption
that v is the same as v'. Noting that a' also appears in
7t*
-) l"* * r' (3.471)

(because r*
can virtually disintegrate into a proton and an antineutron), we can
settle the question of the identity or the nonidentity of z and ,' by examining
whether or not the neutral particle from pion decay Q.a7D can induce a high-
energy neutrino reaction, that is,
u'*n---+e-*p. (3.472)

ln 1962 M. Schwartz and collaborators established experimentally that (3.472)


is forbidden, while
u' t
n-"-> p- + p (3.473)

is fully allowed for y' from pion decay; these experimental facts are in agreement
with the idea that u and a' are different.

fln general, you may introduce as many fields as you like without changing the physical
content, so long as the fields introduced are coupled to nothing.
3-11 WEAK INTERACTIONS AND PARITY NONCONSERVATION t7l

Pion decay and the CPT theorem. As a final example, let us make a comparison
between
7t ------>p tu'-, (3.474)
and
(3.47s)
It is appealing to assume that the interactions responsible for (3.474) and (3.475)
are the same in form as well as in strength, that is, the interactions are invariant
under
,x <----> 1l,t, v € ut , D <---> Ft . (3.476)
In addition it is reasonable to postulate that in (3.475) the field operators r/r"
and rfn, enter in the same combination as in nuclear beta decay, viz. in the com-
bination ifi.,yuryr(l + ry5)+, [which is also equal to -ifr".y^(t + ryr)+J. With
these two hypotheses, the interaction densities may be taken as

ff ,n : ifhurftW"r,ryr(l * ry,)fJ * Hc, (3.477)

for (3.475) and


ff,n": if hft^W',r,ryi(l * "v,)*,'l* Hc, (3.478)

for (3.474). The constant/is assumed to be universal, that is, the same for both
(3.474) and(3.475); wehaveinsertedthefactor hfm,c justtomake (f'zlLrhc)dimen-
sionless. Using the same argument as in A decay, we obtain the following for the
z-e process:
c., _ _+
#_(, ^/*)Ql-ffiQ$)tiil;y'.y{t +,y,)a.t

" h I',0, I
o' *'
l&"*plio#] )(."n l-'t! - ry])],
(3.47e)
where we have suppressed the momentum and spin indices for the free-particle
spinors. The four-gradient acting on the pion plane wave just brings down ipp lh;
because of energy-momentum conservation we get

ipf)u"ryu,y^(l * nyu)a,: _ipf)il",yx(I * ryu)u,

- -iu"(t.p(") + p('))(l
,f .
I y)a,
: n4"cilu(ll ryu)a,, (3.480)

where we have used (3.388) and (3.390) with m,: 0. Let us now work in the rest
system of the decaying pion with the z-axis along the antineutrino momentum
denoted by p. Using the explicit forms of free-particle spinors (cf. Eqs. 3.115 and
3.384), we can easily verify that for p along the quantization axis,
(1 + yu)ol')(p) : 2al,')(p), (1 + y)atD(p) : 0 (3.481)
as mu ' 0, which supports our earlier (l + ryu).ih5-) creates right-
assertion that
handed antineutrinos only. The amplitude for the production of a right-handed
172 RELATrvrsrIc euANTUM MEcHANIcs oF sPIN-+ PARTIcLES 3-1r

(spin-down) electron with momentur p" : -p is proportional to

lW ^l-Ta[,r( - p) ( 1 + ./u) ut')(p)

E"+m"cz { lplc l-m*c' (3.482)


@ : I
1 zE.',
-::-.

as mv ' 0, where we have used


E'-(m2'-m?")c
lpl: v-T'
E"* lplc: ffincz. (3.483)

Note that the use of the normalization convention (3.389) does not cause any
For the other (left-handed) spin state of the electron
trouble in the ffi,
- 0 limit.
we can readily show that
;[')(-pXl + ryi)?,t')(p) : 0, (3.484)

which we could/ actually have guessed from angular momentum conservation.


(The pion is rp/,l..r and the orbital angular momentum cannot have a nonvanish-
ing comp oney(. along p.) The only other thing we must evaluate is the phase-space
factor ./
T(8,+ E.) - ffioc3 4m3,c

Collecting all the factors and using the Golden Rule, we finally get

(3.486)

Similarly
(3.487)

What is more significant, the ratio of (3.a75) to (3.474) is


(m.\'(m'n - m!)'z - l.3x 1o-4.
W - (3.4sg)
W-\*r) 04--m?y--
Note that the muonic decay mode is much more frequent despite the smaller
Q-value available; it is hard to beat the factor (m"lm)' - 2.3 x 10-5.
Actually the fact that the transition probability computed with (3.477) is pro-
portional to mZ is not surprising. To see this we first note that, if m.:0, then

fr"ryury^(l +./5)9, : fr"(l - rs)'rsnrt*,


3-11 WEAK INTERACTIONS AND PARITY NONCONSERVATION I73

would create only left-handed electrons. But the simultaneous creation of a left-
handed electron and a right-handed z in pion decay is strictly forbidden by angular
momentum conservation. It is therefore expected that the amplitude for (3.a73) is
proportional to m*
Although the result (3.488) was first obtained by M. Ruderman and R. Fink-
elsteinf as early as 1949, only two years after the discovery of the pion, for a
very long time the experimentalists failed to find the electronic decay mode; the
quoted upper limit was an order of magnitude lower than (3.488). When it became
evident that the electron and neutrino enter in the combination r|"ry1(l + "/5)+,
in beta decay and also in muon decay, the idea of describing pion decay by Q.a77)
and (3.478) became so attractive that some theoreticians even said, "The experi
ments must be wrong." Truly enough, a number of experiments performed since
1958 at CERN and other places brilliantly confirmed the predicted ratio (3.488).
We mention this example just to emphasize that the power of relativistic quantum
mechanics in making quantitative predictions is not limited to the domain of
electromagnetic interactions.
As a by-product of our calculation we obtain
(f'zl4tthc)- 1.8 x 10-'u, (3.48e)

when the obseryed/pion lifetime is inserted in (3.487). So this is again an example


of "weak'tr'teractions. We may also mention that a similar calculation with Hc
in (3.a78) gives the following result:
l(rt* lr* t a'"):l(tt-
-
T(r* l"* * v'o):l(r-
-
This means that charge conjugation invariance [which demands l(rt* --- p* * v'")
- Itl(tr-
-----+ p- I u'"), etc.] is violated as well as invariance under parity.
is very important to note that the equality of the n* and the r- lifetime holds
despite the breakdown of charge conjugation invariance.
Following E. P. Wigner, we shall define the time-reversal operation which
reverses both momentum and spin. This definition of time reversal agrees with
our intuitive notion of the "reversal of motion."$ Consider now the product

fThe original calculation of Ruderman and Finkelstein was based on the parity conserv-
ing interaction if(hlm"c)(7f"l7xr)f"ysyrf,. The numerical result for the ratio of the
two decay modes is unchanged since
.l."vu ,tt .h, : ll$.vsry.r(l * 't){, * ryu)9,],
"}"YrYl(l -
and the calculation for the emission of a right-handed z proceeds in the same way as
that for the emission of a left-handed z.
glf the final state is not described by a plane wave, a much more careful discussion of
time reversal is needed. This is because time reversal interchanges the roles of an incoming
and an outgoing state. For a detailed discussion of time reversal in both nonrelativistic
and relativistic quantum mechanics, see Sakurai (1964), Chapter 4.
174 RELATMSTTC QUANTUM MECHANTCS OF SprN-+ PARTTCLES 3-1 l

+fl$
c, P, r. $
O,* O"- O"- O"-

!ig. 3-10. CPT


direction.
$ $ 4""$
operation applied to pion decay. The gray arrows indicate the spin

Table 3-5
TRANSFORMATION PROPERTIES UNDER CHARGE CONJUGATION,
PARITY, TIME REVERSAL, AND CPT

Helicity
Charge conjugation{ + + +
Parity +
Time reversal +
CPT +
fBy charge conjugation we mean "particle-antiparticle conjugation."

of charge conjugation, parity, and time reversal, denoted by CPT. The relations
(3.490) are completely consistent with invariance under CPT; the decay configura-
tion obtained by applying CPT is seen to be a physically realizable configuration
with the same transition probability (Fig. 3-10 and Table 3-5).
In the formalism developed in this chapter, the invariance under CPT can be
regarded as a consequence of the use of a Hermitian Hamiltonian. (Actually the
pion decay interaction we have written turns out also to be invariant under CP,
the product of charge conjugation and parity, but it is easy to write down a Her-
mitian Hamiltonian that violates CP invariance; for example, the electric-dipole-
moment interaction of Problem 3-16.) In a more axiomatic formulation of quantum
field theory it can be shown that it is impossible to violate CPT invariance without
drastically altering the structure of the theory. The fact that a wide class of quan-
tum field theories is invariant under CPT was first demonstrated by G. Liiders
and W. Pauli; its physical implications have been extensively discussed by T. D.
Lee, R. Oehme, and C. N. Yang. $

$For a more complete discussion of CPT invariance see Nishi-iima (1964), pp. 329-339;
Sakurai (1964), Chapter 6; Streater and Wightman (1963), pp. 142-146.
PROBLEMS 175

PROBLEMS

3-1. Prove Eq. 3.13.


3-2. Consider an electron in a uniform and constant magnetic field B along the z-axis.
Obtain the most general four-component positive-energy eigenfunctions. Show that
the energy eigenvalues are given by
E:@
with n : 0, l, 2, . . . List all the constants of the motion.
3-3. (a) Construct the normalized wave functions for .E ) 0 plane waves which are
eigenstates of the helicity operator. (The momentum p is not necessarily assumed
to be in the z-direction.) Evaluate the expectation values of l.fl and i,yg.fr
: €I .0, that is, J .lr*> .i^h dt * : I Fi,y ut.i+ dt* and I t+i,y ut .i^1, dtx :
J ^I'>.Wd'x.
(b) Construct the normalized wave function for an E > 0 transversely polarized
plane wave whose propagation and spin ((2)) directions are along the positive
z- and the positive r-axes respectively. Evaluate the expectation values of } and
infs'lr: €1.
34. Let r/r(x, r) be the wave function for an E > 0 plane wave.
(a) Show that ryrf(-x, r) is indeed the wave function for the E > 0 plane wave
with momentum reversed and the spin direction unchanged.
(b) Show that ilzg*(x, -r) is the wave function for the "time-reversed state"
in the sense that the momentum and spin directions are both reversed while the
sign of the energy is unchanged.
3-5. Consider the coupled two-component equations (cf. Eq. 3.26)

-i+,of,Dq{nl
OXtr
: -ff+<"t,
iS"gta,@):
oxp -T+,",,
otf;) : (o, i), of) : (a, -i).
(a) Write the above wave equations in the form

(,;&+T)$':0,
where

*':(l:;)
Obtain the explicit forms o1 ,y'u and ryl : ,yi,y|,ytr,yi and check {,y'r,,y',}: 28r, for
lL,u :1, . . .,5. Findaunitary,Sthatrelates thenew (Weyl) set fryi] andthestandard
(Dirac-Pauli) set [.yr] via t', : ^SyrS-t.
(b) Without using the invariance properties of the Dirac theory, discussed in
Section 3-4, establish directly the relativistic covariance of the coupled two-com-
ponent equations. In particular, find 2 x 2 matrices S(E) and S(') such that

+@,L)'.(x'): S(E,r) +@,1) (x),


where x and x' are related by a Lorentz transformation along the xr-axis. Discuss
176 RELATrvrsrrc eUANTUM MECHANTcs oF sprN-+ pARTrcLEs

also the covariance of the coupled wave equations under parity; $@,r-)'(16,) : ?,
where x' -- (-x, ict),
3-6. (a) Reduce
fi(r') (p,) o k4t(")(p) : i uc')+ (p,) ny
r uo) (p)
to the form XG')+9X$) where uo)(p) and uc')(p') are posilive-energy free particle
spinors and 1(sl and 1(s'l are the corresponding two-component Pauli spinors.
AssumelPI:lP'l.
(b) The interaction of a neutron with the electromagnetic field can be represented
at low energies by the phenomenological Hamiltonian density
ff int : -(*lel hl2m"c)llF,tro p,.|il, rc : -1.91.
Using the expression obtained in (a), compute the differential cross section for the
scattering of a slow neutron by an electrostatic field in Born approximation.
Interpret your results physically. Show, in particular, that even a very slow neutron
can have an attractive short-ranged (83(x)-like) interaction (known as the Foldy
interaction) with an electron. Show also that if this interaction is represented by
an equivalent spherical potential well of radius ro : (ezf 4rnt"cz), then the depth
of the potential is 4.08 keV.l
3-7. Using the uncertainty principle, N. Bohr argued that it is impossible to prepare
a beam of free electrons with all spins pointing in the same direction by means
' of a Stern-Gerlach type experiment or, more generally, by a selection mechanism
that takes advantage of the classical concept of a particle trajectory.$ Justify Bohr's
thesis from the point of view of the Dirac theory of the electron. Would this still
be true even if the electron had a large anomalous magnetic moment?
3-8. Consider the unitary operator
rr _ lffi +IEl , poc.pc
-TET--m
where f Ef is to be understood as a "square root operator," 4Q\I7V\V, which
justgivesJwasitseigenvaluewhenitactsonthewavefunctionfor
a free-particle plane wave.
(a) Show that the application of U to the wave function for a positive- (negative-)
energy plane wave results in a wave function whose lower (upper) two components
are missing.
(b) The above transformation (first considered in 1948 by M. H. Prycelf) can be
regarded as a change in the representation of the Dirac matrices. Show that the
operator in the usual representation that corresponds to x in the new representa-
tion is
_ _. ,
\z ichBd ic3h$(u x p)p czh(a x p)
^_^-_If4_ryWFAg
(The operator X is sometimes called the mean position operator.)

fAlthough the Foldy interaction is much weaker than the nuclear interaction, it can be
detected experimentally as thermal neutrons are scattered coherently by high-Z atoms.
$The original argument of Bohr can be found in Mott and Massey (1949), p. 61.
llThis transformation is a special case of a class of transformations considered extensively
by L. L. Foldy and S. A. Wouthuysen, and by S. Tani.
PROBLEMS I77

(c) Prove
)i : *crp llEl.
(Note that the expression for X is free ftom Zitterbewegung, but the nonlocality
of X is the price we must pay.)
3-9" At some instant of time (say, r : 0) the normalized wave function for an electron
is known to be

r(x,o) :h(l)r"*,.
where e, b, c, and d are independent of the space-time coordinates and satisfy
lalz *lbl'+ lcl'f ldl':1. Find the probabilities of observing the electron
with (i) E> O spin-up, (ii) E> 0 spin-down, (iii) E <O spin-up, and (iv) E <0
spin-down.
3-10. Consider a Dirac particle subject to a (three-dimensional) spherical well potentiat
V(r) : -Vo 10 for r 1rs,
V(r) : g for r ) ts.
(a) Obtain the exact four-component energy eigenfunctions for j : * ("even")
bound states, where "even" means even orbital parity for the upper two components.
(b) Set up an equation that determines the energy eigenvalues.
(c) What happens if the strength of the potential is increased so that Zo becomes
comparable to or larger than 2mc2?
3-11. Discuss how the numbers of nodes of the radial functions G(r) and F(r) of the
hydrogen atom are related to the quantum numbers n, j, and l.
3-12. Consider a positive-energy electron at rest with spin-up. Suppose at I :0 we
apply an external (classical vector potential) represented by

(where a is space-time indep"fi, ;";"r:,L"as for a unit vector in the positive


z-direction). Show that for t > 0 there is a finite probability of finding the electron
in a negative-energy state if the negative-energy states are assumed to be initially
empty. In particular work out quantitatively the following two cases: ha K2mc2
and hc,,t
= 2mc2.
3-13. (a) Prove
42
2 uf)@)u,F)*(p) : e[rzfr(p)z$r+(p) + of;'(-p)uf)r(-pI
: ulr,t Ellmc,).
(b) Using the above relation, prove the equal-time anticommutation relation
between and ./^[(x) (3.409).
^lr"@)
3-14. Consider an allowed pure Fermi B+ decay. Only the vector interaction contributes,
and it is legitimate to replace,f p by ryn.Without using the trace techniques to be
introduced in the next chapter show that the positron-neutrino angular correlation
is given by
|* (alc)", cos d,
178 RELATrvrsrrc eUANTUM MECHANTcS oF sprN-+ pARTrcLEs

where 0 is the angle between the momenta of e+ and z.f Compute also the expecta-
tion value of the positron helicity, and interpret your result, using angular momen-
tum conservation.
3-15. We showed in Section 2-3 that in atomic physics, favored radiative transitions
take place between states of opposite parities, that is, parity must change. Mean-
while, we know that the fundamental electromagnetic interaction conserves parity.
Resolve this paradox.
3-16. (a) Suppose the electron had a static electric dipole moment analogous to the
magnetic moment. Write a Hamiltonian density that represents the interaction
of the electric dipole moment with the electromagnetic field and prove that it is
not invariant under parity.
(b) Show that the electric dipole moment interaction would lead to a mixing
(in the quantum-mechanical sense) between the 2st and 2pi states of the
hydrogen atom. From the fact that the observed and the calculated Lamb shift
agree within 0.5 Mc, obtain an upper limit on the magnitude of the electric dipole
moment of the electron. (Caution: the relevant matrix element vanishes if the
nonrelativistic wave functions for 2sl and 2pl are used.)

{Because ,y+u(p) : r(-p), it is easy to see that the angular coefficient would be just op-
posite in sign if the interaction were scalar. Historically the vector nature of Fermi transi-
tions was first established by J. S. Allen and collaborators, who showed that the positron
and the neutrino tend to be emitted in the same direction in the B+ decay of A35 (almost
pure Fermi).
CHAPTER 4

COVARIANT PERTURBATION THEORY

+1. NATURAL UNITS AND DIMENSIONS


In relativistic quantum theory it is most convenient to use units in which action
(energy times time) is measured in h, and length divided by time is measured in c.
This system of units is referred to as natural units. When we work in natural units
(as we shall for the remaining part of this book) the symbol me may mean not only
the electron mass but also any one of the following:
a) reciprocal length
*"(:ii^'): 3T6-x +o=Em' (4.1)
b) reciprocal time
*"(:#):Izq'+=mra' (4.2)
c) energy
m.(: m.cz): 0.511 MeV, (4.3)
d) momentum
m"(: m"c):0.511MeV/c. (4.4)
In natural units the fine-structure constant is simply
.u-€2 I
- (4.s)

In the "pure" electrodynamics of electrons and photons, apart from mu there


is no other constant that has the dimension of mass or reciprocal length. So, from
purely dimensional considerations, we may argue that the cross section for any
electrodynamic process which is of order e2 in the amplitude must be of the order of
d2 /e2\2 1
mz: \u)d:'6' " (4.6)

re being the classical radius of the electron. Indeed, apart from the numerical
factor 8113, (a.6) is precisely the Thomson cross section computed in Chapter 2.
Working with a and m,, we can form other constants which are familiar from
atomic physics. For instance, the Bohr radius is given by
as - llAmu, (4.7)
so that

", kt -
:1,
ro
#. 137 (4.8)

The Rydberg energy is simply


Ry- : a2 m"f2: 13.6 eV. (4.e)

179
180 COVARIANT PERTURBATION TI{EORY 4-r

In high-energy physics a typical energy scale is set by the rest energy of the pion,
(139.6 MeV for r!
ffi' : (4.10)
[l3s.o Mev for f .

A typical length is then given by


llmn - JT x l0-13 cm, (4.r r)

which is accurate to 0.07 % if the charged pion mass is used. Since the dimen-
sionless constant analogous to a is of the order of unity in strong-interaction
physics, a typical cross section in collisions of strong-interacting particles is ex-
pected to be
llm'o - 2x I0-2Bcm2 : 20 mb, (4.r2)
which is, in fact, roughly equal to the total cross section for r!-p collisions at high
energies. Suppose there is a short-lived state that decays via strong interactions.
The decay width of such a short-lived state must be of the order of the pion rest
energy. As an example, we may mention that the decay width of the 750-MeV
p meson is about 100 MeV.I A decaying state whose decay width is 100 MeV has
a mean lifetime of 6.58 x 10-24 sec, which is a typical time scale of strong inter-
actions. As we mentioned in Section 3-11, the square of a weak-interaction coupling
constant is smaller than that of a strong-interaction coupling constant by a
factor of about 101a. As a result, the time scale characteristic of weak interac-
tions is about l0'4 x 6.6 x 10-24 sec. Indeed the mean lifetimes of particles like
K mesons and A hyperons are in the vicinity of l0-8 to 10-10 sec.
In any practical calculation we can work with ft : l, c: I throughout. At
the very end of the calculation we may insert, if we wish, the numerical values for
h and c in the appropriate places to make the dimension right. In going from
natural units to the usual units all we need to remember are:
h:6.58 x 10-22 MeV-sec, (4.13)
and
hc : 1.973 x 10-11 MeV-cm. (4.14)

For instance, take the example of Io decay (lo * A * .y), which the reader may
have worked out in Problem 2-5. The mean lifetime in the system of units used in
Chapter 2 turns out to be
r(mn* mt)tc'
_
' _- -----Vhk-' (4.15)

{More quantitatively, if the p decay interaction (po tt* * z-) is represented by


-
ff :
int s+,(ot, ft\* - * r,r)
where 6p, 6"n, and 0L are the field operators for po, tr- , arrd zt+, then we can show

r(po* r+*n-):(fl@i#
which gives (g2/4ft) :2 when I : 100 MeV.
a-2 S.MATRIX EXPANSION : INTERACTION REPRESENTATION 18t

If we compute the same quantity using natural units, we, of course, get

(4.r6)

Now k may mean the wave number of the photon Erfhc, the photon energy -8",
the photon momentum Erf c or the photon angular frequency c,: : Erlh; likewise
m^may mean ffincz, (hlmnc)'r or (hfmnc')-t. But we need not worry about all this.
All we should note is that when rnr, r/t>, and k in the formula (4.16) are expressed
in MeV, we must multiply the right-hand side of (4.16) by h:6.58 x lO-22 MeV-
sec to make the expression dimensionally correct. Noting that (mn * mr)c' : 2307
MeV, and E, : 74.5 MeV, we obtain
t}tlZW\'6.58 >!10_ _"_\{eV-sec _
":-4-\zsNlw) w 2.g x tg_,esec. (4.t7)

It is a complete waste of time to keep track of all the h and c at each stage of the
calculation, to convert the value of the hyperon mass in grams, or to express e in
terms of (gram)trz(cm)3r2(sec)-1, etc. It is so easy to get numbers if we just work in
natural units.
To conclude this section we shall make a few remarks on the dimensions of the
field operators. The Hamiltonian density ,ff or the Lagrangian density I must
have the dimension of energy divided by volume:
naturar unitst
lgl : energy/(length)3 (mass)a. (4.r8)
In the free-field Lagrangian density the boson field f (or A) appears in the form
(0 $ I 0 x )'z, (mc I h)'z $'z, l(l I c)A Al 0 tfz, etc. This means that
naturar units>
[dl : \rErrcrgltengt-E mass. (4.1e)
Note that c"@ which appears in the plane-wave expansion (2.60) indeed
has the dimension of ",lenffiglh- which is expected since the creation and
annihilation operators defined in Chapter 2 are dimensionless. According to
(3.360) the Dirac field r/r appears in the Lagrangian density as h{r,yr(0$l0x) and
mcz W. Hence
: \r4(t-teng@T naturar unitst (mass)t/'.
Wnl (4.20)
Noteagainthat"@whichappearSintheplane.waveexpansion(3.392)
is dimensionally correct since the creation and annihilation operators as well as
the free-particle spinor rz(*)(p) are dimensionless. In constructing interaction
densities it is worth remembering that expressions like ir|"yr.$A,,and {a automati
cally have the dimension of the Hamiltonian density in natural units, while an
expression like {3 must be multiplied by a constant that has the dimension of mass.

+2. S-MATRIX EXPANSION IN THE INTERACTION REPRESENTATION


Interaction representation. In this chapter we wish to discuss a variety of physical
processes assuming the validity of perturbation theory. The usual time-dependent
perturbation theory used in Chapter 2 is not too corlvenient for treating processes
182 COVARIANT PERTURBATION THEORY 4-2

involving relativistic particles because neither the space-time coordinates nor the
energy-momentum variables appear in a covariant manner. It turns out that within
the framework of the Hamiltonian formalism manifestly covariant expressions for
transition matrices can be most readily obtained if we use a representation known
as the interaction representation. The utility of this representation in quantum field
theory was first recognized by S. Tomonaga and by J. Schwinger in 1947.
In the usual Schrirdinger representation the time development of a state vector
O(t)(t) is given by the Schrodinger equation
i@@rs)l7t) : ,grsr6<sr, (4.21)

where the superscript (S) stands for Schrodinger. We can split gcsr into two parts,
lI(s)_F16*)+F/f), (4.22)

where f15*) is the Hamiltonian in the absence of perturbation, and f/f) stands for
the perturbation Hamiltonian. In field theory f/6s) is hken to be the free-field
Hamiltonian while 1/9t)is the space integral of the interaction Hamiltonian density.
Consider now the transformation

O(r) :
"rrfs)tqtsr, (4.23)
O(t) : ri nf,) tg<sr r-ia[s)r,
where O(s) is an operator in the Schrddinger representation. The above transfor-
mation can be regarded as a change in the representation-from the Schrtidinger
representation to what is known as the interaction representation. The symbols
O and O which have no superscripts stand for the state vector and the operator
in the interaction representation which correspond respectively to O(s) and O(s)
in the Schr<idinger representation. The time development of Q(r) is given by

'AQ = ilinfreirfs)r*,., f sir8s)r


'U #]
1 4n3s)t7*r, i Hfil s-inos)r eiafs)torsr
(4.24)
where H, is the =;n,rrrtot)rorrr
perturbation Hamiltonian in the interaction representation. Mean-
while, noting that O(s) and Ocoincide at t :
0, we have

O(t) : 4n$s)t O(0)e-m$s)t, (4.2s)


from which follows the equation of motion for O,
O: il//.f), Ol: ilHo, Ol, (4.26)

since I/[t) and Ho are identical. It is amusing to note that (4.24) would give a
constant state vector just as in the Heisenberg representation if H r were zero
and that (4.26) resembles the Heisenberg equation with I/o replacing the total
Hamiltonian. We could have anticipated all this, since the transformation (4.23)
would be identical with the familiar transformation that relates the Schrtidinger
representation to the Heisenberg representation if ,F1pt were replaced by f/tsr.
4-2 S-MATRIX EXPANSION: INTERACTION REPRESENTATION 183

The differences among the three representations are summarized more sys-
tematically as follows :
a) In the Heisenberg representationthe state vector is completely time independent,
whereas the operator which is time dependent satisfies the Heisenberg equation.
b) In the Schrtidinger representation it is the state vector that carries the entire
time dependence.
c) In the interoction representation the time dependence is split into two parts: the
time dependence of the noninteracting systems is carried by the operators, while
the remaining part of the time dependence due to the interaction is taken up by
the state vector. For this reason the interaction representation is sometimes
referred to as the intermediote representation.
In quantum field theory the field operator in the interaction representation
satisfies the equation of motion (4.26) which involves flo only. This means that the
field operator satisfies the free-field equation even in the presence of the interaction
provided that we use the interaction representation. This is indeed gratifying since
from the previous chapters we already know a great deal about the explicit forms
of the free-field operators. We can, for instance, use the plane-wave expansion
(3.392) for the electron field.
As for the interaction Hamiltonian, H, differs from IIF) in that the field oper-
ators occurring in H, are time-dependent free-field operators.

U-matrix and S-matrix. We now propose to obtain a formal solution to the differ-
ential equation (4.24). First, let us define an operator (I(t, t)by
Q(r) : U(t, t)Q(t), (4.27)
where O(tr) is the state vector that characterizes the system at some fixed time to.
Clearly we have
U(to, to1 : 1. (4.28)
Equation (4.24) is equivalent to
i(AlAt)U(t, t) : H,(t)U(t, t). (4.29)
But the differential equation (.29) and the boundary condition (4.28) can be
combined to give a single integral equation

U(t, to1 : 1
- ,l',"dt H,(t)(J(t, to). (4.30)

Let us solve this by the well-known iteration procedure. We have

u(t, to) - I- i
-, f ,'"dt,H,(t,)u1t,, t,)f
t',"dt,H,trl [t
- 1- i
J',"dtrHr(r,) + (-t)' I',"0r, I','"*rr^t,)Hr(tr)+
.. .

+ (-t)" fr"or, I'r'"*, f ,",'dt,Hr(tr)Hr(tr)... Hr(t,)+ ... (4.31)

They physical meaning of U is as follows. Suppose the system is known to be


in state i at
to. The probability of finding the system in state/at some later time r
184 COVARTANT PERTURBATION THEORY 4-2

is given by
(Or I U(t, to)@') l' : I U r'(t, t)12 .
I
(4.32)
The average transition probability per unit time for i "---, f is

, -- !,|(l,Q,to) - 8ro l', (4.33)

since the "1" in (a.31) would be present even if the interaction wereabsent. Although
this ratio has a rather complicated time dependence when the time interval t - to
is small, it approaches a definite limit os /s---)-oo arld t-*oo, &s we shall see
explicitly in working out actual physical problems in the next section. This moti-
vates us to define what is known as the S-matrix by
S: U(*, -*). (4.34)

From (4.27)we see that the S-matrix connects A(-) and @(-"") according to
O(*): S@(--). (4.35)

The explicit form of the S-matrix can easily be read from (a.31):
S,- S(o) *,Sttr + S(t) * Srtl + ...
- I - t J:_ dt,H,(t,) + (-t)' I*_*or, !'_*atrn,Q)H,(t,) + "'
+ (-t)" ll__or, !'_*0,, Jll dt,H,(t,1z,(t,) "' H,(t^) +" '. (4.36)

Whether or not the expansion (4.36) is useful depends on the strength of the
interaction. In quantum electrodynamics the S-matrix expansion turns out to
converge rapidly so that we can obtain results that agree extremely well with
observation just by considering the first few terms; this, of course, is because the
dimensionless constant a is small compared to unity. In the realm of strong-
interaction physics, however, the situation is not so fortunate because the relevant
expansion parameter is of the order of unity.
At this stage it is perhaps of some interest to discuss the connection between the
formalism developed here and Dirac's time-dependent perturbation theory ex-
tensively used in Chapter 2. First,let us look at the formula(2.107),which describes
the time development of the coefficient ct(t). We may imagine that the crforma
column matrix that corresponds to the state vector O in the language of this seq-
tion; the differential equation (2.107) is then seen to correspond precisely to (4.24).
The coefficient cr(r) is nothing more than the matrix element of the expansion
(4.31) taken between the initial (k) and the final (7) state i c5') : Drl corresponds to
the leading term, l, of (4.31) t c|D(t) given by (2.109) is readily seen to correspond
to the second term in the expansion (4'31) with ro : 0 if we recall that
Q I H r(t)l k> : (.j I erE"t H<;t "t I k)
"-tn
: <i IHF)lk)sr@rn)t. (4.37)
A similar remark holds for higher-order terms. From this point of view the basic
idea of interaction representation is as old as Dirac's time-dependent perturbation
theory.
+2 S-MATRIX EXPANSION : INTERACTION REPRESENTATION l8s

There is, however, an important difference between the perturbation method


we learn in nonrelativistic quantum mechanics and the perturbation method we
shall develop in this chapter. In the nonrelativistic theory, using (4.37), we first
integrate oYer tl, t2, Qtc. in (4.36) to obtain a transition matrix T* that goes as
follows:
Sr, : Dro
- 2ri8(E, - E)77i, (4.38)

rrt:nlqJ+4ffi+... (4.3e)

(where we have assumed for simplicity that l?(s) is time independent) and sub-
sequently perform a space integration of the type

H r') : QIV) [ e-t','*H$) eiln'x dax (4.40)

etc. This is not how we shall proceed in this chapter. Instead, we propose to per-
form the space and time integrations simultaneously in a covariant manner. As a
result, it becomes possible to obtain a manifestly covariant expression for the
transition matrix even though neither the basic equation (4.24) nor the expression
for the interaction Hamiltonian

H, = t ffr^"d'* (4.41)

is covariant. We shall see how this method works in the following sections.

Unitarity. We shall finish this section by discussing the unitarity property of the
S-matrix:
,SSt : l, @.a2a)
S+S: l, (4.42b)
or equivalently
X Sr"Sf,: D1i, @.a3a)
S;rS"r :8ri. (4.43b)
*
Although there are several ways of proving this important property, wo present
here a straightforward proof (to second order) based on the explicit form of
the S-matrix expansion (4.36). We assume that fI, is linear in some coupling
parameter ),; for instance, in the case of
Hr: -ie I {r.yr$Au d3x
the parameter ), is just e itself. The expansion (4.36) can then be regarded as a
power series in ),; S(0) varies as )u0 : l, S'(1) as ),, S(2) as ),2, etc. Meanwhile the
unitarity equation @a3a) can be written as
> (st') + st? + st') + .. .xs[?- + sflj* + s[?J* + . . .) : 8ru. @.44)
This must hold as an identity relation for any value of ),. So the coefficient of
each power of )" on the left-hand side of (4.4a) must vanish if the unitarity relation
186 covARrANT pERTURBATIoN THEoRy 4-2

is to be satisfied. Thus
> sflsfg* - 8ru - 0, (a.ala)
> (sfls[]J* + st?stg-) - 0, (4.45b)

> (sfls'?)* + sf)st?;-) + > st?stl. - 0, @.45c)

etc. The first expression @a5$ is trivially satisfied, since


S!o) : 81,, Sf?l* : 8,,. (4.46)

Because of @.aQ the unitarity relation to order f (4.45b) now becomes

r
J:_
dt(i I H,(t)lf>* - t J:_ dt(f I H,(t)l t> : 0. (4.47)

This is satisfied provided that H, is Hermitian.


tz: It tz: tt

/1-axis l1-axis

Fig. 4-1. Region of integration in (4.49) and (4.50).

Before we proceed to prove @a5c) we must prove a very important identity


for S('):
s(2) - (-t)' Il_*or, ['_*a,rn,Q,)H,(t,)
: (- i), Il_*
or,
t*,,dt,
H,7t,) H r(t,). (4.48)

To prove this we rewrite the first form by interchanging the variables of integration,

s(2) - (-t), [*_-rr, ['_*atrnrQr)Hr(tr).


(4.4e)

We are supposed to fix t, andintegrate over tr from -oo to tz; we then integrate
over t2 from -oo to oo. This means that in a two-dimensional trtrplane we must
integrate H{t)H//,) over the region above the line tz: tr as indicated in Fig.
a-l(a). But it is clear from Fig. a-l(b) that we can write (4.49) equally well as

s(2) - (-t)' Il_*0,, t*,*rr,(t,)H,(t,), (4.50)

which proves (4.48). We shall call the first and the second forms of (a.a8) respec-
tively "Form 1" and "Form 2" for S(2).
S.MATRIX EXPANSION : INTERACTION REPRESENTATION 187

Using this important identity for ,S(2) and (4.47) already demonstrated, we see
from @.a5e) that in order to prove the unitarity relation to second order in ),,
it is sufficient to prove that

; J:_
dt(flH,(t)tr) J]_ dt(nlH{t)li>
: - (- r),Ul _0,, Ji,at,p H,(t,) H,(t,) f)* + ll*at,l'_*or,<f lH,(t,)H,(r,)
I I I r>],
(4.s1)
where we have "Form 2" for s['J* in @.a5Q and "Form l" for sfr] in (4.45c).
Using the hermiticity of H r(t) and inserting a complete set of intermediate states
4l" )( n I : I between Hr(t') and Hr(tr), we readily see that the right-hand
side of (4.51) is reduced to

? J__
o,,U:, dt, t S'
;_
at,f(fl H,(t,) I D e I H {t,) | i>, (4.s2)

which is the same as the left-hand side of (4.51). This method can be generalized
to prove that the S-matrix is unitary, in the sense of (4.43a), to all (finite) orders.
A similar proof can be carried out for (4.43b).
To understand the physical significance of the unitarity property we have just
verified to second order let us go back to the probabilistic interpretation of the
U-matrix given earlier. Suppose the system is certain to be in state iatt:
Accordingto (4.32) the probability of finding the system in some particular -oo. final
state f at t : cp is then given by I Sr, 12. When we consid er all possibte final states,
the sum of the probabilities at t : ca must add up to one:

Xf lSnl' : l. (4.53)

This requirem'ent is guaranteed if the S-matrix is unitary. As we emphasized in


Chapter 3, in relativistic quantum mechanics the probability density of observing
a single given particle integrated over all space is no longer constant since particles
are freely created or annihilated. For instance, when we start with e* * e- with
momenta p and
-p, z possible final state may be not only e* + e- with p, and
-p' but also 2ry, 3,y, e* * e- * ry, etc. Yet probability conservation still holds
in the sense of (a.53) provided we include in the sum all possible states into which
the initial state can make energy-conserving transitions.
We would like to emphasize the role played by the hermiticity of the interaction
Hamiltonian in proving the unitarity relation for the ,S-matrix. Conversely, we
might argue on the basis of (a.al that probability conservation in the quantum
theory demands that the interaction density be taken to be Hermitian. In fact, we
can take probability conservation or the unitarity requirement as the starting
point. Assuming that S(1) is given by
-i times the time integral of Hr, we may
ask how S(2) must look if unitarity is to be satisfied. In this way, instead of deriving
the S-matrix expansion from the Hamiltonian formalism, we may derive the S-
matrix expansion from the unitarity requirement for the S-matrix and a few other
general principles. This approach, advocated particularly by N. N. Bogoliubov
r88 COVARIANT PERTURBATION THEORY 4-3

and collaborators is an interesting alternative point of view, but we shall not


discuss it any further.t
In terms of the Z-matrix defined by (a.3S) the unitarity relation reads
i(Tn - T{) -- 2tt E E(Er - E,)TXrT"i. (4.s4)

From this relation it


is possible to derive the optical theorem which relates the
total cross section to the imaginary part of the forward scattering amplitude (cf.
Eq. 2.219). Since the derivation of this important theorem from (4.54) is treated
in most modern textbooks on nonrelativistic quantum mechanics, we shall not
give it here.S
A somewhat different application of the unitarity principle will be discussed in
Section 4-7 where we shall show how (4.45c) can be used to relate vacuum
polarization to pair production by an external field.

4-3. FIRST-ORDER PROCESSES; MOTT SCATTERING


AI\D HYPERON DECAY
Matrix element for electron scattering. We are now in a position to apply the S-
matrix formalism developed in the previous section to concrete physical problems.
Consider first the simplest problem in relativistic quantum mechanics, the scat-
tering of a relativistic electron by an external (unquantized) potential. Although
this problem can be treated in an elementary manner just as well by the usual
Born approximation method (cf. Section 3-5), we shall deliberately use the language
of quantum field theory to illustrate some of the techniques and formalisms which
will be useful for tackling more complex problems.
As in the unquantizedDirac theory, the basic interaction density is taken as
ff ,n, : -ie{nyr*Al!', (4.5s)

where Af;) stands for an external (unquantized) potential. For example, Af(x)
may be a time-independent Coulomb potential
A : 0, Ar : -ieZl4rr (4.s6)

generated by an infinitely heavy nucleus (assumed to be fixed in space) or a time-


dependent three-vector potential
A : (0,0, a cos col), A+:0, (4.s7)

where c is space-time independent. Although the form (4.55) is familiar, we have


to be a little careful with what is meant by the expressionfur*.First,let us rewrite
it, using the decomposition Q.afi):
n'r'"t'
: 3;",h[;i;f'#i1;rTf;;.1 rr-, + .]r-,.h!-,) (4 58)
fsee Bogoliubov and Shirkov (1959), pp.20G226-
(1962)' p. 866.
$Merzbacher (1961), pp.495499; Messiah
4-3 MOTT SCATTERING AND HYPERON DECAY 189

As we explained in Section 3-10, the current operator must be so constructed that


its vacuum expectation value vanishes. This requirement is not satisfied by (a.58)
because (.|t.'+f))o is infinite. However, the expression

@)"u($f,'+b*' + ^7t-)^l^1.) - ^l^l-)^F *) + .7!-'95-t;, (4.se)


which differs from (4.58) only by a c-number function (vr)"B{fr!*),.11-)}, does
satisfy the required property since fr(*) acting on the vacuum state vanishes. In the
literature, to distinguish (4.59) from (4.58) one often denotes (4.59)by $"yr',9:.
We shall not use this notation but always understand F,y r*to be (4.59) rather than
(4.58) whenever ambiguities arise. The physical meaning of each term of (4.59)
has already been discussed in Section 3-10. We merely recall that the first term
destroys an electron-positron pair; the second destroys an electron and creates an
electron, hence it is appropriate for describing electron scattering; the third de-
stroys a positron and creates a positron (positron scattering); and the fourth
creates an electron-positron pair. The four possibilities are schematically shown
in Fig. 3-8.
To compute the transition amplitude from state i to a particular final state l,
we must consider Sr, 8yi since the leading term S(0) 1 simply represents the -
-
amplitude for a physical situation in which nothing new happens. Since Sr,
is just S(1) to lowest nonvanishing order, we consider
8ro
-
st'J : -r(*,l-i, I o', $"y,$Alpl*,), (4.60)

where the four-dimensional integration is over all space and time. In our particular
problem, 06 is an electron state with p : (p, iE) and si @r is also an electron state
with p' : (p', iE') and s/. Denoting by b* and b'* the corresponding creation
operators, we have
@r
l0),: @r : 6'i l0>.
b* (4.61)
It is intuitively obvious that out of the four terms in (4.59) only rp!-tr/r!*) contrib-
utes. This can be verified rigorously. Take, for instance, fit.)+F). A typical
matrix element will look like (0 lb'db" b+ | 0) so far as the creation and annihila-
tion operators are concerned, where we have denoted a typical electron annihila-
tion operator contained in ,rl,nl+r by b", a typical positron annihilation operator in
.p!.t bV d. But

<0 lb'db" b* I 0) : <0 lb' b" b+dl 0) : g, (4.62)


since the positron annihilation operator d acting on the vacuum vanishes. Thus
it is sufficient to consider the .p!-tql*) term only. Furthermore, we claim that it is
legitimate to replace f(*) by '1ffiV ts{s)(p)eie'r, r}t-r bv .J miEVbt+ u$')(pt)e-tn"'.
To prove this rigorously, let b" and b"'+ be a typical annihilation and a creation
operator contained in rfit+r and r|<-r respectively. We must consider
l0> : (0 l{0,, b,,,+}{b,,, bn} I 0>,
(0lb,b,,'+ b" b+ (4.63)
which vanishes if the momentum and spin indices of an annihilation operator
(creation operator) in 9t*t1q(-)) do not coincide with those of the initial (final)
190 COVARIANT PERTURBATION THEORY 4-3

electron. Thus, in the language of quantum field theory, the lowest-order scattering
is viewed as the simultaneous annihilation of the incident electron and creation of
the final outgoing electron at the same space-time point at which the potential
Af:, it operative. This is shown schematically in Fig.4 2(a). In contrast the second-
order amplitude, which we shall not consider here, is represented by Fig. a-2(b).
The anticommutator tb" , b+] in (4.63) is unity
when the momentum and spin indices of b"
coincide with those of br. The case is similar
for fb', b"'+j.So we finally get

st'/ : -, Id^.1^f haa)1p)s4n"of )


", -lr[-# u@ (p) sin' rf n t, r*r. (4.64) (a) (b)

Specializing now to the coulomb potentiat il1,-;i-3;rSiill #fflelectron


case (4.56) and abbreviating tG')(p') and
u(")(p) by u' andu,we obtain

st'J : -2ri8(E- t)[ ##rt|'vnu), (4.6s)

where we have used

I an*.nfr rt(p-p').r : 2rg(E - E,)


! a'4-tZrf 4ftr)s-i{v-ro.*
: -2tti}(E - E,)(Zellp - p,l,). (4.66)
The results (a.6a) and (4.66) are precisely what we expect when we solve the prob-
lem in an elementary manner using the Born approximation method (cf. Section
3-s).
The transition probability is lSt'J l' divided by the time interval t - to that ap-
pears in (a.33). If we naively square (4.65), we get the square of the 8-function that
expresses energy conservation. The meaning of this can be understood if we set
t : Tl2, to : -T12, andlet T --) oo. By the D-function in (a.65) we actually mean
rT /2

- E) --lrrm(tl2r) ) _ruri@-n')tt,J1r.
D(E' (4.67\

But this evaluated at E' - E - 0 is just Tl2r. We therefore sett

l2r8(E' - E)l' :2ttT\(E' - E). (4.68)

fA more "rigorous" way of proving (4.68) is given below:

I E(E, - E)1, - t ^l(rn")l':,, ,r<"'-rlr 6l'


:1sl"!ffi$a\,
: (Tl2n)6(E' - E),
where we have used (2.115).
+3 MOTT SCATTERING AND HYPERON DECAY l9l

Because of this, lSrtl'lT is seen to be independent of T. To get the differential


cross-section element into dA, we must multiply lS;rf lf by the phase space
factor and divide by the incident flux density which, in this case, is just lpllEV.
Thus
do:E#lffil(ff)
:t####D(E,_ E)d,p,

:#ffiaa, (4.6e)
where we have used

d3p'8(E'
- E): ffiaO
: lplEdO, (4.70)
with lp'l : lp l.

