QFT PDF
QFT PDF
Ronald Kleiss
2
0 Introductory remarks 19
0.1 Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
0.2 Basic tools . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
0.2.1 Units and fundamental units . . . . . . . . . . . . . . . 20
0.2.2 Planck units . . . . . . . . . . . . . . . . . . . . . . . 21
0.2.3 Charges . . . . . . . . . . . . . . . . . . . . . . . . . . 22
0.2.4 Conventions . . . . . . . . . . . . . . . . . . . . . . . . 23
0.3 The P 4 Hall of Fame . . . . . . . . . . . . . . . . . . . . . . . 26
0.4 Exercise . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3
4 CONTENTS
2 On renormalization 65
2.1 Doing physics : mentality against reality . . . . . . . . . . . . 65
2.1.1 Physics vs. Mathematics . . . . . . . . . . . . . . . . . 65
2.1.2 The renormalisation program : an example . . . . . . . 66
2.2 A handle on loop divergences . . . . . . . . . . . . . . . . . . 68
2.2.1 A toy : the dot model . . . . . . . . . . . . . . . . . . 68
2.2.2 Nonrenormalizable theories . . . . . . . . . . . . . . . . 71
2.3 Scale dependence . . . . . . . . . . . . . . . . . . . . . . . . . 72
2.3.1 Scale-inpendent scale dependence . . . . . . . . . . . . 72
2.3.2 Low-order approximation to the renormalised coupling 75
2.3.3 Scheme dependence . . . . . . . . . . . . . . . . . . . . 77
2.3.4 Theories with more parameters . . . . . . . . . . . . . 78
2.3.5 Failure of the dot model . . . . . . . . . . . . . . . . . 79
2.4 Asymptotics of renormalisation in ϕ4 theory . . . . . . . . . . 79
2.5 The method of counterterms . . . . . . . . . . . . . . . . . . . 81
2.5.1 Counterterms in the action . . . . . . . . . . . . . . . . 81
2.5.2 Return to the dot model, and a preview . . . . . . . . 82
2.6 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
15 Appendices 437
15.1 Perturbative (non)convergence issues . . . . . . . . . . . . . . 437
15.1.1 Punishment at the singular point . . . . . . . . . . . . 437
15.1.2 Borel summation . . . . . . . . . . . . . . . . . . . . . 438
15.2 More on symmetry factors . . . . . . . . . . . . . . . . . . . . 441
15.2.1 The origin of symmetry factors . . . . . . . . . . . . . 441
15.2.2 Explicit computation of symmetry factors . . . . . . . 442
15.3 Derivation of the diagrammatic sum rules . . . . . . . . . . . . 445
15.4 Alternative solutions to the Schwinger-Dyson equation . . . . 448
15.4.1 Alternative contours for general theories . . . . . . . . 448
15.4.2 Alternative contours for ϕ3 theory . . . . . . . . . . . . 449
15.4.3 Alternative contours for ϕ4 theory . . . . . . . . . . . . 450
15.5 Diagram counting . . . . . . . . . . . . . . . . . . . . . . . . . 450
12 CONTENTS
15
16 LIST OF FIGURES
15.1 Exact and asymptotic number of tree diagrams in ϕ3/4 theory 454
15.2 Computing the average symmetry factors . . . . . . . . . . . . 460
15.3 Multiplication table for the Clifford algebra . . . . . . . . . . 478
17
18 August 31, 2019
Chapter 0
Introductory remarks
0.1 Preface
In what follows, whatever is correct I owe to many other people ; that which
is wrong I managed on my own. I am perpetually in need of, and grateful
to, those pointing out typing or thinking errors in these notes1 .
1
I cordially invite all and sundry to do so. The P 4 Hall of Fame collects the names of
friends who have helped me in learning about, formulating, contemplating, or execrating
one or several issues.
2
Labori inutilis manus salubris (attr. to S. Partlic)
3
Rigor mathesis protractus mortis rigorem inducet (attr. to S. Partlic).
19
20 August 31, 2019
m3
G = 6.67408(31) 10−11 .
kg sec2
The truly, ultimately fundamental units of mass, length and time that can
be recovered from c, ~ and GN are then the Planck units :
r
~
MP = = 2.1764 10−8 kg ,
r Gc
~G
LP = 3
= 1.6162 10−35 m ,
r c
~G
TP = 5
= 5.3912 10−44 sec . (1)
c
These values are outrageously far removed from the typical scales of particle
phenomenology. We may interpret this as an indication that in what follows
the gravitational interaction will not play any part. In fact, in any case we do
10
This has the unfortunate consequence of making it impossible to check the (at least)
grammatical correctness of an expression by dimensional analysis, and even leads to erro-
neous statements about ‘classical limits’ and the like : see our discussion of particle masses
in chapter 6.
