QFT Ii
QFT Ii
QFT Ii
230A – Part II
J. Gunion
The S-matrix
• Ideally, we would like to solve the coupled differential equations that result
from introducing interactions among the free fields, as for instance E&M
interactions found by the minimal substitution rules.
• An exact solution has not been found, but we have made much progress
using a perturbative approach. So far, we have been fortunate that nature
has always allowed such a perturbative approach to be fruitfully compared
to experiment.
QED, in particular, has passed extremely precise tests because we can
compute very precisely due to the small size of α ∼ 1/137.
• The solution to the full problem can be formulated using the Dyson
expansion which is ideally suited for obtaining perturbation results systematically
L = L0 + LI (1)
with
1
L0 =: ψ(x)(i∂ γµ − m)ψ(x) − (∂µAν )(∂ µAν ) :
µ
(2)
2
(using the Lorentz condition formulation for the Aµ fields) and
LI =: eψ(x)Aµγµψ(x) : . (3)
S
S hB, t|O |A, tiS =H hB, t|O H (t)|A, tiH . (9)
d
i O I (t) = [O I (t), H0] (14)
dt
d d h i
i |A, tiI = i e+iH0(t−t0)|A, tiS
dt dt
+iH0 (t−t0 ) +iH0 (t−t0 ) d
= −H0e |A, tiS + e i |A, tiS
dt
= −e+iH0(t−t0)H0S |A, tiS + e+iH0(t−t0)H S |A, tiS
= e+iH0(t−t0)HIS |A, tiS
where U 0 is a unitary operator that takes us from the H.P. to the I.P. for
any operator (not just the Hamiltonian).
† †
[φI (~ y , t)] = [U 0 φH (~
x, t), π I (~ x, t)U 0, U 0 π H (~
y , t)U 0]
This in turn implies that we will 2nd quantize the I.P. fields in a manner
that is essentially identical to what we did in the free field case.
– Note: there is a subtlety since the asymptotic states in the I.P. can have
a different mass and different normalization than the asymptotic states
in the free-particle case.
This difference arises because the asymptotic states must contain not
just the free particle behavior, but the full collection of virtual processes
associated with the presence of interactions (think bare line + line with
all possible insertions, loops, ....).
This is a subtlety that we will eventually address, but not in detail until
we come to renormalization theory.
• At this point, MS simplifies the notation a bit and writes in the I.P.
d
i |Φ(t)i = HI (t)|Φ(t)i (19)
dt
J. Gunion 230A – Part II 8
where
HI (t) ≡ eiH0(t−t0)HIS e−iH0(t−t0) = HII (t) ; (20)
that is, we will be dropping the I superscript reminding us that everything
is in the I.P.
• It will also be important to note that if HIS is a product of fields in the S.P.
(i.e product of operators in the S.P.), then HI will be the same product of
fields/operators where all the fields/operators are in the I.P. (This is shown
by inserting a whole bunch of 1’s in between the operators in the form
1 = e−iH0(t−t0)eiH0(t−t0).)
• If we start with a state at some initial time, |Φ(ti)i = |ii, the solution of
Eq. (19) gives the state |Φ(t)i at any other time t.
Since HI (t) is hermitian, this time development is given by a unitary
transformation. Accordingly, it preserves the normalization of the state:
• This formalism we are going to develop will not be appropriate for bound
states. It will apply to the situation where we consider an initial state in
in terms of which X
|Φ(∞)i = |f iSf i . (26)
f
The unitarity of the S-matrix can be written in this basis as
X
|Sf i|2 = 1 , (27)
f
as follows from
• We must solve
d
i |Φ(t)i = HI (t)|Φ(t)i (29)
dt
with initial condition |Φ(−∞)i = |ii. Equivalently we must compute
Z t
|Φ(t)i = |ii + (−i) dt1HI (t1)|Φ(t1)i . (30)
−∞
This form obviously satisfies the above differential equation and has the
correct initial boundary condition at t = −∞.
The perturbative series is based on the iterative solution of this equation
where we plug in for |Φ(t1)i the same form as given above for |Φ(t)i and
so forth.
Z t
|Φ(t)i = |ii + (−i) dt1HI (t1)|ii
−∞
yielding (as t → ∞)
∞
X Z ∞ Z t1 Z tn−1
S= (−i)n dt1 dt2 . . . dtnHI (t1)HI (t2) . . . HI (tn) .
n=0 −∞ −∞ −∞
(32)
• Now comes a very crucial trick that partly motivates the use of time
ordering. The above expression can be rewritten as
∞ ∞ ∞ ∞
(−i)n
X Z Z Z
S= dt1 dt2 . . . dtnT {HI (t1)HI (t2) . . . HI (tn)} .
n=0
n! −∞ −∞ −∞
(33)
Here, the time ordering operation has been generalized to include an
arbitrary list of operators. It is necessary to insert the T instruction, since
in the original form, Eq. (32), the HI operators (which do not commute
with one another when evaluated at different times) were very definitively
Let me show how one arrives at this using the 2nd order term as an example.
We write (dropping the subscript I for the moment):
Z t Z t1
dt1 dt2H(t1)H(t2)
−∞ −∞
Z t Z t
= dt2 dt1H(t1)H(t2) interchange of integration order
−∞ t2
Z t Z t
= dt1 dt2H(t2)H(t1) relabel t1 ↔ t2
−∞ t1
Z t Z t
= dt1 dt2T {H(t1)H(t2)} no change because of T instruction .
−∞ t1
(34)
Z t Z t1 Z t Z t1
dt1 dt2H(t1)H(t2) = dt1 dt2T {H(t1)H(t2)} (35)
−∞ −∞ −∞ −∞
∞ ∞ ∞ ∞
(−i)n
X Z Z Z
S = dt1 dt2 . . . dtnT {HI (t1)HI (t2) . . . HI (tn)}
n=0
n! −∞ −∞ −∞
∞
(−i)n
X Z Z Z
= d4 x 1 d4 x 2 . . . d4xnT {HI (x1)HI (x2) . . . HI (xn)} .
n=0
n!
(37)
R
where the d4x integrals are over all space and all times.
Wicks Theorem
• The only way to make this language clear is to give a specific example.
MS chooses e−(p) + γ(k) → e−(p0) + γ(k0) at lowest possible order in
perturbation theory.
p)a†s(~
|ii = c†r (~ k)|0i (38)
|f i = c†r0 (~
p 0)a†s0 (~
k0)|0i . (39)
• So, now the trick is to go to the form of S given in Eq. (33) and to use
an n such that there are enough HI ’s in between hf | and |ii as to get a
non-zero result. If you insert the minimal number, i.e. use the smallest n,
then this defines the “tree-level” amplitude.
Since hf | has a different photon in it than does |ii, you will need two
A fields, one to annihilate the initial photon and one to create the final
photon. This means you will need two HI ’s.