Spin sum and projection operator. The problem is now reduced to that of evaluating
lil',ynu lt. For a transition from a particular momentum-spin state to some other
momentum-spin state we may use the explicit forms of the free-particle spinors,
just as we did in discussing weak interaction phenomena (cf. Section 3-l l). Quite
often, however, we are not interested in the transition probability into a particular
spin state, since what is readily observable is the angular distribution of scattered
electrons without any regard to their spin orientations. We must then sum over
the final spin states. This spin sum can most readily be performed using a trick
originally introduced by H. Casimir. First, recall that | il',ynu l2 really means
lil' nl oul' : u+ ry ary au' il' ,y
au
: tt u(ry n) a, u', il'uQy n) u. u,. (4.7t)
So we must evaluate the spin sum )?,=ruf')(p')tf')(p'). To do this we take ad-
vantage of the following matrix relation:

2
2

s:l
u!@)uf)(p) : K-try. p * m)l2mf.B. (4.72)

The proof of this relation goes as follows. First, for the free-particle spinors that
represent the electron at rest, we have

0+(;).
2

2 4t'r(0)rz(s)(O) : 00 100)
s=1

i),'
1

I
(4.73)

0
t92 COVARIANT PERTURBATION THEORY 4-3

Now consider the "inverse'o Lorentz transformation of Section 3-4 (cf. Eq. 3.165)
which "boosts" the electron at rest.
ac)(0) ----+ z(s)(p) :,Sil.2(')(0),
(4.74)
;t'r(0) ---+ t(s)(p; : ;ts)(0)s"o,..
Multiplying (4.73) by S#. from the left and by ,56. from the right, we readily
obtain (4.72) with the aid of
S;.t.ryr,S*" : yr cosh 1
- iry.p sinh 1
: _iry. plm (4.7s)
(cf. Eq. 3.139). The matrix(4.72) can be regarded as the projection operator for the
electron spinor rz(')(p) in the sense that when it acts on z(')(p), we merely obtain
z(')(p) itself, whereas when it acts on the positron spinor o(')(p), it gives zero.
This is evident from (3.391). It is not difficult to show that the analog of (4.71) for
the positron spinor is

(4.76)

where p is the momentum-energy four-vector of the physical positron state (not


that of the negative-energy electron state whose absence appears as the positron
state). Note that

i trf,tp)tf)(p) - of)(p)o[o(p)l : 8oB. (4.77)

(Contrast this with Problem 3-13.)

Cross section for Mott scattering. Let us go back to electron scattering. We may
further suppose that the incident electron beam is unpolarized. We can then sum
over the initial spin states using Casimir's trick once again and divide by two.
So we are led to consider

i
1l
A 21il',vnul'
: ;4 ? 45"(px^r )u,uli')(p')t$')(p'X'vn)B"uf;)(p)

:t 'U'P#!)*(=t#t))
: * m)1,
(ll8m2)Tr(-t.y. p" + m)(-i,y'p (4.78)

where the symbol Tr stands for the "trace" (the sum of the diagonal elements)
andp" is defined by
P'l : -pL, pL' : pL- (4.7e)

To evaluate a trace of this kind it is useful to note that for any pair of four-vectors
a* and b* we have
Tr[(.y. a)(,f .b)l: Tr(ry'b)("y's)l: 4a'b. (4.80)
4-3 MOTT SCATTERING AND HYPERON DECAY t93

This can easily be proved sinCe

f poprf ub, * Vrbryrap : 2\urarb, :2a.b, (4.81)


and
Tr(l) : 4. (4.82)
Recall also that the ry-matrices are traceless so that a term like Tr(.y.p) vanishes.
Quite generally the trace of the product of an odd number of ry-matrices vanishes
as we can prove most elegantly, using ry5 which anticommutes with every ny, (p:
1,2,3,4):
T'(,WH) : Tr(.y3lpnt, . . . nt,,tp)
odd number : Tf(%rypnf, . . . rlrrfprls)
- -Tr(.yprlu ''' rl,f p). (4.83)
Thus

*} zW',ynul' : #r-oo" - p + 4m')

: 1

o^J*(4n.p' + 4E' + 4m')

: #(t, - lp 1, ,in,
$)
. (4.84)

Collecting everything we obtain the differential cross section for the scattering of
unpolarized electrons :

# : I r, ov ffiwD( r - ff sin' $), (4.85)

where we have used


lp - p' l' : 4lp l' sin' (012), F : I pllB. (4.86)
It is seen that, as ?, --) 0, Eq. (4.85) is reduced to the famous Rutherford cross
sectionf
(do\ _ t (Za)'
\aol -T (4.87)

It is also interesting to note that for the scattering of the charged


Klein-Gordon
particle we simply have (4.85), with the B2 sin2 (012) term omitted, as the reader
may easily verify.
The problem we have been treating was discussed, in 1929, without recourse to
perturbation theory by N. F. Mott using an E > m Coulomb wave function
(involving hypergeometric functions) obtained earlier by W. Gordon and C. G.
Darwin. For this reason the scattering of a relativistic electron by a Coulomb
field is referred to as Mott scattering.$ Our perturbative treatment is expected to
be valid at sufficiently high energies with relatively low values of Z, or more quanti-
tatively for ZlQ37p) < l.

fGoldstein (1951), p. 84; Merzbacher (1961), p.228.


$The reader who is interested in the nonperturbative treatment of Mott scattering may
consult Mott and Massey (1949), pp. 78-84.
t94 COVARIANT PERTURBATION THEORY 4-3

Finite charge distribution. In representing the potential by An: -iZel4tr we have


tacitly assumed that the nucleus is pointlike. We shall now show that at sufficiently
high energies a large-angle Mott scattering may be used to probe the finite size of
the nucleus. Let p(x) be the charge density of the nucleus. The l/r potential charac-
teristic of a point nucleus is now modified as follows:
Ze
_ I (ss-, p(x')
-4;T4-GJ-^lx-xT' (4.88)

Hence, calling the momentum transfer q


{: P'- P, (4.8e)
we see that the analog of (4.66) is

:2rig(E, - D ! d'x I o'*, ffi


I o-, Anst(n-o).,
: 2rig(E, - E) o'*' e-ts'x' p(xt) t a"ffi
I
: -2trii(E' - Dffir<q), (4.e0)

where the function F(q), called the form factor of the nucleus, is essentially the
Fourier transform of the charge distribution

F(q) : t eG)e-'n'*dsx
(4.e1)
-Ze
For a point nucleus where we have p(x): -Ze\@(x), F(q) would be identically
equal to unity in agreement with our earlier result. Evidently the cross section is
given by
@oMq : @olddL)*".1F(q) l', (4.e2)
where (do@A)4n1 stands for the expression (a.85). If lql' - 4 |p l2 sin2 (012) is not
too large, we can expand the exponential in (a.91) as follows:
q# + ...)e(x)d'x
F(q): +J (t -,q.x -
- 1-$rqt,+ ..., (4.e3)

where we have taken advantage of the fact that p(x) is spherically symmetric.
[Clearly f(Q depends only on lq l2 whenever the charge density is spherically
symmetric.l The quantity
(r'): tr'OO'.llpd'x (4.e4)

that appears in the expansion (4.93) is known as the mean square radius. The rela-
tions (4.92) and (4.93) have actually been used experimentally to determine the
size of the nucleus, particularly by R. Hofstadter and coworkers.

Helicity change; spin projection operator. In deriving the differential cross section
(4.85) we assumed that the incident electron beam is unpolarized. We also know
4-3 MOTT SCATTERING AND HYPERON DECAY 195

that the final expression (4.85) is easy to arrive at, once we master Casimir's trace
method. However, it is sometimes instructive to work out how the spin direction
changes as the electron undergoes scattering. For this reason let us evaluate the
S-matrix element between definite helicity states. Clearly factors like mlE and
l/lp - pl2 are the same as before. We are concerned with il'rynu, where u and u'
are now eigenspinors of helicity. Let uG)(p) and rz(-)(p) stand for the free-particle
spinors for electrons with momentum p and helicities * I and I respectively.
-
Choosing the z-axis along the initial momentum p, we find that r,r(*)(p) and z(-)(p)
coincide, of course, with rz(t)(p) and rz(2)(p). We must now construct the helicity
eigenspinors rz(r)(p'), where p' is no longer along the z-axis. We can do this most
easily by using the rotation operator of Section 3-4 (cf. also Problem 3-3a). If the
scattering plane (the plane determined by p and p') is chosen to coincide with
the xz-plane, all we need to do is rotate the electron state of helicity + I (- 1)
moving in the positive z-direction by the scattering angle d around the y-axis,
that is,*
z(*)(p') : s'-lu(r)(p) : (."r * - iI, rtr +)r.r(')(p)

0
cos
z
.0
sln
lE+ m z
\2m
h^! (4.e5)

Similarly
#-'i"$
.e
-sln z
0

ue(p'):E
cos
z (4.e6)
#r*"!
-;+r"o'$
With these helicity eigenspinors it is now straightforward to evaluate the ampli-
tudes for "helicity-nonflip" scattering and "helicity-flip" scattering, illustrated in

fThe term S;ul rather than S"ot appears because we are rotating the physical state rather
than the coordinate system.
t96 COVARIANT PERTURBATION TI{EORY 4-3

(a) (b)

Fig. 4-3. (a) Helicity nonflip scattering, (b) helicity flip scattering. The gray arrows
indicate the spin direction.

Fig. 4-3(a) and (b). Apart from a common proportionality factor, they are simply

tt( i(p),y n,(*)(p) : u#[r", + * (#;)' "", +]


: + "ot !, ( 4.e7a)

and
u6(p),ynLz(*)(p) : Tf-,io ! + (rl+;)' rtr' +]
.0 (4.e7b)
- -slnz
Two limiting cases of @.97) are of interest. First, if the electron is nonrelativistic,
the ratio of the probability of helicity nonflip to that of helicity flip is cot'z(012)
which is recognized to be the ratio of the probabilities of observing &r s3 - !
electron with spin in the direction p' and in the direction -p'f. This is expected
since the electrostatic field of the nucleus cannot change the spin direction of a
very slowly moving electron. On the other hand, for extreme relativistic electrons
the probability of helicity flip to that of helicity nonflip goes to zero,as(mlE)'
for a fixed d. This is again not surprising if we recall from Section 3-l I that (a)
apart from a minus sign, chirality coincides with helicity for extreme relativistic
fermions and (b) the rya interaction connects states of the same chirality. To sum-
marize, the spin direction does not change in the scattering of slowly moving elec-
trons, while it is the helicity that is unchanged in the scattering of extreme relati-
vistic electrons. As is evident from the derivation, this statement holds quite
generally for the lowest-order scattering of a spin-f particle by any potential that
transforms like the fourth component of a vector field.
Summing the two transition probabilities for helicity flip and helicity nonflip, we
see that the lowest-order cross section for Mott scattering of a longitudinally
polarized electron is the same as that of an unpolarized electron since

I a(*)(p')"yner(*)(p) l' + I ue(p),ynu(*)(p) f : (#)' ror'


$ + sin' $
-- h(a, - 1p l, sin, f), (4.e8)

f See, for example, Merzbacher (1961), pp. 276-277.


4-3 MOTT SCATTERING AND HYPERON DECAY 197

which is completely identical to (4.84). This does not mean, however, that we can
never detect electron polarization by Mott scattering. It can be shown that if the
scattering of a transversely polarized electron is computed to higher orders, the
differential cross section does depend on sin {, where { is the azimuthal angle
measured from the polarization direction of the incident electron.
When we obtained (a.8a) using the trace method, it was unnecessary to use the
explicit form of the free-particle spinor. In computing the transition probability
between particular helicity states, it is also possible to use the trace method without
recourse to the explicit forms of the helicity eigenspinors. All we need to do is
construct the projection operator for a particular spin state just as we can com-
pute the sum taken over only the electron states (as distinguished from positron
states) once we have the electron projection operator (4.72).
In the Pauli spin formalism, it is well known that
1*o.fl
---z- (4.ee)

is the projection operator for a two-component spinor whose spin direction is ft.
Motivated by this, we may try
1+>.f,
---z- (4.r00)

in the relativistic theory. However, as we emphasized in Section 3-3, a free-particle


spinor is, in general, not an eigenspinor of 2.fi. Moreover, this form is not mani-
festly covariant. So, remembering that Ir and BZn: t,yr,yr have the same non-
relativistic limit, we may try instead
I ! iryury.w
----2 ' (4.r01)

where w is a four-vector such that in the electron-rest system it is a unit (three-)


vector with a vanishing fourth component:
W l,r* frame : (ff, 0). (4.102)
More generally,
w2 - lwl' - l*ol' : l, w. p :0 (4.103)
in any frame. The physical meaning of w is such that when we consider a Lorentz
transformation which brings the electron to rest, then in that rest frame, (4.101)
is the projection operator for the electron spin state whose spin points in the direc-
tion ft. If we are interested in constructing the spin projection operator for a trans-
versely polarized electron, then the four-vector w is again given by
ly(trans) : (fi, 0). (4. r04)

This is clear because, quite generally, a three-vector is unchanged under a Lorentz


transformation in a direction perpendicular to the direction of the three-vector.
The reader who has worked out Problem 3-3(b) will recognize that the use of
i,yr^t.li is completely reasonable for a transversely polarized electron. On the other
198 COVARIANT PERTURBATION TIIEORY 4-3

hand, if we want the projection operator for a longitudinally polarized state with
helicity f 1, we mustuse
n(ronc) : (nfr, tlPl\, (4.105)
\m ml
which evidently satisfies (4.103). In fact, it is not difficult to verify that 2.f and
(i,yu^t.fiE ,yu,ynlpDlm have the same effect when they act on helicity eigenspinors.
-
It is important to note that iry. p and iyrnl. w commute:

l,y. p, fs,l.wl : -ys(y. p,f.w * "y.w,y


- p)
: -2rysP'lY
: 0, (4.106)
where we have used (4.81) and (4.103). It is therefore possible to construct a sim-
ultaneous eigenspinor of -i,y- p and !u"l- w with eigenvalues la and I respectively.
We shall denote such an eigenspinor by u@)(p).To summarize the various free-
particle spinors defined in this text, the description of momentum eigenstates
based on a(1) and u@ makes use of the eigenvalue of the 2 x 2 Pauli matrix o,
for the upper two components in the Dirac-Pauli representation. The description
in terms of a(*) and z(-) is characterized by the eigenvalue of the helicity operator
I.f. Finally the description in terms of z(')(p) makes use of the eigenvalue of
iryunl.W.
In practical calculations we often need to evaluate uf,)(p)nf)$) without the
spin sum (cf. Eq. 4.71). In analogy with (4.27) we expect that this is given by the
product of the electron- (as distinguished from the positron-) projectiort operator
multiplied by the covariant spin-projection operator (4.101):

(4.107)

This assertion can be proved in a simple manner as follows: First , (4.107) is evi-
dently true for the spin-up electron at rest,
00
eqtx+'*, :ll
- f;
: iryuryr\ 00
00
: tilft 00
'1..,
: l,; 000). (4.108)

\:/
Considering a rotation which transforms a unit vector along the z-axis into an
arbitrary unit vector fl, we obtain

u@ (o) rt@)(o) : (*u-xt#t*) (4.10e)

for any at-rest spinor with spin along li. We can then "Lorentz-boost" (a.109) to
get (4.107) in the most general case just as we obtained (4.72) from (4.73). As an
4-3 Morr ScATTERTNG AND HypERoN DEcAy 199

example, we may mention that we can evaluate I t3)(p),ynu(*)(p) l2 without re-


course to the particular form of the helicity eigenspinors
uG (p' ),yn
) ( * )
(p) l' : ilE') (pXryn )u, uf ' t
(pt ) ilf)( p' Xry, )u . ug) (p)
: r'1,^@# gi+tf, ,,%@ si+@1,
I rz

(4. I 10)

where, since we are considering longitudinally polarized states with helicity plus
one, lt and w' are given by (4.105). A straightforward (but tedious) calculation
shows, as it should, that this is nothing more than (4.97a) squared.I

Pair annihilation and pair creation. So f6i w.e have just talked about electron scat-
tering by an external potential. It is importbnt to note that an external potential
is capable of not only scattering electrons (or positrons) but also annihilating or
creating electron-positron pairs. Take, for instance,
e- * €+ --) vacuum. (4.111)
Although this cannot take place in the absenqe of a-:poiential because of energy
momentum conservation, if there is a time-dbpedderri potential whose angular
frequency is equal to the sum of the electron and the positron energies, we expect,
in general, a finite transition probability for (4.111), as we shall see in a moment.
The initial state @r is an electron-positron state represented by
@i : I e-e*) : bts-r+1n-)d(*)*(p*) I 0), (4.n2)
where (p-, s-) and (p*, s*) characterize the electron and the positron state respec-
tively. We have, to lowest order,

s!'i : -r(o l- i, I o,r{,r,^hA,l"-..). (4.1 l3)


This time the only term in (a.58) that contributes is .|!*)(vr).u^hh*', which annihi-
lates an electron-positron pair. Using the now familiar argumento we readily get

st') : -, I d'.lffir(q)(p*)e' "''fr,l1$u@(p-)e"-''f4,6) (4.I t4)

with p* : (pr, iE!): (pr,in/Ti;F +@. This expression is strikingly similarto


the first-order S-matrix for electron scattering given by $.6Q; the only difference
is that the adjoint wave function for the outgoing electron is replaced by the adjoint
wave function that represents the annihilated positron. This similarity is not acoinci-
dence. After all, in the hole-theoretic description pair annihilation is nothing more
than the "scattering" of a positive-energy electron into an empty negative-energy
state. We shall come back to this similarity between (4.64) and (4.114) when we
discuss Feynman's space-time approach in Section 4-5.
It is clear that S!! for pair annihilation vanishes for a time-independent poten-
tial (for example, the static Coulomb potential) because of E(E* + E-) that arises

{In this particular example, it does not pay to use the covariant spin-projection operator.
It is much faster to work with (4.95) as we have done.
200 covARrANT pERTURBATIoN THEoRy 4-3

as a result of the t integration. This, however, need not be the case for a time-
dependent potential such as (4.57) with o - E* + E-, since the integrand now
contains a term whose time dependence goes as exp [-i(E* -l E- - @)t], which
resultsin 8(E* + E- - ,) when integrated. Physically this difference arises
because a time-dependent potential can transfer energy to, or absorb energy from,
the charged particles, whereas a time-independent potential (such as the static
Coulomb potential) can transfer momentum to the charged particles but cannot
exchange energy with them.
We may remark that a process inverse to (4.11 1),
vacuum -----> e* * e-, (4.115)
can be treated in a completely analogous manner. In Problem 4-5 the reader is
asked to compute the probability per unit time per unit volume that the time-
dependent potential (4.57) will create an electron-positron pair in vacuum. We
have to admit that the potential (4.57) is not a very realistic one. In this connection
we should mention that the electromagnetic excitation and the de-excitation of
spin-zero nuclei of the same Z (for example, 0tu:--- 016* where 016* stands for an
excited J : 0 state of 016 with an excitation energy AE of 6.05 MeV) can be repre-
sented by a phenomenological potential of the type
A: 0, Ao:/(r) cos (A'Et), (4.116)
f(r) :0 for r > nuclear radius
so far as their effect on the electron-positron system is concerned. This effective
time-dependent potential offers a convenient framework within which to discuss a
process like
016* _+ Qtu + O* * O-, (4.n7)
which, from the point of view of the electron-positron system, is just pair creation
(4.115). In nuclear physics a process such as (4.117) is known as internal conversion
with pair creation, a theory of which was first formulated by H. Yukawa and
S. Sakata in 1935.t

Hyperon decay. Mott scattering, which we have treated in detail (or more generally
any problem that involves an external potential), is actually not such a good
example for illustrating the covariant perturbation method. This is because the
very presence of an infinitely heavy nucleus prompts us to select a particular
Lorentz frame in which the nucleus is situated at the origin. The lack of covariance
is reflected by the fact that the D-function that expresses energy conservation is not
accompanied by the E(3)-function that expresses momentum conservation. Phy-
sically this arises from the fact that an infinitely heavy nucleus can accommodate
any amount of momentum. If, instead, we treat the electron plus the nucleus
as a whole as a quantum-mechanical system, the formalism automatically gives both
the conservation of total energy and the conservation of total momentum (cf.

IWe shall not discuss this interesting phenomenon any further. The reader may consult
Akhietzer and Berestetzkii (1965), pp. 567-569.
4-3 MOTT SCATTERING AND HYPERON DECAY 20t

Problem 4-13, where electron scattering by a proton is treated without the assump-
tion that the proton is infinitely heavy.)
To illustrate the techniques encountered in covariant perturbation theory let us
now reconsider the decay of a A hyperon (treated earlier in Section 3-l l). Using
the interaction density (3.432), we see that the first-order ,S-matrix element for this
decay process is

s!'J : -,\"- rll o'.+hfr(g + s'v,)^/,^ n) |

: -i Io', #orl^[,r,e-,,,.,)(g + s,,yu)l^Eo"ne*n."] ,


: .l'-%- .l -^ -
-i(2r)ng(n)(pn - pn
ZW ^l ffi ^l f"vo{s * s'ry)un.
- pr),1-T-,,1
(4.118)

Note the appearance of a four-dimensional delta function that expresses both


energy and momentum conservation. It is convenient to define a covariant matrix
element denoted by til ro with the kinematical factors 8(n)(pn - pn
- p), \@,
etc., eliminated:

s"ri : Er, - i(2r)a8<at7pt'- pn


- nr^lffiW$-ll/1t, (4.lle)

where 81; (which comes from 5tor; is zero in this case and,tr n is to be computed to
only first order in g and g'. With this definition,
il ,u : ilr(g * g'ryu)un, (4.120)
which is invariant under Lorentz transformation.
We may be interested in computing the decay distribution of the final tt-p system
arising from a A hyperon whose spin state is characterized by the four-vector w.
If we are not interested in observing the polarization of the decay proton, we must
sum over the proton spin states. So we are led to consider

ptoton spin proton spin

:r'[ts* -s*'ryu)(=t#-)
x (s *r,r,>(it#-)(I_|#'*)] srzt)
The evaluation of the trace is facilitated if we note that
(4.t22)
,.,,,Ji|]"',H;i",*,].,;,,1,."
The only terms that survive in 4mnmntimes (4.121) are
t l s l' Tr[( - i
"y.
p,) (-
.p^)] + t l s l' Tr(m,m n),
i,y

**l s' I' Tr[(*ry'X-iry ' pr),vu(-i"t' p)1, (4.123)


itIS' l' Ttl(-ry r)mrryu*nl + iS *g' Tr[(- i'y' p)'yuinyrny'wffinl,
** g' * g Tr[(-"yuX - i,y' P )i,y u"y' wffi nl,
COVARIANT PERTURBATION THEORY 4-3

where we note that a term like Tr[(-,y)mo(i,yu"t.w)(-i,y'p")] vanishes because


w. pr,: 0. We can "kill" the ry5 in (4.123)by noting thatryu.yrrys : -?lp and ry! : 1.
We then use (4.80) and (4.82) to obtain
4mrmn
p.oton sprn
* 2Re(gg'*)(2po.wm). (4.124)
Without committing ourselves to any particular Lorentz frame we can compute
the differential decay rates as follows:

dw : ltrro)^a,n (pn- p, - p,)r(r*L)(#V)W)ffiffi,,.A"o,k( r,t,.


(4.12s)
The appearance of both V dsprl(2tt)3 and V d3p,l(2tt)3 can be understood if we recall
that we have not yet used the constraint imposed by the momentum conservation
implied by the delta function. We can handle the square of the four-dimensional
delta function just as in@.67) and (4.68), remembering that

p,,) : - po)'x], (4.126)


Dt"(pn - po
lt* lllQr)3| I nl'rexp
- [i(p^ - P,

which is VlQr)' if pn - pn - pp : 0, hence

I D(n,(p,,
- pn - p)f :
#E(n)(p^ - p, - p,). (4.t27)

The expression (4.125) is now reduced tol

dw : (2")^(+;)18b,)ldA"r"*f{+^,^^)*o,F.0', l/il r,l'8'n'(po - pn


- pp).

(4.128)
Note that the normalization volurne V no longer appears. We note that apart from
llEnthe expression (a.128) is completely relativistically invariant. To prove this
assertion it is sufficient to note that

{!:!*'ffi
'u
: !,,,00^o
: Jo,'o
8(P' + m') dnP' (4'l2g)

wherepo is understood to be an integration variable in contrastto E : JEPM


which is fixed once p is given.$ Furthermore, the dependence of dw on l/En is
readily understandable from the point of view of the special theory of relativity
which requires that a faster moving A must decay more slowly.

fA generalization of this formula, to any decay procpss, can be found in Appendix D.


$For any timelike vector (for example, the four-momentum vector of a free particle)
ihe sign of the zeroth component has a Lorentz invariant meaning.
4-3 MOTT SCATTERING AND HYPERON DECAY 203

The expression (4.128) with ) ll/ rol2 given by (aJz$ gives the differential decay
rate for a polarized A hyperon in any Lorentz frame. We can now work in the rest
system of the A hyperon so that p: pp :
-pn, p,r:0. Taking advantage of the
delta function that expresses momentum conservation, we can immediately
integrate over the pion momentum to obtain

dw '"@-\4/\n) l' I \ szttt,'s


:2-lpl'dlplda(ryt\ t /,/,olr g(mo- E, Er),
- (4.130)

which is in complete agreement with the Golden Rule when we realize that mrlEo
and ll2E, combine with | to give Vl(H'r)rul2. Meanwhile, remembering that
"// rol'
w is simply a unit (three-) vector in the A rest system, we have

4mrmoZ ltr rrl' : 2llgl' (Enmo * mrmo) * lg' l' (Enmn - ffirmn)
* 2Re(gg'*)(p0.fi)], (4.131)
which leads to the decay angular distribution (3.447) obtained earlier. If we are
interested in the decay width we must integrate (4.130) over the directions of p.
Thus
l(A --> rc- * P): +e)'Hlono.o,],',n lll rtlz momo, (4.132)
where we have used
dlplS(m^ - En
-Eo):d@M:ffi
EnEo
(4.133)
lPlm"
We may parenthetically remark that (4.132) is a special case of a very useful for-
mula applicable to any two-body decay process:I

r(r --+ 2 + 3): +(*)' H I onrinFspin fI (2mrc,"^i). 6134)


l"t/ r,l, external
fermlon
Combining (a.l3l) with (4.132),we finally get

r(A :
->,,- *p) lW + W(ffi;)lMW (4 r3s)

in complete agreement with (3.451).


The reader is now in a position to work out the angular distribution and the decay
rate of almost any decay process that can be described adequately by a first-order
---- A + ./ (given as Prob-
^S-matrix. As exercises, he may try G) -> e* + e- and Io
lems 4-6 and 4-7).

fThe product
i)
-#a"-' orr(22?rut"'
"*r"
arises because the normalization constant for a boson differs from that for a fermion
knTEV and 1/miEV respectively; cf Eq. (4.119)1. This difference would have been
absent had we normalized the free-particle spinor by r:to(p)rzt"r(p) :2m8,r,.
COVARIANT PERTURBATION THEORY

M. TWO-PHOTON ANNIHILATION AND COMPTON SCATTERING; THE


ELECTRON PROPAGATOR

S-matrix for two-photon annihilation. We shall now proceed to discuss processes


that involve second-order matrix elements. The fundamental relations on which
everything will be based are the two alternative forms of the S(2)-matrix derived in
Section 4-2 (cf . Eq. a.a8). If we use "Form 1," we have

s'(2) - (-t)' I*_*Orr l'_*Or, Hr(tr)H,(tr)


: (-i)' I on*, I,,.,,dLxzffindxl).f in(xr), (4.136)

where the four-dimensional integration over x, is to be performed only for tr l tr


With "Form 2" we have
sr(2) - (-t)' I-__or,li,o,,
H,(t,)H,(t,)

: (-i)' I o'*, I ,,r,,daxz ff in(x),ff 6r(xr) (4.137)

with r, later than tr.l


Without further preliminary remarks we wish to consider a specific physical
process that can be described by considering Stzr. We shall first treat the problem
of two-photon annihilation in detail and show how the notion of the electron
propagator arises in a concrete physical situation. The discussion of second-order
processes which involve the propagation of virtual photons or mesons will be
deferred until Section 4-6.
By two-photon annihilation we simply mean that
e* + e- +2y. (4.138)
The initial and the final states are written as
:
@i I e-e*) : bts-r1n-)nd(*'(P*)* I 0>,
(4.13e)
Or : l2"y) : a[,.,a[,",10).
Needless to say, S(0) and 5tt) give no contributions. So, assuming that S can be
approximated by 5(z), w€ write Syi in "Form 1" as follows:
: (-
S"ri e)'
I r,
On
I r,.,,da
x2 (2,y1 .}(",)ryr1h (x') Ar$(x z),t,nh(xr) A,(x r)l e- e*).
(4.140)

fln the literature "Form l" and Form 2" are often combined as follows:

s'i
(2)
- le D' pl Il *0,, Il * 0,, PIH /t,) H /t,)1,
where the symbol P stands for an operator that arranges the product of factors as follows:
t1] t2'
PIA(I,)B(t,)j: [1(!),t.(."), ':^
lB(tr)A(t) ]f tz ) tr.
This P-product (Dyson chronological producr) should not be confused with the I-product
(Ilick time-ordered product), to be introduced later.
TWO.PHOTON ANNIHILATION AND COMPTON SCATTERING 205

This particular problem (treated to this order) involves only real photons, and it is
therefore legitimate to regard A, as the (quantized) transverse electromagnetic
field discussed in Chapter 2.1 So far as the electron field operators are concerned,
the final state is just the vacuum state; likewise, so far as the photon field is con-
cerned, the initial state is the vacuum state. Since the free-electron field and the
free-photon field operate in different subspaces, we can write (4.140) as

S.ri : (-e)' I On*, I


,".r,dax,
(01fix,)ryr./r( xr)fi(xr),y,*(*r)le-e*)
x (2"y I Ar(x,) A,(x,) I 0). (4.t41)
From Chapter 2 we already know how to handle (2,ylAr(xr)A,(xr)10). Noting
that Ar(x,) can create either photon I or photon2 atx,, we get

(2,y I A r@,) A,(x,)l 0> : l(# 6(;') s - tx''


") (# ef,"> e- it,'',)

. (#ef) s-i'tz'a')(#rS"" r-T''")f, g.r42)


where ef;') simply stands for
uff ') : (6t",), 0) (4.t43)
1ryi1fu6tat) perpendicular to k,. We have, of course, lk, I : o)t> since we are talking
about a real photon.
Let us now concentrate on the matrix element of the electron field operators.
If we rewrite +.yr^h in (4.141) using (4.59), it may appear at first sight thatthere are
as many as sixteen terms. Fortunately most of them give rise to no contribution
when taken between the e-e* state and the vacuum state. For instance, take
.l!-)(",)^hl.)("').7!.)(xr){n5*)(xr), which annihilates two electrons and two posi-
trons. So far as the electron-positron system is concerned, there are only two
particles in the initial state, none in the final state. Hence the matrix element
of such a product does not contribute to this particular process. More generally,
since we have to go from a two-particle state to a no-particle state, there must be
three (*) and one (-). Furthermore, it is legitimate to replace .7"(x,)gB(x,)
by .F!.)(x')./rl*)(x,) since .7!-)(x,) and/or 9!-)(x,) acting on (0 | from the right
give zero. (Physically this is reasonable since .p!-)(x,) and/or fb-)(",) would give
rise to at least one electron and/or one positron created at t, which survive(s) at
t : |-oo.) So out of the sixteen terms we need to consider only two terms:
(y ) "
B
(y,), u.7!* )
(
"'
) tl* ("' ) .7 |,-
) )
(", ).h5. (", )
)
(4.144)
and
) " B (y, ), 0.7!* ( x, ) 9l* (x, ) gS - ) .| !. 1x, ;.
- ("y )
(4.t4s) ) t

Note that the ordering of rp!*r1xr) and *l-)(*r) and the associated minus sign in
(4.145) are chosen to agree with those of the third term of (a.59). We propose to

fOne may argue on the basis of Appendix A that the complete interaction Hamiltonian
must also contain the instantaneous Coulomb interaction. However, the Coulomb in-
teraction is irrelevant if we are computing two-photon annihilation to order e2 in ampli-
tude.
206 COVARIANT PERTURBATION TIIEORY 44

move the positron annihilation operator ./!*r in the first term (4.144) so that the
product .|!l)(x')f$*)(xr) may immediately annihilate the electron-positron pair in
the initial state:

(ry ), B
(y,), a r7!* (x,
)
)
,./n
!*
)
("' ) .7 !- (x,
I
) rfn6*
)
(x, )
:
).u(ry,)rutl*)(x')fr|-)(xr)frf)(x')^/r5*)(xr). (4.146)
(y

As for the second term (4.145), we interchange x, and x, and make the following
changes in the summation indices ltr *
u, dB *.y8 so that the second term may look
as similar as possible to the first one. The product (4.145) now becomes
(ry,)ra (ryr)"u )("r).1n5* )(xr) rlnb- )(",).7!* )(x,
- 0f* )
:
-(ry,),B(ry,)ru.7f*)(xr)./^b-)(x,).|!*)(x,)r/nS*)(xr)
. (4.147)
It is legitimate to interchange freely the integration variables x, and x, in this
manner provided that we keep in mind that t, is now later than tr. This forces us
to use "Form 2," which after all was originally obtained from "Form l" by inter-
changing the integration variables [cf. (a.a9) and (4.50)]. The photon-field matrix
element is unaffected by these changes because (4.142) is symmetric under xr e xz,
p* v. Thus we have

S.r, : (- r)' I On*, I On*, (2,y I Ar(xr) A,(xr)l 0)(ryrLu(y,)ru


x )(",).7!.r(x,)95*)(xr) e*)0(t,
[(0 l.hb.)(",).71. I e-
- tr)
- (01 Qli,@)^h!-,(x,).Ill)(r,).h5*)(x,) le- e*)O(t, - t,)1, (4.148)
where the 0 function is defined by

d(,):{; li3 (4.14e)

There still remains the task of simplifying the electron-field matrix elements.
This can be done by inserting a complete set of intermediate states Z"ln)(nl
between fl*)(",).lf)@r) [or r]!*r(xr)fl-)(x')l and .h.'("')^/rS*)(xr). Evidently the
only contribution comes from the vacuum state. We have
(0 I tF)(x')fr!-)(x,).7!*)("')^/r$*)(",) I e- e*)
: (0 I gb.)(x,).7f-)(",) I 0>(0 | .|!-)(x,)f5*)("n) I e- e*)

: (0 ltb.)(",).1!-,(",) rc>lr[#rf;.)(p*)e,o.'",]l^t#rzf;-)(p-)e' "-',,). (4.150)


An analogous reduction can be performed for the second term. Putting everything
together we finally obtain

sri : (-e), I on*, ! n,r,V#ef) s-t*,',')(#ef,s e-*"',,)

* {:',: i,}]lJ #r'ls'';1p*;eir" '']


x (ry,).p[(0ltl*)(x').7f'(",)10)0(t'- t,) - (0lf!.'(",)9!-'(x')10)0(t, - t')]
(4.lsl)
X (ry,),a
l^l#"r-'7p-1r'o-"f,
+4 TWO-PHOTON ANNIHILATION AND COMPTON SCATTERING

1 .(a2) o-ilc2'x2
t/ZazY
,r)

'("r)lo)
- e(^Y p)"p

r Tm - in-.r,
gf +) (", ) - rF#, u P
5ei
-' z
ll+)("r)---
\lE*v"""
(a) tt > t2

I ro"t -ik,.r.
'
t/-2uzY

(ol7l+t (r r) 9F) (x, )l o)


-
-e("v)ta X2

Qf,'@') -- #*yDo'oo*'"

(b) t, ) tr

Fig. 4-4. Time-ordered graphs for two-photon annihilation.

where we have abbreviated the second term of Eq. (4.142) by the expression
{k, <-+ kr, d, -, az}.
Expression (a.l5l) has both familiar and unfamiliar aspects. First, the appear-
ance of (t1l6-rV)e(f) s-*,''', etc., is familiar from Chapter 2, where we showed
that this is expected whenever a real photon characterized by (k,, a,) is emitted at
xr. Similarly JmlWuetp-'xz and J@*V6dp*', simply represent the annihila-
tion of the incident electronat x, and the incident positron at x, just as in (4.114).
It is clear that of the four electron-positron operators we started with, we have used
up two [.I!*)(xr) and to annihilate the incident positron and the incident
^/r6*)(xr)]
electron. The remaining two operators r/rB(x1) and fir(rr) (not used in annihilating
the incident particles) are seen to be poired and sandwiched between the vacuum
stqtes; this is really new. When l, is later than tz,we are instructed to use the first
208 COVARIANT PERTURBATION THEORY

vacuum product, in which f!-'("r) first creates an electron at exactly the same
space-time point (x2) at which the incident electron is annihilated and one of the
photons is emitted. The appearance of f!*)("t) in the same vacuum product tells
us that the electron created at x2 is subsequently annihilated at xr, the point at which
the incident positron is annihilated and the other photon is emitted. Thus for
h 7 tz, the vacuum expectation value (0 l9!.'("t)^7f,-'("r) l0) characterizes
the propagation of a virtual electron from x2 to xr. This is shown in Fig. M(a),
where the "internal line" that joins x, and xr represents the virtual electron going
from x, to x1. Similarly, when t2) ty then (01fr!.)("r)tl-'("')10) represents
the propagation of a virtual positron from x1 to x22 4S shown schematically in Fig.
H(b). Note that in both cases the action of the creation operator [r/r!-)(x,) or
precedes that of the annihilation operator [91.)("') or r]!*t(x2)l as de-
"7!-,("r)l
manded by the "causality principle."

Electron propagator. The-major breakthrough in the evaluation of the matrix


element is achieved when we combine the two vacuum-expectation-value terms in
(4.151) into a single expression. We propose to consider together the propagation
of a virtual electron from x, to x, for tr ) tz and the propagation of a virtual
positron from x, to x, for tz ) tr (not separately as in the old-fashioned perturba-
tion method). With this in mind we define what is known as the time-ordered
product (Wick time-ordered product) of the fermion operators as follows:

r(t"@)fiu(x')) _[ ^h"@)FB@')
for t>t' (4.1s2)
t-fru("') "lr"(x) for tt > t.
Since the (*) part [the (-) part] of the field operator gives no contribution when
it acts on the vacuum state from the left (from the right), we have

:I (0l./r!.)(x)fb-)(x/) l0> for t ) t',


<0 I 7(.h"( x)fie@')) | 0> (4.1s3)
t-(0 l^,lrl.)("')ft-)(r) l0> for t' ) t,
which is precisely the form that appears in (4.151). Let us now compute (4.153).
For / ) //we must consider
(o | 9!.)(x)fi!-)(rcl) | o>
: +(',3,^l#r(')(p)r,rf'(p) exp (ip.x) p*J #ut''r+(p/)r7f;''(p') exp (-ip'."') l0)
: fi* Iry ? 'f'(P)af)(P) exP liy'(x - r')l
: #lt|ery'p - i,vaQE) * ml'Bexp[rp'(x - x') - iE(t - /)1, (4.rs4)

where we have used


(0 ;br"1n;D(s)+(p/) l0> : 8nn,8,,,, (4.15s)
the well-known rule

ltl-i?: I#, (4.156)


44 TWO.PHOTON ANNIHILATION AND COMPTON SCATTERING 209

and(4.72). Similarly for f > t we get with the aid of (4.76),


1fr1.,(",)ft-)(r) 0>
-(0 I

: - #* I+ 4 - x) - iE(t'- r)]
,,$)au(p) exp [ip.(x'

tn _(i,y. p + *, . t*n [ip'(x' - x) - iE(/ - t)l


I _-- I d'!.f
fu* ) t \ 2m ).u_l
: #[t#Lrv'p - i,yn(_;l) + ml.Bexp[,p'(x - x') - i(-E)(t - t')1,

(4.1s7)
where in obtaining the last line we have changed the sign of the integration variable
(p --"---
-p) so that the whole expression may look similar to (4.154). We shall now
demonstrate that the vacuum expectation value (4.153) is equal to the four-dimen-
sional integral

(4.1s8)

where e is an infinitesimal positive real quantity and pz : lp l' p3 with po not,


in general, equal to JT|PM. we first integrate over p0 :
-

- 6 I o'o Il-*oo, (4.r5e)

Regarding the integrand as a function of the complex variabl" po, we note that it
has poles at two points,

po:Jff+r* -t8:E-iE, (4.160)


po:-ffi +tD --E+i8,
withE:(el2E)whereE,asusual,isunderstoodtoa"ffi,incontrastto
po, which is variable. We are, of course, supposed to integrate along the real axis
in the po-plane. This can be done using the contour-integral method. If t > t', we
must use the semicircular path indicated in Fig. a-5(a) since the po that appears in
exp [-?o(t - /)j cannot be positive imaginary. This means that for t > t' we just

Im(p6)-axis Im(ps)-axis
I

Po: - E+ta
Po: - E*i6--l
-Re(ps)-axis --Re(ps)-axis
X
Po: E-i6

(a\ t>t/t/ (b)t/>t


Fig. 4-5. The po integration in (4.159).
210 COVARIANT PERTURBATION THEORY 4-4

pick up the pole contribution from Po : E - t8. Hence (4.159) is equal to

wItE=iy.p-i,ya(iE)*ml,pexp[ip.(x-x,)-iE(t-t)](4.161)
for /) /', in complete agreement with (4.154). Similarly, by considering the contour
indicated in Fig. 4-5(b) we readily see that the four-dimensional integral is equal
to (4.157) for t' > t.
To summarize,we have proved that

(0 lr(.h.(x)frB(x')) l0> : - *-t, I o^oq+#+ exp [,p.(x - x')] (4.162)


regardless of whether t is loter or earlier than t' . Note that (4.162) calledthe "electron
propagator" is a 4 x 4 matrix which depends on the four-vector difference x - x'.
In the literature (4.162) is often denoted by -*(S'(x - x')),8 (where ,F stands
for Feynman).f The electron propagator is undoubtedly one of the most important
functions in twentieth-century physics; together with the covariant photon propct-
gator to be introduced in Section 4-6, it has enabled us to atttack a variety of
physical problems which we could not manage earlier. Although its explicit form
in the coordinate (x) space can be obtained in terms of a Hankel function of the
second kind, it is more useful to leave it in the Fourier integral form (4.162).In the
next section we shall show that it is one of the Green's functions associated with
the Dirac equation.
Coming back to two-photon annihilation, we find that the expression for the
S-matrix element (4.151) now becomes

s-rc : (- e)'
l on *, l on *, (# r@ ) r- tk 1' r') (
rralL
el") e-
*'' u)

/ m- ^t,-/. /x,)fi,(x,))
\T / \\rn\z \ / f n
* h/ EVD,e"''")(v,).u(0 lr(.lu(
0)(ry,), ,(^[#uaein-'") I

<--+ kz
-' [k' *---> ar)
]
[o,
:1-,1,!o^*,Ion*,,_ikt.t1-ikz.r2€@l)e@2)

y r-iP+.a,+' o-.,, or r(- 6 I O^ n


W)r,,
o{:"=o;,}, (4.163)

where in next to the last line the usual rule for matrix multiplication is understood.
It is indeed gratifying that the vacuum-expectation value that appears in (a.163) is
manifestly covariant. We may come to appreciate this point even more if we note
that (a) the notion of t, being earlier or later than /r is not a Lorentz invariant

fsometimes the symbol (S"(x - x')).8 is used for (4.162) itself. Our definition of Sr
in F. J. Dyson's original paper on this subject.
agrees with that given
+4 TWO-PHOTON ANNIHILATION AND COMPTON SCATTERING 2lt
concept when x1
- xn is spacelike; (b) unlike the anticommutator {"|r"(x),*B@)},
the electron propagator does not vanish for a spacelike separationt (as can be
shown by examining the explicit form of the electron propagator in the coordinate
(x) space). In other words, the separation shown in Fig. U(a) and (b) is not a
Lorentz invariant procedure. To put it in a more pictorial way, a physical process
in which we visualize a virtual electron as propagating from x, to xr may appear
in some other Lorentz frame as a process in which we visualize a virtual positron
as propagating from x, to xz.$ Yet when Fig.44(a) and (b) are considere d together,
the two vacuum-expectation-value terms combine into a manifestly covariant form.
The fact that we get a mathematically simple Lorentz covariant expression when
we consider Fig. M(a) and (b) together suggests that we should take the following
viewpoint. Instead of saying that the virtual particle is an electron that goes from
xrto x, when tr ) tz and a positron that goes from x, to x, when t2 ) ts we might
as well say the following: (a) It is always an electron that goes from x, to x, regard-
less of whether It is later or earlier than tr, (b) if t, happens to be later than /,,
the electron simply goes "backward in time." (We say this with the understanding
that the electron waye function for the virtual electron is that of a negative-energy
electron because of the appearance of a(')(p) l: !y(3,a,(-p).) From this point of
view the positron that propagates forward in time [e.g. the virtual positron in
Fig. aa@)l can be considered as being equivalent to a negative-energy electron that
propagates "backward in time." This interpretation (proposed originally by E. C.
G. Stueckelberg in 1941) forms the basis of Feynman's space-time approach (to be
discussed in the next section).ll
Given an internal line segment that joins x1 and x21 wo must still specify whether
the electron goes from xrto xz or from xrto x,. This can be done simply by drawing
an "arrow." If the arrow points from x, to xr, we understand it to be one of the
following two situations: either the electron goes forward in time from x, to x1
(t, > tr), or the positron goes forward in time from x1 to xz (t, > t,). If we do not
ask whether /1 is earlier or later than tr, Figs. a-a@) and (b) are topologically
equivalent. So they are collectively represented by a single graph Fig. a-6(a) where
t, can now be earlier or later than /,. It is also convenient to draw an arrow for an
external line which represents a real (as opposed to a virtual) electron or positron.
If the direction of the arrow coincides with the direction of increasing t (upward),
then the external line represents a real electron. If the arrow direction is opposite to
the increasing I direction, then the external line represents a real positron. All this

fFeynman (1961b), pp. 83-86, gives an enlightening discussion as to why the electron
propagator need not vanish for a space-like separation.
$In this sense we agree with "purists" such as G. Kiill6n who remarks, "We should like
to warn the reader against too pictorial an interpretation of the diagrams in terms of
particles propagating from one space-time point to another since such pictures sometimes
lead to serious misunderstanding of the physics involved."
llThe first statement in the literature proposing that a virtual positron appearing in the
intermediate states can be treated as a negative-energy electron except for the reversal
of the time ordering is found in a 1930 paper of P. A. M. Dirac. See also our earlier dis-
cussion of Thomson scattering given in Section 3-9.
212 COVARIANT PERTURBATION THEORY 44

Photon 2 Photon I

(a) (b)

Fig. 4-6. Feynman graphs for two-photon annihilation, where t2 ca;rr be earlier or later
than /t.

is clear from Fig. 4-6(a). The "arrow rule" can be concisely summarized by saying
that we associate with the action of r/r(x)[r](x)l an arrow toward (leaving) the
point x regardless of whether the action of the field operator represents annihila-
tion or creation, whether the particle created or annihilated is real or virtual, and
whether the particle is an electron or a positron.
The introduction of the arrows makes it simpler to keep track of the ordering
of the various factors in the S-matrix element (4.163). All we need to remember is
"follow the arrow." We start with the wave function for the incident electron to
be annihitated at xr, -€nf, to represent the interaction with the radiation field at xr,
the electron propagator that takes care of the propagation of the virtual electron
from x2to x, (with the understanding that the virtual particle'is actually a positron
if t, > tr), -enl*to represent the second interaction with the radiation field at x1,
and finally the (adjoint) wave function to represent the incident positron to be
annihilated at x,. If we subscribe to the "follow-the-arrow" prescription, there is no
longer any need to write the matrix indices explicitly.
We have argued that Figs. +a@) and (b) can be combined into a single graph,
Fig. 4-6(a). However, we must still make a distinction as to whether the interaction
of the electron-positron system with the radiation field at xr represents the emission
of photon 1 or of photon 2. For this reason there are actually two topologically
inequivalent graphs we must consider, as is clear from Fig.4-6(a) and (b). In terms
of the ,S-matrix element (4.163),the term denoted by kt <--+ k2, d4 e d2 corresponds
to Fig. 4-6(b). Figure +6(b) is sometimes referred to as a crossed diagram.