11
You would be surprised how difficult it is for even well-trained physicists to put the ~
and c back, when asked to do so in a hurry !
12
Taken from P.J. Mohr, D.B. Newell, B.N. Taylor, CODATA recommended values of
the fundamental physical constants: 2014, Rev. Mod. Phys. 88, July-September 2016.
The uncertainty in the last two digits is bracketed.
22 August 31, 2019
0.2.3 Charges
The electrostatic charge is adopted to the Gaussian system, so as to have
no truck with the ‘permittivity of the vacuum’ and suchlike : that is, two
charges e1 and e2 separated by a distance r feel a mutual Coulomb force F~
characterized by
1 | e1 e2 |
| F~ | = .
4π r2
√
This implies that the charge has the dimensionality of ~c. It follows that, if
we choose the proton charge as the unit charge e, the fine structure constant
e2
αe =
4π ~ c
1
αe = ,
137.035999139(31)
√ kg1/2 m3/2
e = 0.30282 ~c = 5.3844 10−14 .
sec
13
Meaning that it has the same value in all possible systems of units ! An alien civiliza-
tion in outer space will find the same value.
14
Taken from P.J. Mohr, D.B. Newell, B.N. Taylor, CODATA recommended values of
the fundamental physical constants: 2014, Rev. Mod. Phys. 88, July-September 2016.
The uncertainty in the last two digits is bracketed.
August 31, 2019 23
0.2.4 Conventions
Forgetting phases
If two quantities A and B are equal up to a disregarded overall complex
phase we shall write
AlB .
The metric
By convention, the Minkowski metric15 has the form
Kronecker’s symbol
The Kronecker symbol is defined by
α 1 if α = µ
δ µ= .
6 µ
0 if α =
15
In the usual Cartesian coordinate systems. Since we are not considering curved space-
time in these notes, we shall adhere to this simplest of coordinate system throughout,
except when discussing phase space integration where polar coordinates often come in
handy. But there we shall not refer to the metric.
24 August 31, 2019
Note : in Minkowski space, Kronecker symbols tend to carry one upper, and
one lower index. Kronecker symbols with two upper or two lower indices are
slightly suspect and to be treated with care unless they refer to some other
label such as color.
The antisymmetrizer
Using Kronecker symbols we can build the following antisymmetric objects :
µ1 µ1 µ1 µ2
= δ ν1 , = δ µ1 ν1 δ µ2 ν2 − δ µ1 ν2 δ µ2 ν1 ,
ν1 ν1 ν2
µ1 µ2 µ3
= δ µ1 ν1 δ µ2 ν2 δ µ3 ν3 + δ µ1 ν2 δ µ2 ν3 δ µ3 ν1 + δ µ1 ν3 δ µ2 ν1 δ µ3 ν2
ν1 ν2 ν3
− δ µ1 ν1 δ µ2 ν3 δ µ3 ν2 − δ µ1 ν2 δ µ2 ν1 δ µ3 ν3 − δ µ1 ν3 δ µ2 ν2 δ µ3 ν1
and so on. Here we encounter all signed permutations16 of the lower indices.
These are compuationally handy as we see below.
In order not to lumber ourselves with too many explicit Lorentz indices, we
shall use the following shorthand :
µ (a, b, c) means µνρσ aν bρ cσ and so on.
Finally, although it falls outside the scope of these notes, it is useful to note
that the Kronecker and antisymmetrizer symbols are fully-fledged tensors
in a much wider class of spaces than Minkowski space, but the Levi-Civita
symbols themselves are not. For instance, if we move from Cartesian to
polar coordinates, say, the Kronecker symbol remains unaffected but the
Levi-Civita is changed.
0.4 Exercise
Excercise 0 Would-be exercise
This might have been an exercise.
28 August 31, 2019
Chapter 1
1.1 Introduction
29
30 August 31, 2019
The function S(ϕ) is called the action of the particular quantum field theory :
in a sense, it defines the theory. For the probability density to be acceptable,
S(ϕ) must go to infinity sufficiently fast as |ϕ| → ∞. The normalization
factor N is defined by1
Z
−1
N = exp − S(ϕ) dϕ . (1.2)
1
If not explicitly indicated otherwise, integrals run from −∞ to +∞.
2
You are here approaching a career decision ! You may decide simply to measure
the value of ϕ : in that case you have decided to become an experimentalist rather than
a theorist.