• For now, let us not worry about these two extra guys and focus only on the
fields necessary to get from the initial state to the final state.
The only non-zero contribution to
hf | −e : ψA/ ψ :x1 −e : ψA/ ψ :x2 |ii (40)
will come from the structure (notice I am not yet specifying whether the
fields below are at x1 or x2 – we will come to the appropriate mixtures):
− −
hf |ψ A−ψ +A+|ii = h0|as0 (~
k0)cr0 (~ p)a†s(~
p 0)ψ A−ψ +A+c†r (~ k)|0i (41)
where the A+ has an a in it to “kill” the a†s(~ k) used to define |ii and ψ +
has the c in it to “kill” the c†r (~
p) that is also part of defining |ii. Similarly,
−
the c† in ψ will “kill” the cr0 (~ p 0) and the a† in A− will “kill” the as0 (~
k0).
By “kill” I mean the operation (to give one of the 4 kills)
ai(~l)a†s(~
k)|0i = [ai(~l), a†s(~
k)] + a†s(~
k)ai(~l) |0i = δisδ~l,~k|0i + 0 . (42)
X 1
A µ+
(x2) = q µ
i (~l)ai(~l)e−il·x2 . (43)
i,~
l
~
2V |l|
• Note: The ordering which makes the killing simple, as given above, is the
normal ordering of the required operators.
Thus, in general, we will want to expand the S-matrix as a sum of normal
products. The method for doing so is due to Dyson and Wick.
+
where the reordering is such that all absorption operators (i.e.
components) are to the right of all creation operators (i.e. − components).
[A+, B −]+ ,
for two fermion fields
AB− : AB := (45)
[A+, B −] , otherwise
where the next to last line follows since h0|A− = 0 and B +|0i = 0. The
commutator case follows a similar sequence of steps:
AB =: AB : +h0|AB|0i. (49)
: AB : = : (A+ + A−)(B + + B −) :
= A+ B + + A− B + + A− B − − B − A+
= −B +A+ + A−B + − B −A− − B −A+
= − : (B + + B −)(A+ + A−) :
= − : BA : , (50)
T {A(x1)B(x2)} = A(x1)B(x2)
= : A(x1)B(x2) : +h0|A(x1)B(x2)|0i
= : A(x1)B(x2) : +h0|T {A(x1)B(x2)}|0i (51)
T {A(x1)B(x2)} = ±B(x2)A(x1)
= ± [: B(x2)A(x1) : +h0|B(x2)A(x1)|0i]
= ± [± : A(x1)B(x2) : ±h0|T {A(x1)B(x2)}|0i]
= : A(x1)B(x2) : +h0|T {A(x1)B(x2)}|0i , (52)
i.e. exactly the same result since the ± sign changes cancel one another.
In short, we have
: ABCDEF . . . J KLM . . . :
= (−1)P AKBCEL . . . : DF . . . J M . . . : . (56)
T {ABCD . . . W XY Z}
In the maximal contraction cases, if there are matching fields then no field
will be left over.
The proof of this theorem is by induction and can be found in the old book
by Bjorken and Drell on Relativistic Field Theory (2nd volume). Another
form of the inductive proof appears at the end of Peskin and Schroeder,
section 4.3. We will not go through the proof here, but will give some
illustration below of how it works in a specific case of interest.
• Finally, we must figure out what the correct procedure is for the case of
interest where we are dealing with
T {A(x1) : B(x2)C(x2) :}
= (A+ + A−)(B +C + + B −C − + B −C + + C −B +)
= A−(B +C + + B −C − + B −C + + C −B +)
+(B +C + + B −C − + B −C + + C −B +)A+
+[A+, B −]C − + B −[A+, C −] + [A+, B −]C + + [A+, C −]B +
= : ABC : +[A+, B −]C + [A+, C −]B
= : ABC : +h0|T {AB}|0iC + h0|T {AC}|0iB
• You should note that this result is critically dependent upon using : HI :
(i.e. normal ordering the interaction Hamiltonian from the beginning) rather
than HI . In the latter case, you would include equal time contractions in
applying Wicks theorem. Using normal ordering for HI , in a certain sense,
obscures what is really happening.
Peskin and Schroeder discuss the situation without normal ordering HI . In
this PS approach (also quite common in other texts), all the contractions of
fields in the same HI with one another are kept. Such contractions, as we
already know, are associated with infinities. We have discussed how such
infinities can be thrown away in the linear free particle context (i.e. when
considering just one H0 in the free particle case). It is quite another matter
to show that they don’t matter when considering HI multiply repeated
hf |S|ii = h0|as0 (~
k0)cr0 (~ p)a†s(~
p 0)Sc†r (~ k)|0i (62)
with the part of S that we keep being the minimal version capable of
yielding the process. As described earlier this is what MS calls S (2), i.e. the
2nd order component:
e2
Z Z
S (2) = − d4 x 1 d4x2T {: ψ 1A/ 1ψ1 :: ψ 2A/ 2ψ2 :} (63)
2
J. Gunion 230A – Part II 33
and, in fact, only certain components of this S (2) that emerge in the Wick
expansion survive: namely, those with two contracted fermion fields with
the others left over to “kill” the creation and annihilation operators defining
the initial state and the hermitian conjugate of the final state:
2Z
e
Z
S (2) 3 − d4x1 d4x2 : ψ 1A/ 1ψ1ψ 2A/ 2ψ2 :
2
e2
Z Z
− d4x1 d4x2 : ψ 2A/ 2ψ2ψ 1A/ 1ψ1 : (64)
2
From the symmetry under 1 ↔ 2 above, it should be obvious that these
two terms are equal and that we need only evaluate one of these two terms
with a factor of −e2 — let us take the 1st term to evaluate.
−
• In this 1st term we must employ the particular fields components ψ 1 3 c†
and ψ2+ 3 c. We also need one A+ 3 a and one A− 3 a†, leaving:
Z Z
−
S (2)
3 −e 2
d x1 d x2 : ψ 1 A/ −
4 4
1 ψ1 ψ 2 A/+ +
2 ψ2 :
Z Z
− − +
−e2 d4x1 d4x2 : ψ 1 A/ + ψ
1 1 2ψ A/ 2 ψ2 : (65)
X 1
ψ +
(x)c†r (~
p)|0i = p ct(~ q )e−iq·xc†r (~
q )ut(~ p)|0i
q
~,t
2V Eq~
1
= p p)e−ip·x|0i
ur (~ (66)
2V Ep~
q ), c†r (~
after using [ct(~ p)]+ = δrtδq~,~p.