Covariant matrix element for two-photon annihilation. Let us now perform the
indicated integrations in (4.163). We propose to perform the space-time integrations
first. Omitting those factors which do not depend on xr and xr,we have

! Onr,
I On*,exp[i(-k, * p* * q).x,] exp li(-k, + p- - q).x,\
: (2r)aEnlg - (-p* * k,)l(2tt)nEnls - (p- - k,)l (4.164)

for Fig. a-5(a). One of the (2tt)a will be absorbed in the defining equation of the
covariant //-matrix to be given below, and the other (2o)n will be made to cancel
4-4 TWO-PHOTON ANNIHILATION AND COMPTON SCATTERING 213

with l/(2r)4 in the Fourier integral expression for the electron propagator. We
now perform the indicated energy-momentum (4-space) integration which simply
kills one of the E(a) functions (say, the second one) in (4.164) with the result that
4 is replaced everywhere by p- - k2. This amounts to saying that both energy and
momentum are conserved at each vertex, in sharp contrast to the analogous situa-
tion in the old-fashioned perturbation method according to which energy is
not co\served in the intermediate states.f There survives a four-dimensional 8-
function that expresses overall energy momentum conservation. The crossed
diagram Fig. a-6(b) can be treated in a similar manner; clearly q is replaced by
p- - kt in this case. As in the A-decay
case we define the covariant matrix element
til nby
S.r, : 8r, - i(2tt)a8tn)(p- I p* - k, - Orr ,./y'/ rt,
(4.16s)
where E7i is again zero for this particular process. We end up with the remarkably
simple expression

-iZ ,, : (-e)rrr*)(p.)[y .e<',>


=-,i!'(P-=a F)-,!
m=-f .e@,)

* "v' u@Mv'er"'r]z(*)(p-). (4.166)

The first and second terms correspond to Fig. +6(a) and (b) respectively. Note
that the intermediate momentum of the virtual electron is qr : p-
- k, or p- - kr
The relative sign of the two terms is positive as expected from the Bose-Einstein
nature of the photon; we get the correct result automatically since we have used
a formalism in which the radiation fields commute.

Matrix element for Compton scattering. From this single example of two-photon
annihilation, it is perhaps not completely obvious that the various electron field
operators in the ^S-matrix necessarily combine in such a miraculous way that we
always end up with an elegant expression involving the covariant electron pro-
pagator. For this reason, before we proceed to obtain the cross section for two-
photon annihilation, let us work out the second-order matrix element for another
process, Compton scattering. This time the S-matrix element analogous to (4.140)
is

S.ri : (- e)' I nn *, I,,.,,daxr (,y' e-' fr(",)ry, $@r)A


1 r(4)fr(xrh,{( xr)A,,(xr)lrye->,
(4.167)
where lrye-) and | ,y'e-') symbolically represent the initial and final y * e- states.
Remembering that I, is later than tr, we argue that r/^(x,) cannot create a positron
(since there is no positron in the final state) and that $(xr) cannot annihilate a
positron (since there is no positron in the initial state). Thus .h(x,) and r|(xr) can be

lWe shall come back to this important difference between the old-fashioned method and
the modern (covariant) method in Section 4-6.
214 COVARIANT PERTURBATION THEORY 4-4

replaced by ft+r(x,) and frt-'("'). Furthermore there must be two (*) and two (-)
since we start with one electron and end up with one electron. The products of the
electron-positron operators that give rise to nonvanishing contributions are
then seen to be
(y ), B (ry )" u fr !- (x' ) ^/r[* (x' ) .]
) )
l,- I
(x, )
)
(x, ) (4.168)
^lS*
and
(t ) B
(v,), u fr !* (x' ) 9b. ("' ).7
) )
!'-
)
( x, ) 9$- (", )' )
(4.16e)
"

When sandwiched between the initial and the final electron states, (4.168) becomes

("y r). p ("y,)ya (e


-' I .7!- )(x, ) 9b. )(x' ).7F r(xr) f5* )(x') Ie
-
)
: (ry).8(y,)"0(e-' l0t-,(",) I 0>(0 l.hl.'(x').7!-'(",) I 0><0 1.hS.)(x,) I e-)
: l,r/hu!,e-to"")(.y,).B(0
lf!.)(",).F ?@,)l 0Xry,),u
l^l h^r'o''"f.
(4.170)

As for (4.169) we interchange xr t-+ xz, p e v, and aB +-+ ry8. Clearly

(,y r), p (r )r o r| [,* (x, I


)
f 5. )
(", )'.71- 1v' ) r/n!- (x
t )
r )
: (,t,) B (t ) ru[
" - fr !*
)(x,).]t- )(x, ) t5* )(x,) tl- )(x, )
* Vt*'("'X95. )(x,), .7!- )("')] fl- )("')l
: (ryr)"B (ry,) ro [ fr t- )(x' ).]t. )(x, )./rl- (x' )
- ^/n5*
(x, ) ) )

)(x,)) )(",
f- fi\.) (r rxfS.)(;,), .|!- ^/rb-
I. (4.t7t)

Recall now that [f$*)(xr), .]!-)(x')] is just a c-number function of x, - x2 and that
r|!*r and ,r/r!-r respectively annihilate and create positrons. This means that the
second term of (a.171) gives no contribution to Compton scattering.l Sandwiching
(4.I71) between the initial and the final electron states, we obtain
-' )(", ).7!* )("r) )(r' )(",) -
)
(y ). B(y,)"a (e | - fr!- f b- ) 9S. Ie

: l^lhu!,e-tn""](.v,),u(0 I -fr!.'(",)fb-)(x,)
l0)(ry,)"u
l^{#^ruo'"),
(.172)

where we should keep in mind that t, is now later than ft. Note that just as in the
annihilation problem some of the field operators are used to annihilate and create
the incident and the outgoing electron while the remaining operators are used to
create and annihilate a virtual electron (in the case /, > tr) and a virtual positron
(in the case /2 7 tr).Considering the two terms (a.170) and (4.172) together, and
taking advantage of the fact that the photon-field matrix element is symmetric
under p e v, xr e x2e we see that the two terms combine to form a single expres-
sion valid regardless of whether Ir is later or earlier than /, :

l^/ hr*e-tn"")
(.r,)"u (0 l z(tu(x,)fr,(",)) I 0)(ry,),u
l^l #^r"' "f. @.173)

fTo prove this rigorously, note that (0lbtdtd+b+ l0): (01[d', d+]{b',6+}10>:'0 since
the initial and the final electron states are in general different.
4-4 TWO-PHOTON ANNIHILATION AND COMPTON SCATTERING 215

The photon-field matrix element is well known (from Section 2-5, where we dis-
cussed Rayleigh scattering in detail). We therefore get

s"r, : (-e)' I o'*, I o*,l#rate,o',,f1#r{''te-*",,)


" l,tp u!, e- n''
t
"f(ry,)-u (0 I z(.h B(", ) fi,(*,))l 0) (ry,),u l^E u,
"'''
"'1

-, lofk<--->k'l
.- a'l' (4.r74)

The final expression involves the covariant electron propagator just as advertised
earlier. The remaining steps are very similar to (4.164) through (4.166). When the
indicated integrations are performed, we end up with the ufl-matrix element
[defined as in (4.165)]:

-L/il 1t, : (- e)"(")(p') [r. ut"' ffif . e@'t

(4.17s)

This expression can be visualized by drawing Feynman diagrams (Fig. 4-7).Again,


as Feynman put it, "the result is much easier to understand than the derivation."

(a) (b)

Fie.4-7. Feynman diagrams for Compton scattering.

Feynman rules. If the reader has closely followed the derivations of (4.166) and
(4.175) and, in addition, worked out a few similar problems, he will easily con-
vince himself of the following set of prescriptions for writing
-Ltil 7t First, draw all
possible topologically inequivalent diagrams. So far as the external fermion lines
go, for each real (as opposed to virtual) electron or positron annihilated or created,
associate the following free-particle spinors:
Annihilated electron: u(')(p)
Annihilatedpositron: o(')(p)
Created electron: z(')(p)
Created positron: o(')(p)
,
For the photon use eff). Write -€nlp at each vertex. (Note that the factor i that
appears in -ie$,yr*A, cancels with the -i of (-i)" that appears in the nth-order
2t6 COVARIANT PERTURBATION THEORY

S-matrix element.) For a virtual electron (or a positron) represented by an internal


line r-><write
-i,y. Q * m
(4.176)
M+W=E)',
where 4 is the four-momentum of the virtual particle when the internal line re-
presents the propagation of an electron going forward in time. (The case when the
virtual fermion is a positron is taken care of automatically.) Assume energy-momen-
tum conservation at each vertex. If the energy-momentum of the virtual electron
is already fixed by energy-momentum conservation (as in our examples of two-
photon annihilation and Compton scattering), there is no q-space integration
needed, nor is the factor llQtt)a necessary. On the other hand, if the virtual four-
momentum is not fixed, a "nontrivial integration" remains to be performed; we
must integrate with daql(2t)a J As for the ordering of the factors, just "follow the
arrow." In addition, there are certain sign factors which will be explained later.
This would be a logical place to discuss the connection between the old-fashioned
perturbation method based on energy denominators and the modern covariant
perturbation method based on Lorentz-invariant denominators. However, since
this connection can be seen in a simpler way in processes in which spin-zero virtual
bosons propagate, and since spin is an inessential complication in the illustration of
this point, we shall postpone this topic until Section 4-6.

Cross section for two-photon annihilation. The remaining part of this section is
devoted to a detailed discussion of two-photon annihilation and other related
processes. The relation between the -rZ -matrix element for two-photon annihilation
(4.166) and the differential cross section can be obtained just as in the A-decay
case treated in the previous section. The only new feature is the flux density, which
is indicative of how fast the two beams hit each other; it is given by antlV, where
?."1 is the relative velocity of the electron and the positron as observed in the
coordinate system in question. In analogy with (4.127) we have

do : (2rr)n
*,#fitfl r,l'(2m)'#h,er-b1,^,u'n'(p* * p'- k, - k,)'
(4.177)
This formula is good in any coordinate system in which the two incident beams are
collinear. In the center-of-mass (cru) system

?re,:S+H:W (crnr) (4.178)

with lp I : lp* I : lp- l. It is worth keeping in mind that tr,"1 can exceed unity,
that is, the relative velocity can be greater than the speed of light. In the so-called
laboratory system, in which the moving positron collides with the electron at rest
(p- : o), we simply have
arct: lp,llE. (lab). (4.t7e)

lWe shall encounter nontrivial integrations in Section 4-7.


TWO.PHOTON ANNIHILATION AND COMPTON SCATTERING 217

Itis interesting to note thata,"1E*,8- is the same in any Lorentzframe in which the
two incident beams are collinear since we can readily prove
(4.r80)
Using (4.128) and (4.180), we see that (4.178) is Lorentz invariant. This is reason-
able since the integrated cross section must be the same regardless of the coordinate
system we use. Using (4.178) and the analog of (4.133), we can readily work out the
differential cross section in the cu-system:

(#)
"-
This expression is a special case of a
-- + h#
vfi
i+i t il,,t' (2*)'
useful relation:
(4.181)

(#),-: + ## #* ltr r'l'"".H'(22?tu""'i) (4'182)

applicable to any reaction of the type I t 2


- 3 + 4.In the lab system we may be
interested in the differential cross section for the emission of one of the photons,
say photon 1. As usual, we can immediately integrate over k, because of the three-
dimensional delta function that expresses momentum conservation. In evaluating
8(E* + E- - G)1 - o2) d'kr, all we need to know is 0E710lk, I with 0, and {, fixed.
Using the kinematical relations

lp*-k, l: :lkrl:rr,
(4.183)
m*E*:(Dt*atz,
we obtain
(7Er \ _0(rr+ lp*-k'l)
\aTE;l/,,,,b, w
_r , t-
lk, l-lp. lcosd, _m(m+E*) (4.r84)
0)2 0) r (Dz

So the differential cross section in the lab system is

(#),^,
:TTQ?'ffi
h ffi,t'// "
t'

T@Yffi@f"Jl/z"l'(2*)'' (4.18s)

It turns out that in this particular problem the evaluation of l,% ,o l'? is simplest in
the lab system. First

-: im), :
p (0, 0, 0, ("), 0).
e(o) (e (4. r86)
This means that ,y, p
- anticommutes with ry. €(o), hence
(- i"y . p -)("y .u{a)) z{s)(p- ) : - (ry . e@) muG) (p -), (4.187)
where e(") may fg 6(cr) 91u(cvr). Moreover
(p- - kr,r)' - -m2 - 2p-.k,,,
- -m2 | 2ma1,2. (4. I 88)
218 COVARIANT PERTURBATION THEORY 4-4

Thus (4.166) now becomes

.ty'l rt : tr' affi *'Y',-(d !&T''@)fu. (4.18e)

If the initial electron and positron are unpolarized, we must evaluate

* ; lll ,,f : +,I- ot-'rn Q+ rnPG) 5G') Ql'G-)


: ,lvnorv^(-''' o#
t *)"(=r#)), (4.reo)

where we have denoted the expression inside the bracket of (4.189) by O. The
factor t appears because we must average over the four spin states of the e-e*
system. In evaluating ryn Or,yn it is important to remember that
,y{,y.q)+yn: Tr[y.a t ,y4(iq)*l,yr: -f ,e (4.ler)
for any four-vector a, whose space components are purely real and whose fourth
component is purely imaginary. The relation (4.191) can be generalized as follows:

rya[(ry.a)(,y-b)(,y-c) .-- (,y-l)]r,yn: (-ry.1) .- -(-ry.c)(-"y.b)(-"y-o). (4.192)


For the trace arising from the square of the amplitude for Fig. a-6@) we must
evaluate
Tr[(ry. e@)ry.kr,f..rar))(f{. p* I m)(,y..(cr)t. krry.e@'t1Gi,y. p- * m)1. (4.193)
First, terms linear in m are zero because the trace of the product of an odd number
of the gamma matrices vanishes by (4.83). Terms proportional to m2 also vanish
because
,f .kzry.e@r)ry.6'(ar)ry. kz : kZ : 0, (4.te4)
where we have used
(,y.a)(,y.a): a', (4.195)
which is a special case of (a.81). To evaluate the remaining terms we keep in mind
that (4.195) can be used to "ki11" the gamma matrices; so the trick is to cleverly
rewrite the various factors using (4.81) until we reach the stage at which we can
take advantage of (4.195). Specifically
Tt(ry. e@)ry.kzry..(c')t. p+f .utar)n7. krrf -e@')ry. p-)
: TrW.kr(2e<o,',. p+
- p+f ..(cr);t. u@t)r.kr(26<',>. p- - ,y. p_ry.e(",))t..(o,)1
ry.

p-) + Tt(ry. kzy- p*nl.kr,y- p-)


- -)r(ar). p*Tr(,y.kr.y.e@'),y.kzI.
: .p
-)6@) *TrW. kr(2e{",> . kz
- I. kzy.u(''))ry. p_l
* Tr["y. kr(2p *. kz - I' kzy' p *),y' p -]
16(e("')p*)(e(o'r. kr)(kr- p-) + 8(kr. p)(kr. p-)
(4. le6)

where in the last step we have sssd 6(ar) ' P+ : u@r) '7r, and kro p* : kr' P- which
follow from energy-momentum conservation and the transversality condition for
the photon polarization vector. The trace that arises from the interference of Figs.
a-6(a) and 4-6(b) can be obtained in a similar manner using various tricks. The
TWO.PHOTON ANNIHILATION AND COMPTON SCATTERING 219

trace arising from the square of the amplitude for Fig. 4-6(b) can be written imme-
diately just by interchanging the photon labels in (4.196). The net result can be
shown to be

*,t-;il,,r:#lff* fi+ 2-4(e@').er''r)z] . G.lg.)


This leads to the differential cross section for two-photon annihilation:

4(er''1..t"")'], (4.re8)

obtained by P. A. M. Dirac in 1930.


Let us now specializeto the low-energy limit of (a.198). With
(D : 0)r : a2, k - kr - -k2, lp*l: B*m (4.ree)
(valid when the incident positron is very slow), we get

(#),^,: &rt - (e(o'r'6(ar))zl' (4.200)

If the photon polarization is not observed, we can sum over the photon polariza-
tion states. Choosing the z-axis long k : k,, one can orient the polarizationvector
in the x-direction or in the y-direction. In this way we readily construct Table 4-l
for I - (e(o'r '6@z)12. The polarization sum evidently gives rise to just 2. To obtain
the total cross section for pair annihilation we must note that the final state in-
volves two indistinguishable particles. Hence

o^,-+1ffi)r' (4.201)
So we have

otot
- ftz x 2r :4p. (p. < l) (4.202)

for the spin-average total cross section. The cross section varies inversely as the
velocity of the incident positron. This is characteristic of an exothermic reaction.
Although the cross section becomes infinite as €* ,0, the number of annihilations
per unit time remains finite as the positron beam slows down in matter since the
flux is proportional to B*.
Coming back to (4.200) and Table 4-1, we note that the transition probability
vanishes when e(1) and e(2) are parallel. This simple result, we argue, must have
a simple explanation. To see this and other interesting features let us treat the
problem of.low-energy pair annihilation using the explicit forms of the free-particle

Table 4-1
THE VALUES OF 1 - (6t"'r.s@D1z IN (4.200)

6tc,) parallel to x-axis parallel to y-axis


6tcr) par&llel to x-axis 1
6to') parallel to y-axis 0
220 covARIANT pERTURBATIoN THEoRY 44

spinors. As the positron velocity goes to zero,the -// -matrix element becomes

../il 7t: Qe'zl2ma)O(*)(O)[.y. r(a)r.kr.y.s@'t * "y.


u@z)r-k;y.e@'>1rt'-l(0). (4.203)
We can replace cy.e@)r.k2.y-sr") by y. r(at)y.kay.6(az) sitlss the photon polari-
zation vectors have no time components and 1fl+f i vanishes when taken between
,r'-r(0) and t(*)(0). The procedure is similar for the second term. Remembering
(4.199), we obtain, for the expression inside the bracket of (4.203), the following:

-' 1"'1,; Il;il;' ..;i; ::; ?;:


: -2k.(6(ar) y e@"))rynryu,
.,1,, 1
-,, (e(.,, x k)l
(4.204)
where we have used
(Y.aXv'b) : a'b + i2'(a x b), y.L : irya,y52.a. (4.205)
We now show that the e-e* state with a zero relative momentum in the triplet
spin state cannot annihilate into two photons. Because of rotational invariance it is
sufficient to prove this statement for the.fr : 1 state represented by
(t)"_(t)". : b(1)+(0)d(')f(0) l0>. (4.206)
The relevant free-particle spinors are rz(1)(0) and o(1)(0). In accordance with (a.20a)
we evaluate

(u"ynryuu)rriplet, r3:r : o(t)+(0)ryual(t)(0) : (0, 0,0, -,,(-,qi-*,(i) -0.


(4.207)
On the other hand, for the singlet state symbolically represented by
QIJT)l(t)"-(l)". - (j)" (t)".1 : ol,Jz)lbQ+(0)dQ'+10) - b@+(0)dl)n(0l l0>,
(4.208)
we havel

(u+ ry uu)"inrr." : 0, ,, o,(-j-i-t


;fu(0, (i)

L -' -' -'r(--?i-,,(;)


- lz\")to.o.o
_ /.tL. (4.20e)
-N

fNote that this is one place where the minus sign in a(t)(p) : -u(4'(-p) is necessary
and important.
4-4 TWO.PHOTON ANNIHILATION AND COMPTON SCATTERING 221

To summarize, we get
Itr,ol?"ipr"t : 0,

I tr ,r l3i.,sru, : 6/T 1Zm@)' ea l2k.(e (o') x 6taz) ) lz


: (2eaf mz)ll - (e("'r'6taz))z]. (4.210)

These results are, of course, in agreement with the low-energy limit of (4.197) since
+ > l.r/ilrol': tl.4pl?,,or", + +1./Znl?nrru,. (4.2tt)

Symmetry considerations. We have shown that as p+ 0, (a) the triplet state is


-
forbidden to annihilate into two photons, and (b) the singlet state can annihilate
into two photons provided €(4r). 6@z) I :[ l. We shall now demonstrate that these
features follow from very general symmetry considerations and are, in fact,
independent of the perturbation approximation we have made. Let us first list
possible two-photon states with zero total momentum. On the assumption that
k (: kr : -kr) is in the direction of the positive z-axis,6(ar) 61 €(o') must lie in a
plane parallel to the xy-plane. For photon I the two polarization vectors e(r) and
e(2) can be chosen to be unit vectors in the positive x- and y-directions respectively.
For photon 2 with momentufl -k, e(r) can again be chosen to be in the positive
x-direction, but then e(2) must necessarily point in the negative y-direction, because
the negative z-axis, the positive x-axis, and the negative 7-axis taken in that order
form a right-handed coordinate system. (If the reader is confused, he may look
at Fig. 4-8.) In other words, we use the phase convention

.rrr1k) : €(')(-k), €(2)(k) : -612)(_k), (4.2r2)


where we have explicitly exhibited the dependence of e(") on the photon momentum
k or -k. Two-photon states in which the polarization vectors are parallel must be
linear combinations of the following two states:
a[,,a11,,l0), a[,ra+-y,rl0). (4.213)
Likewise, states in which the two polarization vectors are perpendicular can be
formed out of a[,ra\^,r10) and a[,ra\y,rl0); it turns out, however, to be more

z-axrs z-axls

€(2)(_k)

€(2)(k) y-axis €(1)(_k)


/-axls

(a)

Fig. 4-8. Choice of polarization vectors: (a) photon l, (b) photon 2.


COVARIANT PERTURBATION THEORY 44

convenient to consider the following orthogonal linear combinations:

ft{o1,,olr.,,
* a[,,alp,,) lo), (4.2laa)

I '* r
(4.2t4b)
-L1t{o[,'olr,, - afl,ral1,') l0).
The two-photon states (4.213), (4.214a), and (4.214b) all satisfy Bose-Einstein
statistics automatically because the a+ commute.
The two-photon states we have constructed have defininte transformation prop-
erties under parity. First recall that the transverse electromagnetic field transforms
AS

A(x, r) t -A(-x, l) (4.2ts)


under parity. If we go back to the plane-wave expansion of the field operator
(2.60), we readily see that with the phase convention given by @.212), the above
transformation can be accomplished if the individual creation and annihilation
operators transform under parity as follows:
o[.*T Z(-I)"a!y,., au,,?(-l)"a-u,. (a : 1,2). (4.216)
An immediate consequence of this is that (4.213) and (4.214b) are even under
parity but (4.214a) is odd under parity. Meanwhile, as shown in Section 3-11,
the e-e* system with p* - p- : 0 is required by the Dirac theory to have a nega-
tive parity. It then follows that so long as parity conservation holds in the annihila-
tion process, the two-photon system that results from the p* - p- : 0 e-e* system
must necessarily be represented by a state vector of the type (4.214a) with per-
pendicular polarization vectors. This agrees with our earlier'result based on per-
turbation theory.
Next we shall discuss the consequences of charge conjugation invariance. Neutral
systems such as a single-photon state and an e-e+ state can be eigenstates of the
charge conjugation operation C explained in Sections 3-10 and 3-ll.By this we
do not mean that all neutral systems are C-eigenstates; states such as a neutron and
a hydrogen atom cannot be eigenstates of C, since under C they are transformed
into an antineutron and an "antihydrogen atom" (a positron bound to an anti-
proton), etc. The eigenvalue of the charge conjugation operator is known as the
charge conjugation parity or C-parity. We recall that charge conjugation inter-
changes b(')+(p) 41d lfts)t1p) (cf. Eq. 3.a0a). Consider again & p+ : p- : 0 e-e*
system. The triplet state (4.206) transforms as

b(r)r(0)d(1)*(0) lQ) "h"rge "onj d(r)+(0)b(1)*(0) l0>


: _b(1)r(0)d(rx(0) | 0>. (4.217)

Thus the C-parity of the triplet state is odd. Similarly, we readily see from the
following that the C-parity of the singlet state (4.208) is even:

0 I JT)lb1O+ (0) d?'t10) - b@+ (0) do)*(O)l I 0>


charse c,,ni,
Q I JT)lb(r)+(0) d(2)+(0) _ b@+(o) dQ)*(O)l I 0>. (4.219)
TWO-PHOTON ANNIHILATION AND COMPTON SCATTERING 223

(Note that the anticommutation relation between b+ and dt plays a crucial role in
obtaining these results.) We may mention, without going into a proof, that a
consideration of this kind can be generalized to an e-e* system of a definite orbital
angular momentum (without the restriction p+ : p- - 0); the C-parity of an e-e*
system is given by (-l)'*s where / and S stand respectively for the relative orbital
angular momentum and the total spin (S : I for the triplet, S : 0 for the singlet).
Meanwhile the electromagnetic field A, is odd under charge conjugation since 7,
changes sign, but the interaction density
-j rA, must remain invariant. This means
that the creation and annihilation operator for the transverse electromagnetic field
must change sign since A contains al,. and a1, linearly :
^
+ charge con.i chatce coni
at,* -r-:::> -o[r..o, 6k,o
-Gk,o.
--__-> (4.2re)
An immediate consequence of this is that a state with an even (odd) number of
photons is even (odd) under charge conjugation:
ol,,o,a[^o,... Q{r,,o^10) "hu'e" "oni, (- l)"A{r.,o,QIr*o, . . . AI^,,,10>, (4,220)
no matter how the momentum and polarization of each photon are oriented. Re-
turning now to the problem of two-photon annihilation of a p* : p_ : 0 e-e*
system, we see that
triplet (p- : P+ : 0) * even number of photons,
(4.221)
singlet (p- : p+ : 0) --x+ odd number of photons,
so long as the basic interaction is invariant under charge conjugation. Our earlier
perturbation-theoretic result (4.210) is, of course, consistent with (4.221). As for
three-photon annihilations, the perturbation-theoretic calculations were performed
to order e3 in the amplitude with the result that the p+ : p- - 0 singlet state
cannot annihilate into three photons but the p* : p_ : 0 triplet state can.
L. Michel, L. Wolfenstein, and D. Ravenhall showed that these features are a
special case of (4.221), which follows rigorously from charge conjugation invariance
regardless of the validity of perturbation theory.
It is also instructive to consider the selection rules im- I t
posed by angular momentum conservation. This time it is
A t
simpler to analyze the polarization state of a two-photon # ffi
system (kr : -kr) using the circular polarization lan- I
guage. Since the spin component of the photon along I
I
I

the propagation direction is either *1 or


-1, and the I
orbital angular momentum cannot have a nonvanishing I I
r

component in the propagation direction, we see that the t #


total angular momentum component along k : kr : -kz T *on tr vt.r.
denoted by Jris either 0 or *2. Let us consider the "/, : 6 I I
case in detail. As is evident from Fig. 4-9, the two pho-
tons are
'both
right-circularly polarized or both left- Iig: 4-?' The /s :0
circularly polarized; the two possibilities are represented :;:n:i*
:T;o;XTj$:
by the state vectors V** and Vrr. Consider now a 180" indicate theipin direc-
rotation about the x-axis. We denote this rotation opera- tion.
224 covARIANT pERTURBATIoN THEoRY 4-4

tionby 9'(z). For the field operator we have

Ar(xr, xz, xt, t) a'("1 , Ar(xr, -xz, -xr, t),


*""' , (4.222)
Ar,r(xr, xz, xt, t) -Ar,r(xr, - xz, -xy t).

The individual creation operators of photons with momenta along the positive
or the negative z-axis must then transform ast
_+ @
+ rer) _+ _+ 4 rQt)
Ai,z- _t (4.223)
Ail,r A k,r, A-k,2.

Using this property and the commutation relations fot a+, we get

vnn : Ier JT)(a\,,,, -


I ial,,,)lte u,'/TXaln, - ia\1^,,)1 I 0>
Qtr)
>K-U JT)@tr, - iaf,)llev JT)(alu, , - ia\v,,)] I 0>, (4.224)

that is, Vo* is even under Qr(r). Furthermore, it is a well-known fact that any
system with,f, :0 transforms under proper rotations like the spherical harmonics
Yl(Q,d) or the Legendre polynomial P"(cos d). The transformation property of
Pr(cos d) is particularly simple for I
q'ut '(n):
Pr(cos 0) ,P"(-cos 0): (-l)'&(cos d). (4.22s)

This means that *u*, which is even under frt(n), cannot have J -- 1,3,5,...,
where "I is the total angular momentum of the two-photon system represented by
*o*. The case is similar for V;5. $ Since & J : I system cannot have Jt : {2 and
the two-photon system whose relative momentum is along the z-axis has been
shown to have either .,I, : 0 (Vnn and/or V or J, : 2 (9 and/or Vr*), we have
"") ""
proved that
J:Isystem-+2photons. (4.226)

This far-reaching selection rule (based only on angular momentum conservation)


was first deduced by C. N. Yang in 1950.11 From this general result we see that
the "forbiddenness" of the two-photon annihilation of the triplet P- : P+ : 0
e-e* system follows from angular momentum conservation as well as from charge
conjugation invariance. Table 4-2 summarizes the requirements imposed by the
various selection rules.# These selection rules, of course, apply equally well to
systems having the same C, P, and J quantum numbers as to the e-e* system.

fThe fact that a\u,zdoes not go into -a\u,zis due to the orientation of the linear polariza-
tionvectorsfora*,r10) and a\y,rl0) along the positive y- and the negative y-direction
respectively [cf. our phase convention (4.212).1
gAitually,ifitre two-photon system is to be a parity eigenstate, we must consider (llJZ)
(Vnn * Vz,z). Thg parity considerations, however, are irrelevant in a discussion of the
angular momenturtr selection rules.
llThis selection rule was used to rule out the spin-one assignment for a neutral pion very
early in the history of pion physics.
fsome of the results of Table 4-2 can be obtained just as simply without using the lan-
guage of quantum field theory [cf. for example, Sakurai (1964), pp. 15-16, p. 45]. We have
deliberately used field-theoretic language throughout mainly to illustrate how the trans-
formation properties of the field operators are related to those of the state vectors.
TWO.PHOTON ANNIHILATION AND COMPTON SCATTERING

Table 4-2
SELECTION RULES FOR THE ANNIHILATION OF AN ELECTRON.POSITRON SYS-
TEM WITH F+ : p- : 0
The statement "no by C(P, J)" means that the process is forbidden by charge conjugation in-
variance (parity, angular momentum conservation).

2-photon 2-photon
3-photon
annihilation annihilation
(et"'l ;16(a,)) (e(cr)
annihilation
1 e(4,);

Triplet
(C:-1,J:1) NobyCandJ NobyCandJ Yes

Singlet
(C:fl,J:0) NobyP Yes NobyC

Polarization measurement. Let us now look more closely at the two-photon annihi-
lation of the singlet state. We have already shown that parity conservation requires
that the two-photon system must be represented by the state vector @21a{. Sup-
pose there is a device that is sensitive to the linear polarization direction of one of
the outgoing photons, say photon 1 with momentum k. Equation @.2laa) tells us
that there is equal probability for observing the photon polarization in the x-
direction as in the y-direction. This is not at all surprising since there is nothing to
disturb the cylindrical symmetry about the direction k. We shall next consider a
more sophisticated experimental arrangement in which photon polarizations are
measured in coincidence. Imagine observer A and observer ^B who specialize in meas-
uring the polarization of photon I (momentum k) and that of photon 2 (momentum
-k) respectively. Suppose observer I finds that for a particular annihilation event
the polarization direction of photon 1 is in the x-direction; then, according to
(4.214a), observer ^B must find that the polarization vector of photon 2 has no
component in the x-direction. This is quite a striking result if we realize that the
two polarization measurements can be made at a widely separated distance long
afrler the two photons have ceased to interact. In fact the alleged correlation is
expected even if the two measurements are made at a spacelike distance so that
there is no way of communicating to observer B the result of the measurement made
by observer r4 before observer,B performs his measurement. Since photon 2 appears
unpolarized when the polarization of photon I is not measured, it appears as if
photon 2"gets to know" in which direction its polarization vector must point at the
very instant the polarization of photon I is measured. In other words, having
measured the polarization direction of p[oton 1, we can predict with certainty the
result of measuring the polarization direction of photon 2 (which can be very
far away from photon l), despite the fact that both polarization directions are
equally likely for photon 2 when the polarization measurement of photon I is not
carried out.
If the reader finds all this disturbing, it is only natural. Some of the greatest
minds in twentieth-century physics have worried about this problem. In the liter-
ature a peculiarity of this kind is known as the "Einstein-Podolsky-Rosen paradox."
COVARIANT PERTURBATION THEORY

In his famous autobiography, A. Einstein wrote on this subject as follows:


But on one supposition we should, in my opinion, absolutely hold fast: the real factual
situation of the system S, is independent of what is done with the system S, which is
spatially separated from the former.
To analyze this paradox a little more closely, let us ask what constitutes a
quantum-mechanical system in this particular problem. When we measure the
polarization direction of photon l, the quantum-mechnanical system on which
we make a measurement is not the single-photon state a*," l0> but the composite
two-photon system represented by the state vector @.21a$.In quantum mechanics
it is wrong to regard a two-photon system as two separate single-photon systems
even if the two photons are a million miles apart so that they cannot possibly
interact. Any measurement we make on one of the photons is to be regarded as a
measurement made on the entire two-photon system. When observer ,4 finds
with certainty that the polarization vector of photon I is in the x-direction, what
he actually learns is that the two-photon system is in afl, ,a!t,rl0) rather than in
a[,ra\y,, l0>. Any subsequent measurement observer B may perform must be
consistent with the finding that the system at the time of the first measurement is
represented by the state vector a[,ra\y,rl0); from this point of view it is perhaps
not too peculiar that observer -B necessarily finds that the polarization vector of
photon 2 is in the y-direction. The Einstein-Podolsky-Rosen paradox ceases to be
a paradox once we accept the view that the polarization measurement on photon
1 is actually a measurement to determine whether the composite two-photon system
is in a[, ,a\u,,l0) or in a[,ra\y,, l0>.

/
'/ ,' rotation
:i:::f :r::::i::::::ij:::!:it:i::

Scatterer Scatterer
0
Fig. 4-10. Coincidence measurements on the photon polarizations.

We may mention that there have been experiments to determine whether or not
the photon polarization vectors are correlated according to (4.214a). Following
J. A. Wheeler, we consider an experimental arrangement, schematically shown in
Fig. 4-10, where the annihilation photons are scattered by atomic electrons. As is
evident from (2.170), Thomson scattering is an excellent means of analyzing the
direction of the incident photon polarization; for instance, at 0 :90o, the differ-
ential cross section varies as sin2 f, where { is the aximuthal angle of the scattered
photon direction measured from the incident polarization direction. Since the
energy of the I ray is as large as the rest mass of the electron, it is better to use an
analogous expression based on the Klein-Nishina formula (3.344); however, it is
not difficult to show that the strong dependence on the polarization persists even if
E, = m. By measuring the coincidence counting rates using the arrangement
shown in Fig. 4-10 and then repeating measurements with an arrangement where
4-4 TWO-PHOTON ANNIHILATION AND COMPTON SCATTERING 227

the right-hand half of Fig. +-10 is rotated 90o around the propagation direction
k, we can test whether there is, in fact, the perpendicular correlation indicated by
(a.2Iaa). The experimental results of C. S. Wu and I. Shaknov have fully confirmed
that there indeed is such a correlation effect of the annihilation gamma rays as
predicted by (a.2l4a).

Positronium lifetime. We shall now go back to the expression for the B*


-,0 limit
of the total cross section (4.202) to show how this quantity is simply related to the
lifetime of an e-e* bound state, known as positronium, whose existence was estab-
lished experimentally in a series of elegant experiments performed by M. Deutsch
and coworkers in the early 1950's. We first recall that ororu*, which has the dimen-
sion of volume divided by time, characterizes the reciprocal of the positron mean
lifetime when there is one electron per unit volume. So
R: o61'U*p (4.227)
is the reciprocal of the positron mean lifetime when the electron density is p. For
instance, for positron annihilation in a medium in which there are N atoms of
atomic number Z per unit volume, the p that appears in @.227) is NZ. When we
want to calculate the lifetime of a bound e-e* system, the electron density (relative
to the position of the positron) is just the square of the bound-state wave function
evaluated at the origin; for instance, for the ground state of the positronium we get
p : lt"(x : 0) l' : lllrr(2ao)3f, (4.228)
where we have used the fact that the Bohr radius of the positronium is twice that of
the hydrogen atom. In practice, since the electron (positron) velocity in the bound
system is of the order of r-fr times the speed of light, we can use limB.-e o1s1 i1
(4.227) to compute the positronium lifetime. For the ground state we get
I(n : l, ts 2ry) : lim 4 o$;r'otrr* | 9,,(* : 0) l'
-

=]gY@hn
: $d.sm, (4.229)
where the factor 4 is due to the fact that the singlet state does the whole job in
(4.211). This gives for the mean lifetime of the n : | 15 state,
Tsingret :21(aum) 1.25 x 10-10 sec, (4.230)
=
as first calculated by J. Pirenne. The'S state which is forbidden to decay into two
gamma rays is expected to be much more long-lived since the amplitude for three-
photon annihilation must involve another power of e, as is evident from Fig.
4-ll. Without evaluating the complicated diagrams shown in Fig.4-ll we may
guess that the lifetime of the 3,S state must be about 137 times longer. The actual
answer (due to A. Ore and J. L. Powell) turns out to be
Ttriplet 9o
== 1115 (4.231)
rsinsret M;:qa-
for the n : l,s wave bound states.
COVARIANT PERTURBATION THEORY 4-4

Fig. 4-11. Examples of Feynman diagrams for three-photon annihilation.

Cross section for Compton scattering. Earlier in this section we obtained the til -
matrix element for Compton scattering as well as that for two-photon annihilation
(which we discussed in detail). It is interesting to note that the tr-matrix elements
for two-photon annihilation and Compton scattering (Eqs. 4.166 and 4.175)
are related by the following simple substitutions:
krr r(o') + -k, €@), kr, ,(or, --> kt , e@'),
(4.232)
P--->P, P*--> -P''
As for the free-particle spinors we must interchange u(")(p*) and t('')(p') [as well as
z(-t(p-) and u(')(p)1, but this change is automatically taken care of (apart from
a minus sign) by @.232) when we compute 1..y'/ 7r,1'z by the trace method; this is
because
_iny.p*+m_-iny.p'+m (4.233)
zm-zm
Comparing Figs. 4-6 and4-7 ,we see that this set of transformations simply amounts
to changing the external lines corresponding to the (incoming) positron and one of
the (outgoing) photons, photon 1, which appear in Fig. 4-6,into the external lines
corresponding to the outgoing electron and the incoming photon which appear in
Fig. 4-7.In other words, all we need to do is "bend" some of the external lines.
Furthermore, simplification completely analogous to (4.187) and (4.188) takes
place if we compute the -til-matrix element for Compton scattering in the lab
system. For this reason X,," I ft rol'for Compton scattering can be readily written
once we have the expression (4.197) for two-photon annihilation. Apart from the
sign change due to (4.233), the only other change we must note is a factor two
coming from
i2**Z
s+ 5s'
(4.234)

since initially there are just two fermion spin states. So

*p ltrr,r:#[,,t* *_ 2a47et"''e<"'r;'] . (.23s)

In the lab system the relative velocity is just unity (in natural units), and in com-
puting the phase space factor, we must note that
+
0(E' _ at
-Z([FT-: - c'l cos I -rr rr _
-----E7-
ruo')
: Er7'
ma)
(4.236)
4-4 TWO.PHOTON ANNIHILATION AND COMPTON SCATTERING 229

k, e(")

Fig. 4-12. Bremsstrahlung.

which is not difficult to obtain if we note the kinematical relations


E': ,
-,t _ ma (4.237)
ur
-m a r11 - gss01'
Putting the various factors together, we get the famous Klein-Nishina formula
(3.344).If the incident photon is unpolarized, and if we do not observe the polar-
ization of the outgoing photon, we can sum over the final polarization states and
average over the incident polarization directions; this can be done using the
techniques we discussed in connection with Thomson scattering (Section 2-5).
The net result is

#:+(*)'(#* *-sin'd) (4.238)

Note that because ar' is not equal to or (unless ol K m) but is dependent on the
scattering angle via (4.237), the angular distribution is no longer forward-back-
ward symmetric as it is in Thomson scattering. Formala (4.238) has been checked
at a variety of energies. Already in the late 1920's W. Friedrich and M. Goldhaber
had measured the angular distribution of ry-rays with .8, : 0.17 m (^, : 0.14 A; to
show that the observed deviation from the Thomson formula is precisely what one
expects from formula (4.238). The agreement achieved here is one of the earliest
quantitative triumphs of the Dirac theory.

Bremsstrahlung and pair production. To finish this section we briefly mention two
other processes which can be treated using the electron propagator: bremsstrahlung
and pair production (in the Coulomb field of a nucleus). This time the A, that ap-
pears in the S-matrix expansion can be split into two parts,

Au: A? + Atq), (4.23e)


where Alf it simply the Coulomb potential (4.56) considered in connection with
Mott scattering and Ap is the quantized transverse field mentioned in Chapter 2.
For bremsstrahlung there are two topologically inequivalent graphs shown in
Fig. 4-12, where the symbol x stands for a first-order interaction with the Coulomb
potential. The essential point is that the annihilation of the incident photon in
Compton scattering is now replaced by a first-order interaction with the Coulomb
potential. As in the Mott scattering problem there is no over-all momentum con-
230 COVARIANT PERTURBATION THEORY

servation (as far as the electron-photon system goes) sincq the nucleus (assumed
to be infinitely heavy) can take up any arbitrary amount of momentum. For this
reason the S-matrix element cannot be written in the form of a four-dimensional
delta function times a covariant matrix element ,til. However, the S-matrix
element is just as simple; the reader is urged to work out the details (Problem
4-10a) to obtain

s.r, : -i2ttZe3 ffiE1r,'* r - E)m,-*_W


k)
'--]t6v'€'TrY'- t!*.p(o) +
x t(")(p')1,.-'r'(p N.F@ ==iry
.(p' f^k) +-mr.lz(,)(n)
(4.240)
Once we have this expression for bremsstrahlung we can readily write the matrix
element for pair production corresponding to Fig. 4-13 using substitutions analo-
gous to (4.232). The formulas for the cross sections of bremsstrahlung and pair
production were both derived in 1934 by H. A. Bethe and W. Heitler; they can
be found in Heitler's book along with detailed comparison with experiments.l
In Problem 4-10(b) and (c) the reader is asked to work out the cross section for
bremsstrahlung when the incident electron is nonrelativistic and consequently
the energy of the emitted photon is small compared to m.

Fig. 4-f3. Pair production (in the Coulomb field of a nucleus).

We may note that in practice the Compton scattering cross section must be
multiplied by Z (the number of atomic electrons) while the pair-production cross
section varies as 22. At low energies it is also important to consider the photo-
electric effect which is due mostly to the ejection of K-shell electrons. According to
Problem 2-4 the photoelectric cross section for a hydrogen-like atom varies in-
versely as the fifth power of the Bohr radius, hence the photoelectric cross section
due to inner shell electrons is proportional to 25. As for the energy dependence,
the photoelectric effect diminishes rapidly as a-T/2 for o K m (which can also be
seen from Problem 2-4 if we note that lkr l2 is proportional to ar) and as l/or for
@ >m. The Compton cross section takes the constant Thomson value for o K m
(but is still large compared to the binding energy of an atom) and then drops
c,r
rather rapidly as o exceeds m; integrating (4.238) over angles, we obtain the total

fHeitler (1954), pp. 242-268.