Similarly, we find
− 1
c†t (~
X
0 0
p )ψ (x) = h0|cr0 (~
h0|cr0 (~ p ) p q )e+iq·x
q )ut(~
q
~,t
2V Eq~
1 0 +ip0 ·x
= p ur0 (~
p )e h0| . (67)
2V Ep~ 0
X 1
A µ+
(x)a†s(~
k)|0i = p µ
t (~
q )at (~
q )e−iq·x † ~
as(k)|0i
t,~
q
2V |~
q|
1 ~ −ik·x|0i ,
= q µ
s (k)e (68)
2V |~
k|
and
X 1 †
~0
h0|as0 (k )A µ−
(x) = h0|as0 (k ) ~0 p µ
t (~
q )a q )e+iq·x
t (~
t,~
q
2V |~
q|
1 0 ik0 ·x
= q µ
s0 (~k )e h0| . (69)
2V |~
k0|
So let us use these results to first evaluate the contribution of the first term
in the part of S (2) that appears in Eq. (65). Using h0|0i = 1, we obtain,
dropping temporarily the √21V ... factors:
i 4 q/ +m
h Z i
0 ~0 2 4 4 0 0 4 4
= ur0 (p s0 (k ) −e
~ )/ d q 2 (2π ) δ (k + p − q )(2π ) δ (q − p − k)
(2π )4 q − m2
s (~
/ k)ur (p
~)
2 4 4 0 0 q
/ +m 0 ~0
= −e (2π ) δ (p + k − p − k)ur0 (p
~ )/
s 0 (k ) i 2 2
s (~
/ k)ur (p
~)
q − m q =p +k
1
×q q q q (70)
2V Ep
~ 2V Ep
~0 2V |~
k| 2V |~
k0 |
+ − 1 X 1 ~ † ~0 −ik·x+ik0 ·y
[φ (x), φ (y)] = √ [a(k), a (k )]e
2V ~
ω~kω~k0
k~k0
1 d3~
k
Z
V →∞
→ e−ik·(x−y)
2(2π)3 ω~k
≡ i∆+(x − y) . (71)
1 1 −ik·(x−y )
We could have simply written i∆+(x − y) =
P
2V k ω~ e
~ . When
k
0
we reexpressed as a contour integral, the dk would still have been a
3~
continuous integral, but the (2dπk)3 would have been replaced by V1 ~k . And
P
(the dx01dx02 would still be over all time and you would have Dirac delta
functions from these integrals still). The final stages of the calculation
would have then been:
ix1 ·(k0 +p0 )−ix2 ·(k+p)
h Z Z i
0 ~0
2 4 4
ur0 (p
~ )/ s0 (k ) −e d x1 d x2 ψ1 ψ 2 e s (~
/ k)ur (p
~)
i q
/ + m −iq·(x1 −x2 )
h Z Z
0 0 2 4 4 0
X
= ur0 (p
~ )/ ~
s0 (k ) −e d x1 d x2 dq e
(2π )V q 2 − m2
q
~
i X q
/ +m
h Z i
0 2~0 0 0 0 0 0
= ur0 (p s0 (k ) −e
~ )/ dq (2π )δ ([k + p − q ] )(2π )δ ([q − p − k] )
(2π )V q 2 − m2
q
~
n o
2
× V δ~k 0 +p
~ 0 ,~
δ~,~
q q p+~
k
s (~
/ k)ur (p
~)
2 0 0 0 q
/ +m 0 ~0
= −e (2π )V δ ([p + k − p − k] )δp
~ 0 +~
k 0 ,~
p+~
k
ur 0 ( p
~ )/
s 0 (k ) i 2 2
s (~
/ k)ur (p
~)
q − m q =p +k
1
×q q q q (73)
2V Ep
~ 2V Ep
~0 2V |~
k| 2V |~
k0 |
a − ~b) .
V δ~a,~b → (2π)3δ(~ (74)
In any case, putting aside the normalization factors and exact version
of three momentum conservation delta function, the structure we have
obtained makes a lot of sense. We observe an over all coupling strength
of e2, overall 4-momentum conservation δ functions, and a very specific
algebraic structure that can be associated with a momentum-space Feynman
diagram as shown in the first diagram of the figure.
/0 1 3456
1 /(0
- .
!
8
5 19
. -
/20 7 /(0
"#$&%('*))+, /#5$&%'*))+:,
i 4 q/ +m
h Z i
0 ~ 2 4 4 0 4 4 0
= ur0 (p s (k) −e
~ )/ d q 2 (2π ) δ (p − k − q )(2π ) δ (q + k − p)
(2π )4 q − m2
0
s0 (~
/ k )ur ( p
~)
2 4 4 0 0 q
/ + m 0 0
= −e (2π ) δ (p + k − p − k)ur0 (p s (~
~ )/ k) i 2 /
s 0 (~k )ur (p
~)
q − m2 q =p−k0
1
×q q q q (75)
2V Ep
~ 2V Ep
~0 2V |~
k| 2V |~
k0 |
q/ + m i
i = (76)
q2 − m2 + i q/ − m + i
• One of the key ingredients above is the ieγµ “vertex”. Algebraically, you
can think of this factor as coming from:
1. −i for each power of HI appearing in the required level of S-matrix
expansion;
2. −eγµ as appearing in each HI in association with constructing the A/
appearing therein.
This vertex is associated with each interaction of a photon and a fermion
whether these are real or virtual.
In the above case, the initial or final e− is real, the photons are real, but
the intermediate e− is virtual.
Regardless, one writes down (assuming incoming or outgoing electrons, and
not positrons)
ur (p)ieγµus(q)µ
t (k) (79)
for each vertex, even when one of the particles is virtual.
X i
µ
s 0 (~k 0
)ur 0 (~
p 0
)ieγµut(q) ut (q)ieγ ν ur (~
p) ν ~
s (k)
t
q2 − m 2
q/ + m
= µ
s 0 (~k 0
)ur 0 (~
p 0
)ieγµi p)νs (~
ieγν ur (~ k)
q 2 − m2 q =p+k=p0 +k0
• There are a few more rules that we can only get by considering a few other
processes. For the rules for external positrons, we need the appropriate
“killing” operations. For an incoming positron contained in |ii we have:
+ X 1
ψ (x)d†r (~
p)|0i = p dt(~ q )e−iq·xd†r (~
q )v t(~ p)|0i
q
~,t
2V Eq~
1
= p p)e−ip·x|0i
v r (~ (81)
2V Ep~
q ), d†r (~
after using [dt(~ p)]+ = δrtδq~,~p.
Similarly, for an outgoing positron in |f i, we find
1
d†t (~
X
0 − 0
h0|dr0 (~
p )ψ (x) = h0|dr0 (~
p ) p q )e+iq·x
q )vt(~
q
~,t
2V Eq~
Note how these differ from the corresponding electron results. Where we
have a u(~ p) for an incoming electron we find v(~ p) for an incoming positron,
and where we had u(~ p) for an outgoing electron, we find v(~ p) for an
outgoing positron. As a result, the v and u spinors both appear at the end
of the line directed in the direction of electron (i.e. fermion vs. antifermion)
flow. For instance, an incoming positron has the arrow of fermion flow
directed in the outgoing direction, and one writes v.
i 4 q/ +m
h Z i
= v r (p
~)/ ~0
s0 (k ) +e
2
d q 2
4 4 0 4 4 0
(2π ) δ (k − p − q )(2π ) δ (q + p − k)
(2π )4 q − m2
0
s (~
/ k)vr0 (p
~ )
2 4 4 0 0 0 q
/ + m 0
= +e (2π ) δ (p + k − p − k)v r (p s0 (~
~)/ k) i 2 2
s (~
/ k)vr0 (p
~ )
q − m q =k0 −p
1
×q q q q (83)
2V Ep
~ 2V Ep
~0 2V |~
k| 2V |~
k0 |
Note how, in going from the 1st line to the 2nd line above, I moved the
+
ψ 1 α all the way to the right, past the [. . .]αβ matrix and past ψ2−β . This
is allowed since I retain proper Dirac matrix contractions so long as I keep
track of the α index sums later, as I do. A similar statement applies to
moving the ψ2−β over to the left.
The final expression above can be mapped to the second of the Feynman
This figure will be confusing at first. The arrows show the direction of
fermion flow. But we must remember that the incoming (antifermion)
e+ has momentum p entering the diagram from the bottom right, and
the outgoing e+ has momentum p0 exiting the diagram. I.e. anti-
fermion momenta and fermion flow are oppositely directed. Momentum
conservation for the virtual propagator is correctly given (for the direction
of fermion flow) by k0 − p for the 2nd diagram. That is, we always write
the Feynman propagator in the canonical form where q is given by the
momentum carried in the fermionic direction, rather than the antifermionic
direction.
I will not work out the details of the 2nd contribution, which gives you the
first (left-hand) diagram of the above figure. Using our Feynman rules we
A homework problem will be to verify this result in the same kind of detail
as given for the right-hand diagram.
• One important note on the above. We were careful to keep the sign
resulting from strict application of Wick’s theorem, which ended up giving
us a +e2 overall sign for this positron case as opposed to the −e2 overall
sign for the electron case.
Such an overall sign is not actually observable, since we will eventually
square these amplitudes.
• However, one should be very careful with signs in cases where different
diagrams correspond to the exchange of identical fermions in the final state
An example of this, which is also useful for getting the correct result for a
virtual photon propagator, is provided by e−e− → e−e− scattering.
where two changes of sign occurred when we passed two fermionic operators
past one another in order to get to the explicitly normal ordered form given.
We must now adopt a convention for our |ii and |f i states. We use p, r
and k, s for the incoming guys and p0, r 0 and k0, s0 for the outgoing guys
p)c†s(~
|ii = c†r (~ k)|0i , |f i = c†r0 (~
p 0)c†s0 (~
k0)|0i ,
⇒ hf | = h0|cs0 (~ p 0) ,
k0)cr0 (~ (86)
p)c†s(~
ψ2+ρc†r (~ k)|0i
X 1
= p ct(~l)ut(~l)ρe−il·x2 c†r (~
p)c†s(~
k)|0i (87)
~
2V E~l
t,l
1 h
p p e−ik·x1−ip·x2 us(~
k)β ur (~
p)ρ
2V Ep~ 2V E~k
i
−e −ip·x1 −ik·x2
p)β us(~
ur (~ k)ρ (91)
− −
h0|cs0 (~
k0)cr0 (~
p 0)ψ 2 δ ψ 1 α
1 h
+ik0 ·x1 +ip0 ·x2
= p p e us0 (~
k0)αur0 (~
p 0)δ
2V Ep~ 0 2V E~k0
0 0
i
p 0)αus0 (~
− e+ip ·x1+ik ·x2 ur0 (~ k 0 )δ . (92)
−e2 1
Z Z
(2) 4 4
hf |S |ii = q q q q d x1 d x2 ×
2 2V Ep~ 0 2V E~ k0
2V Ep~ 2V E~ k
" #
0 0
+ik ·x1 +ip ·x2 0 0 0 0
+ip ·x1 +ik ·x2 0 0
e us0 (~ ~ )δ − e
k )α ur0 (p ~ )α us0 (~
ur 0 ( p k )δ
µν
−iq·(x1 −x2 ) −g
i
Z
4
(γµ )αβ d qe (γν )δρ
(2π )4 q 2 + i
" #
−ik·x1 −ip·x2 −ip·x1 −ik·x2
e us (~ ~)ρ − e
k)β ur (p ~)β us (~
ur (p k )ρ (93)
Of the four terms in the above expression, two are equal to corresponding
ones of the remaining two. This will cancel the 21 appearing in e2/2. This
always happens in QED. We saw another example of this in our earlier
calculations. There is equivalence under interchange of the x1 and x2
vertices (i.e under H1 ↔ H2). Let us focus on two inequivalent terms. We
begin with (dropping some external factors for the moment) taking the 1st
" #
+ik0 ·x1 +ip0 ·x2
Z Z
4 4 0 0
d x1 d x2 e us0 (~ ~ )δ (γµ )αβ
k )α ur 0 ( p
" #
µν
−iq·(x1 −x2 ) −g
i
Z
4 −ik·x1 −ip·x2
d qe (γν )δρ e us (~
k )β ur ( p
~)ρ
(2π )4 q 2 + i
µν
−ig
4 4 0 0 ~0 0
= (2π ) δ (p + k − p − k )us0 (k )γµ us (~
k) ur 0 ( p
~ )γν ur (p
~) (94)
q 2 + i q =k0 −k=p−p0
where we did the d4x1 and d4x2 integrals to get (2π)4δ 4(k0 − q − k)
and (2π)4δ 4(p0 + q − p) respectively and used one of these to do the d4q
integral. Now, we take the 2nd term of the 1st large bracket and the 1st
term of the 2nd large bracket. We have
" #
+ip0 ·x1 +ik0 ·x2
Z Z
4 4 0 0
d x1 d x2 −e ~ )α us0 (~
ur0 (p k )δ (γµ )αβ
" #
µν
−iq·(x1 −x2 ) −g
i
Z
4 −ik·x1 −ip·x2
d qe (γν )δρ e us (~
k)β ur (p
~)ρ
(2π )4 q 2 + i
µν
−ig
4 4 0 0 0 0
= ~ )γµ us (~
−(2π ) δ (p + k − p − k )ur0 (p k) us 0 (~k )γν ur (p
~)(95)
q 2 + i q =p0 −k=p−k0
• So, by combining the results for the various processes we have computed
1. Draw all possible tree-level (lowest possible order) diagrams that can give
rise to the process of interest making use of just the fermion → fermion
+ photon “vertex”.