4-5 FEYNMAN'S APPROACH TO THE ELECTRON PROPAGATOR 23t

Photoelectric Compton Pair production


effect effect dominant
dominant dominant

-0.01 0.05 0.1


hu, MeY
Fig. 4-14. Relative importance of photoelectric effect, Compton scattering, and pair
production.

Compton cross section


ocn,,,p : orhnm
+#("t* * +) , cD > m. (4.241)

It is evident that at high energies the pair-production process is the most important
one. More quantitatively, Fig.4-14 (taken from R. D. Evans's Encyclopedia article)
illustrates which of the three processes is dominant in a particular energy region
for a given value of Z.

+5. FEYNMAN'S SPACE.TIME APPROACH TO THE ELECTRON


PROPAGATOR
In the previous two sections we showed how to compute the matrix elements for
some simple processes using the language of the second-quantized Dirac theory,
In this section we shall demonstrate how the results we have rigorously derived
from quantum field theory can be obtained more quickly by relying on a more
intuitive method, due to R. P. Feynman, who based it directly on the particle wave
equation of Dirac. We present this alternative approach in some detail since it
provides an intuitive understanding of the essentials of the field-theoretic results
derived in the previous two sections.

Green's function. Since Feynman's approach is based directly on the wave equation,
we first review the way of solving an inhomogeneous differential equation in
mathematical physics. Take the trivial example
V'd(*) : -p(x). (4.242)
As every "child" knows, the solution to Poisson's equation (4.242) with the
boundary condition f(*) ----- 0 as lx I * oo is given by

d(x) : J
O(*, x')p(x') d'x', (4.243)
232 covARrANT pERTURBATToN THEoRy 4-5

where
(4.244)

The function G(x, x') is called the Green'sfunction for Poisson's equation and satis-
fies
V2G(x, x') : -5t:)(x - x'). (4.24s)
In other words, instead of solving (4.242) directly, we first solve the unit source
problem (4.245); once we obtain the solution to this simpler problem, the solution
to the more general problem (4.242) can be written immediately because of the
superposition principle.
In relativistic electron theory we are interested in solving the Dirac equation

(r,*- + m)^h : ie,v, A,nlr i (4.246)

here (and throughout this section) rf is a c-number wave function, not the quan-
tized electron field. The right-hand side of (4.246) is very much like the source
p(x) in (4.242).In analogy with electrostatics we may argue that it is too difficult to
tackle (4.246) directly; instead, let us first consider the simpler unit source problem;

(rr&+ *)K(x, x'): -i8(4)( x - x'), (4.247)

or, writing the Dirac indices explicitly, we have

[(,y)"B@l7xr) I mS,BlKBr(x, x'):


-i8(4)(x - x,)8or. (4.249)
The function K(x, x') is the Green's function for the free-particle Dirac equation
just as G(x, x') is the Green's function for Poisson's equatio.n. In complete analogy
with (4.243) we expect that a solution r/n to the more complex differential equation
(4.246) satisfies

f(") : -J K(x, x')e,y rAr(x,)*(x,) dax,. (4.24e)

That this is indeed the case can be proved by direct substitution:

(rr*+ ^)[-J "," , x')e,y*Au@)*{*') o^*'f


x' ) e'v' A'(x', ) "r(x' ) d a
x'
:!,':,::; (4.2s0)
Note, however, that even if we add to (4.249) any solution to the free-wave equation
(e : 0), the differential equation (4.246) will still be satisfied. To be more explicit,
let us consider

9(") : to(x) - I *@, x')e,yrAr(x')*(x') dnx', (4.2st)


where 9'(x) is a solution to the Dirac equation when e : 0. We can easily verify
that (4.251) satisfies $2aQ just as well. In a scattering problem, for instance,
may represent an incident plane wave which would be present even if there
^/rr(x)
were no interaction.
4-5 FEYNMAN,S APPROACH TO THE ELECTRON PROPAGATOR 233

There is an important difference between the differential equation in electrosta-


tics (4.242) and the differential equation for the Dirac electron (4.246). The right-
hand side of (4.242) does not contain d; so we could write the solution (4.243)
immediately in a closed form. In contrast, in the case of (4.2aQ the function rf
itself appears on the right-hand side; as a result, (4.251) is an integral equation.
So, even if we obtain an explicit form of K(x, x'), as we shall in a moment, we cannot
write the exact solution to (4.2aQ in a closed form. However, if a perturbation
expansion in powers of e can be justified, we can obtain an approximate solution
to (4.246) accurate to any desired power of e by the now familiar iteration method:

9(") : to(x) + t an*' K(x, x')l_.e.yrAr(x')lgo(x')

+ ! an x' t dn r" K(x, x')l-ery rAr(x'))K(x', x")l-ey,,A,(x")l*o(x")


+ I d^*' I dnx" t dnx"' K(*, x')f-e,yrAr(x')lK(x', x")
X l- e,y, A,(x")lK(x", x"')1,- ey1A1(x"')]./ro(x"') + - . . (4.2s2)
.

Our next task is to obtain an explicit integral representation of K(x, x'). First,
because of translational invariance (in both space and time)it is clear that K(x, x')
is a function of x - x' only. So without loss of generality we can set x' - 0. We
solve (4.247) using the four-dimensional Fourier-transform method just as we
solve (1.36) using the three-dimensional Fourier-transform method. Let us define
K(p) such that
K(x,0) : lll(2r)^1[ fr1p1rre'r d|'p. (4.253)

If we substitute this expression for K(x,O) in(4.2a7) with x' : 0, we get

ltlQr)al J fiv. p i dk(p)e'o'' dnp : -r5rar(x)


: -lil(2tr1n1 t e*'' dnt. 6254)
So the differential equation (4.247) is now reduced to a simple algebraic equation
in momentum-space,
(i^t.p*m)K(p):-i, (4.2ss)
or
k@):-il?,v.p*m), (4.2s6)
where ll!'y. p * m) is understood to be a 4 x 4 matrix such that if we multiply
with (i.y . p + m) fromthe right or from the left, we obtain the 4 x 4 identity matrix.
Clearly we can write (4.256) as

. i(-i,y.p*m)
y' - _ (i"r.p
Rfrl: _-i.y.P*m (4'257)
+ m)(-i"y.p + m): ffi'
since (i,y- p)' : -p2. Going back to coordinate (x) space, we get, using (4.253),

K(x,0) : -6Io^o++Pe'o" (4.258)


234 covARrANT pERTURBATToN THEoRy 4-5

or' more generalll


x,)=: _ (-i"y. p * nt)
x,
& I nro jafi-' -ip.<r_,,t (4.2se)

The integrand of (4.259) is already familiar from the previous section. When
: X.E: + JlpP + *,
we integrate along the real po-axis, there are poles at po

K(x, x'): -
6 [ o'o eip'(x-x') Il_lt,ffi . 6260)
The particular form of K(x, x') depends on the particular manner in which we go
around the poles in the complex po-plane. That this kind of ambiguity exists is to be
expected on physical grounds, since we have not yet specified the boundary condi-
tions to be used in connection with the differential equation(4.247).

(a) Kr"1(x, x/) (b) Kp(x, x/)

Fig. 4-15. Prescriptions for the integration (4.260) in the complex po-plane.

K. and Ku.. We shall now consider two types of K(x, x') corresponding to different
boundary conditions. First, we shall look at what is known as the retarded Green's
function with the property

K,.,(x,4
{1 3, ','.',, (4.261)

In the complexpo-plane, K.", corresponds to the choice of contour indicated in Fig.


a-15(a). When t) t', we must close the contour in the lower half-plane; hence
we pick up the contributions due to both poles.In contrast, when t I t', we receive
no contribution from either pole. Alternatively we may use the contour indicated in
Fig. 4-15(b). It is evident from Fig. 4-5 and Eq. (4.159) that K(x, x') then coincides
with the expression (0lf(rf(x)fr("')) l0> : -(*)S"(x - x'), discussed ex-
tensively in the previous section. The Green's function obtained by following the
prescription shown in Fig. 4-15(b) is denoted by Ko(x, x'). In other words

(4.262)
We already proved in the previous section that if we integrate in the clockwise
direction around po : +.E, we obtain (4.154), while if we integrate in the counter-
clockwise direction around po: -E, we obtain (4.157). For K,", with t > t'
4-5 FEYNMAN'S APPROACH TO THE ELECTRON PROPAGATOR 23s

we must go around both E and E in the clockwise direction. We therefore obtain


-
Kn(x, x')
f lim) (mlEv)lu(s)(p)t(s)(p)stn'@-t't + o(')(p)D(s)(p)e-'p'(r-r')l for t ) t',
Jt--- P,s
-- ),
[0 for t' > t. (4.263)
In contrast, for K" we have the old result

:= m [ztsrlp);<'r(p)sre'@-"') for t ] t'' (4'264a)


Kr(x, x,)
lT i,Ztr
tt-r,',,nrrrs)(p)s-ir for t, > t. (4.264b)
.@-r't1

Note again that po in (4.259) through (4.262) is, in general, not equal to E
: whereas P, in(4.263) and (4-264) is identical with E
"/T|FW,
The burning question is: Which of the Green's functions, K,"" or Kr, is the
appropriate one to use in connection with the perturbation expansion (4.252)?
In the Schriidinger theory we can show that the correct Green's function is the one
that vanishes for > t' t; this is evident intuitively because the wave function at
t can be influenced by the presence of a potential acting at / if / is later than t',
but not if I is earlier. (This point will be clearer when the reader works out Problem
4-l l.) So we may be tempted to use K*, in the relativistic electron theory as well.
We shall now demonstrate, however, that the use of K*, is incompatible with
the hole-theoretic interpretation of the Dirac theory.
Let us consider a potential problem assuming the validity of first-order pertur-
bation theory. If K in the perturbation expansion (4.252) is taken to be K,*r, we
have, with the aid of (4.263),

lh(") =,lo(x) + t an r' Kn (x, x')l- e'v rAf)(x')l$o(x')


: rho(x) + >p' X clil,(l)t'@V u@(!)ere"el
2

s'= I

+> i
p's/:1
t[l],(l)t,{mW 1s<s')7Pr1r-in"n1, (4.26s)
where

"[{],(r) - -e I o'*' I'__0, "/@V ,-tn"n'trG)(p)ry,.,1r0( x')Af)(x'),


(4.266)
c[.!,(t) - -e
!o'r' I'_*or ,6W rin"r'r(s')(p)ryr.iho(xt)Af;i(xt).

Let us assume, for definiteness, that rfo(x) is the normalized wave function for a
positive-energy plane wave characterized by p and s. According to the usual inter-
pretation of wave mechanics, lc[{],(r) l'zgives the probability for finding the electron
at t in a positive-energy state characterized by p', s' (assumed to be different from
p, s) when the electron is known with certainty to be in state p, s in the remote past.
In fact, from Section 4-3, we see that c[i],(*) is precisely the first-order ^S-matrix
for the transition of an electron characterized by (p, s) into another state character-
ized by (p', s') (cf. Eq. 4.64). This is reasonable because a positive-energy electron
state is known to be able to make a transition into some other positive-energy
state in the presence of an external potential. On the other hand, c$ll,(r) is much
COVARIANT PERTURBATION THEORY 4-5

more disturbing. Recalling that ?r(s)(p)e-ip'" is a negative-energy wave function


because of the relation (cf. Eq. 3.384),

0(t,r)(p)€-tp.o Trrn,t,(-p)ercp).x-i(-E)t (4.267)

(where E is positive), we see that for a large value of t there is a finite probability
for finding the electron (initially in a positive-energy state) in a negative-energy
state. (See Problem 3-12 for a specific example of this kind.) Thus the use of K".
leads to the catastrophic result that in the presence of an external potential a posi-
tive-energy electron can make a transition into a negative energy state, in violent
contradiction to the hole theory.
Let us now use K, in the perturbation expansion (4.252). To first order we get,
with the help of (4.264),

9(") - to(i + t an *' K,(x, x')l-e,y*Af)(x')l"lro(x')

p/ s/:l

+> i
p' s/: I "lll,(4t
JWV al)(pt)e-ie'.nf, (4.265)
where

c[tl,(t) - -e I O'*' I'__0,' "W ,-rn"r'6G)(p)ryrfo( x')Af;)(x'),


(4'269)
cfll,(r) : *e o, *,
I Ii or, ,@ ,ip",'0$')(p,)rr+o(*,)A?(x,).
Note that this time the limits of the t' integrations are different for c(+) and c(-).
Everything is satisfactory as /
-r oo. Letting .ilro stand again for a normalized
plane-wave solution characterized by p and.t, we see that c$il,(*) is identical to the
,S(t)-matrix element for the scattering of an electron initially in (p, s) into another
electron state characterized by (p', s'), just as in the case of (4.266). However, the
undesirable feature of (4.266) is now absent; c[;J, does go to zero as /----- oo when
K" is used. Physically this means that the positive-energy electron cannot make
an energy-conserving ("real" as opposed to "virtual") transition into a negative-
energy state even in the presence of a time-dependent external potential. Thus the
catastrophic transition into a negative-energy state is now forbidden.
On the other hand, the behavior of (4.269) &S /----+ -oo is somewhatpuzzling.
Although cf{J, goes to zero as /----+ -€, c[t],(-oo) does not, ingeneral, vanish.
As a result, *(t: -oo) is not equal to is quite strange; in fact, as we
^fo.This
shall show in a moment, its ultimate physical interpretation requires a new
postulate concerning the meaning of the Dirac wave function. Fortunately some of
the results obtained in Section 4-3 help us to understand the physical meaning of
c$ll(-oo); we note that, apart from a minus sign (which is not significantf), c[;J,

tHadweused d+b+|0) rather than b+d+ l0) as the initial state in discussing e- * e+
vacuum (Section 4-3), the S-matrix element (4.114) would have been opposite in sign. -
4-5 FEYNMAN'S APPROACH TO THE ELECTRON PROPAGATOR 237

(--) is nothing more than the first-order S-matrix element (4.114) for e- * e*
-----) vacuum in the presence of Af
when fo is set equal to a normalized (positive-
energy electron) plane wave and (p', s') is identified with the momentum-spin
index of the annihilated positron. This encourages us to identify
cS},@JWVr<''t1pt)e-ip"o as the wave function for an available empty negative-
energy state (with -p' and r' : 5 - s') into which the incident electron is capable
of making a transition in the presence of Af;); as the time goes on, such a state
becomes rarer and rarer because of pair annihilation, and, as /----- oo, c[1J,, in fact,
goes to zero. Alternatively we may identify I c[;J,(r) l2 as the probability for finding
a positron in state (p', s') which will certainly be annihilated by the time t - q.
To compute the S-matrix for pair annihilation, we simply go backward in time and
ask, what is the probability amplitude for finding, at t : - @, the positron which is
doomed to be annihilated at later times. The nonvanishing of cf;J, for large nega-
tive values of /, which ffi&y, at first sight, seem rather puzzling, is seen to be quite
welcome if we realize that the external potential can not only scatter electrons
but also annihilate electron-positron pairs.
In order to fully appreciate the physical significance of each term in (a.268)
and (4.269) it is instructive to examine the properties of the integral operator

I o'*' Kr(x, x'),yn (4.270)

acting on a Dirac wave function whose space-time coordinate is x'. Suppose the
time ordering is "ordinary" in the sense that t > tt. When the operator (4.270) acts
on a free-particle positive-energy wave function, we simply get the wave function
for the same momentum-spin state evaluated at a later time:

I o'*' K(x, x')rynJ@ u@(p)ete',


: 2 @lE'v)rs)(p')u@+(p')6iW a(')(p) ,'o"' xt e-ip"t'+ip'r'
F I nd3
: JmW r@(p)sto'r, (4.271)

where we have made use of the orthogonality property of the free-particle spinor
(3.106). On the other hand, when the integral operator @.270) acts on a negative-
energy wave function (4.267), it obviously gives zero so long as t > tt [recall
ut')+(-p)a(s)(p) : *r.rt''r+1-n)4G o' n,(-p) - 0]. Similarly, when the time
ordering is "opposite," that is, l' > t, the negative of the integral operator (4.270)
acting on a negative-energy wave function gives the wave function for the same
negative-energy state evaluated at an earlier time; when (4.270) with r' ) / acts
on a positive-energy wave function it gives zero. To summarize:

!o,,,Ko'(x,x,)"ynIffi:,-'',,E]:,:,;:.,I:{%,::],;,,,:::,u,,,,,-,.
(4.272)

In other words, the integral operator (4.270) carries positive-energy solutions


forward in time and negative-energy ones backward in time.
238 COVARIANT PERTURBATION THEORY 4-5

Returning now to the first line of (4.268), we first note that


-e,yrn[o(xt)Af;i(xt)
represents the interaction of the incident electron with the external potential at x'.
As a result of this interaction, we expect both positive- and negative-energy waves
originatingat x'. Looking into the future, the Green's function Kr(x, x') with t > {
carries positive-energy electron states forward in time. Looking into the past, the
Green's function Kr(x, x') with t < { carries negative-energy electron states
backward in time. But we argued earlier that the probability amplitude cfl for
such a negative-energy state represents the probability amplitude for the corre-
sponding positron state which is doomed to be annihilated at later times. We can
therefore visualize pair annihilation as the scattering of a positive-energy electron
into a negative-energy state which propagates backward in time (Fig. 4-l6b) in
complete analogy with ordinary scattering (Fig. 4-l6a). The positron which is to
be annihilated at x' is simply an electron "scattered backward to earlier times."
From this point of view a negative-energy electron going backward in time is
equivalent to a positron going forward in time. This equivalence which was
suggested in the previous section by examining the propagation of a virtual electron
(positron) is now seen to make sense also for a real electron (positron).

- etpAl!)
yr: (x/, it/) (x/, it/)

x: (x, it)

Fig. 4-16. Graphical interpretation of (4.268): (a) ordinary scattering (t ) t'), (b) pair
annihilation (t < t').

So far we have treated electron scattering and pair annihilation in detail. Con-
sidering in (a.268) to be a negative-energy free-particle wave function, we can
^/r'(x)
formulate positron scattering and pair production to first order in the external
potential in an entirely analogous manner. In connection with electron scattering
and pair annihilation we have seen that, with r/.'(x) taken to be a positive-energy
wave function, the use of K" (rather than K.".) ensures the boundary condition
that in the remote future r/n has no component of the type p3)r-ia't.In the for-
mulation of problems of positron scattering and pair production, the correct
boundary condition which we must use is that in the remote past r/. has no com-
ponent of the type z(') eip'' i this is because neither in positron scattering nor in
pair production should there be a positive-energy electron going forward in time
at t : With {no taken to be a negative-energy wave function, it is easy to see
-co.
that the use of K" automatically guarantees the above boundary condition.
More generally, to determine the wave function at all times in the Feynman theory,
it is necessary to specify the positive-energy components in the remote past and the
4-5 FEYNMAN'S APPROACH TO THE ELECTRON PROPAGATOR

negative-energy components in the remote future. This is because the positive-


energy electron propagates forward in time while the negative-energy electron
propagates backward in time. In contrast, we may recall that the Schrtidinger
wave function can be completely determined at all times once we know the wave
function in the remote past.

Feynman's approach to Compton scattering. The utility of Feynman's view that the
electron paths can "zigzag" forward or backward in time can be fully appreciated
when we consider more complex problems. For this reason we now examine the
second'order term in the perturbation expansion (4.252). For definiteness consider
the case of Compton scattering. The Ar(r) now represents the equivalent time-
dependent potential, discussed in Chapter 2, corresponding to the emission and
absorption of a photon (ll,W)ef)s-tt''" for emission and (U,J'2&)ef)rt*''
for absorption. We look at (4.252) to second order with rfo taken to be the wave
function that represents the incident positive-energy plane wave characterized by
(p, s). With Kset to Ko we get

t(") = JmW + t an *' t dn *" Kr(x, x')l-e,yr(U,\/T;nef)


y{s)(p)ein'e sit'',1

x Kp(x' , x")l_ery,(I\ffi)ef,) s-ix"'"llJmW v{s)(p)stn'a"1


L rI k <----> k'l(
- lo *- a'l' (4.273)

where we have omitted a term linear in e which cannot contribute to scattering.


As /--)oo, we can replace the first of the two Green's functions, K"(x,x'),by
(a.264a) since / is necessarily later than t'. Thus
y:, yts)(p)ein.x + > ) c$ll4*)t\/mffi uo')(p)ere'.r1, (4.274)
f(") "/W
where

c[i]4*) : I On x'
I dn x" (,r/Wry ;rs'r(pr) e-ie"r')l-e"y*el nQrV)ulf) ero',|

x Ko(x' , x")l-ey,(U,t6T) el"') s-i're"t"l6@ u{st(g')stn'r")


( k <---> k'1t
-L lo
)
*- a'l'
(4.275)

When we recall that Kr.(x', x") is equal to (0 lT(h@)fi(x")) l0), the coefficient
c[i](-) is seen to be completely identical to the second-order S-matrix (4.174)
for Compton scattering which we derived in the previous section, using the language
of quantum field theory. That we can derive the correct matrix element in such a
simple manner is the main advantage of Feynman's space-time approach.
In Feynman's language we can read the various factors in the integral (4.275)
from the right to the left as follows. First, the incident electron characterized by
(p, s) emits a photon (k' , d') at x" . Subsequently the electron goes from x" to x' ; if
the time ordering is "ordinary," that is, l' > t",then the first form of Kt given by
COVARIANT PERTURBATION THEORY 4-5

@.26a$ ensures that a positive-energy virtual electron goes forward in time, as


shown in Fig. a-fi@). On the other hand, if the time ordering is opposite, that is, .
t" > tt, then K"(x, x') carries a negative-energy virtual electron originatin g at x"
backward in time to x'; this is shown in Fig. 4-17(b). Meanwhile, using the field-
theoretic formalism, wo have already shown that (0lZ@n(x')il@,,))10) rep-
resents the propagation of a positive-energy electron from x" to x' when the time
ordering is "ordinary" and the propagation of a positive-energy positron from x'
to x" when the time ordering is "opposite." Comparing the two points of view and
recalling that K"(x', x") is identical with (0 lT(^h@)fi(x,,)) l0) we see once
again that negative-energy electrons going backward in time represent positrons
going forward in time. Finally at x',the virtual electron (positive energy or negative
energy depending on whether t' > t" or t' < t") absorbs the incident photon
(k, a) and goes on as a positive-energy electron (p', s') propagating forward in time.
As we emphasized earlier, a great virtue of covariant perturbation theory is that
we can combine the terms corresponding to two different time-ordered graphs,
Fig.4-17(a) and (b), into a single simple expression.

Kp(x, x/)

(a) t/>t // (b)trt 2tr

Fig. *17. Examples of time-ordered graphs for Compton scattering. Note that (a) and
(b) together correspond to a single Feynman graph (Fig. 4-7a).

We have seen that the results which we rigorously derived in the previous two
sections from quantum field theory can be written rather quickly when we use
Feynman's space-time approach. There is no doubt that the calculational steps
become shorter and simpler when done in Feynman's way. This is perhaps not so
surprising since Feynman's method which is based directly on the covariant wave
equation of Dirac does not conceal the relativistic invariance of the theory at each
state of the calculations. In contrast, since the customary Hamiltonian formalism,
which was used to derive the calculational methods of the previous two sections, is
based on the SchrtrdingerJike equation (4.24) which singles out the time coordinate
in the very beginning, we must work somewhat harder to arrive at the same coyar-
iant matrix element (for example, by combining separately noncovariant pieces
in a clever way).
In this text we first demonstrated how to derive the matrix elements for some
simple processes, using quantum field theory, and then showed how the same
matrix elements can be obtained much more quickly when we use the space-time
approach of Feynman. It is true that we can hardly overestimate the simplicity and
+5 FEYNMAN'S APPROACH TO THE ELECTRON PROPAGATOR 24I

the intuitive appeal of Feynman's approach. We must, at the same time, admit,
however, that the logical structure of Feynman's method is somewhat more in-
volved. This is because a formalism based on the single-particle wave equation is
used to describe processes such as pair creation, pair annihilation, and even vacuum
polarization, which involve more than one particle at a given time. In fact, the
careful reader might have already noted that the physical interpretations of (4.268)
given go beyond the usual interpretation of wave mechanics in which the square of
the Fourier coefficient represents the probability for finding a single particle in the
state in question at a given time. For this reason we prefer to regard Feynman's
graphical method as a convenient pictorial device that enables us to keep track of
the various terms in the matrix elements which we derive rigorously from quantum
field theory. Our philosophy closely parallels that of F. J. Dyson who in his 1951
lectures at Cornell University made the following remarks:

In thiscourse we follow the pedestrian route of logical development, starting from the
general principles of quantization applied to covariant field equations, and deriving
from these principles first the existence of particles and later the results of the Feynman
theory. Feynman, by the use of imagination and intuition, was able to build a correct
theory and get the right answers to problems much quicker than we can. It is safer
and better for us to use the Feynman space-time pictures not as the basis for our cal-
culation but only as a help in visualizing the formulae which we derive rigorously
from field theory. In this way we have the advantages of the Feynman theory, its cor-
rectness and its simplification of calculations, without its logical disadvantages.

Historically it was S. Tomonaga and J. Schwinger who first recognized the


importance of the interaction representation for performing covariant calculations
in quantum field theory. Using this representation, they were able to isolate in an
unambiguous way the finite (observable) part of the radiative corrections to the
basic ry, vertex (a topic to be discussed in Section 4-7). Subsequently, in two famous
1949 papers, "Theory of Positrons". and "Space-time Approach to Quantum
Electrodynamics," R. P. Feynman developed a set of "rules" which are based
directly on the Dirac and Maxwell equations but not (explicitly) on quantum
fleld theory; using his "rules" he could obtain the correct results much more quickly
and efficiently than Tomonaga and Schwinger. Meanwhile, starting with the
S-matrix expansion of the field operators (as we have done in this book), F. J.
Dyson demonstrated that the Feynman rules are direct and rigorous consequences
of quantum field theory and that the approach of Tomonaga and Schwinger is, in
fact, mathematically equivalent to the simpler approach of Feynman.
Whether the reader prefers to derive the Feynman rules rigorously from quantum
field theory or more intuitively from the wave equations i la Feynman, there is no
doubt that the graphs and rules which Feynman taught us to use have had profound
influences on a number of areas of physics-not only quantum electrodynamics
and high-energy (elementary particle) physics but also nuclear many-body prob-
lems, superconductivity, hard-sphere Bose gases, polaron problems etc. In less
than two decades the graphical techniques of Feynman have become one of the
most indispensable and powerful tools in attacking a broad class of problems in
theoretical physics.
242 COVARIANT PERTURBATION THEORY +6

4-6. M9LLER SCATTERING AND THE PHoroN pRopAGAToR; ONE-MESON


EXCHANGE INTERACTIONS

Scalar meson exchange. In the previous two sections we learned how to compute
second-order processes which involve virtual electrons and positrons or, more
generally, virtual spin-] particles. In this section we shall discuss processes which
involve virtual photons or, more generally, virtual bosons. (Processes that involve
the propagation of both a virtual fermion and a virtual boson at the same time will
be treated in the next section.) Our main discussion centers on the electron-electron
interaction brought about by the exchange of a single "covariant photon." Let us,
however, first treat the simpler problem of the exchange of a spinless meson.
We shall consider a spin-| particle of mass rn which we call a 'onucleon"; the
corresponding field, denoted by 9(x), is assumed to satisfy the Dirac equation and
to be quantized according to the rules of Jordan and Wigner. In addition, we shall
consider a spin-zero neutral particle of mass p which we call a "meson." The
field associated with the meson, denoted by d(x), is assumed to satisfy the Klein-
Gordon equation and to be subject to the usual quantization rules (cf. Problem
2-3).The "nucleon" and the "meson" interact via Yukawa coupling:
ffrnr:GWO. (4.276)
We do not claim that our "nucleon" and "meson" bear any similarity to the nucleon
and the meson observed in nature; as we emphasized in Section 3-10, the observed
low-mass mesons are pseudoscalar mesons which do not admit a coupling of the
type (4.276) because of parity conservation in strong interactions.
Let us consider the elastic scattering of two (identical) nucleons to order G2
in the amplitude. Initially we have two nucleons characterized by (p,, s,) and
(pr,sr); the two final nucleons are characterized by (pi, s{) and (pj, sj). Since
there is no meson in the initial state or in the final state, the initial and the final
state are the vacuum state so far as the meson operators are concerned. The
second-order matrix element ("form 1") reads

s!,J : eic), I orr, I,,.,,daxr(Nl, NLl0,,@r)gB(x,)0 ,(xr)ga(x,) I N,, /r,)


x (1)"B(l),u(0 ld(x,)d(x,) I 0), (4.277)
where
I N', AI,> : btbl l0>, I Ni, Ni> : bi+ b'L+ l0> (4.278)
with the obvious abbreviation'bl : D('')+(p,), etc. Although the ordering of
bi+ and bL+ is somewhat arbitrary, with the choipe @.278) we get s(0)
- I when
state 1 and state 2 coincide with state l' and state 2/ respectively. In (.277),
{r^h
is written as .7"9u(l),B so that our method here can be immediateiy generalized
to the ryr(or the ryu) coupling case. This time, since there are two nucleons in the
initial state, both 9("') and rf(xr) must be used to annihilate the incident nucleons;
similarly both fr("') and fr(x') must be used to create the outgoing
-nucleons.
The terms containing 'P'r+r and gt-r make no contributions since there are no
antinucleons to be annihilated or created. Hence we can replace the product of
+6 MOLLER SCATTERING AND THE PHOTON PROPAGATOR 243

the fermion operators in (4.277) by

g t a@ ) {,(x,) g(",) ---+ .}!- )(x,) f!* )(x,) fr !- )(",) )(",). (4.27 9)
"(x,) ^h5*
Note that if we are computing the S-matrix to order G2, there are no virtual nucle-
ons; every one of the fermion operators in (4.277) is needed to annihilate and
create the real nucleons in the initial and the final states. As'for the meson opera-
tors, we can obviously replace (0 ld("')d(x,) l0) by (0 ld(+)(x')d(-)(x,) I 0),
where 0(+)(d(-)) is the annihilation part (the creation part) of the meson field:

d(*' : 4 #akeik'',
(4.280)
d,-r:4#ale-ik.,
witho:Jw.Physically(0|d(+)(X')d(_)(x,)l0)meanSthatavirtual
meson is created at x, and subsequently (note tt ) t) annihilated at xr, as is clear
from our experience with the fermion case in Section 4-4.We shall say more about
this vacuum expectation value later.
Coming back to the fermion operators, we now move r|nt*) to the right and
.h(-) to the left so that the product of the annihilation operators may immediately
annihilate the nucleons in the initial state:
.|!- )(x,) .l!. )(",).7f )(x,) rf$* )(x,)
: -{t|-)(x,).7t-)(x,)tb*)(x,)^h5*)(x,) * .l!-)(x,)[tl.)(",), fr!-'(",)],,h5*)(x,).
(4.281)

It is not difficult to convince oneself that the last term of (a.281), which is just
.7;-'(r,)f5*)("r) times a c-number function of xr and xr, does not contribute to the
scattering problem in which we are interested; it just gives rise to the amplitude for
a process in which nucleon I and nucleon 2 proceed in space-time independently
of each other even though one of them suffers a self-energy interaction.t Our next
task is to evaluate acting on ll/', 1Vr> from the left and
"hl.'("').h5*)(xr)
the right. Let b6 and bp1 stand for
-.p!-)(x,)0!-'("') acting on (N{, Ni I from
typical annihilation operators contained in the fields ft")("') and gr+r(xr) re-
spectively. So far as the creation and annihilation operators are concerned, a
typical term arising from lf', Nr> looks like
^l'1.)("')fi*)(xr) I
b 1qb Plbr bI to'

-
iffi#ifii$''il' " (4.282)

where we have used formula (3.414). Similarly we can compute the Hermitian

fThe diagram corresponding to this is an example of what is known as a disconnected


graph.
244 COVARIANT PERTURBATION THEORY 4-6

conjugate of
-rp!-)( x')Fl-)(x,) acting on (i{{, N! | from the right as follows:

l-biDbiDbir b''r l0>l* : [(8r,rrl8r.<r,r - 8<,,r,,8,.ryr) I 0>]t, (4.283)

where bis and bizl stand for typical annihilation operators contained in [r]!-)(x,)l+
and [r]f)(xr)l+ respectively. The physical meaning of (4.282) can be seen as follows.
The operator Tfnt+r(xn) can be used to annihilate either nucleon 2 or nucleon l;
if r/r(*)(xr) annihilates nucleon 2, then ft*'(",) must necessarily annihilate nucleon
l, and vice versa.

(olo[f,\o[,-]lol
--__ ____
1014$1,4[/,101

(a) (b)

Fig. 4-18. Examples of time-ordered graphs in nucleon-nucleon scattering (t, > tr).

Let us focus on the case where th(*'(xr) and 9t+r1x,) annihilate nucleon 2 and
nucleon 1 respectively; this corresponds to the second term of the last line of @.282).
Looking at (4.283) we see that there are still two possibilities: (a) .[<-t1xr) creates
nucleon 2', and 4.t-t(xr) creates nucleon l', or (b) .F(-)(x2) creates nucleon l', and
qt-r(xr) creates nucleon 2'. These two possibilities can be visualized by drawing
time-ordered graphs as in Fig. 4-18, where we have recalled that
(0 l{(.t(x')d(-)(xr) l0) gives the propagation of a virtual meson from x, to x1.
It is very important to note that (4.283) tells us that we associate a minus sign with
Fig. 4-18(b). From our experience with the scattering of two identical particles in
nonrelativistic wave mechanics we recognize that this is precisely the minus sign
associated with an antisymm etrized final state as demanAeO Uy thi Pauli exclusion
principle. Our formalism based on the second-quantized Dirac theory automatically
gives the full force of Fermi-Dirac statistics. We also note that no other factors such
as JT, etc., are needed; the only other thing we must keep in mind is that in obtain-
ing the total cross section we integrate over only half of the solid angle_s, just as in
the case of two-photon annihilation (cf. Eq. 4.201). We must still consider the
case where annihilates nucleon l, and rfnb*)(",) annihilates nucleon 2
^/n[*'(xr)
[cf. theirst term of @.282)1. However, since we are integrating over both x, and xr,
we may as well make the interchanges x1 € xz, dB +_> ryE so that 95*)("r) annihi-
lates nucleon 2, and rhl.)(x1) annihilates nucleon l, just as before. We must, of
course, keep in mind that Ir is now earlier than t, ("form 2"). Noting that the
product of the first term ot @.282) with (1) <+(2) and(4.283) with (1,) * (2,) is
MOLLER SCATTERING AND THE PHOTON PROPAGATOR 245

equal to the product of the second term of (4.282) and (4.283), we see that the
S-matrix (4.277) becomes

S!'i : (-iG)' I on *, I no *,
X lulrs-in;'arurriarr')(ilLe-ro;'rzsrgtnz'u) - (uLe-tri'tturrtnrtt)(ille-tP'rrzuzeinz'1271
x [(014(*)(x,)0t-r(x2)10)d(r, - h) + (0ld(.)(x,)dt-r(xr)10)0(t, - rJ].
(4.284)
From (4.284) we observe that the problem is now reduced to that of evaluating
the vacuum expectation value of a time-ordered product which we newly define as
follows:
(0 lr(d(x)d(x')) I 0) : l0(x)d(4 I 0) for t } t"
o) for r > ,1. .
{(0 (4'28s)
l(o ldrod(x) I

Unlike (4.152), the T-product of two boson field operators is defined without any
minus sign even when the time ordering is "opposite" (t' > t); in the ,S-matrix
occurring in practical problems the various field operators not used in annihilating
or creating real particles always arrange themselves in such a way that we have the
combination (4.153) for fermions and the combination (4.285) (or its analogs)
for bosons. Clearly (4.285) is equal to
(o ld(")d({10) : (0 ld(.)(")d(-)(x/) | 0>
: l'g iA - k'.x')l(Tlauaf,,l0)
E hexp [i(/c.n

: (2o)' JI ryexp
2a [ik.(x - x') - iol(t - t')] (4.2s6)

for t ) t'. Similarly for---L-


t' 2 t, (4.285) reduces to
:
(0 I d(x')d(x) I 0) (0 ld(.'("')d(-)(x) | 0>
:# I#exp [ik'(x - x') - i(-a)(t - t'))' (4'287)
Again, just as in the fermion case [cf. (4.154) through (4.162)), we get, forboth
t> / andt' > t,
(o lr(f(x)d(",)) I o> : -& I o^o
F+;+ (4.288)

with
k2: lkl, - lkol,, (4.28e)
where ko is a variable, not, in general, equal to ar : \4WT7. Having studied the
fermion case in detail, we can easily understand the physical meaning of (a.288)
(known as the meson propagator). When / is later than t' a virtual meson created
at x' propagates from x' to x and then gets annihilated at x; when l' is later than
t, a virtual meson propagates from x to x'. Thus the combination of the vacuum
expectation values indicated by (a.285) automatically keeps track of the correct
"causal sequence" of a virtual process in which a meson is emitted by one of the
nucleons and is subsequently absorbed by the other nucleon, a point emphasized
COVARIANT PERTURBATTON THEORY

by E. C. G. Stueckelberg prior to the advent of Feynman's theory. We again


mention that the notion of r being later or earlier than ttis not a Lorcntz invariant
concept when x and x' are spacelike and that the meson propagator does not vanish
outside the light cone (x - x')' > 0l; yet the combination (4.285) is Lorentz
invariant as is evident from its explicit Fourier integral form (4.288). In the litera-
ture one often defines A"(x
- x') so that
to,@- xl: - 6[o^o#" (4.2e0)

The previously defined S" is related to A, via the expression

lS"(x - x')f,p: (r,#,- *),uoo(x - x') (4.2et)

with m replacing p in the expression for Ao.

I
Kp.nF+A-a (

(a) (b)

Fig. 4-19. Feynman graphs for nucleon-nucleon scattering: (a) direct scattering,
(b) exchange scattering.

Substituting (4.288) in(a.28Q and integrating over xl and xr, we easily obtain the
explicit form of the covariant ,4-matrix defined, as usual, by

Sr,:8ru- i(2r)aE(n)(p' Ip,-pi-ot, frrr.


(4.2e2)
To order G2 we getfrom (4.284)

-ifr ,r: vt"lrco,


(-iG)2 | ==,
(uiu,-J)(4Lur\ .= ("Ly:)@i"r\ I
,=-
-V11'1 p' - iel ilu., - pi)' + p' --l-d)'
@'2e3)

just obtained (4.166) from (4.163) in the problem of two-photon annihilation.


as we
The various factors in @.293) can be read immediately from the Feynman graphs
Fig. 4-19 where, in contrast to the time-ordered graphs Fig. 4-18, the virtual
meson is emitted either at x2 or at x, (and subsequently absorbed at xr or at xr),
depending on whether tr) tz or t2> fr. Again the "rule" for writing
-i.ril is
remarkably simple. Just as we associate (-i,y. q + m)lli(q' * m' - tu)l with each
virtual spin--| fermion of four-momentum 8, we associate llli(q' * p'- ie)l with
each virtual spinless meson of four-momentum g. Recall that in computing the
energy-momentum of the virtual particle we assume energy-momentum conserva-
4-6 MOLLER SCATTERING AND THE PHOTON PROPAGATOR 247

tion at each vertex. (One may argue that the four-momentum of the virtual meson
in Fig. aa9@) irp, - p', (: pL - pr) or pi -p, depending on whether the meson
is emitted by nucleon I or nucleon 2;but in the boson case this is of no importance
since the meson propagator is a function of 42 only.) Because of the scalar coupling
(4.276), we have (-iG) at each vertex in place of the (-e,y) appropriate for the
electromagnetic interaction. In addition, we multiply the result by + I for Fig.
a-D@) ("direct" fermion-fermion scattering) and - I for Fig. 4-19(b) ("exchange"
fermion-fermion scattering).
Before we discuss the more complicated problem of photon exchange, we shall
briefly outline two alternative methods for deriving the ,.//-matrix element (4.293).
First, we show how we may have guessed it from semiclassical considerations with
a nonquantized meson field. According to Yukawa's idea, the preseilce of nucleon
1 generates a ("classical") meson field which we denote by dtt'. Whentheinteraction
Lagrangian is given by the negative of @.276), the field d(') satisfies the inhomo-
geneous Klbin-Gordon equation

Ed"'(x) - tn'+"'("): Gs(')(x), (4.2e4)


where rrtr(x) is the scalar invariant constructed out of the initial and final wave
functions for nucleon l, viz.
s(r) - @ttt, u, gt<n'-o"t''. (4.2es)
We solve the differential equation (4.294) assuming that the space-time dependence
of 0tt) is also given by et<n'-ni)'". We then get

d"'(") : -6=fi av^ffitt'|1 u, sr{n'- n;t'r (4.2e6)

Since we are interested in the interaction between nucleon I and nucle'on 2, we


compute the effect of nucleon I on nucleon 2. The Hamiltonian for the interaction
of nucleon 2 with the field {(1) (generated by nucleon 1) to lowest-order is

Hint: c J O"'{")s(2)(x) dBx

c J O,',{") (4.2e7)

But from the S-matrix expansion (4.36) we see that (-i) times the time integral of
11,n. is to be identified with the (S
- 1) matrix element to lowest order. Therefore
we get, to order G2,

s!'J : -(- ro''


* I d' , ffiVexp [i(p, * p, - pi - pl)'x]
: (Zr)a g,n,(p, * p, - pi - pi)

" ffi, g.2sl)


248 COVARIANT PERTURBATION THEORY 4-6

in complete agreement with the first term of (4.293) which represents "direct
scattering." We can obtain the second term of (.293) by taking into account the
antisymmetrization rule for the final-state two-fermion wave function.
There is yet a third method for obtaining(4.293) which goes as follows. As in the
first method, we use this time a formalism in which the meson field is quantized, in
the sense that we talk about the emission and absorption of a virtual meson; how-
ever, the ,S-matrix is computed according to the rules of the old-fashioned second-
order perturbation theory based on energy-difference denominators. The scattering
process is visualized as taking place in two steps. First, the initial state lN,, Nr>
makes a transition into either of the following two types of intermediate states:
(a) an intermediate state made up of a virtual meson, nucleon l, and nucleon 2'
and (b) an intermediate state made up of a virtual meson, nucleon 1', and nucleon 2.
The intermediate state of either type then makes a transition into the final state
lNl, N). The existence of two types of intermediate states, of course, arises from
the fact that there are two time-ordered graphs corresponding to the single graph
in Fig. a-D@); type (a) intermediate states arise when nucleon 2 first emits a
virtual meson which is subsequently absorbed by nucleon 1, as represented by the
time-ordered graph Fig. 4-18(a), whereas type (b) intermediate states arise when
the virtual meson is emitted by nucleon I and subsequently absorbed by nucleon 2.
According to the rules of Chapter 2 the emission and absorption of a meson is
equivalent to an interaction of the nucleon with the time-dependent (scalar) poten-
tials ,,rhlTaV st*'a. \ll!1en nucleon 2 emits a meson of momen-
"@e-ik'& and
tum k, energy a, to become nucleon 2' (without any accompanying change in nu-
cleon l), we associate with this process a time-independent matrix element

f/i,";i*1 : G
[,dt*alu,st$z-oL-kt'a
: G fi!2u2V go,,o;*y, (4.zee)

where co satisfies the energy-momentum relation

Similarly, when nucteon , to become nucleon t', we


associate with this process "or;;#*eson
F{lo'):G A'1u1V6p,*1r,o1. (4.300)

In addition, the rules of the old-fashioned perturbation theory tell us that we must
divide the product of the perturbation matrix elements by the "energy denomi-
nator," which is
(E,*E,)- (8,* El)-^
for a virtual process involving a type (a) intermediate state. In this way we obtain
4-6 MOLLER SCATTERING AND THE PHOTON PROPAGATOR 249

the effective Hamiltonian matrix element between the two nucleons in the form
f1f,1us).iletnis; f/r(6rs),Pte'"is;
?( (E' * Er) (E' * ED + (E, t Er) (E', * Er) a
- - ^' - -
:
#"'*"'n"*oiil'1u'it!2u2

"lz;h+4+-,]
: (2rt)r8,r,(p, * p, pi pj)G zillurillu,
- -
"'.[L2a(E,
I - I l. (4.301)
- Ei - or) ' 2co(E1
- El - cr)J'
where we have let V --+ oo. The s(2) matrix element differs from the above expres-
sion simply by a factor (-2ri) times D(E * Ez E|
- - ED.' We now observe
that the bracket in the last line of (a.301) can be reduced to a remarkably simple
form. Taking advantage of over-all energy conservation, we can replace E, Ei
by -(^E', E).Hence
-
-
or-( Er- + (n *(E,-Ei
I
-w, -Ei - ar) e)t E
E'r)
I
:- I
lk l' +
- (E, -
tL' E'r)'
I
(4.302)
(pr-Pi)'+ s,"
Apart from a factor i this is precisely the propagator for a spin-zero meson in
momentum space. We thus see that (4.301) times
-2zti D(8, * Ez - Ei - Ej) is
equal to the direct-scattering term of the',S(2)-matrix element computed earlier
using the covariant method (cf. Eqs. 4.292 and 4.293). Note once again that by
combining two noncovariant expressions we have succeeded in obtaining a single,
more concise, covariant expression.
The above calculation illustrates the basic differences between the (prewar) old-
fashioned method based on energy denominators such as E,
- Ei - ar and the
(postwar) covariant method based on relativistically invariant denominators such as
(p, - p)' + p'. ln the old-fashioned method momentum is conserved at each
vertex, as indicated by the Kronecker deltas in (4.299) and (4.300). However, the
intermediate state is not assumed to be an energy-conserving one; indeed, it
is precisely the energy difference between the initial and the intermediate state
that appears as the energy denominator. On the other hand, in the covariant
method, where energy and momentum are treated on the same footing, we get
automatically both energy conservation and momentum conservation at each
vertex because of the appearance of a four-dimensional integral of the type
tdax expli(p, - pi - k)'xl. The exchanged virtual meson, however, is notvisualizid
250 covARrANT pERTURBATIoN THEoRy

(in the sense of a Feynman diagram) as satisfying the energy-momentum relation


6 : nffi' instead, we have, in general,
k'+ p': lkl'- lkol'* ti +0. (4.303)

Indeed, it is precisely the difference between kz : (pt - p)'z and - p'2 that appears
as the denominator of the covariant meson propagator in momentum space. From
this example we can appreciate the real advantage of covariant perturbation theory
in which we insist that the space and the time integrations be performed simul-
taneously.