In so doing, keep track of the fermion line direction. An incoming
fermion line must be traced continuously to an outgoing fermion line (or,
equivalently, to an incoming antifermion line).
2. For each vertex, write ieγµ, where a different Lorentz index µ must be
used for each vertex and this µ must be connected either to the µ on a
virtual photon propagator or to the µ of an external photon polarization
vector.
3. For an outgoing photon γ, p, s, write ∗s (~ p), where I here allow for a
complex polarization basis.
4. For an incoming photon γ, p, s write s(~ p).
5. For an incoming e−, p, r write ur (~ p).
6. For an outgoing e−, p, r write ur (~p).
7. For an incoming e+, p, r write v r (~
p).
8. For an outgoing e+, p, r write vr (~p).
9. For the positron cases, note that the momentum direction will be opposite
the fermion arrow direction.
q/ + m
i (97)
q2 − m2 + i
where q is the momentum in the direction of the fermion arrow and the
value of q is obtained by momentum conservation at the vertices.
11. For an internal photon propagator, write
−g µν
i (98)
q2 + i
• It is the last of the Feynman rules above that we have yet to derive using
an explicit example.
We choose to use the virtual e− loop correction to the photon propagator
as our basic example. The basic Feynman diagram that corresponds to
our calculation is below. There, I use the label p for the free momentum
associated with the loop.
"! !
(99)
(100)
µ+ † 1 µ −ik·x
A (x)as (~
k)|0i = q s (~
k )e |0i ,
2V |~
k|
µ− 1 µ ik·x
h0|as (~
k )A (x ) = q s (~
k )e h0| . (101)
2V |~
k|
q
derived earlier, dropping the two 1/ 2V |~k| factors. If I interchange the x1
and x2 variables and relabel dummy indices (µ ↔ ν, α ↔ ρ and β ↔ σ),
then it is easy to see that the 2nd term is identical to the first term.
~ (2) ~ 4 4 1
hγ(r, k)|S |γ(r, k)i = (2π) δ (k − k) q q M ,(103)
2V |~
k| 2V |~
k|
with
" #
4
d p q/ + m p/ + m
Z
M = +e2 Tr i
/ r (~k)i
/ r (~k)
(2π) 4 2 2
q − m q = p+ k 2
p −m 2
(104)
Note that this result has the opposite sign compared to what you would
write down just naively using ieγµ and ieγν for the two vertices and i qq/2−m
+m
2
and i pp/2−m
+m
2 for the internal fermion propagators.
Note also how the Dirac structure is written by starting to the left with
the p + k propagator and then working “backward” to the p propagator as
shown in the earlier figure.
In this final form, we have also used the Trace notation. The Dirac index
structure was such that one ended up with a continuous connection of
• We will consider colliding two particles in the initial state |ii with momenta
pi = (Ei, p
~i), i = 1, 2. (Sorry about the confusion of notation; I am simply
following MS here.) The final state |f i is assumed to contain N particles
with p0f = (Ef0 , p
~f0 ), f = 1, . . . N .
1/2 Y !1/2
X X Y 1 1
Sf i = δf i + (2π)4δ 4( p0f − pi) M
f i i
2V Ei f
2V Ef0
(105)
The δ 4 function in the above is obtained in the limit of an infinite time
interval T → ∞ and an infinite volume, V → ∞. For finite T and V , we
X X lim X X
4 4
(2π) δ ( p0f − pi ) = T →∞ δT V ( p0f − pi)
f i V →∞ f i
lim Z +T /2 Z X X
≡ T →∞ dt 3
x exp ix · (
d ~ p0f − pi) (106)
V →∞ −T /2 V f i
P 0 P
replaced by δT V ( f pf − i pi). In deriving the cross section, it is helpful
to take T and V finite temporarily.
• For finite T and V , we can define the transition probability per unit of time
as
w = |Sf i|2/T (107)
h P 0 P i2
which will contain the factor δT V ( f pf − i pi) . For one of these
P 0 P
factors, we revert back to the continuum result δT V ( f pf − i pi) →
4 4
P 0 P P 0
(2π)
P δ ( f pf − i pi). For the other factor, we can take ( f pf −
i pi ) = 0 and use δT V (0) = T V , keeping the finite volume and time
interval.
This is the result for transition from state i to a single final state f . It is
of course vanishingly small in the continuum limit of V → ∞ as the final
state possibilities become a continuum or possibilities.
~f0
Y V d3p
. (109)
f
(2π)3
The factors of V , above, cancel those in the denominator in the final state
part of the expression for w, leaving one V in the numerator from the δT V
and two V ’s in the denominator from the initial state killing operations.
h~
p|~ p)a†(~
pi = h0|a(~ p)|0i = δp~p~h0|0i = 1 , (110)
where we have used the finite volume result for the commutator,
p), a†(~
[a(~ p 0)] = δp~p~ 0 . (111)
And, the incident flux is vrel/V where vrel is the relative velocity of the
colliding particles. To see this, picture a box of volume V traveling with
velocity vrel towards a fixed (for convenience) box of volume V . Orient
the traveling box so that one face is parallel to the front face of the fixed
box. Start at t = 0 with the front face of the traveling box touching the
front face of the fixed box. If the facing faces of the two boxes have area
1 ~f0
Y V d3p
dσ = w
[ vVrel ] f
(2π)3
X X 1 Y ~f0
d3 p
= (2π)4δ 4( p0f − pi) 0
|M| 2
.(112)
f i
4E1E2vrel f
(2π)32Ef
Note that this differs from the result in MS (with the extra 2m factors
for fermions) by virtue of the different normalization for fermion fields and
operators. Using my normalizations (which are the same as essentially all
modern treatments), there is no distinction in the basic normalizations for
bosons and fermions.
~f0 are
Note that the δ 4 function above implies that not all of the p
• The above result holds in any Lorentz fame in which the colliding particles
move collinearly.
In such a frame, and assuming the particles are moving in opposite direction
to one another (it is also ok if one is a rest), the relative velocity is given
by
|~ | |~ |
p1 p2
vrel =
+
E1 E2
1
= |E2|~p1| + E1|~
p2|| . (113)
E1E2
Let us compare this expression with (remember opposite directions for the
colliding particles are assumed)
2 2 2 2 2 2
(p 1 · p 2 ) − m 1 m 2 = (E1 E2 + |~
p1 ||~
p2 |) − m1 m2
2 2 2 2 2 2
= E1 E2 + |~
p1 | |~
p2 | + 2E1 E2 |~
p1 ||~
p2 | − m1 m2
2 2 2 2 2 2 2 2
= (|~
p1 | + m1 )(|~
p2 | + m2 ) + |~
p1 | |~
p2 | + 2E1 E2 |~
p1 ||~
p2 | − m1 m2
In short, we have
2 2 2 1/2
E1E2vrel = (p1 · p2) − m1m2 . (115)
This last result, substituted in the general form of dσ, implies that dσ is a
d3 p
Lorentz invariant. (Recall that (2π)3~2E is a Lorentz invariant.)
p
~
E1E2vrel = |E2|~
p1| + E1|~
p2|| = |~
pcom|Etot , (116)
E1E2vrel = |E2|~
p1| + E1|~
p2|| = |~
p1|m2 , (117)
• Since the center of mass system is so common and since we often consider
2 → 2 processes, it is useful to give some results specific to this case.