Matrix element for Msller scattering. We are now prepared to make an attack on
our main problem-the scattering of two relativistic electrons, commonly known as
Mlller scattering after C. Mlller, who first treated this problem in 1931. First,
recall that the basic Lagrangian density of quantum electrodynamics is

st : -ir*r,, - fi(r,fr+ *)+ | ie{r.yugA,. (4.304)


Using the techniques of Appendix A applied to the Dirac electron case, we can
deduce from this Lagrangian density the total Hamiltonian as follows:
H-Ho*H,, (4.30s)
where

Ho: t IqE(r) 12 * lB,r, y)d'x + J Vtv.v + m)gd3x, (4.306)


and

Ht : - tu [frvf.A'' ) d3 x+
+ I n' r, S a' r,(fi'#)r-*:9it)-"'
(4.307)
BV (.I^yrf)xr,r wo mean fr(*', t)rynf(xt, r); the superscript (I) stands for "trans-
verse." In this method only the transverse part of A, is treated dynamically, and
the basic interaction (4.307) is seen to consist of the interaction of the transverse
electromagnetic field with i :
irfi^yrlr and the instantaneous Coulomb interaction
between charge densities. We quantize only the transverse part of Ar according
to the method of Chapter 2 so that

AtD(x) : p, ayoetk', 1 e\4 a[.e-t*'n). (4.30s)


#? frfrl",
When the Dirac field is also quantized, both il^y* and {nyn* are to be understood
as operators properly antisymmetrized so that (0lQ.y-lh l0) vanishes (cf. Eq.
4.59). The S-matrix expansion of Section 4-2 can now be carried out with f1,
taken to be (4.307).
Because the interaction Hamiltonian consists of two terms, the problem of
the interaction between two electrons is a little more involved than that of the
nucleon-nucleon interaction brought about via the exchange of a scalar meson.
4-6 MOLLER SCATTERING AND THE PHOTON PROPAGATOR 251

Fortunately the first term of (a.307) is seen to be completely analogous to the space
integral of the scalar coupling density (4.276). Therefore, so far as the exchange of
transverse photons is concerned, everything we have done with the nucleon-nucleon
interaction via one-meson exchange can be carried out just as well for the Mlller
interaction provided that we replace the identity matrix by ,y, and the coupling
constant G by -ie. Since the steps analogous to (4.277) through @.28a) are
essentially the same, we can immediately write the analo g of @.28a)

Sftrun*r:(_ e)r lOn*, IOn*,


X - t o i' n
(ilt2 e - in i' a z' n't,
l(tt', e "' ry
r tt r rt " "') z
r . u 2 et n

(u|e-ro6'ur.uretn''r,17il1e-to|',,r.uzernz'rt)1<0lf@tL)(xr)A|D(xr))10>
-
(4.30e)
where the Z product of transverse electromagnetic fields is defined as in (4.285).
Let us now evaluate the vacuum expectation value of the Z-product. First, if t > t',
we get

<0lT(AlLt(x)A\tt(x')) I 0> : (0 lAltt(x)A1l(x/) | 0)


: + ? ? 4 Z ffi 'r)
eta') ei
&'' - *'''' t (o av'a[''' o)
I I

:1,
Vi 4 *u["'r5o) e'r'(n-r') (4.310)

Comparing this with (4.286), we note that the only new feature is the polarization
sum ) zo:t efo) ej"). This summation can be readily performed by recalling that
6(t),6(2), and k/lk I form an orthonormal set of unit vectors. Suppose we define
€(3) - k/lk l. (4.31l)
Clearly e[") with dF 1,2,3 can be regarded as an element of the orthogonal
matrix that transforms three unit vectors along the usual x-, !-, and z-axes into the
new set 6(r),6(2), and e(3): k/lkl; hence we expect to obtain the orthogonality
conditionf

X t[")e-!o) :
a:l
E'i (4.312)

or, equivalently,
2

)
q:l
ef")6,('r - 8ri - k&jllkl, (4.313)

The remaining steps in the evaluation of the vacuum expectation value are com-
pletely analogous to (4.286) through (4.288). We get

<0|T(A'L)(x)A!Ll(x,))l0>:hJo^o(u,,-,ffi)ffi.(4.3l4)
fSee, for example, Goldstein (1951), pp. 97-98.
2s2 COVARIANT PERTURBATION THEORY

Inserting (4.314) into (4.309) and integrating over dax, and dnxr, we obtain (to
order e2) that part of the covariant -6-matrix which is due to the exchange of
transverse photons as follows:

-i//,'i^,u:ffi,,1l@,,^yu,)(u,,,yu,)-(u,,,t.iu,)@L,t.Qu,)1

#%KaLv
u,)'(il',^v uz) - @L^t'tt' ur)(ili^t' u,)f, 8' (4.315)

where
4:Pr- P't:PL- Pr, Q' : Pr - PL:Pl- Pr, (4.316)

and fl and {' are unit vectors along q and q' iespectively. This is as far as we shall
go with the exchange of transyerse photons for the time being.
Let us now look at the second term of (4.307), which represents the instantaneous
Coulomb interaction. If we are computing the electron-electron interaction to
order es in the amplitude, we need to consider this term only to first order. Using
the rule for writing Stt', we get

slQoou : -'+ I " I o' *, I 4' r,(e'',


ell(F''n:L)*o{fut9)*,ll t', e,)

: -'+ I on*, I on*,8(r, - ,,>

(4.317)
for the Coulomb contribution. The manner in which the fermion fields operate on
the initial and final states is the same as before except that this time we need not
worry about whether /1 is earlier or later than t, since the whole interaction is instan-
taneous. The factor ] just cancels with a factor 2 arising from the fact that$(xr) can
annihilate electron 1 as well as electron 2 (cf. Eq. 4.282 and 4.283). Hence (4.317)
becomes

s!!o'rr : -rr' lonr, Ion*r#!;''],,


X tn i'
r r, etp'' n') X (ttl s- t n i' n')
[(Ai s- " ry
n " ry ny rstn "'

(frLe-tet'"' nlnurrtn''",)(iltre-inl'rz, nuzernz'a211. (4.3 l8)


-
Let us look just at the direct term since, apart from a minus sign, the exchange term
is obtained from the direct term by interchanging the final-stite labels pi<-> p!,
ri <-> sl. For the space-time dependence, it is convenient to rewrite the product
of the plane waves as follows:
exp(-ip''.xr * x, - ipl. x, * ipr. xr)
ipr.
: exp l-i(pL - p)'(x, - x1)l exp li(p' * Pz - p',
- p)'x'1. (4.319)
Since the dax, integration is over all space-time, instead of integrating over dnxr,
we can integrate over daxr, where xg : x2 - xr. The /r-integration just kills the
8-function, whereas the dsxr-integration gives the three-dimensional Fourier trans-
form of the Coulomb potential, lllpl - p, 12. The dax'-integration gives (2rt)n
times the four-dimensional D-function that appears in the definition of the til-
4-6 MOLLER SCATTERING AND THE PHOTON PROPAGATOR 253

matrix element. Thus, we obtain

- iu// g.',) : e1@ffi - <otlr'yniY-'1,iy tl


: (-"),[ry#@ - au)lut'v +u)f. (ul'v
@320)
We should emphasize at this stage that neither -4111"') nor flfr""t), taken
separately, is covariant. We shall now show, however, that when we sum the two,
we end up with a compact expression which is not only covariant but also much
simpler. To see this explicitly, we first rewrite the direit-scattering term of (a.320)
as follows:

1- ry,(ui"Y n ry.)l/-,'Y n u,) : (-e)'(itti.yp)(il|,y q l' - lqol' - iefi(q'z I


^",)ll t' )6-14
- ie)
:: (-r)'I
-(tri"ynu,)(ilI
nl+ur) -TqT-
+
(iq)' (,ui,yau)(ilLrynur)
i(q'z
- ie) lq l' i(q'z
- ie) l
(4.321)
Meanwhile, because of the properties of the free-particle spinor (3.105), we have
uiry'qu': -tfiili,y'(p' - pDlu' :0 (4.322)
or
iq oil'r'Y aut : - ul^l' Iur. (4.323)
Similarly, remembering (4.316), we get
iqoilLrynltz : -u'r^f 'q'uz. (4.324)
We therefore find
(iqr)' (tt'r,y aur)(u'r,y aur) : (ili^t . gur)(il|^t . eur), (4.32s)
and(4.321) now becomes

7- ryr(ri'v nryllwry nu,) : (- e),@WW I (- e)rffi


(4.326)
This leads to a remarkable result; the second term of @.326) cancels exactly the
second term in the first bracket of (4.315). A similar cancellation also takes place
in the exchange-scattering amplitude. Hence

-i,// 1t : -i(,tr(t.ra.s) 1 .tr11."t))

:- ((-e)' (ui"yrur)(u'rryrur) (u|"yrur)(ili,yrur) f


--rrflffi- ift;=Ef=id @.327)

The final result (4.327) is simple and covariant; so is the basic Lagrangian (a.30a)
of quantum electrodynamics with which we started. This suggests that a somewhat
awkward formalism has been used to derive the simple matrix element (4.327).
Indeed we could actually have guessed (4.327) much more quickly from semi-
classical considerations analogous to (4.294) through (4.298), that is, by arguing
that the presence of electron I induces a ("classical") four-vector potential lyt
254 COVARIANT PERTURBATION THEORY 4-6

satisfying (in the Lorentz gauge)

aAfP : -ie@nlry rurst(nr - ni)'x (4.328)


which interacts with electron 2.It was precisely this kind of reasoning that led C.
Mf,Iler, in 1931, to write the matrix element for electron-electron interaction.
The quantum-electrodynamical derivation given here is essentially that of a 1932
paper of H. Bethe and E. Fermi except that we have used the perturbation-theo-
retic formalism based on covariant denominators rather than that based on energy-
difference denominators.

Covariant photon propagator. It is also possible to derive (4.327) more elegantly by


constructing a manifestly covariant quantum theory of the Maxwell field. The
result (4.327), when compared b (a.293), strongly suggests that this can be accom-
plished if we quantize not just the transverse parts of A, but also its longitudinal
and fourth components. Instead of starting with the formalism of Appendix A in
which only the transverse parts of A, are treated as dynamical fields, let us suppose
that we can take all four components of Arto be four independent HermitianKlein-
Gordon fields of zero mass subject to the usual rules of field quantization. We
expand A, as follows :

At,:0lJT) I c=r
Note that the a indices now run from I to 4. The polarizationvectors are so chosen
that
e(1,2) _ €o,2)* : 0) With ilo,2) I k, fi(r) I fi(z).
(ii{r,r,r
e(3) - €(3)* - (k/lkl,0); e(4) - :
-€(4)* (0,0,0, i) (4.330)
Note the orthogonality relation
4

2 ,li'rlf'*
a:l
:6pu. (4.331)

The creation and annihilation operators are assumed to satisfy the following:

lau,o, a[,,"): Drk,Do*,, lau,o, ak,,o,f : Ia[r,,,, alr,,.f :0, (4.332)


just as in the transverse case. We may call a[,, and a[,nthe creation operators for a
longitudinal and a timelike photon respectively. In this formalism we forget about
the instantaneous Coulomb interaction; the electron-electron interaction is
visualized as taking place via the exchange of four types of photons-two transverse,
one longitudinal, and one timelike. To compute the -ril-matrixelement for Mlller
scattering all we need to do is evaluate <0lr@p(x)A,(x'))10). This can be
easily accomplished because of (4.331) and (4.332); we obtainl

(4.333)

fln the literature one often defines Dr(x - x') by

* o,r* - x,) : -# I W: lT !n,r* - x,).


4-6 MOLLER SCATTERING AND THE PHOTON PROPAGATOR 255

which is called the covoriant photon propagator. Using this, we can immediately
write the covariant matrix element (4.327). Thus, to obtain the covariant result, it
is simpler to use a formalism in which all four components of A, are quantized.
There are, however, two important features which we overlooked in the previous
paragraph. First, in treating all four components of A, as kinematically independ-
ent, we have completely disregarded the Lorentz condition; actually this turns out
to be not too serious because the source of the electromagnetic field is conserved.
Second, we have assumed that An@) is a Hermitian field operator, whereas we know
from the correspondence with classical electrodynamics that the expectation
value of An@) must be purely imaginary. To overcome these difficulties a rather
elaborate formalism has been developed by K. Bleuler and S. Gupta to justify
what we have done in the preceding paragraph. The proposal of Bleuler and Gupta
is, among other things, to modify the definition of a scalar product in Hilbert space
by introducingthe notion of indefinite metric (negative probabilities); as a result,
(A^) becomes imaginary despite the hermiticity of the field operator An. We
shall not discuss this elaborate method in any further detail since we have already
learned how to arrive at the correct covariant matrix element (4.327) by using
the alternative noncovariant method. A good discussion of the Bleuler-Gupta
formalism can be found in many books.f

photons photons
(instantaneous) (two types) (four types)

a) (b) (c)

Fig. 4-20. Exchange of "covariant photons."

In any case it is convenient to visualize the covariant matrix element (4.327)


as representing the exchange of four kinds of "covariant photons." Looking
back at the Coulomb contribution (4.326), we can say that the so-called instantane-
ous Coulomb interaction is actually equivalent to an interaction due to the emission
and subsequent absorption of a time-like photon and a longitudinal photon, con-
sidered together. Taking into account the possibility thatthefour types of "covari-
ant photons" can be exchanged between two electrons, we can completely forget
about the existence of the instantaneous Coulomb interaction. This is indicated
pictorially in Fig. +20. From now on by a photon we will mean a covariant
photon.
It can be shown that this equivalence of the instantaneous Coulomb interaction
and the exchange of a timelike and a longitudinal photon (taken together) holds

fSee, for example, Mandl (1959), pp.57-71; Kiill6n (1958), pp. 181-204.
COVARIANT PERTURBATION THEORY +6

not just for this particular problem of electron-electron interaction treated to


order e' in the amplitude but also for any problem in quantum electrodynamics
so long as the source of the electromagnetic field satisfies the continuity equation.
Thus we can add to our previous list of "rules" for writing -i// the following
prescription:
In a process which involves the propagation of a covariant photon, just insert
8*"/[i(q'- ,e)] (4.334)
in the same way as we insert llli(q' * p'- tu)l for a spinless meson and
(-i"y.q I m)lli(q' + m2 - ie)l for a spin-f fermion.
We have argued that a virtual photon can be visualized as having four states of
polarization. On the other hand, we know that a real photon, or a free photon,
has only two states of polarization. At first sight this may appear disturbing; after
all the so-called real photon should be the limit of a virtual photon where lq I

approaches 40. As a matter of fact most real photons of physical interest are,
strictly speaking, virtual in the sense that they are emitted at some place and
absorbed at some other place. To analyze this feature for the particular problem of
electron-electron interaction, we simply note that as lqlapproaches lqol, the first
term of @.321), which represents the exchange of a timelike photon, tends to cancel
the second term of @.321), which, because of @.325), represents the exchange of
a longitudinal photon; this cancellation becomes complete when the photon be-
comes real, that is, when I q I : lqo l. If one wishes, one may say that the photon
always has four states of polarization regardless of whether it is virtual or real,
but that whenever it is real, the timelike photon and the longitudinal photon always
give rise to contributions which are equal in magnitude but opposite in sign.

Cross section for Msller scattering. Let us now compute the cross section for Mlller
scattering. The connection between the differential cross section and the ,,y'/-matrix
element can readily be obtained as in the case of two-photon annihilation. Quite
generally, for any elastic fermion-fermion scattering we have in the cu system
1 I (4mrmr)2
(%)"-: 4 (4tt)'z E?o,
I // r,l', (4.335)

which is a special case of (4.182). If the initial electron beam is unpolarized, and if
we are not interested in observing the final spin states, we must sum over the
initial and final spin states and divide the result by four. Our task is to compute
a, b, and c in
I t.r .L/
.LJ
+srsi
^ E2 s; \pr=pF-r(p,t-pff-ffi
(4.336)
For the coefficient a we have
I
O:j +s1sis2si
: ffrtf("-)(p,)ry"A("-)(pi)"yrl
Tr [A(-)(pr)ry,A("-)(p'r),y ul, g.337)
L6 MOLLER SCATTERING AND THE PHOTON PROPAGATOR 2s7

where we have used the abbreviation


A@-)(p,) : -(iry.pr * m)12m, (4.338)
etc. The coefficient b can be obtained from (4.337) by the substituti on pi pL.
Finally for the coefficient c we must evaluate the long trace -
)
C:a+s1si } }
(tt'1,y ru )(uiry rur) ([trry, u'r) (fi ,ry,u!)

I ss
: nL/4 (ilt,y ru r) (ft r"y, ui) (ui,y ru r) (u rry, u'2)
Z 51 si } }
:
+Tr[A(-)(pi)vrA,"-,,pr)ry,A,u-,(pi)ryrA("-)(p,)ry,]. (4.339)
The evaluation of (.337) is facilitated if we use (4.80) and its generalization
Tr (.y. a)(^y. b)(,y. c)("y. d)l : 4l@. b)(c. d)
- (a. c)(b. d) + (a. d)(b. c)1,
(4.340)
which can be readily proved. The evaluation of (a.339) may look more formidable,
but actually it is not so bad if we use

"Y
: -2y'a,
*bt'a)ry,
r p("t' a)(,y' b),y, : zK"y a)("y. b) + ("y. b)("y. a)l :
- 4a. b,
f , (f . .
a) (,y b) (,y c)
"t, - - 2 (y c) (,y b) (,y a),
. . . . (4.34t)
where we are to sum over the p-indices .These formulas are not difficult to prove
using the basic anticommutation relation for the gamma matrices; for insLnce,
the proof of the second formula of
@3a\ goes as follows:
: l- ("y, a,),y, * 2a rlry y b 7,y,
ry, (ry, a,) (,y y r
^b ^)
: -(y. a)(-,y^b^I, * 2br),yr * 2(,y.b)(,y.a)
: 2l@. q)(y. b) + (,y. b)(,y. o)l
:4a.b, @.342)
where we have used ryrry, : 4 (p-indices summed). We shall not evaluate these
traces in detail since in Problem 4-13 the reader is asked to work out similar
expressions for electron-proton scattering. The net result is that we can express
o,b, and c in terms of three Lorentz-invariant scalar products

Pt'Pz: pl'pl,
Pt'P't: Pz'P'2, (4.343)
Pz'pl. Pt'PL:
Actually among the three scalar products only two are independent since the three
quantities,
: -(p, + pr)z : -(pi -l p'r)' : 2m2
s 2pr. pr,
-
t : -(Pr - p)' : -(p', p)' :2m2 * 2pr. pi,
-
u : -(p'
- pL)' : -Qi - p)' : 2m2 I 2pr. p'r, (4.344)
258 covARrANT pERTURBATToN THEoRy 4-6

turn out to satisfy one relationf


J + / | u:4m2. (4.34s)
To prove (4.345) we first verify it in the cu-system, where E : Er : Ei : Ez: EL,
P: Pz: -Pro P' :PL: -p|,lpl: lp'l:
(p, * Pr), : _E?"r: _(28)' _ _4|pl' + *r),

'i;,1'1,, : ; . I',', :1|l l,ll ; .], 3l


I (4.346)
Since s, t, and u are invariant scalars, the relation @.345) which has been verified
in the cu-system is valid in any coordinate system.
Even if we do not evaluate the traces (4.337) and (4.339), we can immediately
write the nonrelativistic limit of the differential cross section for Mlller scattering,
because frl"yrut = 0 and ill,ynut = 1 for Jr : si and 0 for s1 ;tsi in thenonrelati-
vistic limit (p : p' : 0). using (4.327), (4.335), and (4.346), we readily get the
spin-average cross section

(d,\ rE rA [ 1 r I I
(4'347)
U-O/"" l6€aLsina (012) ' cosa (012) sin'z(012)cos'(01\)'

where B is the velocity of one of the electrons in units of c (not the relative velocity
of the two incoming or outgoing electrons). Clearly the first term of @3a7) is just
the Rutherford term; the second term is necessary because the whole expression
must be symmetric under 0 -- tt - 0 for the elastic scattering of two identical
particles. The third term, which is due to the interference of direct scattering and
exchange scattering, clearly shows the effect of the Pauli principle, which tends
to reduce the cross section at 0 =, zt 12. We may mention that the exact expres-
sion for the scattering of two nonrelativistic electrons differs from the above
perturbation-theoretic result in that the third term d @.347) is multiplied by cos
{(alZ\log [tanz (012)l]. The extreme relativistic limit of the Mlller scattering
cross section is also relatively simple to obtain. In evaluating the indicated traces in
(4.337) and (4,339), 1r"-'(p,), etc. can be replaced by prl2m, etc.; we then get
-i.y.
ld'\ nnrfr(m\'[l+cosa(012),1+sina(012), 2 I
\7ol"" - T-\E-l l-W- - - cos'@7T- -
(4.348)
Effective potential; the Breit interaction. In nonrelativistic wave mechanics (as well
as in classical mechanics) we introduce the notion of a potential to be used in
conjunction with the Schrodinger equation. In contrast, in relativistic field theory we
visualize the interactions as arising from the exchange of quonta.We may naturally
ask how the two concepts are related. It is not difficult to establish the desired
connection for slowly moving particles if we are concerned just with lowest-order

fThe variables s, t, and u are sometimes called the Mandelstam variables after S. Man-
delstam, who studied the analytic properties of meson-nucleon scattering amplitudes
as functions of -(po + pN)z, -(p, - p',)',and -(p, - p'r)'.
4-6 MOLLER SCATTERING AND THE PHOTON PROPAGATOR 259

processes. According to the Born approximation in the Schrtrdinger theory the


differential cross section for the elastic scattering of particle I (mass mr) andparticle
2 (mass mr) is given in natural units byf

(#)",:\ry ! e-',"*v(x)sto'* o'*l , (4.34e)

where 2."u stands for the reduced mass,

: ff|' f/1"
ffired, ----:,L, (4.350)
tTlr t tTtz

and p and p' are the initial and finial momenta of one of the particles (say, particle 2)
in the cu-system. This formula can be easily generalized to cases where the particles
have spins and the potential V depends on momentum (-tV) and spin. Coming
back to (4.335), we see that for slowly moving fermions the covariant matrix
element .,til Tris related to the differential cross section as follows:

(#)"-=l(#) (ffi)*,,1' (4.3s1)

Comparing @3aD and (4.351) we obtain

,./il 7t : J Vr-,o'* dt *. (4.3s2)

with
: pi - p,
q (4.353)
where we have chosen the phase d @.352) in such a way that a repulsive force
(positive potential) corresponds to a positive ,til 7i in conformity with the definition
of the covariant matrix element (4.292). See also (4.38) and (4.39). Fourier-trans-
forming (4.352) we get

v:#l*r,eiq'xclsq. (4.3s4)

This formula enables us to construct an effective three-dimensional potential


(to be used in connection with the Schrodinger equation) once we know the
nonrelativistic limit of the covariant matrix element. The essential point is that the
so-called potential is nothing more than the three-dimensional Fourier transform
of the lowest-order .til-matrix element.$ In the case of scattering of two identical
fermions the "//-matrix element that appears in (4.354) is just that part of the
covariant matrix element which corresponds to direct scattering since the use of
antisymmetric wave functions in nonrelativistic wave mechanics automatically
takes care of the contributions due to exchange scattering.

fSee, for example, Dicke and wittke (1960), pp.293-295; Merzbacher (1961), pp.226-
227.
$In higher orders, however, a more careful consideration becomes necessary since the
iteration of lower-order potentials also gives higher-order matrix elements.
COVARIANT PERTURBATION THEORY 4-6

Let us see how the rule expressed in (4.354) works in the case of Mlller interac-
tion. In the cvt-system the direct term of (.327) is reduced to
e'z (ui'y u r) (ui'y u r) e' (ui ^t u r) (u'r't u r)
vlofi
@.ir*ct)
7r7/
--r--Tqln-
n n
- . (4.3s5)

At very low energies il'r,yau, can be replaced by 8,,,i, ui^lu, byzero; similarly for
ilLryrur.Hence the insertion of (4.355) into (4.354) immediately gives the repulsive
Coulomb potential e'zl(4rr), as expected from our earlier considerations. To see
the physical meaning of the second term of (a.355), which begins to be important
when B is not too small compared to unity, we first recall (cf. Eq. 3.215) that

nll ar') x-(pi - p')]rr,,r


u,,"tu, - *{sot[-'(n' * * (4.3s6)

Note that this vector is "transverse" in the sense that it has no component in the
direction pr - pi. Calling the first and second terms of (a.356) respectively "current"
and "dipole," we see that the second term of (a.355) gives rise to a current-current
interaction, a dipole-dipole interaction, and a current-dipole interaction, As an
example, we shall consider the dipole-dipole interaction in detail. Remembering
e : pi - pz: -pi * p,, we can readily take the Fourier transform of the dipole-
dipole matrix as follows:
e2 I (o"' x q).(o"' X q)sio'x
-@)W dtq

: (h)z(o('\) x v) -(o,', x v) (*J-r)


:-#.vx l#.o(#)] GssT)

We recognize that l(eoz)l2m) x V(l/4rr)lis precisely the vector potential gener-


ated by the magnetic moment of electron 2; hence (4.357) is just the interaction
energy between the magnetic moments of the two electrons expected from classical
considerations.f As for the current-dipole interaction, which goes like
[or0r X (pi - p,)].(p, * pi), we can use the identity

I '-,,"*f,#".[x x (-iy)]eto'xdtx
: -io'(p x n) [vr-us'*dsx (4.358)

(valid for any spherically symmetric V) to prove that it gives rise to a spin-orbit
force (cf. Problem 4-14). The spin-orbit potential that arises in this manner should
not be confused with the spin-orbit potential in the Thomas sense, discussed in
Section 3-3. The Thomas interaction is actually buried in the first term of (4.355)

fSee our earlier discussion on the hyperfine structure of the hydrogen atom given in
Section 3-8 (especially Eq. 3.327).
4-6 MOLLER SCATTERING AND THE PHOTON PROPAGATOR 26t

as we may readily verify using


io(L) x (pi x p,)r
[t * + ----C;ff)r(sr).
(ui"you,) (4.3se)
=1tsi)+ ffi
As a result, the spin-orbit interaction between two electrons is three times as strong
as the Thomas interaction:

v,-s) : _#+#(*",)".(o"' i ""'). (4.360)

Prior to Mdller's work, G. Breit worked out all these correction terms to the
Coulomb potential using essentially classical arguments and applied them to the
He atom; hence they are collectively known as the Breit interaction between two
electrons.

pLd,t t\t),t

vLt)'t

(a) (b)

Fie.4-21. Bhabha scattering: (a) direct diagram, (b) annihilation diagram.

Bhabha scattering; positronium. Having studied the Mlller amplitude in detail,


we can readily obtain the amplitude for electron-positron scattering, or Bhabha
scattering (so called after H. J. Bhabha who treated this problem in 1935). To
order e2 the products of the electron-positron operators that give nonvanishing
contributions are
.|!. )(x,) )(", r(x,) r/'!* )(x,),
)
^h!- 0!-
'fi!-r(r') tl,.'(",).7t' r(x,)fS- '7*r1"1-]u.7t-'(",)^f5*)(xr).7!.)(",)9!-)(",),
.[!. (x, )./, t' ;',;;-r (x,) rf 5- (x,),
) )

"flu{;.)(x,)^hS-)( x,)$?(x,)g[,")(x,).
^F!-)(x,)g!-)(",) $1.)(*,)*,S*,(",)
, (4.36r)
We can see that the first two terms give rise to the fermion lines in Fig. 4-21(a).
(Note again that for the second term we interchange x, € x, and redefine the
Dirac indices so that the second term looks similar to the first one.) The third and
fourth terms are somewhat novel. Take, for instance, .|!-)(xr)95-)(xr) in the third
term; this pair of operators can create an electron-positron pair at xr. Similarly
.|!")(x').hb*'("t) can annihilate an electron-positron pair at x,. Thus the third and
the fourth terms give rise to the fermion lines in Fig.4-21(b), known as the annihi-
262 covARrANT pERTURBATToN THEoRy 4-6

lation diagrom. The manner in which the covariant photon propagator joints the
products of the Dirac spinors is the same as in the Mflller case for both Fig. a-21(a)
and (b). The only treacherous point is the sign of each diagram. To begin let us
arrange the first term so that the annihilation (creation) operators stand on the
right (left). Apart from an uninteresting term that does not contribute to the scat-
tering process, the first term is just

f!-'(",) {1,-)(*r[rf, r(x,)r/n[*)(x,). (4.362)


This is to be sandwiched between the initial state 6ts-t+(p_)d,'.,*(p*) | 0) and the
final state b( s'-) +
(p,_) 6,';', +(p,.) g). But
|

)(x,){^5+ )(x2)b t(p- s. I *(p*)


I 0)
( s- I t
O!. )4/
: ff1
n v 1D,u exp (ip - x z * lp * x, ) 0),
a.v 1 1rn1 6
. .
I

(0|dt'l'1n}b,'.,:,.,Mu,Bu,,exp(-ip,_.x,-ip1..X')(0l,(4.363)

where u6, il'y, etc. stand for a5'-l1n -), uli-)(p'-), etc.; we therefore associate a minus
sign with Fig. 4-21(a). Since the ordering of the electron and positron operators
in the third term of (4.361) differs from that of the first term by an odd number
of permutations, we must associate a plus sign with the annihilation diagram,
Fig. 4-21(b). As a result the til-matrix element for Bhabha scattering is written
as follows:

-itr ,,: (-r)'[- iI@* -ru)(u'


(A' ,y nl ru) (u' ,y ru)(u'nyru)
P'n)'- iul iIQ* + p-)'- iul I
+ Gs64)
Comparing Fig. 4-21 with the analogous Feynman diagram for Mlller scattering
[or comparing (4.364) with (4.327)], we see that Bhabha scattering and Mlller
scattering are related by the simple set of substitutions

t Pi ----> -P *, (4.365)
-> -P'*,
P

Pz+P-, PL->P'-
(apart from an overall minus sign). From this point of view the annihilation
diagram of Bhabha scattering is related to the exchange diagram of Mlller scat-
tering. Indeed, in the hole-theoretic description the annihilation mechanism of
Fig. 4-21(b) corresponds to a process in which the incident electron undergoes
exchange scattering with one of the negative-energy electrons in the Dirac sea;
hence we expect a relative minus sign between the direct and the annihilation
diagrams. We again see that such sign factors are automatically taken care of if
we use the second-quantized Dirac theory.
Using (4.354), we can derive the potential corresponding to each of the two
terms in (a36Q. In the nonrelativistic limit the first term obviously gives a Cou-
lomb potential since D'ryno: 8r,ri, il'rynu: 8r_r,_, and u'^f a : ft'^lu : 0 for
p+ : p- : p'* : p'- : 0. The sign of the Coulomb potential corresponds to an
attraction since the covariant -y'./-matrix element is negative. This, of course, is
just what we expect from elementary considerations; e- and e* are known to attract
each other. The potential corresponding to the second term is more interesting.
MZLLER SCATTERING AND THE PHOTON PROPAGATOR 263

The reader may show in Problem 4-15 that this annihilation term gives rise to a
repulsive short-ranged [E(3)(x) type] potential that affects just the 3S-state.
At this point we shall briefly comment on how the perturbation-theoretic matrix
elements between initial and final plane-wave states can be used in bound-state
problems. Obviously we cannot directly apply perturbation theory to compute
bound-state energies. However, we can first obtain approximate bound-state
energies by solving exactly the Schrcidinger (or the Dirac) equation with a potential
that is relatively simple to handle (for example, the Coulomb potential); we then
use the usual perturbation-theoretic method to estimate the effect of the difference
between the true (total) interaction and the potential interaction already consid-
ered.{ Let us see how this method works in the case of a positronium. We first
obtain approximate energy levels of the positronium using the Schrddinger (or the
single-particle Dirac) theory with the usual Coulomb potential. Having obtained
the Coulomb wave functions and energy levels we can estimate the effects of the
dipole-dipole interaction, the annihilation interaction, etc., not previously con-
sidered in solving the Schrtidinger equation with the
-e'zl(4zrr) potential. Using
this method we readily obtain the hyperfine splitting of the ground state of posi-
tronium as follows (Problem 4-15):$

E(n : l, 'S) - E(r: l, 'S) : (i I g)a, Ry_, (4.366)

which is numerically equal to 8.2 x l0-a eV. In this formula, Ry- represents
(a'z12)m as usual; the $ is due to the dipole-dipole interaction contained in Fig.
4-21(a), whereas the ! arises from the repulsion (in the 35 state only) caused by the
annihilation force represented by Fig. a-21(b). We may mention parenthetically that
higher-order corrections due to the anomalous magnetic moment, the annihilation
force involving two virtual photons, etc., slightly modify this formula so that the
final result is

E(n :1, '^s) - E(n: t, ,s) : dz R!-14 *+- *("t, * f)] ,

. (4.367)
as shown by R. Karplus and A. Klein. Numerically formub (.367) gives 2.0337
x 105 Mc for the microwave frequency corresponding to the 3,S-t,S energy
difference. This theoretical value is in an exceedingly good agreement with the
experimental value, (2.0338 + 0.0004) x 105 Mc, obtained by M. Deutsch and
collaborators. In any case, even setting aside the small corrections of order dsRy*,
it is abundantly clear that we must take into account the annihilation force of

fAn alternative, more formal, procedure that leads to the same results makes use of
what is known as the bound interaction representation. See, for example, Jauch and
Rohrlich (1955) pp. 30G321.
$Although the hyperfine splitting of the hydrogen atom is smaller than the fine structure
splitting by a factor of about (m.lmr), (see Eq. 3.328), the hyperfine splitting and fine-
structure splittings are of the same order for positronium. For this reason the hyperfine
splitting of positronium is often referred to simply as a fine-structure splitting.
COVARIANT PERTURBATION THEORY

Fig. 4 2l (b) (which has no analog in the hydrogen atom) if we are to obtain the
correct splitting between the two hyperfine states of positronium. The necessity
of the annihilation force has also been established experimentally in Bhabha scat-
tering dt p1ab) 3m.
-
One-pion exchange potential. The relation (4.354) can also be applied to meson
theory to derive so-called one-meson exchange potentials. In the case of scalar
coupling the first term of the Fourier transform of the covariant matrix element
(4.293) immediately leads to the familiar Yukawa potential

r./(s) G2 e-tLr
- -4; r', (4.368)

obtained in Section 1-3 from classical considerations. Note that the sign of the
force between two nucleons due to the exchange of a scalar meson is attractive,
in contrast to the Coulomb case.
It is known empirically that the lightest meson that can be exchanged between
two nucleons is a pseudoscalar z meson. If we use the pseudoscalar coupling
ffrnr: iG$,yu*Q (4.36e)
in place of the scalar coupling, we obtain for the direct scattering part of the co-
variant matrix element the following:

- i,// gt "> : "-,#l yuri$"?';+ (4.370)

Comparing this with the analogous expression in the Mlller case, we see that,
apart from the finite mass of the exchanged boson, the only change necessary is
-elp --- G"ye.In the nonrelativistic limit we just have

il|ntuu,: (x*i,+, -*<,t+ol).:p|)(_: - -x(,,,t+o('1)-:e*<,,t,


)(&*;,,,,)
@371)
and
ilLyuu, y(s'ilt
o!)-:9*<',t
- (4.372)

So (a.370) leads to
NR
il,fin ------ (4.373)

where m and p arc respectively the masses of the nucleon and the meson.f This

fln place of the pseudoscalar coupling we may use a pseudovector coupling


ffin, iF= : a
ft*vt'lr^h:-ax-6'
So far as the one-meson exchange interaction is concerned, the result is identical to (4.373),
provided we make the replacement (G'z l4r) -+ (F'z l4r)(2ml D'.
MOLLER SCATTERING AND THE PHOTON PROPAGATOR 265

gives rise to the potenti::,


_ Gz
r"#(o"''v)1a*I 'v|# (4.374)

It is not difficult to show by direct differentiation that for r + 0, (4.374) can also
be written as

: #(#)' oe,). +] t,,1+,


veq
{,o,',
where Srz is the well-known tensor operatorf
.
[,# +
h* s.37 s)

Srz : l3(o").x)(o(z).x)lr'l - o(t) -6{2). (4.376)


The potential (4.375) is appropriate for describing the exchange of a neutral
pseudoscalar meson, for example, the p-p (or n-n) interaction due to z0 exchange.
In discussing the p-n case, however, we must take into account the fact that the
neutron (proton) can emit a tt-Qt*) meson to become a proton (neutron), and that
the proton (neutron) can absorb a r(r*) meson to become a neutron (proton).
It turns out that the principle of charge independence in nuclear physics requires
that the coupling constant for n € p I tt-, for instance, be related to that for
p e p * ro by a simple numerical factor. As discussed in almost any textbook
of high-energy nuclear physics, the pion-nucleon interaction consistent with charge
independence may be taken to be
ff,n,: iclnfTd"n0.ryuf , * rt6i.frrryr9" * d,(frryufo - fr.vug"y],
(4.377)
where 6r,6.n,and {j1 are the field operators respectively for fr|,8-, and z* (for
example, S"5 annihilates r- and creates z*).$ If we compute the one-pion-exchange
potential using (4.377), we obtain the potential (4.375) multiplied by the isospin
factor (first derived by N. Kemmer in 1938):

T^) .",r, : I I for 7: l,


(4.378)
[-3 for z:0,
where Z stands for the total isospin of the two-nucleon system , T :1 for 1S,
tP,tD,etc., and T:0 (notaccessibleto p-p, n-n) for 35, tP,3D,etc.
As we shall see in a moment, the interaction constant G that appears in (4.377) is
rather large; therefore the use of perturbation-theoretic arguments is, in general,
not very fruitful in meson theory. The one-pion exchange potential which we have
derived, however, turns out to be of some use when we discuss the collision of two
nucleons with large impact parameters or, equivalently, when we consider the
higher / partial-wave amplitudes of the nucleon-nucleon scattering matrix. The

{From the transformation properties of S' under various symmetry operations we infer
that this operator can change the relative orbital-angular momentum of the two-nucleon
system without changing the total angular momentum, parity, and the symmetry of the
spin-wave function; hence it connects, for instance, the 3S, state with the 3D, state. See,
for example, Blatt and Weisskopf (1952), pp. 96-108; Preston (1952), pp. 77-83.
$See, for example, Nishijima (1964), pp. 75-89; Sakurai (1964), pp. 197-222.
266 COVARIANT PERTURBATION THEORY

reason is that the exchange of many pions is unimportant at distances considerably


larger than llp,= 1.4 x l0-t3 cm. Crudely speaking, this is a direct consequence
of the uncertainty principle which requires that the energy-conservation violating
emission and reabsorption of an object with mass M can take place with an ap-
preciable probability only for a time interval comparable to or shorter thanllM,
giving rise to a potential whose range is of the order llM. (A more quantitative
argument shows that the two-pion exchange and the three-pion exchange potential
indeed fall off as e-"'lr and e-t,'lr atlarge distances.) In other words, we expect
that the tail end of the nucleon-nucleon potential is correctly given by the one-pion
exchange potential, a point emphasized by M. Taketani and others. Meanwhile
because of the centrifugal barrier, the higher /-waves are sensitive only to the tail
end of the nucleon-nucleon potential. Therefore from the higher /phase shifts ofthe
nucleon-nucleon scattering amplitude we can estimate the pion-nucleon coupling
constant G, as has been done by M. J. Moravcsik and coworkers. The coupling
constant determined in this way turns out to be
G'l4r = 14, (4.37e)
which agrees rather well with the value determined in entirely different ways, for
example, from the dispersion relations for pion-nucleon scattering and low-energy
photoproduction of charged and neutral pions.
Another feature of the one-pion exchange interaction worth taking seriously
is its tensor force. It turns out that the exchanges of two pions and other heavy
mesons do not contribute much to the tensor force; hence one-pion exchange is the
main contributor to the tensor force known to exist between two nucleons. In the
case of t,S(T : 0), the sign of the tensor force predicted by the one-pion exchange
potential is attractive since 7{t) .7tz) -
-3. The sign obtained in this way correctly
accounts for the sign of the quadrupole moment of the deuteron (which is "cigar-
like" rather than "tangerine-like"). For this reason, among the various meson
theories proposed in the 1930's, the theory based on pseudoscalar particles of
isospin unity (that is, with electric charges 0, *e) was preferred by J. Schwinger
and others long before the discovery of the pion.

Vector meson exchange. In this section we discussed in detail the propagator


rules for a spin-zero meson and a photon. In nature there also exist spin-one (vector)
mesons with nonvanishing rest masses (p, a, 6, K*); it is therefore not entirely
academic to obtain the propagator rule for a spin-one meson. The analog of (a.33a)
for a finite-rest-mass, spin-one meson turns out to be
8,,, + (qrq,lp') .
(4.380)
i(q' + t"' - iu)
The derivation of this rule from quantum field theory is somewhat involved. We
can, however, make (4.380) plausible by looking at the wave equation for a spin-one
particle coupled to a source jr(cf. Problem l4):

no, -*(*)- r'Q,:i,. (4.381)


4-7 MASS AND CHARGE RENORMALIZATION 267

Taking the divergence of both sides, we get


,a6,, a
tL"ffi: (4.382)
aiJ,.
The wave equation (4.381) therefore becomes

[d, - p'Q*: jr* t^(*)


. 1AA
r tt p' 0x r7x,t'
: /o _ I A2 \. (4.383)
\ou,
trd*^a*,11,.
We may now regard the last line of (4.383) as a "new source" and proceed as in
(4.294)through (4.298).In this way we obtain the spin-one meson propagator(a.380)
in momentum space. If the vector meson is coupled to a conserved source, we can,
of course, forget about the qrq,f y,2 term in (a.380) since (017x,)j, vanishes.
It is evident that if the source of the vector meson is proportional to iSryr$,
then the potential due to the exchange of a vector meson is similar to the photon
exchange potential discussed in connection with the Mlller interaction except
for the occurrence of e-p'lr in place of llr; in particular it contains a repulsive
central potential and an associated spin-orbit potential of the same sign as the
Thomas potential. Proton-proton scattering experiments performed by 0. Cham-
berlain, E. Segrd, T. Ypsilantis, and others in 1956 revealed the presence of a
short-ranged repulsion and a spin-orbit force which are somewhat difficult to
explain using a meson theory based on the exchange of uncorrelated pseudoscalar
pions. It was then conjectured by G. Breit and J. J. Sakurai that a unified explanation
of the short-ranged repulsion and the spin-orbit force would be possible if there
existed vector mesons of masses several times the pion mass. (The posssible existence
of such vector mesons was also conjectured by Y. Nambu and by W. R. Frazer
and J. Fulco to explain the electromagnetic structure of the nucleon.) Subsequent
experiments by B. C. Magli6, W. D. Walker, and others have confirmed that there
do indeed exist heavy vector mesons, known as aro and pt'0, that can be exchanged
between two nucleons.

4-7. MASS AND CHARGE RENORMALIZATION; RADIATIVE CORRECTIONS

The last section of this book will be devoted to higher-order processes in quantum
electrodynamics. We tacitly assume that the Feynman rules which we have derived
from field theory for certain second-order processes can be generalized to more
complicated cases. The field-theoretic proof of the Feynman rules for the most
general nth order S-matrix is not entirely trivial. Just as there are two (2 !) ways
of writing S(2) las indicated bV (a.a8)], there are n! alternative ways of writing S(n).
In any given problem we can cleverly manipulate the time orderings of the various
field operators in the S-matrix element so that the end result can be expressed
268 COVARIANT PERTURBATIoN THEoRY 4-7

simply in terms of the electron and/or the photon propagators and the wave
functions of annihilated and created particles. To achieve this in a systematic way
it is convenient to use a formalism (based on chronological and normal products)
developed by F. J. Dyson and G. C. Wick (discussed in almost any standard text
on quantum field theory).f

Electron self-energy. Before we discuss high-order processes such as radiative


corrections to the external potential problem, it is advantageous to look at the
self-energy of the relativistic electron using the rules of covariant perturbation
theory. The relevant matrix element (to order e2) is represented by Fig. 4-22 in
which the internal wavy line represents the propagation of a covariant photon.
This single diagram represents in old-fashioned language both the instantaneous
interaction of the electron with the Coulomb field created by the electron itself
and the self-interaction due to the emission and reabsorption of a virtual transverse
photon. (We recall that for a nonrelativistic electron a self-energy effect of the latter
type was already considered in Section 2-8.) Note also that when t2) tr, Fig.
4-22 represents a virtual process in which there is an electron-positron pair in the
intermediate state. (Pair creation at x, is followed by pair annihilation at x2.)

Big.4-22
Electron self-energy.

Using the Feynman rules we can write the second-order matrix element for an
electron in state (p, s) to make a transition to state (p', s') in the absence of any
external field as follows:
S.ri :-i(2n)a 8n(p - p')*/(mlEV)(mlE'V)ut''t1r'r I (p)rz"'(p), (4.384)

where ) (p) is a 4 x 4 matrix given by

Note that there is one nontrivial integration to be performed in four-dimensional


momentum space. This integration is expected since the virtual photon can have
an arbitrary four-momentum even if we assume energy-momentum conservation
at the two vertices.

{A particularly lucid discussion of the Dyson-Wick formalism can be found in Mandl


(1959), pp. 87-104. Akhietzer and Berestetskii (1965), pp. 309-316, discuss in detail the
reduction of the most general third-order S-matrix element in quantum electrodynamics.
4-7 MASS AND CHARGE RENORMALIZATION 269

It can be seen, on invariance grounds, that Z @) can be expressed as a linear


combination of the identity matrix and the 4 x 4 matrix ,y. p. It is more con-
venient, however, to consider an expansion of 2 (d about the point _i,y. p : m of
the form
Ifu) : A + (i,y.p * nt)B -l (i"y.p + m)'>r (p), (4.386)

where A, B, and l/(p) do not contain gamma matrices. Clearly, when X fu) is
sandwiched between free-particle spinors as in (4.384), only the A termcontributes.
ln order to understand the physical meaning of Alet us consider a new "fictitious
interaction"
lf t,.,, : 8^fi^lr. (4.387)

Note that this can also be regarded as an additive contribution to the mass term of
the free-field Dirac Hamiltonian. The first-order S-matrix element induced by this
fictitious interaction is just
S"ri:-i(2t)aEn(p-p,)@6mil(s,)(pl)z.',(p).(4.388)
Ifwe replace 2@) by A in (a.38a) and then compare it with (4.388), we readily
seethat A can be understood as a change in the electron mass due to the interaction
mechanism represented by Fig. 4-22 according to the relation.
A :8m. (4.339)
So far as the free electron is concerned, a theory in which we start with m : ffib^,'
(the mass the electron would have if e were zero) and take into account the self-
energy diagram Fig. 4-22, is equivalent to a theory in which we start with m
- flbo,u * 6m and omit the diagram in Fig. 4-22. When the electron is not free,
however, the self-energy diagram has a further effect, as we shall see later. There-
fore, in more general cases, it is better to start with
l7l : l\x vn 6fi1, (4.3e0)
"*
take into account the self-energy diagram, and subtract the effect due to the ficti-
tious interaction (a.387). To put it another way, when we use the observed mass,
ffibu". * 6m, in the free-field equation, the full interaction density is given by the
usual interaction density minus 6^W so that we must consider both Figs. -B@)

(b)

Fig. 4-23. Diagrams to be considered when we use the observed mass in the free-field
equation.
270 CoVARIANT PERTURBATIoN THEoRY 4-7

and (b), where the latter represents the negative of the first-order effect of the inter-
action density (4.387). Comparing this with what we did in Section 2-8, we see that
this subtraction procedure (first formulate d by Z. Koba and S. Tomonag a in 1947)
is a covariant generalization of H. A. Kramer's idea of mass renormalization.
As we shall show in Appendix F, the constants A and ,B in (a.3S6) are both infinite
although \tb) can be shown to'be finite. These infinities arise from the fact that
the energy-momentum of virtual photons, represented in Fig. 4-22, can take an
arbitrarily high value. These divergence difficulties are, in fact, one of the most
unsatisfactory features of present-day theoretical physics. We are forced to argue
that at short distances the emission and absorption of virtual photons of very high
four-momenta are not correctly accounted for by the usual propagator rule. On the
other hand, we do know that the propagator rule works well when the magnitude
of k2 is small. Therefore, to obtain a finite result for ) , we may proceed by artificially
modifying the Fourier transform of the covariant photon propagator according
to the prescription

d:-o ------
rcPyTdc(k'),
(4.3et)

where the function C(k)' (known as the cut-off function) satisfies

C(k'z)=l for lk'l<L',


(4.3e2)
C(k')=0 for lk'l>A'.
In (4.392) the (positive and real) constant A is assumed to be much greater than
the electron mass (and, as we shall discuss later, probably larger than the muon
mass by at least an order of magnitude). There are, of course, many functions
which we can use to represent C(k'); a particularly simple one (which is actually
used in Appendix E to compute l) is

c(k'):
+ k,.
(4.3e3)
^+,
photon propagator
Inserting the modified (4.391) into (4.385), we can express the
(formally divergent) constant I in terms of A. This calculation is done in Appendix
E. The result is

u:tl# (4.3e4)
'"r(#)
That the self-energy of the free Dirac electron is positive and logarithmically
dependent on the cut-off with coefficient 3aml2tt was first demonstrated by V.
F. Weisskopf in 1939.
It is instructive to contrast this result with that for the self-energy effects of a
nonrelativistic electron. In the nonrelativistic (Schrcidinger) theory the Coulomb
part of the electron self-energy is just the expectation value of
(e'z l2)l l@z lx - x' l)l
q-7 MASS AND CHARGE RENORMALIZATION 27t

as x'approaches x:
Esr,i' : +J,,/,*.r')llaffi5ra,(x - r')]+..(*) d'x
e2 f d3o
2(2r)3 J lp l'

: + Iotnt' (4.3e5)

This diverges linearly with lpl.We saw in Section 2-8 that the self-energy of a
nonrelativistic electron due to the emission and absorption of transverse photons
also diverges linearly with the virtual photon momentum (cf. Eq.2.248). A qualita-
tive argument as to why the electron self-energy is much less divergent in the
relativistic case goes as follows: In the Dirac electron theory we must take into
account the creation of a virtual electron-positron pair in the vicinity of the
electron; because of the Pauli exclusion principle, the original electron gets
effectively repelled by the virtual electron. This gives rise to a spread-out charge
distribution for the physical electron, hence a smaller self-energy.f Note that
the characteristic distance at which this pair effect takes place is of the order of
llm, which is 137 times larger than the classical radius of the electron, a distance
in the vicinity of which the classical theory (and also the nonrelativistic Schrcidinger
theory) of the electron runs into difficulty. For this reason it is completely un-
realistic to discuss the electron self-energy within the framework of classical elec-
trodynamics.
Suppose we ask what value of A we must have in order that the self-energy
(4.394) represents a substantial part of the observed electron mass. Clearly we get

L- me2n/34
= l1room, (4.3e6)

which corresponds to a distance of 10-r" cm. This value for the cut-off Iy is ridicu-
lously high; in this connection we may mention that the mass of the entire universe
is estimated to be about 1080m. There will certainly be a lot of "new physics" before
the energy-momentum of a virtual photon reaches such a high value. Thus the
idea of attributing most of the observed electron mass to the interaction mecha-
nism represented by Fig. +22 is a very dubious proposition. In what follows
we assume that even though A : 8m is formally divergent, it is still "small" in the
sense that it is of the order of r+? times the electron mass. Note in this connec-
tion that even if we set A : l0 BeV, I is only 0.015 m.
To understand the physical meaning of the coefficient.B in the expansion (4.386)
it is profitable to consider a diagram in which the self-energy interaction of Fig.

flt is amusing to remark in this connection that the electromagnetic self-energy of a


charged Klein-Gordon particle can be shown to diverge quadratical/y with A.
COVARIANT PERTURBATION TI{EORY 4-7

sr(p) srk) ,srfu)t- iEo)l,srb)

(a) (b)

Fie. 4-A. Modification of the electron propagator.

4-22 now appears as part of a more complicated diagram. To this end let us first
look at an internal electron line represented by Fig. aaa@). To simplify the nota-
tion we define

s.(p):iffi:fr#m), (4.3e7)

where the -ie prescription is implicit. The question which we now ask ourselves
is how S"(p) changes in the presence of a self-energy interaction. Clearly, for every
internal electron line of the kind shown in Fig. a-2a@) we can consider a higher-or-
der diagram represented by Fig. a-za@). This means that the electron propagator
(in momentum space) is modified as follows:
S'(p)*S"(p) * S"(p)[- ; X (p)] ^S.(p), (4.3e8)
where the m that appears in S"(p) is understood to be the bare moss. Before we
evaluate this, we note that for any operators X and I (which need not commute),
we have the useful identity

in:+- Lx"***"*"* , (4.399)

as can be verified by multiplying by X + Y from the right. Let us insert in (a.398)


the expression for ) (p) given by (a.386), ignoring the finite Zr @) term.f We see
that the modification (4.398) amounts to the substitution

# +#eo#-#
(4.400)

where we have taken advantage of (a.399). As stated earlier, both A and B are
dependent on a cut-off parameter and become infinite as the cut-off is made to go
to infinity. Nevertheless the modified propagator [the last line of (a.a00)] is strikingly

flt can be shown that the Zr b) term gives a momentum-transfer-dependent structure


to the electron propagator when p2 is very far from -m2; because of its finiteness, Xt(p)
is not relevant for the purpose of circumventing the divergence difficulties of quantum
electrodynamics.
4-7 MASS AND CHARGE RENORMALIZATION 273

Fig. 4-25. Modification of the photon propagator.

similar in form to the original (bare particle) propagator. First, we see that the new
mass of the modified propagator is m + A, in agreement with our earlier interpreta-
tion of the coefficient l. Second, apart from the above-mentioned change in the
mass, the modified propagator and the original propagator differ by an overall
multiplicative constant I
- B. If we call the modified propagator in momentum
space Sl,fu) and ignore the change in the electron mass, we have

S,(p): (l - B)S"(p) (4.401)

in the vicinity of -iry.p - m. We shall comment on the physical implications of


the constant B when we discuss vertex corrections. At this stage we simply remark
that the factor I B can be shown to be regarded as the probability of finding
-
a bare electron in a physical ("dressed") electron state. The constant I B is
-
known as the wave-function renormalization constant for the electron (to order
d).*

Photon self-energy; vacuum polarization. We are now in a position to discuss the


modification of the covariant photon propagator in a similar fashion. The photon
propagator is modified because the covariant photon can virtually disintegrate
into an electron-positron pair for a certain fraction of the time, as indicated in Fig.
4-25.$ The last diagram of Fig. 4-25 with an electron "closed loop" is known as
a "photon self-energy diagram" in analogy with Fig.4-22. Calling

(4.402)

(where the
-ieFig.prescription is ir:2' ,;;;:;,'we can write the modincation
represented by 4-25 as follows:

8*,D'(q) * 8r, * D,r D r(q)[-ifl^,(q)]8",


Dr(q) D r(q)
: [8r, I Dr(dl-iflu,(q))]Dr(q), (4.403)

f Quite generally, the proportionality constant between ^S1,.@) (in the vicinity of -i.f .p:m)
and Snfu) is denoted by Z, so that Si'k) : ZrS"(p) af
-i,y.p = m.To order a we have
Zz:l-8.
$This modification should not be confused with the modification (4.391) which is an
artificial device to make the electron self-energy finite.
COVARIANT PERTURBATION TIIEORY 4-7

which can be recognized as the analog of (4.398) for the photon. We must now
compute the quantity flr,, which is known as the polarization tensor. We first
write the result and then explain where it comes from:

-in,,(q) : -(- r), I ffir' {ffiv,ffi,v,}. (4.404)


The appearance of the two el'ectron propagators is easy to understand. As in the
electron self-energy case, there is one nontrivial momentum space integration to be
performed since the virtual energy-momenta of the internal lines are not completely
fixed even though we assume energy-momentum conservation at each of the two
vertices. The only features of @.a0$ that are not completely obvious from the set
of rules given earlier are the appearance of the minus sign and the trace symbol.
To understand these new features we may go back to the S-matrix expansion to
order en. As far as the electron operators are concerned, both the initial and the
final state are the vacuum state. So we must consider, for /, ) tz,
(ry ) "
),u (0 I fr !-'(x') {n!* )(x' ) fr !- )(x,) ^/r5- )(x,) I 0)
B(ry

: @ r), B(nt,)ru (0 I ./r!- )(", ) .7,{,- )(rr) {;t+ )(", ).,h5-,("r) I 0>
-- (t,)"8(y,),a (0 lg!.)(x,)^71-)(*,) I 0) (0 l.h.)(x,)f5-)(x,) I 0)
: -(0 I r(fu(x,)./"(x')) I 0) (ry,)"B (0 I z(.lu( x,)fi,(x,)) I 0) (ry,),6. (4.405)
From this we can understand why we must associate a minus sign with the closed
loop;{ the appearance of the trace in@.a}Q is also obvious from the manner in
which the Dirac indices appear in the last line of (4.a0f .
It is important to remark that, unlike 2 @), nr,(q) is not a 4 x 4 matrix. The
general structure of fIr,(q) can be inferred from the requirement of Lorentzcovari-
ance. There is only one four-vector available, namely the four-momentum of the
covariant photon q. So we can write [r,(q) as the sum of an invariant scalar function
of qz times D' and another invariant scalar function times qnq,. For Q : 0, however,
the polarization tensor ff, can be proportional to 8r, only; so it is more illuminat-
ing to write it as
nr,(q) - ,8,,, * 8r"4'il"'(q') I qrq,fI,',(q'), @.4a6)
where D is a constant [not to be confused with Dr(q)]defined byg

[r,(0) : DDu,. (4.407)

Let us now look at the modification of the covariant photon propagator when
all four-components of q are small; as a concrete example, we may consider the
elastic scattering of two very slow electrons due to one-photon exchange. Keeping

fThis can be seen to be a general rule applicable to any closed fermion loop.
$It is obvious that the relation e:0 necessarily implies q2 :0; the converse, however,
is not true.
4-7 MASS AND CHARGE RENORMALIZATION 275

only the leading term DE' of (4.406), we see that (4.403) now becomes (cf. Eq.
4.3ee)

fu-fu*i,<-,osr,)#
(4.408)
i(q' + o)
But this is nothing more than the propagator for a neutral-vector meson of mass
JD coupled to a conserved source (cf. Eq. 4.380). In the static limit it gives
rise to a potential between two electrons that falls off like a repulsive Yukawa
potential exp(- JDr)lr.In reality, however, we know that the observed photon
mass is zero and that the interaction between two slow electrons goes like llr at
large distances; therefore, if the theory is correct, it must give D : 0. Yet, accord-
ing to (4.404) and (4.406), D is given by

D:,i[#l,s )
io2 S d4 p (2p' + 4m'z)
-- 'c I @7V -y7=7;,' (4.40e)

where we have used

E tr,,(0) - D X E,,:4D (4.410)

and (4.341). It is not difficult to convince oneself that almost any "honest" calcula-
tion gives D + 0. In fact, using the techniques of Appendix E with a convergence
factor ltv'lb'+ A')l', one can readily show that D is a positive, real constant
that depends quadratically on A. In other words a straightforward calculation
based on (4.409) gives the following result: If the bare mass of the photon is zero,
the square of the physical photon mass goes like the square of the cut-off.
There have been numerous sophisticated papers written to explain why this
quadratically divergent constant D must be discarded. For instance, W. Pauli
and F. Villars noted that if we subtract from the integrand of (4.404) a similar
expression with m replaced by a very large mass and then integrate with a suitable
convergence factor, we get a much more reasonable result (even though such a
procedure has no meaning in terms of physically realizable particles). Others
prefer to start with a theory of a neutral vector boson of a finite bare masS y'cuare
such that the observed mass becomes exactly zero after the boson mass is renor-
malized, that is, Dpt : - pf;,* (even though this sounds like a very artificial
explanation for the simple observed fact, pfnysicar :0).1 We shall not present
these arguments in detail since they essentially say that the formalism must be set
up in such a way that the observed photon mass is strictly zero.Instead, we ask:
If the quadratically divergent constant D has been removed, does ffr,(q) contain
any interesting physics ?

fThis method is discussed in some detail in Jauch and Rohrlich (1955), pp. 191-192.
276 covARrANT pERTURBATToN THEoRy 4-7

Before we discuss the fI(1) and fI(2) that appear in(4.406), we wish to remark that
there is one property of ff, which we have not yet fully exploited. Suppose we
consider the influence of an externally applied potential on an electron as in
Section 4-3. In addition to Fig. 4-2(a), which is of order e times the parameter
that characterizes the strength of the external potential (for example, Ze in the
Coulomb potential case), there is an additional higher order contribution, repre-
sented by Fig. 4-26, known as a "vacuum polarization diagram," where x stands
for a first-order interaction with the external potential. This diagram gives an
amplitude - 137 times smaller than that of Fig. a-2@). The corresponding
S-matrix element is proportional to

(4.4n)
ry"rr,,@)Atp(q),
where 4f'(q) is the four-dimensional Fourier transform of the external potential.
The appearance of fI' can be understood by noting that the virtual electron-
positron pair in Fig. 4-26 gives rise to an integral identical with the polarization
tensor considered in connection with the modification of the photon propagator
(Fig. 4-25). Comparing g.all) with the direct scattering term of the Mlller am-
plitude, we can visualize Fig. 4-26 as the interaction between e'fIr,,4Y) and the
charge-current of the electron via one-photon exchange exactly in the same way
as we visualize Fig. 4-20(c) to represent the interaction of the charge-current of
electron I with that of electron 2 via one-photon exchange. From this point of view
e'fl,Af) is nothing more than the Fourier transform of the charge-current
induced in the vacuum due to the presence of the external potential. Note that ffu,
characterizes the proportionality between the
induced current and the external potential; that
is why ff is called the polarization tensor. Since
this induced current is expected to satisfy the
continuity equation, we must have
q*r1,,@)tr!f)(g) : 0 (4.412)

in momentum space. This relation must hold for


any external potential so that we have the
important result Fie. 4-26. Vacuum polarization.
QrfI*,(Q) :0. (4.4r3)

From the requirement that the induced current be unchanged under gauge trans-
formations on the external potential, we can also infer that
lr*,(q) Q, : o- (4.4r4)
But, since ff
is symmetric in p, and u (cf. Eq. 4.406), we do not learn anything
new from (4.414). Applying the requirement (4.413) to flru as given by (4.406),
we obtain

q,lD + q'n"'(q') + q'fI"'(q')l : 0, (4.415)


4-7 MAss AND cHARGE RENoRMALTzATIoN 277

from which followsf


Ir(2) _ (_Dlq\ _ Ir0). (4.4t6)
In other words,
frr,(q): (6r,q'-qrq,)II(q'), (4.417)
where we have defined fI(q') to be -fi"'(q').
In order to actually calculate fl(g') it is convenient to set p : v in (4.417) and
sum over the p, indices

fi,n,,Qil
: (n,i,a,, - n,)n(n\ : 3q,rr(q?). (4.418)

Now, if D (which is equal to Xl,:, [rr(0)i4) were zero, we could as well write
44
Z flrr@) (4.4te)

Even though almost any "honest" calculation gives D + 0, the way we compute
II(q')must be consistent with assigning a null value to the integral that appears in
(4.409); otherwise the observed photon mass would not vanish. This means that,
when we actually computefl(q'z), we must use the expression

II(q') : (ll3q') 2lII,,(q) - rl,u(0)l (4.420)

[with flrr(q) given by @.aDa)] rather than (4.418) because that part of [uu(q) which
leads to anonvanishing photon mass must be discarded. Both Z uflrr(Q) and
Xr[rr(0) are quadratically divergent; however, the diflerence that appears in
(4.420) turns out to be only logarithmically divergent. Using (4.404), (4.420),
and the techniques of Appendix E, we can evaluate TI(q') in a straightforward man-
ner. We shall not present this calculation in detail because we later present an
alternative (somewhat simpler) method for calculatin gn@\. The result is
tI(q'):C+q'[II(q'), (4.421)
where

C: #'o'(#) (4.422)

and

q,rrr(q,) : -+ !',ar4t - z)toglt+q"(k; ')f. (4.423)

Note that the constant C is logarithmically divergent with the cut-off A, whereas
flr @') is completely finite. When 4' is small, (4.423) gives

n(q,): -ht+- hffi)+ l (4.424)

{Many textbooks state that D must vanish as a result of the requirement (4.413) or (4.414).
This argument, however, is not completely satisfactory since on the basis of Lorentz
invariance and charge conservation (or gauge invariance) alone, we cannot prove that
lI(2) has no Pole at q2 : g.
278 COVARIANT PERTURBATION THEORY 4-7

Historically speaking, the fact that D and C are respectively quadratically and
logarithmically divergent was already recognized in 1934 by P. A. M. Dirac, W.
Heisenberg, and R. E. Peierls. The structure of flr@\ was also discussed in the
1930's by R. Serber, M. E. Rose, and W. Pauli. We have rederived these old results
using the language of covariant perturbation theory.

Fig.4-27. Change in the Msller scattering amplitude.

We are now in a position to examine the physical meaning of C and flr. Let us
recall that we have been discussing the modification of the photon propagator due
to an electron closed loop. The effect of this modification on the Mlller scattering
amplitude is represented by Fig.4-27. Clearly this change amounts to
n
W
t

- r, (n,ry,u I
#1u,, - ftt'r]{aLv,u,)
: (8,,q' q,q)JC * q'fIr(q')l|@L,t,ur)
-
e,(il1ry,",1#1u,, -
: e2(r - gfuiry-u)J!,1'vrur)
lq" * ie,rrrln\@i"yrur)(uiry,ur), (4.425)
where in the last step we have taken advantage of the fact that the qrq, term does
not contribute because of (.322). Let us treat the case where q2 is small. Evidently
only the first term of the last line of @.a2\ is important. It can be seen that the 2
constant C is completely analogous to the constant ,B discussed in connection witht
the electron self-energy diagram. The modified photon propagator at small q2
differs from the original propagator by a factor (l
photon propagator in momentum space }r,D'u(q), we have
- C); calling the modified
D,(q): (1
- C)D,(q) (4.426)
in the vicinity of q': 0. This is quite analogous to (4.401). The constant I
- C
is known as the "wave-function renormalization constant" for the photon (to
order a).f
Instead of regarding I
- C as a correction factor to be applied to the photon
propagator, we can alternatively regard it as a correction factor to be applied to the

fln the literature the proportionality factor between D'r'(q) (in the vicinity of q2 :0) and
Do@) is quite generally denoted by Z, so that DL.@) : ZtDr.(q) atq'- 0. To order a we
haveZt:l-C.
4-7 MASS AND CHARGE RENORMALIZATION 279

coupling constant. From (4.425) we notice that at very small momentum transfers
the modified Mlller amplitude is the same as the lowest-order Mdller amplitude
except that the electric charge is decreased (note C > 0) as follows:

ez ______>
er(I _ C). (4.427)

The charge of the electron which we physically measure may be deduced from the
strength of the interaction between two electrons at small momentum transfers.
This means that the observable physical charge €obs is to be identified vv116 tzf-- 6'
times the bare charge eo*o (the electric charge which we would measure if the
closed-loop effect were absent);
gobs: a/-l- Ceon"". @.4,8)
Thus, as far as small q2 effects are concerned, the modification represented by
Fig. 4-27 simply deueases the effective coupling constant. What is observed to
be 11137.04 is not 4^,"148 butrather (l - C) times that quantity. Just as in the
case of the electron self-energy nobody knows what e5*"" iS since nobody knows how
quantum electrodynamics must be modified at small distances. If we employ the
usual cut-off procedure, we see that even with a cut-off of order 10 BeV, the de-
crease in the value of e2 f 4r can be computed from (4.422) to be only
- 3y .
The basic idea which we have discussed in the preceedingparagraph is known as
charge renormalization. The observable physical charge denoted by €obs in
(4.428) is often called the renormalized chargejust as the observed mass given by
ffib^,. * 8m is called the renormalized mass. Historically speaking, at the first
Solvay Conference convened after the empirical discovery of the positron, P. A.
M. Dirac presented an argument to show that the observed electronic charge
must be smaller than the bare charge by an amount -lll37 times the bare charge
because of the creation and annihilation of virtual electron-positron pairs.
At this stage one may argue that the electronic charge can be measured not only
by the interaction strength between two electrons but also by the response of the
electron to an externally appliedfield.It is easy to see that in this case, just as before,
the appropriate correction factor to be applied to the bare electric charge 3 *-1- g
To see this for the particular case of the Coulomb potential we notice that the S
matrix element for Mott scattering (4.65) must be modified because of the vacuum
polarization graph as follows (cf. Fig. 4-26):

Zezil',ynu ,
- (-e)(u',y*u)(-iC - iq'zIIrl(q'}un - QrQ) (Ze\
- -Jql- - \iqTl
_ _Zez(l -_9)fi'ryp, _ qrflt(n\(_W), (4.42e)

where we have used q' :lql' and omitted the factor -2zti 8 (E - E')
J@IEVWW) which is common to every term in@.a2\. We see from (4.429)
that both (-Ze) and e get multiplied by 'JT - C, which, of course, is equivalent
COVARIANT PERTURBATION THEORY 4-7

to saying that we multiply the electron charge bV Jl=T and the external Cou-
"fr=T.It is amusing to note that this decrease in the effec-
lomb potential bV
tive charge of the nucleus is in agreement with the qualitative discussion of the
vacuum polarization, given in Section 3-9, according to which the virtual electron-
positron cloud tends to neutralize the positive electric charge of the nucleus.
To understand the physical meaning of If/ let us turn our attention to the second
term of the last line of @a2\, which, according to @.a2Q, is just
-Zerale|rmz)
when lq l' is not too large. Since the effective interaction potential between the
ntrcleus and the electron (assumed to be nonrelativistic) is just the three-dimensional
Fourier transform of @.a2\, we see that the Coulomb interaction has been modi-
fied as follows:

7ef;^u _, _Ze?o"*{l
-@ - C) _ Z$:*ao:ng,r,(*)
T----rfrfr
(4.430)

When the electron is scattered with a very small momentum transfer, the form of the
potential "experienced" by the electron is just that of the usual 1/r potential, since
scattering with a small momentum transfer is sensitive only to the outer region of
the potential; only the magnitude of the charge is decreased by a universal factor,
a change taken care of in any case by charge renormalization (cf. Eq. 4.428). On
the other hand, according to the intuitive treatment of vacuum polarization given
in Section 3-9, when the momentum transfer becomes somewhat larger, or,
equivalently, when the impact parameter becomes smaller, the electron is expected
to start penetrating the polarization cloud with a resulting increase in the effective
interaction strength. The I function term in (a.a30) is seen to represent precisely
this effect. Note that the sign of this 8 function interaction corresponds to arl
attraction (as it should from the simple argument of Section 3-9). We may mention'
that this term gives rise to an energy difference between the2s! and the 2p! states
of the hydrogen atom since the expectation value of the I function potential is
zero for I + 0 states and finite for / : 0 states. eualitatively we get

A,Ent : -#l.h(o) p : -#Ry-0,0, (4.43t)

as first shown by E. A. Uehling in 1935. The energy shift (4.431) contributes


Mc to the total Lamb shift of + 1058 Mc. To the extent that the experimental -27
and the theoretical value of the Lamb shift agree to an accuracy of a few tenths
of 1 Mc, we see that the vocuum polarization effect is a real physical phenomenon
which must be taken seriously. We wish to emphasize that the formally divergent
constant C is not directly observable (because it is absorbed in the definition of the
observed electric charge), whereas we do get finite numbers for physically observ-
able phenomena such as the Uehling effect.
We should, of course, consider also the vacuum polarization effects due to vir-
tual p,*yt- pairs, rr*7t- pairs, pp pafts, etc., but these are much less important
+7 MASS AND CHARGE RENORMALIZATION 28t

because ilr(q') at small 42 is inversely proportional to the mass squared of the


charged particle; for example, the vacuum polarization effect due to p* p- pairs is
(205)' times /ess important than that due to e*e- pairs. This means, in particular,
that the Uehling effect in a muonic atom (an atom in which one of the orbital elec-
trons is replaced by a negative muon) is due mainly to virtual e*e- pairs , not p,+ 1t-
pairs. We might mention that measurements of the X-ray energies emitted by
high Z muonic atoms (performed by V. Fitch, J. Rainwater, and others) indeed
confirm the existence of the vacurim polarization term. The vacuum polarization
effect also plays a significant role in the electromagnetic corrections to low-energy
proton-proton scattering.
Let us now go back to expression (4.423). Following the mz -- m2 - fe prescrip-
tion, we can readily show thatfl(qz) possesses an imaginary part given by
Im fl(q') : q'ImfIr(q')
: (2alr)1,* z(r - z)r0lz(I - z) -l (m,lq\l
: (al3)lt - (2m'lqz111l a@ffil(-q' - 4m'). (4.432)
In other words, the polarization tensor becomes complex when q' < -4m2.
The reader who has worked out Problem 4-5 will note that, apart from 2lal2l2a2,
(4.432) is identical to the probability that the external three-vector potential
A: (0, 0, c cos ra,t), a ) 2m, creates an electron-positron pair per unit volume,
per unit time, where ar' corresponds to
-q'.This similarity is not accidental. To
understand this point we go back to the unitarity equation @.45c) of Section 4-2.
Let the initial and the final state be the vacuum state in the presence of the externdL
potential A : (0, 0, a cos ort). The unitarity relation @.a5c) now becomes
sf?). + ,sf?) : -> l,st")lr,
n
(4.433)

where i stands for the vacuum state. Sfi) is then the amplitude for the vacuum to
remain the vacuum with two electromagnetic vertices. But, apart from certain
kinematical factors which the reader may work out in Problem 4-16, Sf?) is
essentially -ifI(q') evaluated at q' :
-a)2, since the only way to have two elec-
tromagnetic interactions and still end up with the vacuum state is to have an elec-
tron-positron pair in the intermediate state. This means that the left-hand side of
(4.433) is essentially the negative of the imaginary part of fI(q'). We now argue that
the S[? that appears on the right-hand side of (4.433) must represent the amplitude
for the external potential to create an electron-positron pair and that state n can
only stand for an electron-positron pair state; this is because there are no states
other than e*e- states which are connected to the vacuum state viathefirst-order
S matrix. The elements Sj?) and SIJ/ are graphically represented in Fig. +28.
Thus the unitarity equation (4.433) essentially says that the imaginary part of fl(q'z)
is proportional to the probability for pair creation; this is precisely the relation
which we inferred by comparing (4.432) to Problem 4-5.
We argued earlier that the polarization tensor fI' is essentially a proportionality
constant that relates the induced current to the external field. From this point of
COVARIANT PERTURBATION THEORY 4-7

x
/
So,Q)(Vacuum --> Vacuum) Slt(Vacuum -> e-e+)
(a) (b)

Fig.4-28. The S-matrix elements in (4.433).

view fI(42) can be regarded as the dielectric constant of the vacuum. That this
dielectric constant becomes complex for q2 1
-4m2 is to be expected on physical
grounds, since the Maxwell field loses energy because of the creation of real pairs.
We are now in a position to discuss an important relation between the real and
imaginary parts of [(q'z). Let us first recall that e'zfrr,(q)Af,(q) is the Fourier
transform of the induced charge-current due to an externally applied potential
whose Fourier transform is Afi1q1. In other words, fIr,@)i{f)(q) characterizes the
response of the vacuum to the applied field. Since this response must be causally
related to the applied field, an argument completely analogous to the one used to
derive the Kramers-Kronig (dispersion) relation for the scattering of light may
now be repeated (cf. Eqs. 2.203 through 2.209).In particular, considering a three-
vector potential which takes the form of a E(r) function pulse, we see that Jhe
"causality principle"f requires that

In(n'lls,=-.,€-t'tda
: o for t< o. (4.434)

This in turn implies that we can write a dispersion relation for fl(q') with respect
to the variable : G,
^
Re[(q,)lq,:_.,:+''J]-ffiImfI(q,,)|q,,:_.,,.(4,435)
We can rewfite this using the variable ar':

Re r(q,) lqo=_.,: +r, I dr,(d++ '?+;) rrn n1q,z)lq,o=_.,


(4.436)

Since the imaginary part of fI vanishes for co'z < 4m2,we have the final result

Re fr(4,) : +r'J],*ffr' (4.437)

We remark once again that Im fI (-s) is essentially the probability for pair creation
in the vacuum by the external potential of Problem 4-5 with arz s. The function:

fThe connection between the vanishing of the Fourier transform of II(q2) and the causality
principle can be discussed in a more rigorous manner when one uses the Heisenberg
representation of the electron field operator. See Kiill6n (1958), pp. 279-285.
4-7 MASS AND CHARGE RENORMALTZATION 283

n@') that appears on the left-hand side of (a.a37), however, is supposed to char-
acterize more complicated phenomena such as charge renormalization and the
Uehling effect. In order to calculate these (less elementary) vacuum polarization
effects, it is sufficient to solve Problem 4-5, which is one of the most elementary
problems in relativistic quantum mechanics. We can then use the dispersion relation
(4.437) to evaluate Re fl(q'z). This powerful method based on unitarity and causality
has been advocated particularly by G. Kiill6n, and by R. N. Euwema and J. A.
Wheeler.
To see how this dispersion theoretic approach works in detail, we compute the
coefficient C responsible for charge renormalization.In this case we can set e' : 0
in (a.a37) to obtain a sum rule for the renormalization constant as follows:

c : fl(o) : +7t P, J|"- ,rlm rl(-s). ( 4.435)


t-., s

According to (4.432), which the reader is urged to prove in Problem 4-16 from
the unitary relation (4.432) and Problem 4-5, we have

lg I- r(-s)
:
1l* [-s
Im n'(-s)] : dl3. (4.43e)

Substituting this asymptotic value in (a.a38), we immediately see that C is a lo&-


rithmically divergent constant that depends or s,,.,,x, the square of a cut-off energy,
as follows:

:
C
#bg ('u-) * a finite constant. (4.440)

Since a finite constant is of no physical significance when it appears together with


a divergent expression, we see that (4.440) is in exact agreement with (4.422), which
was obtained by evaluating a much more complicated integral. Note also that
this approach offers a direct (and rigorous) explanation of the sign of the charge
renormalization. The constant C must be positive because the probability for pair
creation is positive definite; as a result the correction facto r to be applied
to the bare charge (or to the electromagnetic potential) is less "/Rthan unity. Instead
of (4.437) it is also instructive to write a once-subtracted dispersion relation for
II(q') by applying the now-familiar analyticity argument to fI(q'z)lar rather than
to II(q') itself. This time the dispersion integral is completely convergent, and we
can readily obtain the finite part of fI(q'), denoted earlier by q21r , responsible for
the Uehling effect, etc. This is left as an exercise (Problem 4-16).
Before we leave the subject of vacuum polarization, it will be worth while to
make some remarks on the effect of a free radiation field on the vacuum. When
A<;'"a1 represents a classical free radiation field satisfying the D'Alembertian equa-
tion (with no source term) and the Lorentz condition, there is no induced current
e'Ir,,Af,''"d) because
(q'Er, - qrq)A!''"t) : o. (4.44t)
Thus for free photons or for traveling electromagnetic waves in empty space,
there is no vacuum polarization effect that depends linearly on the electromagnetic
284 COVARIANT PERTURBATION THEORY 4-7

field. This agrees with our common-sense idea of the vacuum. When nonlinear
effects are considered, however, this is no longer true. Two photon beams can
scatter each other because of the polarization of the vacuum brought about by a
"square diagram," as represented in Fig. 4-29, a possibility first discussed by H.
Euler.f The cross section for this process, known as the scattering of light by light,
can be computed to be about 4 x 10-31 cmz at @
- ffi, which is too small to be
observed experimentally. When two of the photon interaction vertices in Fig. 4-29
are replaced by interactions with the Coulomb potential of a nucleus, we get a
diagram for a process known as Delbriick scattering (after M. Delbriick who first
argued that such a process is to be expected on theoretical grounds). This is shown
in Fig. 4-30. Delbri.ick scattering, whose angular distribution is characterized by a
very sharp forward peak, has been observed experimentally by R. R. Wilson.

Fig. 4-29. Scattering of light by light. Fig. 4-30. Delbriick scattering.

Radiative corrections. We argued earlier that the multiplicative factor I


- C can
be regarded either as the correction factor to be applied to the covariant photon
propagator or as the correction factor to be applied to d: e2f4tt. Since the cor-
rection factor I
- B for the electron propagator is quite analogous to I - C,
we might ask whether this factor also contributes to the renormalization of the
electric charge. To answer this question, for instance, for the particular case of an
external potential problem, it is necessary to consider systematically all possible
diagrams that contribute corrections to the first-order scattering amplitude of
Fig. 4-2(a). There are altogether six such diagrams, as shown in Fig. 4-31, to
order e2 times the strength of the original first-order amplitude of Section 4-2.
They are collectively known as radiative corrections to electron scattering by an
external field. The first one, Fig. 4-31(a), is the vacuum polarization diagram
already discussed. Figure 4-31(b) is referred to as "vertex correction"; it will be
discussed in detail in a moment. Figures 4-31(c) and (d) are the self-energy cor-
rections to the initial and final electron lines. Figures a-31(e) and (f) are necessary
whenever we use the observed (already renormalized) mass for the electron (recall
the arguments given in connection with Fig. 4-23); clearly (e), or (f), cancels with
that part of (c), or (d), which goes as the coefficient A.

lln addition to Fig. 4-29 there are five other diagrams which can be obtained by per-
muting the orders of the photon interactions. Each diagram turns out to be separately
d ivergent logarithmically, but the sum can be shown to be finite.
4-7 MASS AND CHARGE RENORMALIZATION 28s

.\.\"

//
(d) (e) (f)
Fig. 4-31. Radiative corrections to the scattering by an external field.

We shall now look at Fig. a-31(c) more carefully. Apart from the mass correc-
tion, which is taken care of by the counter graph (e), (c) is seen to give rise to a
modification of the free-particle spinor for the incident electron as follows (cf.
Eq. 4.386):
a.,,(p)-___.>tl(,)(P)*vL,un*,(-')|r,,.o-lm)B*('ry.p+m),>'(r)]z.',1p;.
(4.442)
We can safely drop the last I/ term since a"'(p) satisfies the free-particle spinor
equation (3.105). We may next be tempted to cancel (i,y . p * m) with l/(iry . p * m)
to obtain I - B as the correction factor to be applied to the incident spinor.
However, this procedure cannot be justified when we realize that (i,y - p * m)
acting on n(')(p) is zero, which means that the correction factor is indeterminate.
A careful treatment (due to F. J. Dyson who used an argument based on the adia-
batic switchings of the interaction in the remote past and the remote future) reveals
that the right correction factor due to Fig. 4-31(c) is not I - B but I - (Bl2).
We can see this in a nonrigorous way as follows. Recall first that when the electron
appears in an internal line, as in the case of ^Sr,(p), we get I - B as the correction
factor (cf. Eq. 4.401). Now the electron propagator [whose Fourier transform is
,Sr,(p)l is bilinear in the electron spinor; in contrast, in an external electron line the
electron spinor appears linearly. Hence the appropriate correction factor due to
the self-energy interaction mechanism of Fig. 4-31(c) may be inferred to be just
4r-B:l_tB+0(a') (4.443)
A similar argument holds for Fig. 4-3l(d).t
As in the case of the photon propagator, we can regard these multiplicative
factors as corrections to be applied to the electric charge. We then get
e ---> [r - +B - +B * O(a'z)]e = (l - B)e, (4.444)
due to Fig. 4-31(c) and (d).
We must now turn our attention to the vertex correction represented by Fig.
4-31(b). Using the Feynman rules, we see that this diagram modifies the ry, vertex

{The whole problem of wave-function renormalization can be most unambiguously han-


dled using what is known as the "reduction technique," which is beyond the scope of
the present book. See, for instance, Chapter 13 of Gasiorowicz(1966).
286 COVARIANT PERTURBATION THEORY 4-7

that represents the first-order interaction of the electron wi'th the external potential
as follows:

-rYr+-el1, * /Yr(P', P)1, (4.44s)

where Lr(p',p) is a 4 x 4 matrix given by

L,( p', p) : (- 4' I ffi iar,{ffi\, -

- ttit(p-i,y.(p
/\ - k) -f m \^, (4.446)
k)' * m'
- - iell "'
When Lr(p',p) is sandwiched between the free-particle spinors 17' (from the left)
and u (from the right), it is possible to write it as

Lr(p', P)lrr.o,=-^,tr.p=-m : LV, I lY{,(P', P), (4.447)

where L and /v{,(p', p) are uniquely defined by

lim tt' It r(P' , P)u : L{t',y pu, @.aa8a)


p'-p
A{'(p', p) :
lI\r(P', p) - Lr(p, P)ll,r.o,:-m,t7.p:-7n. (4.448b)
We might wonder why limr,-o Lr(p', p) taken between the free-particle spinors
depends just on ry, and not on pr;the reason is that we can always re-express any
dependence on p*by taking advantage of the analog of the Gordon decomposition
(3.203) for the free-particle spinors:

(p' + p)ril' u : 2imil' nlru I i(p' - p),f,t' o,u. (4.44e)

As we explained in connection with the vacuum polarization, the electric charge


that we physically observe may be determined from the scattering of an electron
with a very small momentum transfer (p' --rp). Therefore the correction factor to
be applied to the electric charge due to Fig. a-31(b) is just I + L.I
To summarize the various correction factors to be applied to the electric charge,
we have
__> _ .B)(t + L)e,
e
"re(t (4.450)

where 4R comes from the vacuum polarization graph, Fig. 4-31(a), as dis-
cussed earlier, B from a4l@) and (d), and I * Z from 4-31(b); these factors
I-
are due respectively to the photon field renormalization, the electron wave function
renormalization, and the vertex correction. In quantum electrodynamics, however,
something rather unexpected and remarkable takes place: the coefficients B and
I become equal. To prove this important equality we first rewrite (4.385) and
(4.446) as

I (p) : iez
I@DF(kh,s'(p - k)-t,, (4.4s1)

fQuite generally the ratio of the electric charge corrected by the vertex correction to the
bare charge is known as llZr. To order ez wehave L : (llZ) - l.
+7 MASS AND CHARGE RENORMALIZATION 287

and
L,(p' , p) : - Iffi D,(k)'v,s
"(p' - k)'v rs,(p - k)'v,. (4.452)

But quite generally we have


0S"I0 p*: ,S"(p)ry rSo(p), (4.4s3)
which can be proved by first differentiating both sides of
S.(P) SF'@) : I (4.4s4)
with resp ect to p,
#,tr'ro) *
s"(p)#_[t(try. p * m)l : o, (4.4ss)

and then multiplying by ,S'(p) from the right. Note that(4.453) essentially says that
the insertion of the,y* vertex in an internal electron line without any energy-mo-
mentum transfer is equivalent to the differentiation of the electron propagator with
respect to pp.Using this important identity, we can rewrite limo,-oAr(p',p) as
follows:
LY * : t\y, n r(o', P) l,r' o' = -m, t7' p=
-7

: *1ffio,(t)o,a!#r,. (4.4s6)

Meanwhile, from the definition of .B given by (a.386) and the expression (4.4st)
for )(p), we get
uF')(p)B,y ra"'(p) : t(s)(p) (
- i A> I0 p )u"' (&)
: a,'',fu) [
r Iffi , "(,r)r,Q#o,] r,,,(n)
: y<s,t(p)Lyru,',(p), (4.4s7)
from which follows the equality
B:L. (4.458)
This then implies that
(1
-^BXl+L):l*0(a'). (4.4se)
The modification (4.450) now reads
e------>A-ce, (4.460)
to order a, which means that the charge renormalization is due to the vacuum
polarization only. This is an extremely gratifying result; it shows that the net cor-
rection factor to be applied to the electric charge is independent of the type of
charged Dirac particles. Without such a relation there would be no guarantee
that the renormalized charge of the electron is the same as that of the muon since
B and L are explicitly dependent on the fermion mass (as well as on the cut-off A).
288 COVARIANT PERTURBATION THEORY 4-7

The relation (4.460), which is a consequence of the equality (4.458), tells us that if
the bare charges of the electron and the muon are equal, then their renormalized
charges must also be equal since the same coefficient C appears in both the electron
and the muon case. As shown by J. C. Ward, the statement that the charge
renormalization is due only to the photon field renormalization can be generalized
to all orders in perturbation theory. For this reason, the equality in (4.458) or its
higher-order analog is in known the literature as the Ward identity.f
Presumably a cancellation mechanism of the kind (4.459) is also at work for the
case of the electric charge of the proton which possesses mesonic interactions.
Here the major part of the nucleon wave-function renormalization arises from the
fact that the proton can emit or absorb virtual pions (and other heavy mesons as
well). In this case one must prove that a diagram such as Fig. 4-32(a) does not
alter the charge of the proton when considered together with Fig.4-32(b), etc.
We may mention that the experimentally observed proton charge is equal in magni
tude to the electron charge to an accuracy of one part in 1022.

\,o,o.,

Neutronf
)--r:
..)'
,/'*"(a)
Fig.4-32. Radiative corrections due to mesonic interactions.

Coming back to the "pure electrodynamics" of the electron and the muon,
we must still investigate what kind of physics is contained in tt{"(p',p) defined
by (a.aa8b). A straightforward calculation using the techniques of Appendix
E shows that tr:[converges at high values of lk'z l; however, the integral diverges
logarithmically at low values of I k l' as I kl'
- 0. This is an example of what is
known as the "infrared catastrophe." A possible way to dispose of this difficulty in
a covariant manner (due to R. P. Feynman) is to temporarily assign a very small
but finite mass to the photon; this can be accomplished by modifying the photon
propagator that appears in @.aaQ according to the prescription
11 (4.461)
W=l[1-4Ez-4-y.;'
where ),,,,;n is the "mass" of the photon. With this device virtual photons of very
low energies (c,r 5 tr,,.in) do not get emitted or absorbed. We shall comment on the
physical meaning of this artificial procedure in a later part of this section. The

f In terms of Z r, Z r, and Z, the Ward identity reads Z t : Z, so that eo5* : J\ eou"u.