1 4 0 0 3 0 3 0 2
dσ = 0 0 δ ( p 1 + p 2 − p 1 − p 2 )d p
~1 d p
~2 |M|
pcom |Etot (2π )2 E1 E2
16|~
1 0 0 0 2 0 0 2
= 2 0 0 δ (E1 + E2 − Etot )|~ p1 | d|~ p1 |dΩ1 |M|
64π |~ pcom |Etot E1 E2
#−1
∂ (E10 + E20 )
"
1 0 2 0 2
= 0 0 0 |
|~
pcom | d Ω 1 |M|
64π 2 |~
pcom |Etot E1 E2 ∂|~
pcom
q q − 1
1 ∂ (m01 )2 + |~ 0 |2 +
pcom (m02 )2 + |~ 0 |2
pcom
0 2 0 2
= |~
pcom | dΩ1 |M|
pcom |Etot E10 E20
64π 2 |~ 0
pcom |
∂|~
#−1
0 0
"
1 |~
pcom | |~
pcom | 0 2 0 2
= + |~
pcom | dΩ1 |M|
pcom |Etot E10 E20
64π 2 |~ E1 0 E2 0
0
1 |~
pcom | 0 2
= 2 0 0 dΩ1 |M|
64π Etot (E1 + E2 ) |~
pcom |
0
1 |~
pcom | 0 2
= 2 |~
d Ω1 |M| , (118)
64π 2 Etot pcom |
• The above and other results for differential cross sections apply irregardless
of whether or not we are dealing with identical particles.
However, if there are two or more identical particles in the final state, we
must not duplicate integration regions in obtaining the total cross section.
For example, in the 2 → 2 process just computed, if the final particles
were 2 e−’s, then the case where (θ10 , φ01) = (α, β) is not physically
distinguishable from the case where (θ10 , φ01) = (π − α, π + β). In this
case, we should only integrate over 0 ≤ θ10 ≤ 12 π, or, equivalently, we could
integrate over all θ10 and then divide the final result by 2!:
1 dσ
Z
tot
σcom = dΩ01 0
. (120)
2 4π dΩ1 com
In the more general case of n identical particles in the final state, we can
obtain the correct result for σ tot by integrating over all of phase space and
then dividing by n!.
(As we shall discuss, the e− and µ− are not identical particles. They carry
a lepton identifier quantum number that distinguishes them. See MS for
more details.)
What this means in our theoretical framework is that there are creation and
annihilation operators for electrons that are distinct from those for muons.
|ii = c†e(~
k)d†e(~
p)|0i , |f i = c†µ(~
k 0)d†µ(~
p 0)|0i . (124)
and that the terms of interest then arise as the two cross terms containing
the product of HeI with HµI . Of course, the x1 ↔ x2 2nd term above gives
exactly the same contribution as the 1st term and so we will work with the
equivalent form
Z Z
(2) 2 4 4
S 3 −e d x1 d x2 T {: ψ µ (x1 )A / (x2 )ψe (x2 ) :}
/ (x1 )ψµ (x1 ) :: ψ e (x2 )A (127)
We now insert this between our initial and final states and use the standard
killing operations. Thus, we compute (you will have noted that I am not
attempting to keep spin indices on the creation and annihilation operators
defining |ii and |f i)
" Z Z
(2) 0 0 2 4 4 − ρ −
hf |S |ii = ~ )cµ (~
h0|dµ (p k ) −e d x1 d x2 ψ µ (x1 )γ ψµ (x1 )Aρ (x1 )Aσ (x2 )
#
+ σ + † †
ψ e (x2 )γ ψe (x2 ) ce (~ ~)|0i
k )d e ( p
2 0 ρ 0 σ
= −e uµ (~
k )γ vµ (p ~)γ ue (~
~ )v e ( p k) ×
• So, now let’s insert this into our 2 → 2 cross section expression:
0
|~ |
dσ 1 pcom 2
= |M| (130)
dΩ01 com 64π 2Etot
2 |~ pcom|
dσ 1
0 ~0 0
= ρ
~ 0)
uµ(λ , k )(ieγ )vµ(λ , p
dΩ com 64π 2Etot
2
−i
2
~)(ieγρ)ue(λ, ~
v e(λ, p k) (131)
q2
+ i q = p+ k
~)(ieγρ0 )ue(λ, ~
[v e(λ, p k)]∗ = ue(λ, ~
k)†(ieγρ0 )†(ve(λ, p
~)†γ 0)†
k)†γ 0γ 0(−ieγρ†0 )(γ 0)†ve(λ, p
= ue(λ, ~ ~)
= ue(λ, ~
k)(−ieγρ0 )ve(λ, p
~) (133)
so that we have (exposing Dirac indices in the 3rd line before proceeding)
α β 1 α β β α 1
Tr[γ γ ] = Tr[γ γ + γ γ ] = Tr[[γ α, γ β ]+]
2 2
1
= Tr[2g αβ 14×4] = 4g αβ (139)
2
Tr[γ αγ β γ γ δ ]
= Tr[(2g αβ − γ β γ α)γ γ δ ]
= Tr[2g αβ γ γ δ ] − Tr[γ β (2g α − γ γ α)γ δ ]
= Tr[2g αβ γ γ δ ] − Tr[γ β 2g αγ δ ] + Tr[γ β γ (2g αδ − γ δ γ α)]
= Tr[2g αβ γ γ δ ] − Tr[γ β 2g αγ δ ] + Tr[γ β γ 2g αδ ] − Tr[γ β γ γ δ γ α]
= 8[g αβ g δ − g βδ g α + g βg αδ ] − Tr[γ αγ β γ γ δ ] . (140)
Since the last term above duplicates the starting Tr, we can move it to
Note that if we contract some momenta with some of the free γ indices
we will get things like
Tr[p/ γ β k/ γ δ ] = pαkTr[γ αγ β γ γ δ ]
= 4pαk[g αβ g δ − g βδ g α + g βg αδ ]
= 4[pβ kδ − g βδ p · k + pδ kβ ] (142)
Well, that’s all we need for the moment. The Appendix of MS has more.