4-7 MASS AND CHARGE RENORMALIZATION 289

net result is that, for small values of q : p' - p, tyrr(p' , p) can be written as follows:

tv{,(p',p) l##('"r* - +)] r, * (#)(*)Q.c.,. @.462)


In Appendix E we shall discuss how to derive this formula and show in detail that
the coefficient of q,o,, is indeed (al2r)(ll2m).The important point to benoted at
this stage is that once the constant I is cancelled by the constait B, the observable
effects are completely independent of the divergence of A, athigh values of lk'zl.
To summarize, when 4 is small, apart from the charge renormalization, the
total radiative corrections due to Fig. a-31 (a) through (f) modify the first-order
S-matrix (4.64) as follows:

-r^W)a,,y,uAfi(q)
{r,[r - ##(r"sfr - + - +)]
-=> -e
^Wo' -'\2rr'
(
")(+)o,,q,luA?@), (4.463)
where Q:p-p'and
Il:,(q) : I dn *r-t@'-il'' Aff,(x). (4.464)

Note that the factor -f that multiplies -(al3r)(q'lm') arises from the vacuum
polarization graph (cf. Eqs. 4.424 and 4.429).
We shall first comment on the physical significance of the last term of @a$).
Suppose we consider an interaction Hamiltonian density (cf. Eq. 3.21l)

ffin,: -ffiltrrStl$",,$f
erc 7A\e> : (4.46s)

We can readily compute the first-order S-matrix element for the scattering of an
electron due to the interaction (4.465). The result is

s., : ffi o,,tt dnxe-'<o'-r'' W


^[W@il' !
:'# il' o' uu(t iq')
I xAf;t da

^M@
f-^Wfr'o"'uq'A'f'(q)' (4.466)

which is identical to the last term of @.a$) with rc set equal to (al2r). Meanwhile
we have already shown in Section 3-5 that the interaction (4.465) together with
the usual ie$,y r$Af;) interaction results in a total magnetic moment of the electron
given by (l * rc)(el2m) (see Eq. 3.212). This means that, as a result of the vertex
correction Fig. 4-31(b), the electron appears to have a total magnetic moment
COVARIANT PERTURBATION THEORY 4-7

given by1

t1",: (t * #)(h), (4.467)

as first shown by J. Schwinger in 1948 prior to the precise confirming measurement


of the electron magnetic moment by P. Kusch. Subsequently the fourth-order
(a2) corrections to the electron magnetic moment were computed by C. Sommer-
field and A. Petermann. The theoretical value for the magnetic moment to this
order is

t1.,: [t * #-0.328 (*)']*


: r.ool tss6(#i) (4.468)

This value is in excellent agreement with the latest experimental value


Il + (dlzr) - (0327 + 0.005xalr)'z1(el2m) obtained by H. R. Crane and co-
workers.
We shall now comment on Feynman's method for treating the infrared problem
based on the fictitious photon mass 1,,,,1n introduced in (a.461). As an illustrative
example, we shall show how the dependence on ),,,,6 disappears for the particular
problem of small-angle Mott scattering at nonrelativistic energies provided that
we ask "physically sensible questions." When we compute the elastic scattering
cross section to order (Za)'a using the S-matrix (4.463) with Al:'(q) given by
(4.66), we clearly get

#: (#)' [r - #Lql'loe(*t) * terms independent oru,,,"] , G.46s)

where (doldA)Q) is the (uncorrected) Mott cross section calculated to lowest order
as in Section 4-3. Now, according to Problem 4-10, the bremsstrahlung cross
section per unit energy for the emission of a very soft photon by a nonrelativistic
electron can be written as
dzo _ (do\'o'2o''il.
M-\re/ rffi.' (4.470)

As a consequence of the l/o behavior, the bremsstrahlung cross section integrated


over the photon energy diverges logarithmically at low values of ar; this divergence
would be absent if the photon had a finite mass. The cross section for the emis-
sion of a finite mass photon whose energy is between o) : tru,i., and o : cDg can be
shown to be

f).,,(ffi),,o,," n,.., pr,o,on


do : (#).''# W [t"t (*) - +]' ('47 t)
where c,ro is assumed to be some photon energy at which the low-energy approxima-

fWe emphasize that (4.465) with rc : dl2r is not to be regarded as a "fundamental


interaction." What we have shown is that the fundamental ie$vu*A*interaction gives
rise to a higher-order effect which simulates a first-order effect of (a.465).
+7 MASS AND CHARGE RENORMALIZATION 291

tion we have been making is still valid.I We now make the crucial observation that
in any Mott scattering experiment there is always the possibility of the emission of
photons which are too soft to be detected. In reality it is the energy resolution of
the experimental apparatus that decides whether a particular event with the emission
of a very soft photon should be counted as a scattering event or thrown away as a
bremsstrahlung event; the worse the energy resolution, the more the events which
are counted as "scattering events." In other words, what we physically measure in
any realistic experiment is the probability for elastic scattering plus the probability
for "slightly inelastic" scattering with the emission of a photon with energy
less than AE where AE is the minimum photon energy detectable by the
experimental arrangement in question. Thus the observable scattering cross
section is actually the sum of (4.469) and (4.471) with aro set to AE. But this sum is
seen to be completely independent of the fictitious photon mass ),,,,.,in So that ),-1.
can now be made to go to zero.$ Thus if we ask the "right" question, the so-called
infrared catastrophe is not a catastrophe at all. (In contrast, the divergences which
we have encountered in computing the electron self-energy and the charge renormal-
ization constant are real "catastrophes"; to solve these problems we would have to
drastically modify the present form of quantum field theory.)
We have shown how the elimination of the infrared divergence comes about
only for the particular problem of low-energy Mott scattering to order (Za)'za in
the cross section. In a good-resolution experiment the emission of many very soft
photons can become important, which means that higher-order effects cannot be
ignored. In fact, the whole perturbation argument breaks down when alog(ml\,,,r^)
becomes comparable to unity, which occurs at L,,,in
= s-r37 m = 10-60 lz. Fortu-
nately the elimination of the infrared catastrophe can be achieved to all orders in
perturbation theory, as shown by F. Bloch and A. Nordsieck back in 1937.ll
To summarize, once we renormalize the mass and the charge, the radiative
corrections to electron scattering by an external Coulomb field are completely free
of divergences of any kind. This important point was first demonstrated in 1948
by D. It6, Z. Koba, and S. Tomonaga and by J. Schwinger.

fNote that the factor that multiplies log L,,,1n is expected from (4.470);when we integrate
(4.470) with respect to a; from co,,,in to some energy coo (greater than @n,in), we obtain
exactly the same factor that multiplies log ar,,,1,,. The additive constant log 2 | in
(4.471) arises because
-
a) a finite mass photon can be longitudinally polarized (spin perpendicular to the prop-
agation direction) and
b) the relation I k I : ro no longer holds.
Since these points are rather technical, we shall not discuss them in detail. The reader
who is interested in the derivation of (4.471) may consult Jauch and Rohrlich (1951),
pp. 33G338.
$One might legitimately ask whether the tr,,,1n -+ 0 limit is a smooth one since we know
that the number of the photon spin states must decrease discontinuously from 3 to 2.
Fortunately, in any problem involving fictitious finite-mass photons, the probability for
thd emission of a longitudinally polarized photon turns out to be \fl,6/(lkl, + ii,,r)
times the probability for the emission of transversely polarized photons, a point demon-
strated by F. Coester and J. M. Jauch.
llFor the Bloch-Nordsieck method consult Akhietzer and Berestetzkii (1965) pp.413-422.
292 COVARIANT PERTURBATION THEORY 4-7

Lamb shift. As a final application of @.a$) we shall show how this expression
may be used to obtain a finite answer to the historic problem of the Lamb shift.
As usual, it is convenient to first construct an effective potential that corresponds
to the S matrix element (4.463), where lr(q) is now the Fourier transform of the
Coulomb potential (cf. Eq. 4.66). As is familiar from our earlier discussion on the
vacuum polarization, the 4'-dependent term of (a.a63) just gives rise to a 8(u)(x)
function potential. To find the effective potential corresponding to the anomalous
moment term we may use the results of Problem 3-6 in which the reader is asked
to work out the lowest-order matrix element of the (rcel2m) fi (iornFnn)r/n interac-
tion between free-particle states. Taking the Fourier transform of the matrix
element obtained there and taking advantage of (4.358), we get for the effective
potential corresponding to the anomalous moment interaction

t/(,no,, mom)
- (#)&)l**r*,r*) + #". "]
. (4.472)

Note that the first term (the analog of the Foldy term in the electron-neutron
interaction) affects just s-states while the second "spin-orbit" term affects all but
s-states. To summarize, apart from the leading term
-e'l(4ttr), the effective poten-
tial due to (4.463) is

v@,r, - #("t #-+ - + * *) u,,,,*, + #@.L). (4.473)

If the photon mass were finite, we could immediately make use of the effective
potential (4.473) to compute the Lamb shift. Unfortunately, as it stands, the
expression diverges as ),,,,1n ----- 0; this divergence, of course, stems from the fact
that the original integral (4.446) is divergent at low values of lk'l. At this stage it is
fruitful to recall that Bethe's nonrelativistic treatment of the Lamb shift, discussed
in Section 2-8, is completely free of divergence difficulties of this kind even though
it is in difficulty at high values of the virtual-photon energy. Accordingto (2.253),
(2.255), and (2.259) Bethe's expression for the energy shift of an atomic level A
corresponds to the effective potential

: #t"r (4.474)
v t{,u"
G#?,, E(')(x),
where E, is the energy of an atomic level that can be reached from state A by an
electric-dipole transition. It is now clear how we must proceed to obtain a com-
pletely finite expression for the Lamb shift. We use Bethe's formula (4.474) to esti-
mate the contributions from very solt virtual photons; the nonrelativistic approxi-
mation used by Bethe should be a very good one if the photon wavelength is of the
order of or larger than the atomic radius, that is, for E, { am. For the contribu-
tion from virtual photons of energies higher than am we should use formula(4.473),
which is divergence-free as far as the high-energy contributions are concerned.
However, it is not so straightforward to 'Join" the two expressions since the low-
energy cut-offthat appears in (4.473) is expressed in terms of the mass of a covariant
photon with four polarization states whereas the cut-off that appears in Bethe's
4-7 MASS AND CHARGE RENORMALIZATION 293

calculation is the maximum energy of a zero-mass virtual photon with only two
(transverse) polarization states.f A careful treatment (first done correctly by J. B.
French) shows that the right correspondence between Feynman'S tr,,.i,, and Bethe's
;'Guax) i,
log lOg 2Ef""*> s. (4.47s)
),n,1,,
-+ -
Since the derivation of this result is somewhat technical, we do not wish to perform
it in this book; we shall simply mention that the difference between log 1,o,,,, and
log.E!-"*r is precisely the additive constant that appears in the bremsstrahlung
cross section for a finite-mass photon (4.471).$ Using this result and the effective
interactions (4.473) and (4.474), the atomic level shift is seen to be completely
independent of Z'f""*; or ),,,,ioi for a hydrogen-like atom characterized by n and I
we get

AEn, : #l'ere:E;;+ *- + * +] | f.,(o) t,

t#{,-,'-,JI*r *n,t'd'x ror {',1',!i s.476)


where we have replaced o -L in (4.473) by its eigenvalue. For s-states of the hydro-
gen atom the energy shift (4.476) reduces to

LE,:o : Y * nr*J"ere,+a + +- +* +], G.477)

since the last term of (4.476), due to the "spin-orbit" interaction, is zero. For
I + 0 states of the hydrogen atom we can usg

J#t\,,,1'dsx:ffi (4.478)

to obtain

.' '.'
: #* . +)0 + r) "

LE,*o
^r* 1fl :ror i:t-L' (4.47s)

[L- Dl
Finally we get for the energy difference between the 2st and the 2p$ levels of the
hydrogen atom

E(zs!) - E(zpi) : #nv- [tog ffi + +- + + + * +]'


(4.480)

fTo show that the joining of Feynman's result to Bethe's result is indeed nontrivial,
we may mention that this treacherous point caused a considerable amount of confusion
(even among theoretical physicists of Nobel prize caliber) in the first attempts to obtain
a finite result for the Lamb shift.
gThe reader who is interested in the detailed derivation of (4.475) may consult Feynman
(l96la), pp. 152-157; Bjorken and Drell (19&),pp.173-176; AkhietzerandBerestetzkii
(1965), pp.423429.
294 COVARIANT PERTURBATION THEORY 4-7

which corresponds to 1051 Mc. Various small corrections to formula (4.480)


have been estimated; they include the use of the Dirac wave function in place of
the Schrtidinger wave function, the finite mass of the proton, and the fourth order
corrections (terms of order aa Ry*). The theorettcal value then becomes
(1057.70 + 0.15) Mc, to be compared with the experimental value (1057.77
+ 0.10) Mc.f
Historically formula (4.480) was first derived in 1949 independently by N. M.
Kroll and W. E. Lamb, by J. B. French and V. F. Weisskopf, and by N. Fukuda,
Y. Miyamoto, and S. Tomonaga, who all used the old-fashioned perturbation
method based on energy difference denominators. The derivation which we have
sketched is that of Feynman, which is considerably simpler and shorter. What is
more important, the separation of formally infinite (but not directly observable)
quantities from finite observable quantities is much less ambiguous when we
calculate the various terms in Feynman's way, which makes relativistic covariance
apparent at each stage of the derivation.

Outlook. In this section we have shown how, despite the divergence difficulties
inherent in the present form of the theory, we can extract finite numbers that can
be compared to experimentally measured quantities for certain simple higher-
order processes. One may naturally ask whether this procedure based on the re-
normalization of the mass and charge can be applied to other processes in quantum
electrodynamics and to amplitudes involving even higher powers of a. The answer
given by F. J. Dyson, A. Salam, and others is affirmative; once the mass and charge
are redefined there are no further infinities for any purely q\rantum-electrody-
namical processes to all finite orders in perturbation theory. In fact we have an
unambiguous and workable set of prescriptions that enables us to calculate any
purely quantum-electrodynamical effect to an arbitrary degree of accuracy in terms
of just two parameters, the observed electron (or muon) mass and the observed
charge. $
On the other hand, the very existence of divergences inherent in the theory
makes us suspect that quantum electrodynamics must be modified at short dis-
tances or, equivalently, at high energies. We may recall that, in order to obtain
finite values for the mass and the charge of the electron, we had to introduce cut-
offs. We now wish to demonstrate that such a cut-offprocedure is basically unsatis-
factory. Take, for instance, the prescription for modifying the photon propagator,

u --fu(#P)' (4.481)

fMore recent measurements of the Lamb shift by R. T. Robiscoe give (1058.05 + 0.10)
Mc. This value is somewhat different from the current theoretical estimate (1057.50 *
0.1l) Mc. At this writing it is not known whether this small discrepancy can be taken as
evidence for breakdown of quantum electrodynamics.
$The reader who is interested in seeing how the renormalization program can be carried
out in more general cases to all orders in perturbation theory may consult Chapter 19
of Bjorken and Drell (1965).
4-7 MASS AND CHARGE RENORMALIZATION 295

which was used earlier to obtain a finite value for the electron self-energy. Writing
it as

(4.482)

we see that except for a minus sign the second part of (4.482) is the propagator for
a neutral-vector boson of mass A. But because of the minus sign the hypothetical
vector boson is coupled with -(e'zlLr) in place of (e'zl4r); this means that the
coupling constant is purely imaginary. An interaction density that gives rise to
(4.482) must therefore contain a term which is not Hermitian. Now we have seen
in Section 4-2 that a non-Hermitian interaction results in a nonunitary S-matrix
element. In other words, the modification (4.482) violates probability conservation.
Up to now, despite many heroic attempts, nobody has succeeded in satisfac-
torily modifying the theory without abandoning some of the cherished principles of
twentieth-century physics-Lorentz invariance, the probabilistic interpretation of
state vectors, the local nature of the interaction between j r(*) and A r(x) (that is,
the field operators interact at the same space-time point), etc. It appears likely that
to overcome the divergence difficulties in quantum field theory we really need new
physical principles. P. A. M. Dirac views the present situation as follows:

It would seem that we have followed as far as possible the path of logical development
of the ideas of quantum mechanics as they are at present understood. The difficulties,
being of a profound character, can be removed only by some drastic change in the
foundations of the theory, probably a change as drastic as the passage from Bohr's
orbit theory to the present quantum mechanics.
Just as the modification of classical physics was stimulated by a series of historic
experiments that demonstrated the quantum nature of light and the wave nature of
the electron, it is not inconceivable that some future experiments in quantum
electrodynamics may throw light on the question of how the present theory must
be modified. We may recall in this connection that in calculating finite (cut-off
independent) quantities in quantum electrodynamics we have tacitly assumed that
the cut-off A (which characterizes in a phenomenological way a possible modifi-
cation of quantum electrodynamics at distances of order l/A) can be made to go to
infinity; otherwise the calculated values of the Lamb shift, the anomalous moment
of the electron, etc., would be dependent on A (cf. Problem 4-17). To turn the
argument around, if the quantities we calculate disagreed with observation, we
might gain some hints as to how the theory must be modified. The fact that the
results of observations made so far agree extremely well with the theoretically
predicted values (calculated under the assumption A---* oo) can be used to set a
lower limit on A or, more generally, limits on the validity of the present quantum
electrodynamics at short distances. For instance, from the fact that the observed
anomalous part of the muon magnetic moment (measured by a CERN group)
agrees with the theoretically predicted value to an accuracy of 0.4 \, we may
conclude (using Problem 4-17) that the cut-off A must be greater than I BeV
(or l/A <2 x 10-'a cm).
296 COVARIANT PERTURBATION THEORY

Quantum electrodynamics is not a finished subject. It is true that, thanks to the


success of the renormalization procedure, quantum electrodynamics has come to
enjoy a "status of peaceful coexistence with its divergences" (as S. D. Drell puts
it); yet we are still very far from constructing a complete and finite theory. The
advance in quantum electrodynamics made in the period 1947-1950 is likely to be
viewed by future historians o.f physics as the last successful attempt to extract
physically sensible results from the relativistic quantum theory whose inherent
difficulties at a fundamantal level were already recognized in the early 1930's.

PROBLEMS

a-1. (a) Derive (4.38) and (4.39). Note how the "i€" arises as we attempt to give a
meaning to the integral

gi(z"-zr)t"
J:- dt,
How would (4.39) be modified if Agsl had a harmonic time dependence e+iat'l
(b) Show that the unitarity relation (4.54) implies for a decaying state I t > the
analog of the optical theorem
1

Ttr - -zlmT.,r.
(c) In Section 2-B it was shown that the imaginary part of the energy shift of
an atomic level computed to second order is related via (2.237) to the mean lifetime
computed using first-order perturbation theory. Discuss this connection using
the relation derived in (b).
4-2. Suppose the Coulomb potential transformed relativistically like a scalar field
(rather than like the fourth-component of a vector field) so that the, interaction
of the electron with the Coulomb potential would read
#int: e$^y6<rt , Q@ : -Zel(4rr).
Show that both the shape and the energy dependence of the differential cross
section would be completely different from those given bV (a.85) at high energies
even though the two differential cross sections are identical at nonrelativistic
energies.
+3. lf is intuitively obvious that the amplitude for the scattering of a positron by a
Coulomb field computed to first order is equal in magnitude but opposite in sign
to the electron scattering arnplitude for the same momentum-spin staies. Prove
this statement rigorously. Show also that the interference terms between the first-
order and the second-order amplitude are opposite in sign for electron and positron
scattering. (Note: In reality the electron scattering cross section and the positron
scattering cross section differ significantly for scattering from high Z nuclei, for
example, by as much as a factor of 6 for Z : 8O, 0 :90", Elm : 2.)
H. (a) Prove (4.76).
(b) Show that the spin-projection operator (4.101) can be used without any modi-
fication for processes involving positrons.
PROBLEMS 297

4-5. Prove that the probability per unit volume per unit time that the external potential
(4.57) [A : (0, 0, a cos ot), At :0] creates an electron-positron pair in the vacuum
is given by
R : + (#\(+)' ^, (t *'#)
^l'
ry
What is the angular distribution of electron-positron pairs ?
4-6. The o-meson (whose rest energy is 783 MeV) is an electrically neutral J : l-
particle that decays via strong interactions into 7r+ + r I z0 most of the time.
Since a virtual process
o) --) pair of strongly interacting particles -- > rf

is allowed, the a>meson is expected to decay into e+ * e- some of the time. This
decay interaction may be represented phenomenologically by
ff int, : isfr{y r{
where d, is the neutral vector field corresponding to the cD meson. Experiments
have indicated that
f(ar --+ e* * e-)/f.o. = 10-4,

where ftot is known to be about 10 MeV. Estimate g2l4rt. Express your final answer
in the form
gzl4rt : (dimensionless number) x (#)'.
(Note: In the rest system of the r,l meson the fourth-component of the polarization
vector vanishes, cf. Problem 1-4.)
4-7. Repeat Problem 2-5 by considering the interaction density

ffint: ffi(goo,,+')(+r,,) + H"

and compare the results. Do not make any nonrelativistic approximation. lHint:
Simplify the.fr-matrix as much as you can before you square it and take the trace.l
4-8. Discuss the results of correlated polarization measurements on the two-photon
state (4.214a) when (a) both observers have detection devices sensitive to the cir-
cular polarization, (b) observer A and observer B have detection devices sensitive
to the linear and the circular polarization, respectively.
4-9. (a\ Compute the differential cross section for the scattering of a n* meson by an
unpolarized proton in the crvr-system to lowest nonvanishing order using the in-
teraction density (4.377).
(b) The coupling constant G is estimated from nuclear forces to be Gzl4zr = 14.
Show that if this value is used, the computed doldA at very lowkinetic energies
(= 20 MeV) is in serious disagreement with observation:
@old$)"rserved t (0.1 - 0.2) mb/ster

independent of 0 after Coulomb corrections are made. (Note: This shows that
the use of perturbation theory is unjustified, and/or that the pseudoscalar coupling
theory is wrong.)
4-lO. (a) Derive the ,S-matrix element for bremsstrahlung (4.240). Simplify it as much
as you can.
298 covARIANT pERTURBATToN THEoRy

(b) Show that the bremsstrahlzrg cross section for the emission of a very soft
photon of momentum k, polarization vector 6ta), b1l a nonrelativistic electron
is given by

ffi : #fr (flfr)*",,1 et'r' ( P' - P) I' ,

where (dolddl)u"tt stands for (4.85). Sum over the photon polarizations and in-
tegrate over the photon direction to obtain
dzo _2a lp' - pl' ( do\
V6A;,- fr-- aF \7-ol".,.
lHint: The polarization sum is facilitated if you take advantage of (4.313).I
,1-11. (a) Show that in the nonrelativistic Schrcidinger theory the Green function satis-
fying
/.4 x'):;5ta)(x
vn + *v')c{x, - x')
is given by

G(x,x'): x') t(lplrl2m)(t


#!Orrexplip.(x - - - t')l

: i [ #exp [ip.(x - *'l J]_ * mA


How must we perform the integration along the real ar-axis if G(x, x') is to vanish
for t'> t?
(b) Using the Green's function with the above boundary condition, compute
the transition amplitude for the elastic scattering of a Schrcidinger particle by
a time-dependent potential
{(x, t) : V(x) cos o,t
to lowest nonvanishing order. Show that the transition amplitude we obtain in
this way is identical to the one obtained using the conventional second-order time-
dependent perturbation theory.
+12. (a) In Section 3-9 the amplitude for Thomson scattering was computed using
the hole theory and the old-fashioned perturbation method. Setting p : (0, im),
k : kt : 0 in (4.275) and using the explicit forms of Kp(x, x') given by (4.264),
demonstrate the equivalence of the old-fashioned method to the calculational
method of Section 4-5 for this particular process.
(b) When O. Klein and Y. Nishina derived formula (3.3M) in 1929, the hole-
theoretic interpretation of the Dirac theory had not yet been formulated. They
therefore assumed that the negative-energy states are completely empty and a
positive-energy electron can make a virtual transition into a negative-energy state.
Why were they able to obtain the correct formula?
(c) Using the example of Compton scattering, convince yourself that the minus
sign in (4.2ilb) can be visualized as arising from the Pauli exclusion principle.
+13. It is evident that if the proton behaved as if it were just a heavy muon, the .ft-
matrix element for electron-proton scattering would be given by
//
aoft--7, ez(a$,y ru) (frL,y ru.)
PROBLEMS 299

with 4 - p(v)' - ptt) - p(e)


- p(e)'. Actually the strong interactions modify the
ryp-vertex of the proton current in such a way that frtu,y uup must be replaced by
ail,r p F {q2) * (rcI 2mr) o u, e, Fr(q')l u u,

where r is the anomalous proton magnetic moment in units of el(2mr) and Fr(q2)
and Fr(q2) (which can be shown to be real for q'z > 0) are the Dirac and the Pauli
form factors of the proton normalized so that Fr(0) : Fr(O) : 1. Derive the famous
Rosenbluth formula for electron-proton scattering in the laboratory system (the
proton initially at rest):
( d"\ _ a2 cosz (012)
\dolr^u - 482 sina (012){l + lzE sin'z(012)lmrl]r
x r * &lrrr,rn', * rcF,(q')1z tanz (+)
{tatar
. .'lF,(q\P)l
with
4E2 sinz (012)
q2:
|* (2Elmn) sin2 (012)'

where the incident laboratory energy of the electron denoted by E is assumed to


be much greater then rn". lHint: Simplify the ll-matrix element as much as you
can before you square it and take the trace.l
4-14. (a) Prove (4.358).
(b) Show that the spin-orbit potential between two electrons is given by (a.360).
,t-15. (a) Show that the annihilation diagram of Fig. +21(b) has no effect on the singlet
state of an e+e- system with p* : p- : 0, first, using the explicit forms of the
free-particle (at-rest) wave functions, next, using only symmetry arguments.
(b) Construct an effective potential that corresponds to Fig. 4-21(b). Show that
this "annihilation potential" affects only the 3S state.
(c) Show that the energy difference (hyperfine splitting) between the 35 and rS
states of positronium with n : 1 is given by (a.366).
4-16. (a) Express in terms of II (q2) the S-matrix element (up to second order in e) for
the vacuum to remain the vacuum in the presence of the external potential con-
sidered in Problem 4-5. Using this result and Problem 4-5, show that the unitary
equation (4.433\ indeed gives (4.432).
(b) Applying the usual analyticity argument to tI(q')|, with 42 : -cD2, write
a once-subtracted dispersion relation for II(q2). Evaluate the dispersion integral
to obtain expressions for flt(42) valid for q2 = 0 (cf. F,q. a.42Q, and for q2 ) m2.
+17. Apply the techniques of Appendix E to evaluate the anomalous magnetic moment
of the muon with a finite A. Discuss how the predicted (Schwinger) value of the
magnetic moment would be altered if A were 500 MeV, I BeV, or 10 BeV.
APPENDIX A

ELECTRODYNAMICS IN THE
RADTATTON (COULOMB) GAUGE

The purpose of this appendix is to derive the basic Hamiltonian of electrodynamics


subject to the requirement that the three-vector potential A satisfies the trans-
versality condition. We shall first show that it is always possible to consider a gauge
transformation of the type (1.82) such that in the new gauge the transversality
condition
V.A(x, l) :0 (A-t)
is satisfied. To see this explicitly we just set

A(ne$')(x, /) : 4tota)(x, t) t Yy(x, t),


,,4[""'u)(x, t) : A6"tu)(x, /) * Olc)[ay(x, t)l7tl, (A-2)
where

x(x, t) : + t #{[r.[(ora)( x', t). (A-3)


We then have
!.[(new) : V.(A("rar * VX) : 0, (A-4)
which shows that the transversality condition is satisfied for the new vector poten-
tial. On the other hand, from (1.75), which holds quite generally, we get

Y,Ao - *ffi + +*(r.A + ++) - -p (A-5)

for the zeroth component of lr. When the three-vector potential satisfies the trans-
versality condition, it is seen that lo is determined by Poisson's equation
YzAo(x, t) : -p(x, t), (A-6)
which does not contain any time derivative. The solution to this differential equa-
tion satisfying the boundary condition Ao 0 at infinity is
-
Ao(x,t): * Ior_*'!#i' (A-7)

We wish to emphasize that Ao in (A-7) is an instantaneous Coulomb potential in


the sense that the value of Ao at / is determined by the charge distribution at the
same instant of time; in particular, it should not be confused with the retarded
Coulomb potential
Ao(x, t) : * I o'*'#?l (A-8)
r=t_tx_x,l

301
302 APPENDIX A

which we obtain by solving the D'Alembertian equation for Ao.l To summarize


what we have done so far:
a) we can always choose the vector potential so that the transversality condition
(A-l) is satisfied;
b) if ,4 satisfies the transversality condition, then lo is given by (A-7).
The gauge in which I satisfies (A-l) and Ao is given by (A-7) is called the radiation
(or Coulomb) gauge.
The total Hamiltonian for the charged particles and the electromagnetic field is
given by the space integral of
ff:ff.rrr+ ffrn^rr", I ffin. (A-9)
For the Hamiltonian density of the electromagnetic field we have, from (1.73),
_
/?ettr
sv
09"n, 0A* _ &q
e.r
0(0A*l0x) 0xn
: ;(lB l' + lE l') -
I
iF.YA4

:;(lBl'+ lEl') -
I
pAo + V'(AoE), (A-10)

where the last term of the last line is of no significance since it vanishes when we
evaluate 1ff"^,d'x. As for the interaction part of the Hamiltonian density, we
write it as
ff;',rt: - j*Ar, (A-ll)
whereT, is the charge-current (four-vector) density. Combining (A-10) and (A-11),
we get
I/u,,. 1 Hin : {Zr -, * ,ff i^)dtx
[
: +J c" l' + I r.l')drx - J i .A drx. (A-12)

At first sight this appears to be a very strange result because the pAo term has com-
pletely disappeared. We shall come back to this "puzzle" in a moment.
Our next task is to show that the Hamiltonian (A-12) can be written in a some-
what different form when we work in the radiation gauge. First, we recall that,
quite generally, the electric field E is derivable from

E:-vAo-+# (A-13)

We claim that in the radiation gauge the irrotational and the solenoidal part of
E defined by
E: Err* Er,
(A-14)
V x E't:0, V'E : g.

fNote that the D'Alembertian equation for lo follows from (A-5) when the vector poten-
tialsatisfies the Lorentz condition, not the transversality condition. For the derivation
of (A-8) see, for example, Panofsky and Phillips (1958), pp.212-214.
APPENDIX A 303

are uniquely given by

Ett: -YAo, Er: -+T, (A-15)

where the superscript (I) is to remind us of the transversality condition (A-1).


This can be proved by noting: (a) V x (Vlg) vanishes, (b) the divergenc" o, ,S<rt l7t
vanishes because of the transversality condition (A-l), and (c) the decomposition
of E into E11 and E is unique.
Meanwhile we can show that the space integral of lE l2 is actually given by

I Efa'*: J t El2d3x * t eAod.'x (A-16)

in the radiation gauge. To prove this we first note that

t Era'*: JI
Erl' - 2(vA).Er * | E11l,ld3x

: J t r.'rlzdsx +, I Ao(y.r.r)drx+ J I Ey,lzdsx

: J C Er l' + lE,,l'z)d3x. (A-17)


Furthermore,

J tn, ryd'x: I {v,a,).(YA)dsx


: J
v .(AoYA)dtx - t Aov,Aod',
: pAod'x. (A-18)
J
Combining (A-17) and (A-18), we see that the space integral of lEl' is indeed
given by (A-16). Coming back to the expression for the Hamiltonian given by
(A-12), we note that the pAo-term which, we thought, had disappeared is actually
buried in the space integral of lE l':

f/"".+ Hi':+JCBI'+ lurl,)dtx + [ (+cdo - i.4,',)a'". (A-1e)

Note, however, that the pAo-term now appears with a factor f.


We are finally in a position to eliminate Ao altogether (or equivalently Els)
from the Hamiltonian. Using (A-7), we can write the pAo-term of (A-19) as

t I on,orx - t I o'. I o'*' *W (A-20)

Hence

H",,, * Hint: tInBl'z*lEr l')d' x


* +[ d'x t d'x' & t)p(x', t)
4r lx - x' t- !i'A'''d'* (A-21)
APPENDIX A

or, in terms of A(r), j, and p only,

F/",,, + Hin :}. J (r


v x A(r-) r * | +#l')o'*
- Ji .A*dsx + + I o'* Io'*,ffi. @-22)

In this form the electrostatic potential Ao (or the irrotational part of E) has been
completely eliminated in favor of the instantaneous Coulomb interaction. This
means that we can forget about the dynamical degrees of freedom associated with
As or Er. In quantizing the Maxwell field it is sufficient to quantize just the trans-
verse vector potential A{r) provided we work in the radiation gauge. This procedure
was first advocated by E.Fermi inl932.
From (A-21) and (A-22) we see that the only term in the interaction Hamiltonian
which involves the electromagnetic potential explicitly is the space integral of
-j.A<|. In nonrelativistic quantum mechanics this is given just by the difference
4#,(r"'-+)' -?W (A-23)

As for the Coulomb part, we can use the familiar form


p:>e,grar(x_1oty (A-24)
to obtain

tlo'. 10,*,ffi:BW#fu1 -r2


,lsl e1

+47F6r=75'l.
(A-2s)
The last term which represents the interaction of a charged particle with the
Coulomb field due to the same charged particle is infinite; hence it is usually
dropped. However, in discussing the self-energy of the Dirac electron, it is neces-
sary to include the relativistic analog of this term, as shown in Section 4-7.
APPENDTX B

GAMMA MATRICES

I. Our notation (also used by Pauli, Kiill6n, Rose, Mandl, Akhietzer and Beres-
tetzkii, etc.) is given below:
: 2g *,,
{r r, "1,}
where Itrru: lr2r - - - 4, yL: yri
: I fl zil{f :
"l s a,
fi.r rr^"rl pf ,nl i xnfo

{ryu, ryu} : o,

where p: l, . . . 4, "y[: ,ys, ,y?: l;


a.b : a.b + atbE: a.b - aobo, ap: (a, a) : (a, ias);

rr: -i Bar= (,1_ ;"1 , k : t,2,3,


np lI 0\
nf+: U: (O _1,
DP l0 1\
s: - \t o)'
rf

ct, :|'# : -iYunl, (tn * ,),

6,j: -,rsdn: rr (- Qjk cycric),


= )
ck4 : :
-cue -rsct j - dk= (: "J (ijk cycric),

int,ntr=(* j),

intuntn=,(_1,
;)
For free-particle wave functions

(rr&+ry){n :0, (i,v'p I mc)u("(p) :0.


305
306 APPENDIX B

II. Notation used by Messiah, Bjorken and Drell, Schweber, etc.:


:2gP', p,u :0,1,2,3;
tyr,rl']}
8p, : gP', goo : 1, gu, :
-1, gr, :0 if P * v;
: -"yk;
nlot:"y0,
"yrl
a.b : avbr: asbs - a.h, ctr : (arra),

at, : gpra' : (eo,


-a);
np/I 0\ o*\.
Yo : B: (o nvr : Bdr -/0
-l' \-ou o I
For free-particle wave functions,

?,r,h +ff)^h : o, ("y- p - rnc)yrst(p) :0.


UL Useful relations (our notation only):

@' a)('Y' b) : 2o' b


- (ry' b)('Y' o),
(y.aXy.b) : (a.a)(a.b) : a.b + t>.(a x b),
ryn[(ry. a)(,y.b)
. . . ("y. l)]r,yn : (-ry. l) . . . (-ry. b)(-,y. a),

if a : (a, iar) with ar and a, real, etc.;


nltnlp:4
I
"yr(ry.a)'f p: -2"y.o I

,t,("y. a)("y. bh, :_ro?.. b) + 2(,y. b)("y. al summed ;

?r. fr-inoices
,y
r(ry. a)("y b)("y ch t,
. . : _ 2(,y.cXry. b)(,y . o) I
t'%#P:o'
Tr(ryug):0,
less than 4z (for 47 nonvanishing only if all four are different)
Tr(1,) : g,

where
le : nl p, 6 ttw i,y 5,y p, ,f s,

Tr[(ry. a)("y.b)] : 4a.b,


Tr[(v. a)(,y. b)("y.cX.y. d)l : 4l@. b)(c. d)
- (a. c)(b. d) + (a. d)(b. c)1.
IV. Free-particle wave functions in the Dirac-Pauli representation:

a) For g:nflpfffi7>0,
*\o,z) : [i(p. xlh) - (iltlh)],
^l-#u,r,,,(p)exp
APPENDIX B

where

z"'(p):,rW(r.,rr*l-,
)
\(p' * ip,)l@*mc')l

u,,, (p) : EF( .


r r, -, o,tlrrn .,,,1
\-p,rl@ I mcz) I

b) For E: -\IiF7lfr7 <0,


*\(8,4) : ,ffiu,',0,(p)exp [i(p .xlh)+ (tl EVlh)|,

where

+w(:i;,Y l,j,T,,-),* -,,,),


ncz
\ ;
* "'')
u,'';'Ji;'
#(;'l!,,',
\1
c) For "positron sPinorsl,
: -u,n,(-p), a,r,(p): z(r)(-p).
APPENDIX C

PAULI'S FUNDAMENTAL
THEOREM

Before we prove Pauli's fundamental theorem that guarantees the representation


independence of the gamma matrices, we shall summarize the properties of the
gamma matrices obtained in the main text. In Section 3-5 we showed that we can
construct sixteen 4 x 4matrices denoted by I', with A - l, . . ., l6.They are
(c-1)
each of which satisfies
H:1. (c-2)
Using the anticommutation relation betwear Ip and ry, alone (that is, without
recourse to any particular representation), we can show by direct inspection the
following two additional properties :
a) Define a constant rlauby

|rl, : Tdalc. (c-3)


Then
T,Ea
: + l, - l, +i or -i. (c-4)
b) Matrices Ia and l, either commute or anticommute (that is,Irl, : *Irfr).
Moreover, for every T"(+ l), there exists at least one lo with the property
rBl/IR - -lz. (c-s)
As examples of property (b), we may mention that (C-5) can be checked for
Iz : iryuryrandorras follows:
,yu(irysry)"ys : -irlsnf s,
nfzanfz
- -on. (C-6)
As we already mentioned in the main text (see footnote on p. 106), using
property (b) we can readily prove the following which we call property (c).
c) Any 4 x 4 matrix X can be expanded in a unique linear combination
16
X :2 x.nl^q. (c-7)
To prove Pauli's fundamental theorem we must first demonstrate what is known
as Schur's lemma which is stated as follows:
Suppose that there is a 4 x 4 matrix with the property

lX,yrf :0 for every ,yp 0n:1,2,3,4); (c-8)


then Xis a multiple of the identity matrix.

308
APPENDIX C

The proof of this lemma goes as follows. Using property (c) we can expand X
X: xele * }nxtlt, (C-9)

where lo(* 1) is any one of the 15 11 matrices in (C-l) (excluding the identity
matrix). lf X commutes with every ?p, then it must necessarily commute with
lrwith B :1,...,16:
|BX : Xlo, laxlu - X. (C-10)
Because of property (b), for the particular I, appearing in (C-9) we can find (at
least one) l, such that (C-5) is satisfied. With such lu we get
X:TnXls
: x/sl;l, * }nxcl7xtTo
: -x,tle, + > (;f l)xsla, (c-r l)
C+A

where we have used (C-9), (C-10), and properties (b) and (c). Comparing
(C-l l) with (C-9) and recalling property (c), we conclude that
XA: -Xl : 0. (c-12)
Now I, is arbitrary except that it cannot be the identity matrix. The relation
(C-12) then means that, when we expand Xas in (C-7) all the expansion coefficients
are zero except for the coefficient of the identity matrix. It therefore follows that
Xis a multiple of the identity matrix. (Of course, X canbe identically zero.)
We are now in a position to prove Pauli's fundamental theorem which can be
stated as follows:
Given a set of 4 x 4 matrices fryr] and another set of 4 x 4 matrices {ryi,} with
P:I
{nl r, ,1"} - 28 p,, (c-r3)

{ryr, ryi,} : 281,,, (c-14)


there exists a nonsingular 4 x 4 matrix ,S with the property
,yt, _ ,Sry"S-t. (c-rs)
Moreover, S is unique up to a multiplicative constant.
To prove this we first define Ii in the same way we defined Ir, for example, if
I,e : inlu"ln : inyflznls, then Il : iry|ry'n : iry'rrt'rry'r, etc. It is evident that

ryr; : TnIb (c-16)


with the same Tn&s in (C-3). We consider ^S
of the form
S: E T'FIA, (c-17)
where F is some (arbitrary) 4 x 4 matrix. Multiplying both sides of (C-17) by
li from the left and Iu from the right, we get
USf'" : rytiFr/rB. (c-18)
A
310 APPENDIX C

Because of (C-2), which also holds for the primed gammas,


(c-1e)
We can therefore rewrite (C-18) as follows:

ry,slB (c-20)
AC
But, because of (C-17), we get
llSlu : 5. (c-2t)
Note that this is true for any lr.
We claim that F can be chosen so that ,S + 0. To prove this, suppose S were
zero for eYery F. Choose .Fso that
F,.B :8oo,8Bp, (c-22)
for example, for 4' : l, F' :2;
t0 100\
t:lolooool
o o ol (c-23)

\ooo ol
Clearly the assumption S : 0, together with the choice (C-22), enables us to re-
write (C-17) as
> (rru",(I n)8,, : 0 (c-24)
for the 8e element of the matrix S. Since S is identically zero for every F by assump-
tion, (C-24) must hold for all B'and e, which leads to the matrix equation

? (li)6"'1,
: 0. (C-2s)
This clearly contradicts the linear independence of the l-matrices. So there must
existFfor which ,S # 0.
Our next task is to prove that S is nonsingular. We construct
S': >TbF'lc, (C-26)

where F' is some 4 x 4 matrix. Repeating steps analogous to (C-18) through


(C-21), we get

Moreover, F' can be so chosen that S'is not zero. Consider


S',S : lB^S'I;ry,SlB : IB,S'^Sl/r. (C-28)
This is true for every lr. Hence
[,s"s, Yr] : 0. (c-ze)
By Schur's lemma S'S is a multiple of the identity matrix
S'S : C, (c-30)
APPENDIX C 3il

where c is a constant. Thus


S-l : ,S'/c. (c-31)
SinceS'is not zero, we have shown that S-1 indeed exists. Coming back to (C-21),
we can multiply both sides of (C-21) by 5-' from the right:
I;,S|8'S-t - l, (c-32)
or
- ltu, (c-33)
'sl,'s-r
which is of the form (C-15). We have thus exhibited the existence of ,S with the
property (C-15).
To prove the uniqueness of S, let us assume that there exist St and S, both
satisfying (C-33). We then have
S,ryrSft : SrryrStr (C-34)
or
^Sr
t,Srryp : ryrS2-rS1. (C-35)
Because of Schur's lemma, SttS, is seen to be a multiple of the identity matrix.
Thus
Sr : 4Sz,

where a is a multiplicative constant. This concludes the proof of Pauli's fundamen-


tal theorem.
APPENDIX D

FORMULAS AND RULES IN COVARIANT


PERTURBATION THEORY

I. Definition of the ll-matrix:


Srr : 8ro
- i(2r)a$<tt(Pt^') - Pt*\@,tr y1,

where
for fermion
,,' : [*tlEt
[l/28j for boson.
If. Relation of .& n to transition probabilities and cross sections ("covariant
Golden Rule"):

a) Decay I ----,2 + 3 +...+ n.The differentialdecay ratedwis

d'w:$rlz,f We*,.-ileo-@o)2E,Pr>r^(2r)a$<tt(o'_ P,0,)

To eliminate the 8(a) function first integrate over the (three-) momentum of
one of the final-state particles, and then use

lprl'don
d'prg(n,-f,r'\-
"):w,:ffi,or,,r, lprl'dlprldhr

For 1 ----- 2 + 3 the (partial) decay rate is given in the rest frame of particle I by

r(t 2 + 3): + -bW I on p^^,1// rul, rr effir",,,,i),


-> - spin fernrions

where lp'r I : lp, I : I p, l.

b) Differential cross section for I *2-3+4 +...*n whenthemomenta


of particles I and 2 are collinear:
,1ll:
cto
r 1u;1tr ril'l g (2m'"'n'i)1ffi, &
^,*
x (2r)a6(n,(0, * p, - Ao,)t
: lp' I : lp,
[E r lpt" I
in the clt system (lpt,I l)
'u,,rErE, : in the rest system of particle 2 (lab system)
l*rlprl
l@ ingeneral.
APPENDIX D 3r3

For 1 * 2 3 + 4,the differential cross section in the ctr,t-system is given by


-
(#) : + &* *i il r''',""#.",,t-r""",i)'
t

"-
One may sum over the final spin (polarization) states. If the initial particles are
unpolarized, one may sum over the initial spin states and divide by the number
of the spin states of the initial system.
UI. Rules for writing -i.,t6 1ti
a) External lines
e!") for each absorbed photon (or spin-one boson)
ef;)for each emitted photon (or spin-one boson)
"tt
;d
(If eff) represents circular polarization, it should be complex-conjugated
when it appears in the final state.)
1 for each spinless boson absorbed or emitted ,.,'r,"
rzt"(p) for each absorbed spin-{ fermion /
tt"G) :(-l)s;<s-sr1-p) for each absorbed spin-] antifermion y'
rzt"(p) for each emitted spin-f fermion
4/
a,',(p) : (- l)s4rs-sr1-p) for each emitted spin-{ antifermion ,/
b) Vertex factors

-€f pif: -ie $'Yr*Ar


,ffin
-iG if ff^r: G w6
G,ys if ff rn : iG filu*Q, etc.

In general, take -i ff rn, (more precisely i g n) and replace the field operators
by the appropriate free-particle wave functions. Omit e+tp'' and the factors
already taken care of by external lines (for example, ef;) and zr')(p)) and nor-
malization constants (,rTlnr). The remainder is the vertex factor.l
c) Internal lines

Photon ("covariant photon \, ryfu-; 11

Spin-one boson:
#=6(u",*T)
I
Spinless boson:
WTV4 F--------{

fermion: -i,y.qlm
Spin-$
@w-,4
llf the interaction density involves derivatives of field operators, # i6 difrers from - I in ,
It can be shown (using an argument originally given by P. T. Matthews for the case of
the pseudovector coupling of a pseudoscalar field) that the vertex factor in the Feynman
diagram should be read directly from i gi6 rather than from -i #in . In this connec-
tion we may remark that it is possible to formulate an S matrix expansion using -9ling
(without recourse to ff int). See Bogoliubov and Shirkov (1959), pp.206-226.
314 APPENDTx D

d) Caution
i) Assume energy-momentum conservation at each vertex. Integrate with
daql(2tt)a over an undetermined four-momentum (that is, four-momentum
not fixed by energy-momentum conservation). If the internal four-momentum
is already fixed, no integration is necessary.
ii) For each closed fermion loop take the trace with a factor (- l).
iii) Put together the various pieces so that as we read from the right to the
left, we just "follow the arrow."
iv) Multiply each matrix element corresponding to a particular diagram by
the permutation factor 0". For example, 8" is * I and - I for the direct and the
exchange term of fermion-fermion scattering, - 1 and * 1 for the direct and
the annihilation term of fermion-antifermion scattering. The sign of the matrix
element is of physical significance only when interference between two or more
diagrams is considered or when we are constructing an effective potential.

IV. Properties of free-particle spinors (see also Appendix B, Section IV):


(iry.p f z)zt')(p;:9, r7(')(pXiry. p * m):0,

(-i"y.p * m)u(')(p):0, t(s)(px-i't.p + m):0,


(p' - p),il'o,uu: i(p' 1- p)ril'u + (m' 4 m)fi',y,u,
(p' - p)*il'inlunl*u: (*' I m)il',yuu,

L rf,tp)ag)(p) : - (o#),,
If u@(p)and o(')(p, spinors satisfying
"r;;r.-particle
inf uny.wurrr(p) : tt@)p, inyury.wa,*r(p) = u,d(p),
then

where
til. p :0, w' :lw l' - wB : l, [irf u"y.w, irl.p] : 0,

(ft,0) in the rest frame,


(ff, 0) for transverse polarization,
w- (#' for longitudinal polarization.
#)
(o* ffi,'#) ingeneral.
APPENDIX E

FEYNMAN INTEGRALS; THE COMPUTATIONS


OF THE SELF.ENERGY AND THE ANOMALOUS
MAGNETIC MOMENT OF THE ELECTRON

When the energy-momentum of an internal line in a Feynman diagram is not


completely fixed, there is a nontrivial integration to be performed in four-dimen-
sional k-space. To make an attack on such a four-dimensional integral, R. P.
Feynman invented a set of very clever tricks.
Useful formulas. We first prove

rdnk I i (E-la)
J@Y@'+l -ll);:9;'-s'
r d'k ku _^ (E-tb)
J@fi(k-fT$-"'
from which many other useful formulas can be derived. The second one (E-lb) is
obvious, since the integrand is an odd function of k, and the integration is over
all values of k and ko. To prove (E-la) we start with a simpler one-dimensional
integral
f- dkn
_-f- d(o ,.
)-*v;v-iu:J--W (E_2)

We can handle this integral by the usual contour method. Noting that there are
poles at ko : +(,r/-lnl' + s -iE), we get the result

J--1Er1s-;;1 :mF
f-dkoir (E-3)

(whether we close the contour in the upper half-plane or in the lower half-plane).
We now differentiate both sides of (E-3) twice with respect to s to obtain

f* ,, uf,dko
J-*(k' , ="
s-ie|- :3?--l--^
8 (lkl'*s)F'
. (E-4)

The remaining three-dimensional (k-space) integration is elementary since the


integrand is spherically symmetric. Using

I (k d'k5)oiz _4r lkl,dlkl


- "" Ji- (lk
J lz 1 o l' + s)t'
4tt
(E-5)
- 3s'
we readily obtain (E-la).

315
316 APPENDIX E

We now substitute ls ---, k - p for the integration variable in (E-la) and (E-lb);
we also let s ---- s
- p2. We then get variations of (E-la) and (E-lb) as follows:
ldnk i
J@W:77;57fl', (E-6a)

I dnk kp _ ip*
(E-6b)
I (2r)a (k' - 2k- p* s- ie)r - 32r2(s
- p')'
which are quite useful. Furthermore, differentiating both sides of (E-6a) and
(E-6b) with respect to s and p,, wa obtain more useful formulas:
td4k I I
(E-7a)
JWQc-w-2k.pf s- i.I: W'
I dnk t, _ ip* .
Jww:xrcnry' k (F_7h\
\L-tu)

I @dffi : w#loro,* *r' - p')8,,].


(E-7c)
In practice the denominator of the integrand of a typical four-dimensional
integral is made up of different invariant factors multiplied together, for example,
(k' - 2p.k)(k'z + A')'. Such a product can always be rewritten so that formulas
(E-6a) through (E-7c) are immediately applicable. To see this we make use of
what is known as the parametrization method (also introduced to covariant per-
turbation theory by Feynman.) The simplest relation along this line is
lr'dx (E-8)
ab- Jolax+b(l -,r)lD'
-I

which can easily be proved by considering


1_ I lbdz (E-e)
a6-6-aJ"V
withz - ctx + b(l - x). Differentiating(E-8)withrespecttoaweget
I t xdx
a'b:-'('-)06ary=fi' (E-10)

A slight generalization of (E-8) is


1- ', It s- lt-" ,l
A - " Jo"n Jo al , (E-l l)

the proof of which is left as an exercise for the reader. From this relation we
further obtain

,)*: a l'oa, I',-" orra , OL Ai + td = aVy, etc. (E-,12)


Electron self-energy. Let us see how we can apply Feynman's method to compute
the electron self-energy.According to (4.385), (4.386), and (4.389) we have
APPENDIX E 3t7

where the prescription k2 ----' k2 ie is implicit. But


-
,yrl-i"y.(p
- k) + ml,yt, : 2iry.Q - k) * 4m

- -2i"y.k * 2m, (E-r4)


provided that the whole expression is sandwiched bV Zr"t1n; from the left and
a"'(p) from the right. We can, of course, use (E-l l) to combine the three factors
in the denominator of (E-13), but instead we choose to use

(#X"#*) :J;'d'@+T (E-ls)

and take advantage of (E-10). We then have

Dm:-ie,
,^,
Iffi!"'a,ffi
Id4k ,.1'fn' at), 2xdx(-2i,y.k f- 2m)
-,t- J@),
*, !'oo. l:' .!ffi , @_16)

to which (E-6a) and (E-6b) are immediately applicable when we recall that k2
actually means k2 ie. The k-space integration gives
-
6m : e t',a, S^' a,ffil,.o=_^
:#l',*li'a,ffifO. (E-17)

> m2, as follows:


We integrate over /, remembering A2
("' dt _ I 6,f Lzx I m2(l - x)21
Jo tx+m'(l-x') x',"oL m'(I-x)' I
l, I A2x I
= ; tos lp11;'y_l
:+{.-(#) *,ost#]} (E*18)

(When x is extremely close to zero, the approximation we have made breaks


down but note that we later integrate over x.) The final integration over x is
elementaryi we end up with

Dm-(*)*'"u(#) (E-1e)

up to an additive constant, which, of course, is insignificant when it appears togeth-


er with a divergent expression.f We may mention that this covariant calculation
is several times shorter than the original calculation of the same quantity by
V. F. Weisskopf who used the old-fashioned perturbation theory.

fln fact, the actual value of the additive constant is characteristic of the particular form
of the cut-off function we have used.
318 APPENDIX E

Vertex correction; the anomalous magnetic moment of the electron. As a second


application of Feynman's method, we consider the four-dimensional integral
which we encountered in connection with the vertex correction (cf. Eq. 4. 446).
Setting p' : p'2 : -ffi', we get

L r(P', P)lrr. o' = -^, it. p : -m

: tez I
I Y'nl-
(2tt)a
i'v '(p,'
k'(k'
n:l'y.
tl.i"t 'tp , !) + ml'v,.
"; "k)- t2pt:ft)(ft2@ft)- (E-20)

This integral is divergent at both large and small values of lk'z l. We therefore
modify the photon propagator as follows:

i_-Gfu;XF+,) =ffi
+iY' \P
LL)
k' f tr|,," k, + A2 - J
^,^,^(k,
where ),,,,in is a fictitious "photon mass" assumed to be much smaller than m.
We then have

Lr(P', P)lrr.o':-^,ir.p:-m: iez , (E-22)


I @ Iir,,Or*
with
N,,: "y,[-i^t.(p' - k) + m],yr[-i"y.(p - k) + mlny,, (E-23a)
D : (k' + t)'(k' - 2p.k)(k' - 2p'.k). (E-23b)

Using (4.341), we can simplify the numerator Nr:l


Np : 2"y. p"yrry. p' - + ,y.k"yr"y. p')
2("y. p"yr"y.k

- 2iml,yr,y.(p - + p' 2k) + ,y.(p + p' - 2k),yrl


* 2,y.kyr,y.k - 2,yr*'. (E-24)
As for the denominator we can combine the various factors in D, using (E-12):

+ : u I',* l,-' n, . (E-2s)

We propose to perform the indicated integrations in the order k, t, !, and x. The


k-space integration can be performed immediately if we take advantage of (E-7):

I dnk Np
J (2r)alk' - 2k.(px * p'y) + t(l - x - y)ln

:, iM* (E-26)
-

fThe reader may wonder why we do not simplify further the third term of (E-24) using
,yp@.b) *(,y.b),yu:Zbp. Thereason is thateventuallywewishtoexpress ltp(p',p) in
terms of qro* and ryr, not in terms of pu * p', and ,y r.
APPENDIX E 3I9

where
Mr:2"y. p,yrry. p' - p'y) + ,y.(px I p,y)ryr,y. p,l
2W. p"yr"y.(px *
-2iml,yrry.@ * p' -2px - 2p'y) * ry.(p * p' -2px - 2p,y),y,1
*21,y. (p, + p' y)ry rry.@x i p'Dl
-2lt(l - x - y) - (px * p'y)'1"y, - 2m'yp. (E-27)
We can eliminate p and p' in favor of m, q, and q2, remembering p' : p + q,
p. p' :
-lm' * (q'l2)l and keeping in mind that -i'y. p(-i"y. p') standing on the
right (left) can always be replaced by m,for example,
iry. p' lrisht + -m * i"y. q. (E-28)
We then have

M*:,y,{-2t(l - x-y)-4m,1(x+y), - (l - x-y)l-2qrxy}


* imryr,y.ql-2x + 2(x -f y)yl * im,y.lt,ytl2y - 2(* + Di
* "y.8"tpnt.q(-2xy * 2x + 2y - 2). (E-29)
We further note that the physically observable effects such as the magnetic moment
must be independent of the particular gauge which we use for the external poten-
tial. So we may as well assume that the external potential satisfies the Lorentz
condition which reads in momentum space as follows:
q,A,(q) :0. (E-30)
This means that we can replace I
by -,y. Q,y r(or vice versa)
pnl
- q since the difference
is just 2qrwhichgives zero when contracted with Ar@). Using
nlpf .Q r- -ry.Qryp,
f ' Qf pnl' q -----> - Q'nf , (E-31)
we can write (E-29) as

M, :,yr{-2t(I y) - 4m'l(x + y), (1 x y)l


- x - - - - - 2qr(l
-x-y)j
_, (,y. q,yr :,yr,y. q) l_2m(x + /Xl _ x _ y)1. (E-32)

Recall that A, is given by


L r( P', P)lrr. o' = -*, i t. p : -m

: - #I',n* l,-"a, I)'*^il . (E-33)


It is now clear that when the x-, y-, and /-integrations are performed, A, can be
written asf
Lr(p', P)lrr.o,:-,o,it.p=-m : |vrGr(Q') + (tlZm)q,o,rGr(q'). (E-34)

$This form of A, can be shown to follow from Lorentz invariance, parity conservation,
and charge conservation (gauge invariance) alone.
320 APPENDIX E

As we explained in the main text, the Schwinger correction to the magnetic


moment is given by the coefficient of the q,o,u-term at small values of q'.So let us
calculate Gr(0):

G,(o) : f,o{r*Y !"a* l,-" a, Ii'*,,a, . (E-35)


The /-integration gives
1
dt
J:; lr(l - x-!)+m'(x*y)'l' l-x-y
1 I
X
tI1,*(1 -x-y)+m'(x*y)' A'(l - x-!)+m'(x*y)'
(E-36)

The integral is completely convergent at the upper limit; we can safely let A2 ----+ oo.
We may also be tempted to set trl,,,, :
0, but, because the domain of the x-y integra-
tion includes the point x 0,: !:
it is safer to keep ),i,,i,, 8t this stage. Thus

G,(o) :ry I"* f,-', a, . @-37)

Now this integral is seen to be finite even if l,i,i,, : 0, sinqe

I ,* I'o-"
o, : +, (E-38a)

!',a* l,-' o, #,: - J, "* x dx : r. (E-38b)

Itis therefore legitimate to set tr3,,r., : 0 after all. This would not be the case if the
numerator in (E-37) contained a term independent of x and y; in fact, in integrating
some of the ry, terms in (E-32) it is essential to keep ),1i" to obtain a result that goes
as log(mz/I|,r) (which is responsible for the "infrared catastrophe" as },ili,, -' 0).
With tr3,i. : 0, the integral (E-37) can be trivially performed using (E-38). We get
G'(0) : al2r (E-3e)
'
which is just Schwinger's result.
We emphasize again that the anomalous magnetic moment which we obtain
from (E-35) becornes completely independent of the cut-off A and the fictitious
photon mass 1,,,,i1 &S we let [ ---+ oo and tr,,.i. + 0. If we are just interested in com-
puting the anomalous moment, it is actually not necessary to modify thE photon
propagator according to (E-21). We have deliberately calculated it in this way
because (a) the method we have used also enables us to calculate Gt(q'z) in a
straightforward way, and (b), using this method, we can easily see how the com-
puted value of the magnetic moment of the electron (muon) would be modified if
the cut-off A were not too high compared to the electron (muon) mass (cf. Problem
4-17). As we mentioned in the main text, G,(0) : L (which is also equal to B
because of the Ward identity) is logarithmically dependent on both A and trurini
APPENDIX E 32t

the logarithmic dependence on A is readily seen to arise from (E-32) and (E-33)
since the integrand contains a term that goes like llt for large values of r. On the
other hand, the dffirence Gr(qz) - Gt(0) [which is the coefficient of the ,y, part of
L{,(p' p)] depends only on },*1n (cf. Eq. 4.462).If any strength remains, the reader is
encouraged to complete the evaluation of the integral (E-33) by computing the
coefficient L and Gr(q') - G'(0) (for small values of q').

*
:
BIBLIOGRAPHY

Throughout the book no attempt has been made to quote original papers. Some of the
most important papers on the quantum theory of radiation and covariant quantum
electrodynamics are reprinted in Schwinger (1958).

A. I. ATHIEZER and V. B. BnnnsrErsKII, Quantum Electrodynamics (translated by G. M.


Volkoff). Interscience Publishers, New York (1965).
H. A. BprHn and E. E. SarpnrBn, Quantum Mechanics of One- and Two-Electron Atoms.
Academic Press, New York (1957).
J. D. BronrcN and S. D. Dnnrl, Relatiuistic Quantum Mechanics. McGraw-Hill Book
Co., New York (1964).
J. D. BronrEN and S. D. Dnrlr, Relatiuistic Quantum Fields. McGraw-Hill Book Co.,
New York (1965).
J. M. BLa,tr, Superconductiuity. Academic Press, New York (1964).
J. M. Brarr and V. F. Wntss<oer,Theoretical Nuclear Physics. John Wiley and Sonr,
c New York (1952).
N. N. Bocollusov and D. V. SHrnrov, Introduction to the Theory of Quantized Fields
(translated by G. M. Volkoff). Interscience Publishers, New York (1959).
E. V. CoNooN and G. H. Snonrr.nv, The Theory of Atomic Spectra,2nd ed. Cambridge
University Press, Cambridge, England (1951).
R. H. Dlcre and J. P. Wtrrrn, Introduction to Quantum Mechanics. Addison-Wesley
Publishing Co., Reading, Mass. (1960).
R. P. FeyNua,N, Quantum Electrodynamics. W. A. Benjamin, New York (1961a).
R. P. FrYuuaN, The Theory of Fundamental Processes. W. A. Benjamin, New York
(1e61b).
S. Gastonowcz, Elementary Particle Physics. John Wiley and Sons, New York (1966).
H. GorosruN, C/asstcal Mechanics. Addison-Wesley Press, Cambridge, Mass. (1951).
W. Hnnrux, Quantum Theory of Radiation,3rd ed. Clarendon Press, Oxford, England
(tes4).
J. D. JacrsoN, C/assical Electrodynamics. John Wiley and Sons, New York (1962).
J. M. Jaucn and F. Ronzu.tcn, The Theory of Photons and Electrons. Addison-Wesley
Publishing Co., Cambridge, Mass. (1955).
G. KALrfN, "Quantenelektrodynamik" in Handbuch der Physik (S. Fliigge, Ed.), Band
V, Teil l. Springer-Verlag, Berlin (1958).
323
324 BTBLTocRAPHY

G. KAr.I-t6N, Elementary Particle Physics. Addison-Wesley Publishing Co., Reading, Mass.


(1e64).
C. Ktrrer, Elementary Statistical Physics. John Wiley and Sons, New York (1958).
F. MrNu,, Introduction to Quantum Field Theory. Interscience Publishers, New York
(1e5e).
E. MrnznAcHER, Quantum Mechanics. John wiley and Sons, New York (1961).
A. MEssm,n, Quantum Mechanics (translated by J. Potter). North-Holland Pubtishing
Co., Amsterdam, The Netherlands (1962).
P. M. Monse and H. Frsnnacn, Methods of Theoretical Physics. McGraw-Hill Book Co.,
New York (1953).
N. F. Morr and H. S. W. MnssEy, The Theory of Atomic Collisions,2nd ed. Clarendon
Press, Oxford, England (1949).
K. Nrsnruv,A, Fundamental Particles. W. A. Benjamin, New York (1964).
W. K. H.,rP.qworsrv and M. Prtrrrtys, Classical Electricity and Magnetism. Addison-
Wesley Publishing Co., Cambridge, Mass. (1955).
M. A. PRrstoN, Physics of the Nucleus. Addison-Wesley Publishing Co., Reading, Mass.
(te62).
P. Rourx, Aduanced Quantum Theory. Addison-Wesley Publishing Co., Reading, Mass.
(1e6s).
M. E. RosE, Elementary Theory of Angular Momentum. John Wiley and Sons, New York
(1e57).
M. E. RosE, Relatiuistic Electron Theory. John Wiley and Sons, New York (1961).
J. J. Saru*tr, Inuariance Principles and Elementary Particles. Princeton lniversity Press,
Princeton, N. J. (1964).
S. S. ScHwEBER, Introduction to Relatiuistic Quantum Field Theory. Row and Peterson,
Evanston, Ill. (1961).
J. Scnwncen, Selected Papers on Quantum Electrodynamics. Dover Publications, New
York (19s8).
E. Sncnb, Nuclei and Particles. W. A. Benjamin, Npw York (1964).
R. F. SrRrarnR and A. S. WrcnrM.lN, PCf, Spin and Statistics, and all that. W. A.
Benjamin, New York (1963).
E. C. Tncnuansu, Introduction to the Theory of Fourier Integrals. Clarendon Press,
Oxford, England (1937).
G. WeNrzu-, Quantum Theory of Fields (translated by J. M. Jauch). Interscience Pub-
lishers, New York (1949).
INDEX
INDEX

A anti-Stoke's line, 53
"arrow" (see also internal line), 2ll-212,
Abraham, M.,64
2ts-2t6
absorption of photons by atoms, 3G37
axial-vector density, 105
adjoint equation, 82
adjoint spinor, 82
Aharonov, Y., 16 B
Aharonov-Bohm effect, 1G19 bare charge,279-280
Alexander, J.,132 bare mass; see also mass renormalization
Allen, J. S., 178 of the electron, 69-70, 270-273
a-matrices, 81-82, I 15-1 16 ofthe photon, 274-278
Anderson, C. D., 133 beta decay, 14, 166-167, 177-178
angular momentum; see also spin B-matrix, 81-82
in the Dirac theory, l0l,ll3-114, Bethe, H. A., 70,230,254,292-293
122-124,152-153 Bhabha, H. J.,261
of the radiation field, 3G-31 Bhabha scattering, 261-2&
angular-momentum selection rule; see bilinear covariants, 104-107
selection rule bispinor, 81
annihilation diagram, 261-262 Bleuler, K., 255
annihilation force, 262-2& Bloch, F.,291
annihilation operator Bogoliubov, N.N., 187
for an electron, 145-148 Bohm, D., 16
for a fermion, 28 Bohr, N., 16, 33,176
for a photon,24-27 Bohr-Peierls-Placzek relation, 63
for a positron, 149-151 Bohr radius, 127, 179
anomalous (magnetic) moment boost; see l-orcntz boost
of the electron, 109, I 15, 289-290, 320 Bose-Einstein statistics, 2, 27
of the proton, 110 bound interaction representation, 263
anomalous (magnetic) moment interaction, Breit, G., 116, 261,267
109-1 10, 289-290 Breit interaction, 260-261
anticommutation relation(s) br e m s st r ahl ung, 229-231
of the Dirac field, 153-154
of the Dirac matrices, 81, 83 C
for electron operators, 72, 146
for fermion operators, 28 C-matrix,l4l
for positron operators, 149 c-number field,l47
antineutrino, 166 C parity; see charge conjugation parity
327
328 rNDEX

CP invariance,lT4 for, 138, 229,231


cross section
CPT invariance, 160, 173-174 matrix element for,215
canonical (conjugate) momentum, 3, 5, conserved current
145 in the Dirac theory, 82-83
Casimir, H., 191 in the Klein-Gordon theory, 10-11,
causality, 33, 58-61, 63-64, 154, 208, 7s-78
282-283 constant of the motion, 113-115
Chamberlain, O.,267 continuity equation, t2-13, 7 6-78, 82-83
Chambers, R. G., 17 Coulomb gauge, 21, 301-304
charge conjugate field, 153 Coulomb interaction, 21, 252-256,
charge conjugate wave function, 14V143 303-304
charge conjugation, 142-143, 153, 173, coupling constant (see also fine-structure
222-223 constant), 157
charge conjugation invariance, 17 3, for beta decay,167
222-223 for A decay, 166
charge conjugation parity, 222-223 for zra decay,l73
charge conservation (see a/so continuity for the pion-nucleon interaction,
equation), 10-13 156-157, 266
charge-current density, 1G-l3, 7 6-78, for p decay, 180
107-1 10 covariance, 6
charge-current four-vector; see charge- of the Dirac equation; see Dirac
current density equation
charge operator covariant Golden Rule, 312-313
for the Dirac field, 148, 150-151 covariant Z-matrix, 201, 213, 312-313
for the Klein-Gordon (scalar) field,77 covariant perturbation theory, 179-299,
charge renormalization, 277 -281, 283, 312-314
285-288 covariant photon (exchange), 255-256
Chew, G. F ., 132 covariant photon propagator, 255-256,
chirality, 168-169 273-279
chronological product, 268 covariant spin projection operator; see
circular polarization (vector), 31, 223-224 projection operator
classical description Crampton, S. 8., 130
of radiation phenomena, 35-39 Crane, H. R.,290
classical field, 1-19, 59-62 creation operator
classical radius of the electron, 49, ll9 for an electron, 145-148
Clebsch-Gordan coefficient, 42, 124 for a fermion, 28
Clifford, W. K., 106 for a photon,24-27
Clifford algebra, 106 for a positron, 149-151
closed loop, 273-274 crossed diagram,2l2
Coester, F.,291 cut-off energy (momentum), 68, 270-271,
commutation relation(s) 294-295
for photon operators, 24 cyclotron frequency, 115
of the radiation field, 33
for radiation oscillators, 24
D
of the scalar field, 73
complex conjugation, 9 Dalitz, R. H., 165
complex scalar field, 9 Darwin, C.G., 88, 125, 193
Compton scattering, 138, 213-215, Darwin term, 88, 119
228-231,239-240 Deaver, B. S., 18
INDEX 329

decay rate; see mean lifetime electric quadrupole transition, 44-45


Delbriick, M.,284 electromagnetic interactions, I 57
Delbriick scattering, 284 electron-electron scattering ; see Mgller
destruction operator; ree annihilation scattering
operator electron-positron annihilation; see pair
detailed balance,4T annihilation, two-photon annihilation
Deutsch, M.,227,263 electron-positron scattering ; see Bhabha
dielectric constant scattering
of the vacuum, 281-282 electron propagator, 208-210, 233-239,
dimension, 180-181 271-273
of the field operators, 181 electron-proton scattering, 298-299
dipole-dipole interaction (see also electron scattering
hyperfine splitting), 260 by an external potential, 188-189
Dirac, P. A. M., I , 24,'17, l3l-133, 219, electron self-energy; see self-energy
278-279,295 elementary particle interactions
Dirac equation, 78-86 classification of, 156-157
covariance of, 96-100 emission of photons by atoms, 37-39
free-particle solutions of, 89-94 energy-momentum tensor, 18
nonrelativistic limit of, 85-88 energy shift
plane-wave solutions of, 91-94, 30G307 of an atomic level, 66-69
Dirac field; see also quantization equivalent (vector) potential, 37-39
Hamiltonian of, 145 Euler-Lagrange equation, 5
Lagrangian of, 145 Euwema, R. N., 283
plane-wave expansion of, 145, 155 Evans, R. D., 231
Dirac matrices (see also gamma matrices), exchange scattering, 247
80-85, 104-107 , 305-3 1 1 exclusion principle; see Pauli exclusion
Dirac operators, ll2-ll7 principle
Dirac-Pauli representation ; see gamma external lines, 313
matrices
Dirac sea,132-134 F
Dirac spinor, 81
direct scattering,24T Fairbank, W. M., 18
disconnect ed gr aph, 243 Fermi, E.,20,130, 166, 254,304
dispersion relation, 58-64, 282-283 Fermi-Dirac statistics, 2, 28, 152
Doll, R., 18 Fermi selection rule, 167'
Drell, S. D., 296 Fermi theory of beta decay, 166-167
Dyson, F. J., 210,241,268,285,294 Feynman, R. P., 137, 167, 2ll, 215,231,
Dyson-Wick formali sm, 268-269 288, 290, 294, 315-316
Feynman diagram; see also Feynman
rules
E
for Bhabha scattering, 261
effective potential, 258-261 for bremsstrahlung, 229
Einstein, A., 23, 47, 226 for Compton scattering, 215
Einstein-Podolsky-Rosen paradox, for Delbriick scattering, 284
225-227 for electron scattering by an external
electric dipole approximation, 4l potential, 190
electric dipole moment for electron self-energy, 65, 268-269
of the electron, 178 for modification of the electron
electric dipole transition, 4144 propagator,2T2
330 rNDEx

for modification of the photon G


propagator,2T3
g-factor; see gyromagnetic ratio
for Mpller scattering, 255,278
gamma (7") matrices, 8G-81, 83-85,
for nucleon-nucleon scattering,
257, 305-311
246
pair production, 230 Dirac-Pauli representation of, 80, 84,
for
305-306
for radiative corrections, 285, 288
Majorana representation of, 85, 141
for scattering of light by atoms, 48
standard representation of, 80, 84,
for scattering of light by light, 284
305-306
for three-photon annihilation, 228
Weyl representation of, 84, 169, l7 5
for two-photon annihilation, 212
?5 matrix, 105
for vacuum polarization, 27 6
Gamow-Teller selection rule, 167
Feynman graphs; see Feynman diagrams
Garwin, R. L., 165
Feynman integrals, 315-321
gauge transformation, ll, 14-16, 20, 301
Feynman rules, 215, 241, 246-247, 256,
of the first kind, 11
267-268,313-314
field concept, 1-3 ofthe second kind, 14-16
Gaussian (cgs) units, 12
field operator,29
Gell-Mann, M., 11, 59,167
field tensor, 12
fine-structure constant, 12, 157 , 179
Goldberger, M. L.,59
fine-structure splitting, 88, 127
Golden Rule (see also covariant Golden
Rule), 41
finite charge distribution, 194
Goldhaber, M., 168,229
Finkelstein, R., 173
Gordon, W., 108, 125,193
Fitch, V., 281
Gordon decomposition, 108-109, 286
fluctuations, 32-33, 35-36
gravitational interaction, 1 57
of the electron coordinate; see
Green's function; see also electron
Zitterbewegung
propagator
flux density,7518,83
flux quantization, 17 -18 in the Dirac (Feynman) theory,
232-239
Foldy, L. L.,176
Foldy interaction, 176 for Poisson's equation, 231-232
form factor,l94 in the Schrodinger theory, 298
four-gradient, 6 Grodzins, L., 168
four-vector, 5-6 Gupta, S., 255
gyromagnetic ratio (see also anomalous
four-vector current (density), 10, 7G77,
magnetic moment), 79, 109
82-83, 104-105
Fourier decomposition, 21, ll7, 145,
H
150, 155
Frazer, W. R.,267 Hamilton's variational principle; see
free field, 7 variational principle
free-particle solution ; see Dirac equation Hamiltonian; see also interaction density,
free-particle spinor, 9l-94, 149-150, interaction Hamiltonian
195-196, 306-307, 314 of the Dirac field, 145, 150
French, J. B., 293,294 in the Dirac theory, 81-82, 113
Friedman, J. I., 165 of the Maxwell field, 14
Friedrich, W.,229 of the one-dimensional wave equation, 3
Fukuda, N.,294 of the radiation field,22,29-30,
Fulco, J.,267 302-304
functional derivative, 5 Heavyside-Lorentz units, 12
INDEX 331
Heisenberg, W., 33, 43, 50, 151, 154, 278 isospin, 11,265
Heisenberg representation, ll2, 182-183 It6, D., 291
Dirac operators in, ll2-ll7
Heitler, W., 33, 36,230 J
helicity, 93, ll4-115, 195-196 Jauch, J. M.,291
of the neutrino, 167-170 Jeans, J. H.,23
helicity change (flip) Jordan, P., 28, 33
in Mott scattering, 195-196
helicity eigenspinor, 195 K
Hermitian conjugation, 9
hermiticity of the Hamiltonian, 187 K (electron) capture, 167
Hilbert transform, 61 K meson, 11, 165
Hofstadter, R., 194 K operator,122-124
hole theory,132-144 Kbll6n, G.,211,283
hydrogen Clike) atom, 43-44,88, 125-131 Karplus, R., 263
energy levels of, 88, 127 Kemmer, N., 265
hypercharge, ll Kirchhoff's law, 55
hyperfi ne splitting, 129-130, 263-264 Klein, 4.,263
hyperon decay, 159-166, 240-203 Klein, O., 7, 120, 138, 298
Klein-Gordon equation, 7, 75-78
Klein-Nishina formula, 138, 229
I Klein's paradox, 120-l2l
indefinite metric, 255 Kleppner, D., 130
index of refraction, 62-63 Koba, 2.,270,291
induced emission, 38 Kramers, H. A., 50, 58, 140,270
infinitesimal rotation; see rotation Kramers-Heisenberg formula, 49-50,
infrared catastrophe (divergence), 288, 55-56, 70
290-291,320 Kramers-Kronig relation (see also
instantaneous Coulomb interaction dispersion relation), 58, 63
(potential), 252-253, 301, 304 Kroll, N. M., 294
interaction density; see also interaction Kronig, R., 58-59
Hamiltonian Kusch, P., 109, 290
for beta decay,167
for A decay, 159-161
L
for zr+ decay,lTl Lagrangian; see also interaction density
for p decay, 180 in classical mechanics, 3-5
for the Yukawa coupling, 7-8, 156, of the Dirac field, 145
242,264 of the Maxwell field, 13-14
interaction Hamiltonian1' see also of the scalar field,4,9-11
interaction density Lagrangian density; see Lagrangian,
for atomic electrons, 36 interaction density
for the Dirac electron, 107-109, Lamb, W. E., 71,129,294
154-155, 304 Lamb shift
interaction representation, I 81-l 83 Bethe's nonrelativistic treatment,
intermediate representation; see interaction 7o-72
representation relativistic treatment, 280, 292-294
internal conversion (with pair creation),200 A decay, 159-166, 2AO-203
internal line, 208, 2ll-212,313 Landau, L. D., 170
intrinsic parity; see parity "large" (bilinear) covariant, 107
332 rNDEX

"large" component, 86, 107 mean position operator, 176-177


kderman, L. M., 165 mean square radius, 194
Lee, T. D., 165, 170,174 Meissner effect, 17
level shift; see energy shift meson propagator, 245-246, 266-267
lifetime; see mean lifetime Michel, L.,223
line breadth, 55 mirror reflection, 100
linear polarization vector, 21, 30-31, Miyamoto,Y.,294
5 l-52, 219-222, 225-227 modified propagator
local commutativity (see also causality), for the electron, 272-273
59 for the photon, 273-278
localized state, I l8-1 19 Mpller, C.,250,254
London, F., 16, 18 Mpller scattering, 250-261
longitudinal photon, 254-255 cross section for, 256-258
l-nrentz, H. A., 62,64 matrix element for, 25O-253
Lorentz boost, 101-102, 192 momentum operator
Lorentz force, 15,99, ll4 of the Dirac field, 148, 152
Lorentz transformation, 5-6, 95-99 of the radiation field, 30
improper orthochronous, 100 Moravcsik, M. J.,266
orthochronous, 100 Mott, N. F., 193
Mott scattering, 188-199, 290-291
M cross section for, 192-193
matrix element for, 189-190
//-matrix ; se e cov ariant .til -matrix muon capture, 170
mks rationalized units, 12 muonic atom, 281
Magli6, B. C.,267
magnetic dipole transition, 45
N
magnetic moment (see also anomalous
magnetic moment), 109-1 10, Nd.bauer, M., 18
289-290,320 Nambu, Y., 267
magnetic monopole, 13 natural units, 179-181
Majorana, E., 140 natural width, 55
Majorana representation; see gamma negative-energy solution, 77, 90-94,
matrices lt8-l2l
Mandelstam, S., 258 negative-energy states, 77, 9V94, 104,
Marshak, R. E., 167 llg-121, 1,31-143
mass; see also self-energy neutral scalar field, 6-9, 72-73
ofthe photon, 30-31, 275,288-291 neutrino, 144,166-174
mass renormalization (see a/so self- Newton, T. D., 119
energy), 69-7 l, 268-273 Nishijima, K., 11
Matthews, P. T., 313 Nishina, Y., 138, 298
Maxwell, C., 13 Nordsieck, A.,291
Maxwell equation, 12-14, 169 normal product, 268
Maxwell fields, 12-15 nucleon-nucleon scattering, 242-250,
mean lifetime, 43, 67,296,312 264-267
of excited states of the hydrogen atom, number operator
43-4s for an electron, 150-151
of the A hyperon, 165-166 for a fermion,2S-29
of the zrt meson,172-173 for a photon,24-27
of positronium,22T for a positron, 150-151
INDEX 333

o periodic boundary condition, 22


Petermann, 4.,290
observed charge; seebarc charge, charge photoelectric effect, 23, 73, 230-231
renormalization photon (see also covariant photon and
observed mass; seebare mass, mass
radiation field), 23-31, 256
renormalization photon mass; ree mass of the photon
occupation number, 26 photon self-energy; see self-energy of
occupation number operator; see number
the photon
operator photon states, 26-27, 221-225
Oehme, R., 174
z'+ decay, 172-174
"old-fashioned" perturbation theory
7r meson (see also one-pion exchange),
(see also time-dependent perturbation
9-11, 156,264-265
theory), 248-250 pion; see 7r meson
one-meson exchange, 242-250, 264-267
Pirenne, J.,227
one-pion exchange (potential), 264-266
Planck, M.,23
Onsager, L., 18
Planck's radiation law, 4647
Oppenheimer, J. R., 132 plane-wave expansion; see Fourier
optical theorem, 63, 188 decomposition
Ore, A.,227 plane-wave spinor; see free-particle spinor
orthochronous Lorentz transformation;
Poincar6, H.,'12,64
s e e l-or entz transformation
Poisson's equation, 231-232, 301
polarization
P
of electrons (see alsohelicity), 114-115,
pair annihilation (see also two-photon 196-198,314
annihilation), I 33, 199-200, 237 -239 polarization measurement
pair creation; see pair production of a two-photon system, 225-227
pair production, 133, 200, 229-231, polarization tensor, 273-283
281-283,297 polarization vector; see circular
parametrization method, 3 1 6 polarization, linear polarization
parity (see also space inversion), 42, Pontecorvo, B., 170
gg-100, 103-104, 157-159, l7 5-17 6, positive-energy solution (see also negative-
178 energy solution), 90
of antiparticles, 159 positron, 132-134, 142-143, 207 -208,
of the pion, 163 23s-237
of the positron, 158-159 positron spinor, 149-150
parity nonconservation, 160-166 positroniu m, 227 -228, 263-264
parity selection rule; see selection rule hyperfine splitting of ,263
Pasternack, S., 71 lifetime of ,227
Pauli, W., 2, 11,33,77,78, 132-133, potential ; see Coulomb interaction,
154,275,278 effective potential, equivalent
Pauli exclusion principle, 28, 132-134, potential, one-pion exchange
1.52,224 potential, Yukawa potential
Pauli matrices, 80, 84 Powell, C. F., 9
Pauli moment interaction (see also Powell, J. L.,227
anomalous moment interaction), principal quantum number, 127
109-110 probability conservation, 7 5-7 6, 187, 295
Pauli's fundamental theorem, 83-84, probability current, 7 5-7 6, 83, 105
308-31 1 probability density, 7 5-76, 83, 105
Peierls, R. E., 278 Proca, A., 19
334 INDEX

projection operator renormalized mass; seebare mass, mass


for electron states, l9l-192 renormalization
for positron states, 1.92 representation independence
for spin states, 197-199,314 of the gamma matrices, 83
propagator ; see electron propagator, resonance fluorescence, 53-57
covariant photon propagator, scalar retarded Coulomb potential, 301-302
meson propagator, spin-one meson retarded Green's function; see Green's
propagator function
proton-proton scattering, 267 Retherford, R. C., 71,129
Pryce, M. H., 176 p decay, 180
pseudoscalar density, 105 p matrices, 82
pseudoscalar field, 156 Robiscoe, R. T., 294
pseudoscalar (pseudoscalar) coupling, Rose, M. E.,278
156,264 Rosenbluth formula, 299
pseudoscalar (pseudovector) coupling, Rosenfeld, L., 33
156,264 rotation, 95, 100-101, 194-196
pseudovector density ; see axial-vector Ruderman, M., 173
density Rutherford cross section, 193
"pure" Dirac particle, 110 Rydberg energy, 71,179

a S
4-number field,l4l
S-matrix, 184-188
quadrupole moment
Sakata, S., 200
ofthe deuteron,266
Sakurai, J. J., 167, 267
quantization
Salam, A.,170,294
of the Dirac field, 143-156
scalar field, 5-l l, 72-73, 156
of the radiation field, 29-35
scalar meson exchange, 242-250
quantum field (see a/so Dirac field,
scalar meson propagator, 266-267
field operator, quantization), 1-3
scattering; see also Bhabha scattering,
quantum theory of radiation,2VT4
Compton scattering, Delbri,ick
scattering, Mpller scattering, Mott
R
scattering, proton-proton scattering,
radiation damping, 54 Rayleigh scattering, Thomson
radiation field, 2G-36, 301-304 scattering
radiation gauge, 21, 301-304 of the electron by an external potential,
radiation oscillator, 23 1 10-1 1 1, 188-199, 290_29t

radiative correction s, 284-29 I of light by atoms, 47-63


Rainwater, J., 281 of light by light, 284
Raman, L. V., 53 Schrodinger, E., 7, ll7
Raman effect, 53 Schrodinger equation, 16
Ramsey, N. F., 130 Schrodinger representation, ll2, 182
Ravenhall, D., 223 Schur's lemma, 308-309
Rayleigh, Lord, 23,50 Schwartz, M., 170
Rayleigh scattering, 50-51 Schwinger, J., 109, 182,241,266,
Rayleigh's law, 50 29V29t,320
reduction technique, 285 seagull graph, 4849, 137
renormalized charge; see bare charge, second quantization, 148
charge renormalization Segrb, E.,267
INDEX 335

selection rule spontaneous emission, 38, 4145


for atomic transitions, 4l-42, 44-45 standard representation; see gamma
for beta decay,167 matrices
for two-photon annihilation, 222-225 Stern, O., 110
self-energy; see also mass renormalization Stoke's line, 53
of a bound electron, 65-68, 7V72 strangeness, ll
of the electron, 64-70,268-273, strong interactions, 157, 180
316-317 Stueckelberg, E. C. G., 211,246
ofthe photon, 273-278 subsidiary condition, 19
Selove, W., 132 Sudarshan, E. C. G.,167
semiclassical theory; see also classical Sunyar, A. W., 168
description superconductor, 17-18
of radiation,3T-39
Serber, R., 278
T
Shaknov, 1.,227
)o decay, 73, 180-181 I-matrix, 185, 188
o matrices ; see Pauli matrices I-product; see time-ordered product
singleJevel resonance formula, 56, 74 Tani, S., 176
"small" (bilinear) covariant, 107 r-0 puzzle,165
'osmall" component, 86, 107 Telegdi, V. L., 165
Smekal, A., 53 tensor, 6
Sommerfeld, A., 1,127 tensor density, 105
Sommerfield, C., 290 tensor force, 265-266
space inversion (see also parity), 99-100, tensor operator, 265
156-159 0 function, 206
space-time approach (of Feynman), Thirring, W., 59
232-241 Thomas, L. H., 87
space-time diagram; see Feynman diagram Thomas-Reiche-Kuhn sum rule, 7 4
spectral line Thomas term, 87*88, 261
broadening of, 55 Thomson, J. J., 51
spin Thomson scattering, 51-53, 134-138
of the electron, 78-79, 90, 100-101, three-photon annihilation (of an electron-
113-115,152-153 positron pair), 227-228
ofthe photon, 30-31 time reversal,lT3-174
spin-angular function, 124 time-dependent perturbation theory,
spin-magnetic-moment interaction (see 3941,184
also magnetic moment interaction), time-like photon, 254
36,78-79 time-ordered graph, 207, 240, 244
spin-one meson propagator, 266-267 time-ordered product, 208, 245, 25t
spin operator Toll, J. S., 59
in the Dirac theory, 113-115,152-153 Tomonaga, S., 182, 241, 270, 291, 294
spin-orbit force; see spin-orbit interaction total charge operator; see charge operator
spin-orbit interaction , 88, 260-261, trace (of a product of gamma matrices),
267, 292 192-193,257,306
spin-orbit potential; see spin-orbit transition matrix; see T-matrix
interaction transition probability, 4041, 184, 312
spin projection operator; see projection transversality condition, 20-21, 301
operator for spin states transverse electromagnetic field (see also
spin sum, 191-193 radiation field), 20-22, 301-304
336 INDEX

transverse-photon exchange, 251-252 vertex correction , 284-291, 3 I 8-321


two-body decay vertex factor, 313
transition rate for, 203,312 Villars, F.,275
two-body reaction virtual electron-positron pair, 138-140
differential cross section for, 217, 313 virtual photon, 67, 256
two-component neutrino theory, 167-176 von Neumann, J., 106
two-component wave equation (see also
two-component neutrino theory),
79-80, 91, "1.69, l7 5-17 6
w
two-photon annihilation (of an electron- Waerden, B. L. van der, 79-80
positron pair), 204-213, 216-227 Waerden equation ; see two-component
matrix element for,2l3 wave equation
cross section for, 219 Walker, W. D., 267
Ward, J. C., 288
U Ward identity, 288
wave-function renormalization (constant)
U-matrix, 183-184 for the electron, 273,285
Uehling, E. A., 139,280 for the photon, 278
Uelrling effect, 1.39, 280-281
weak interactions, 157-174, 180
uncertainty relation (for photons), 34-35
Weisskopf, V. F., 54,77, 270, 294, 317
unitarity, 185-188, 281-282 Weyl, H.,132,170
units; see Gaussian units, Heavyside- Weyl equation, 1,69-170
Lorentz units, mks rationalized
Weyl representation; see gamma matrices
units, natural units
Wheeler, J. A., 226,283
Wick, G. C., 208,268
v Wick time-ordered product; see time-
V-Ainteraction, 167 ordered product
vacuum Wigner, E. P., 28, 54,119,173
in the Dirac hole theory,132-133 Wigner-Eckhart theorem, 42
vacuum fluctuation Wilson, R. R., 284
of the radiation field, 35-36 Wolfenstein, L., 223
of the scalar field, 73 Wouthuysen, S. 4.,776
vacuum polarization, 138-139, 27 6-281, Wu, C. 5.,765,227
284-28s
vacuum state,27 Y
van der Waerden; see Waerden
van Kampen, N. G., 59 Yang, C. N., 132,165, 170, 174,224
variational principle, 3-5 Young's modulus, 4
vector; see four-vector Ypsilantis, T., 267
vector meson, 26G267 Yukawa, H., 2, 8, 156, 200
vector meson exchange, 266-267 Yukawa coupling, 7, 156, 242, 264-265
vector meson propagator; see spin-one Yukawa interaction; see Yukawa coupling
meson propagator Yukawa potential, 7 -9, 264-265
vector potential, 13-18, 20-22, 36-39,
301-304 Z
vector spherical harmonic, 45
velocity operator, ll5-ll7 Zit t erbewe gung, ll7 -119, 1 39-140

ABCDD6987

Potrebbero piacerti anche