• We now return to the initial spin average, final spin sum version of our
cross section:
2
−i
dσ 1 1
X
0 ~0 ρ 0 0 ~
= uµ (λ , k )(ieγ )vµ (λ , p
~ )v e (λ, p
~)(ieγρ )ue (λ, k)
2 4
64π 2 Etot 2
dΩ com q + i q =p+k
0
λ,λ,λ0 ,λ
(143)
yielding
p · p0 = k · k0 = E 2(1 − cos θ)
p · k0 = k · p0 = E 2(1 + cos θ)
(p + k)2 = 2p · k = 4E 2 (147)
so that we obtain
2
dσ 1 1
= e4E 4[(1 − cos θ)2 + (1 + cos θ)2]
dΩ com 8π 2Etot
2 4E 2
1 2 2
= (4πα) [1 + cos2 θ]
8π 2Etot
2 16
2
where s = Etot = 4E 2, the latter holding when masses are neglected. Also,
we frequently write
t2 u2
dσ 4π 1 2 1
= (4πα) +
dt s 8π 2s 4 s2 4
" 2 #
2 2
2πα t u
= + . (151)
s2 s s
This is a Lorentz invariant form that actually holds in any frame, since it is
expressed entirely in terms of Lorentz invariants.
A note on dimensions: dσ/dt has ’energy’ dimensions of 1/E 4 in the
h̄ = c = 1 units. In the massless limit, we see that all these dimensions are
provided by the energy scale of the process, as encoded in s. This is always
the case when the underlying theory has a dimensionless coupling constant
such as e, which in turn is always the case if we generate interactions using
the minimal substitution rule. In minimal substitution, ∂µ → ∂µ + iqAµ.
Since ∂µ and Aµ have the same dimension, R q4 must be dimensionless. To
check the dimensions of Aµ remember that d xL must be a dimensionless
action. With L = − 14 (∂µAν − ∂ν Aµ)(∂ µAν − ∂ ν Aµ), it is clear that Aµ
must have the same dimensions as ∂µ in order to cancel the dimensions of
d4x.
• Write
M = α ~ β ~ ~ ~
r1 (k 1 ) r2 (k2 ) . . . Mαβ... (k1 , k2 , . . .) . (152)
The µ(~ k) are gauge dependent objects. In particular, if we write (we keep
to a basis in which our initial choices — you will see why I use the word
“initial” in a moment — for the polarization ’s are real)
XX 1 ~
A µ+
(x) = q µ
r ( k)e −ik·x
ar (~
k)
~
k r 2V ω(~
k)
µ−
XX 1 ~ +ik·xa† (~
A (x) = q µ
r (k)e r k) (153)
~
k r 2V ω(~
k)
µ µ µ
and perform the
gauge transformation A (x) → A (x) + ∂ f (x) with
f (x) = f˜(k) ar (~
k)e−ik·x − a†(~
k)eik·x , then the corresponding change
in µ(~
k) is
µ → µ(~
r k) − ikµf˜(k) .
r (154)
0 = k1αβ ~ ~ ~
r2 (k2 ) . . . Mαβ... (k1 , k2 , . . .)
~ β
= α
r1 ( k 1 )k 2 . . . Mαβ...(~
k1 , ~
k2, . . .)
= ... . (155)
This is true only in the case of Abelian gauge theory (which QED is). In
a non-abelian gauge theory, this final form is not true and only a single
photon can be “removed” at a time — all the other polarizations must be
kept in place for the identity to hold.
• Another point of view that gives this same result is current conservation.
Let us focus on just one of the photons and write M = µ(~ k)Mµ(k).
We know that the photon is seeing an interaction of the form L =
Then
Z
µ µ ik·x
4
k Mµ(k) = d x (−i∂ )e hf |jµ(x)|ii
Z
= − d4xeik·x[−i∂ µhf |jµ(x)|ii]
= 0 (158)
• With this result, we can now simply perform polarization sums for the cross
X 2 X
X= ~ ~ ∗ ~
Mr (k) = Mµ(k)Mν (k) µ (~k) ν ~
r (k) . (159)
r
r =1,2 r =1,2
X
~ ν ~ 1
µ
r (k) r (k) → −g µν
− 2
[k ν µ
k −(k·n)(k µ ν
n +k ν µ
n )] , (164)
r =1,2
(k · n)
where nµ = (1, 0, 0, 0), but the other terms don’t contribute because of
the Ward identity.
• Let us now use this technology for the case of Compton scattering. This
was discussed beginning with Eq. (61) in some detail with the result that
we developed the two contributing amplitudes
q/ + m
p 0)/ s0 (~
Ma = −e2ur0 (~ k0) i 2 / s(~
k)ur (~
p) (165)
q − m2 q = p+ k
/0 1 3456
1 /(0
- .
!
8
5 19
. -
/20 7 /(0
"#$&%('*))+, /#5$&%'*))+:,
We need X X X
1
4
2
|Mrsr0s0 | = 1
4
Mαβ M∗αβ (167)
pol, s,s0 spin, r,r0 spin
where we have written (dropping the r and r 0 indices for the moment)
~ β ~0
Mss0 = α
s ( k) s0 (k )Mαβ (168)
Mαβ = Ma
αβ + M b
αβ (170)
with
p/ + k/ + m
Ma
αβ p 0)γβ
= −ie2ur0 (~ γαur (~
p)
(p + k)2
− m2
p/ − k/ 0 + m
Mbαβ p 0)γα
= −ie2ur0 (~ γβ ur (~
p) (171)
(p − k 0 )2 − m2
e4
Xaa Xbb Xab + Xba
= + − (172)
16 (pk)2 (pk0)2 (pk)(pk0)
where
0 β α 0 ∗
X
Xaa = [ur0 (p
~ )γ (f
/ 1 + m)γ ur (p
~)][ur0 (p
~ )γβ (f
/ 1 + m)γα ur (p
~)]
spin
0 β α 0
X
= [ur0 (p
~ )γ (f
/ 1 + m)γ ur (p
~)][ur (p
~)γα (f
/ 1 + m)γβ ur0 (p
~ )]
spin
!
β α 0 0
X X
= Tr γ (f
/ 1 + m )γ ~)ur (p
ur ( p ~) γα (f
/ 1 + m)γβ ~ )ur0 (p
ur0 (p ~ )
r r0
h i
β α 0
= Tr γ (f/ 1 + m )γ (p/ + m)γα (f/ 1 + m)γβ (p/ + m) (173)
0 α β 0 ∗
X
Xbb = [ur0 (p
~ )γ (f
/ 2 + m)γ ur (p
~)][ur0 (p
~ )γα (f
/ 2 + m)γβ ur (p
~)]
spin
h i
α β 0
= Tr γ (f
/ 2 + m )γ (p
/ + m )γβ (f
/ 2 + m )γα (p
/ + m) (174)
0 α β 0 ∗
X
Xba = [ur0 (p
~ )γ (f
/ 2 + m)γ ur (p~)][ur0 (p
~ )γβ (f
/ 1 + m)γα ur (p
~)]
spin
h i
α β 0
= Tr γ (f
/ 2 + m )γ (p
/ + m )γα (f
/ 1 + m )γβ (p
/ + m) (176)
1. First, we have
α α β 1 α β 1 αβ
γ γα = gαβ γ γ = gαβ [γ , γ ]+ = gαβ 2g = 4. (177)
2 2
2. Next, we have
α ν αν ν α ν ν
γ γ γα = (2g − γ γ )γα = (2 − 4)γ = −2γ (178)
3. Thirdly, we have
ν µ α ν ν µ α
γα γ γ γ = (2gα − γ γα )γ γ
µ ν ν µ
= 2γ γ − γ (−2γ )
µ ν
= 2[γ , γ ]+
µν
= 4g (179)
4. Finally, we need
ν µ σ α ν ν µ σ α
γα γ γ γ γ = (2gα − γ γα )γ γ γ
µ σ ν ν µ σ α
= 2γ γ γ − γ γα γ γ γ
µ σ ν µσ ν
= 2γ γ γ − 4g γ
µ σ ν µ σ σ µ ν
= 2γ γ γ − 2(γ γ + γ γ )γ
σ µ ν
= −2γ γ γ . (180)
α 0 β
h i
= Tr (f/ 1 + m)γ (/ p + m ) γα ( f
/ 1 + m)γβ (/p + m) γ
0
h i
= Tr (f/ 1 + m)(−2/ p + 4m)(f / 1 + m)(−2/ p + 4m)
0
h i
= 4Tr (f/ 1 + m)(/p − 2m)(f / 1 + m)(/ p − 2m)
0 2 0 0 4
n h i o
= 4 Tr[f/ 1/
pf p ] + m −4Tr[f
/ 1/ p ] − 4Tr[f
/ 1/ / 1/p ] + Tr[/p/p ] + 4Tr[f
/ 1f
/ 1 ] + 4m Tr[1]
(
0 0 2 0
= 16 [2(f1 p)(f1 p ) − (pp )(f1 f1 )] + m [−4(f1 p) − 4(f1 p )
)
0 4
+4(f1 f1 ) + (pp )] + 4m
0 2 4
= 32[(pk)(pk ) + m (pk) + m ] , (181)
β α 0
h i
Xab = Tr γ (f
/ 1 + m)γ (/
p + m ) γβ ( f
/ 2 + m)γα (/
p + m)
β 2 0
h i
= / 1 + m)(−2f
Tr γ (f p + 4mf2 β + 4mpβ − m 2γβ )(/
/ 2 γβ / p + m)
"(
= Tr −8(f1 f2 )/
p + 4mf
/ 2/
p + 4mf
/ 2 (f
/ 1 + m) + 4mp
/ (f
/ 1 + m)
) #
2 0
−2m (−2f
/ 1 + 4m) (/
p + m)
"
0 2 0 2 0 2 2 0
= 4 −8(f1 f2 )(pp ) + 4m (f2 p ) + 4m (pp ) + 4m (f1 p ) + 4m (f2 p)
#
2 2 4
+4m (f2 f1 ) + 4m (f1 p) − 8m p 0 , 2nd from m
1st from /
"
2 0 2 0 2 0 2 2 2 2 0
= 4 −8m (pp ) + 4m [(pp ) − (pk)] + 4m (pp ) + 4m [m + (pk)] + 4m [m − (pk )]
4 2 2 0
= 16[2m + m (pk) − m (pk )] , (184)
e4
Xaa Xbb Xab + Xba
= + − (185)
16 (pk)2 (pk0)2 (pk)(pk0)
1 1 1
− = (cos θ − 1) , (188)
ω ω0 m
0
X X ω ω
1
4
2
|M| = 2e 4
+ − sin2 θ . (189)
pol, s,s0 spin, r,r0
ω0 ω
p · k = (p0 + k0 − k) · k = p0 · k + k0 · k = p · k0 + k0 · k , (190)
which, in turn, gives Eq. (188). For the cross section, we return to the
expression
X X 1 Y ~0f
d3 p
4 4
dσ = (2π) δ ( 0
pf − pi ) |M|2 (192)
4E1E2vrel f
(2π)32Ef0
with E1E2vrel = [(p1 · p2)2 − m21m22]1/2. In our case we can identify particle
1 as a photon with m1 = 0 and particle 2 as the target proton. Then we
E 0 = [m2 + (~
k −~
k0)2]1/2 = [m2 + ω 2 + ω 02 − 2ωω 0 cos θ]1/2 (194)
so that
∂E 0 ω 0 − ω cos θ
= (195)
∂ω 0 E0
J. Gunion 230A – Part II 107
implying that
∂(E 0 + ω 0) ω 0 − ω cos θ
= +1
∂ω 0 E0
ω 0 + E 0 − ω cos θ
=
E0
ω + m − ω cos θ
=
E0
mω
= , (196)
E 0ω 0
ω0 E 0ω 0 ω0
dσ ω
= 2e4 + − sin2 θ
dΩ 64π 2ωmE 0 mω ω0 ω
4 0 2 0
e ω ω ω
= + − sin2 θ
32π 2m2 ω ω0 ω
• Finally, we wish to return to the gauge invariance issue. You will recall that
we had the two amplitudes:
q/ + m
Ma = −e2ur0 (~p 0)/ s0 (~
k0 ) i 2 2
/ s (~k)ur (~p) (199)
q − m q=p+k
q/ + m
p 0)/ s(~
Mb = −e2ur0 (~ k) i 2 2
/ s 0 (~
k 0
)ur (~
p) , (200)
q − m q=p0−k
p/ + k/ + m
p 0)/ (~
Ma → −e2u(~ k0) i 2 2
k/ u(~p) (201)
(p + k) − m
0
−
p
/ k
/ + m ~
p 0)k/ i 0
Mb → −e2u(~ 2 2
/ ( k 0
)u(~
p) (202)
(p − k) − m
In Ma, we write
p/ + k/ + m p/ + k/ + m
k/ u(~
p) = (k/ + p/ − p/ )u(~
p)
(p + k)2 −m 2 (p + k)2 − m2
p/ + k/ + m
= (k/ + p/ − m)u(~
p)
(p + k)2 − m2
so that
p 0)/ (~
Ma → −ie2u(~ k0)u(~
p) . (204)
In Mb, we write
p/ 0 − k/ + m p/ 0 − k/ + m
p 0)k/
u(~ p 0)(k/ − p/ 0 + p/ 0)
= u(~
(p0 − k)2 − m2 (p0 − k)2 − m2
0 0 p/ 0 − k/ + m
p )(k/ − p/ + m)
= u(~
(p0 − k)2 − m2
0 −(p0 − k)2 + m2
= u(~
p )
(p0 − k)2 − m2
p 0) ,
= −u(~ (205)
so that
p 0)/ (~
Mb → +ie2u(~ k0)u(~
p) . (206)