3
A clarifying remark is necessary here. In this text, the Green’s functions are simply
defined to be expectation values. This may appear to contrast with the use of Green’s
functions in the solution of inhomogeneous linear differential equations such as are encoun-
tered in classical electrodynamics where one uses them to compute the electromagnetic
field configurations for given sources. The difference is only apparent since, as we shall
recognize, the latter type of Green’s functions are in our treatment simply the two-point
Green’s functions ; and for theories such as electrodynamics, where the electromagnetic
fields do not undergo self-interaction, the two-point functions are in fact the only nonzero
connected Green’s functions. Be not, therefore, misled into thinking that there are some-
how two sorts of Green’s functions. The Green’s function formulation of electrodynamics
will in fact appear as the classical limit of the Schwinger-Dyson equation discussed below.
August 31, 2019 31
This is called the path integral, for reasons that will become clear later. It
can be written as
Z
Z(J) = N exp − S(ϕ) + Jϕ dϕ . (1.6)
The number J, which here serves purely as a device to distinguish the various
Green’s functions, is called a source, again for reasons that will become ap-
parent later. Once Z(J) is known, an individual Green’s function is extracted
by differentiation : n
∂
Gn = Z(J) . (1.7)
(∂J)n J=0
C0 =0
C1 = hϕi the mean
C2 = h(ϕ − hϕi)2 i the variance (1.9)
C3 = h(ϕ − hϕi)3 i the skewness
C4 = h(ϕ − hϕi)4 i − 3C2 2 the kurtosis
and so on. Since W (0) = C0 = 0, the same information about the probability
density is also contained in the field function :
∂ X 1
φ(J) ≡ W (J) = J n Cn+1 . (1.10)
∂J n≥0
n!
with µ a positive real number. For any action, we shall call the part quadratic
in the fields (or bilinear in the case of several fields) the kinetic part. This
action, called the free action, consists of only a kinetic part. The path integral
is now simply computed by
Z
1 2
Z(J) = N exp − µϕ + Jϕ dϕ
2
2 !
J2
Z 2
1 J J
= N exp − µ ϕ − + dϕ = exp . (1.14)
2 µ 2µ 2µ
Computing the path integral is now a much less trivial matter. A possible
approach is to assume that, in some sense, the ϕ4 theory is close to a free
theory, that is, in the same some sense, λ4 is a small number. We can then
expand the probability density in powers of λ4 :
k
1 2 X 1 λ4
exp(−S(ϕ)) = exp − µϕ − ϕ4k . (1.18)
2 k≥0
k! 24
This procedure is called perturbation theory. Having thus reduced the prob-
lem to the previous case of the free theory, we cavalierly10 interchange the
series expansion in λ4 with the integration over ϕ and arrive at the following
expression for the Green’s functions :
G2n = H2n /H0 ,
k
1 X (4k + 2n)! λ4
H2n = − . (1.19)
µn k≥0 22k+n (2k + n)!k! 24µ2
11
The sum starts at 1 since a constant, ϕ-independent term in the action is always
immediately swallowed up by the normalization factor N .
36 August 31, 2019
where in the last lemma we have used partial integration, and the fact that
the integrand vanishes at the endpoints. Symbolically, we may write the SDe
as
∂ 0 ∂
S(ϕ) Z(J) = S Z(J) = JZ(J) . (1.26)
∂ϕ ϕ=∂/∂J ∂J
E4 For our sample model, the ϕ4 theory, the SDe reads12
1
λ4 Z 000 (J) + µZ 0 (J) − JZ(J) = 0 . (1.27)
6
Using the series expansion of the path integral we can express this as a
relation between different Green’s functions :
λ4
Gn+3 + µGn+1 − nGn−1 = 0 , n ≥ 1 . (1.28)
6
This relation may usefully be rewritten as follows :
1 λ4
Gn = (n − 1)Gn−2 − Gn+2 , n≥2 . (1.29)
µ 6
If we start by assigning to the Green’s functions the values Gn = δ0,n ,
then repeated applications of Eq.(1.29) will precisely reproduce the Green’s
functions of Eq.(1.21)13 .
Although this leads to very nonlinear relations between the various connected
Green’s functions this form of the SD equation is actually even simpler to
apply : with φ(J) = 0 as a starting pont, iterating the assignment (1.32) then E6
results14 in the correct form of φ(J), giving the connected Green’s functions
of Eq.(1.22).
1.3 Diagrammatics
1.3.1 Feynman diagrams
An extremely useful tool for computing Green’s functions and connected
Green’s functions is at hand in the form of Feynman diagrams. In this section
we shall first introduce these diagrams and their concomitant Feynman rules.
Only after that shall we prove that these diagrams do, indeed, correctly
describe Green’s functions. Feynman diagrams are constructs of lines and
vertices. A vertex is a meeting point for one or more lines. Diagrams are
allowed in which one or more lines do not end in a vertex but, in a sense sie
wandern ins Blaue hinein : such lines are called external lines. Lines that are
not external lines, and end up at vertices at both ends, are called internal
lines. Diagrams may be connected, in which case one can move between any
two points in the diagram following lines of that diagram ; or they may be
disconnected, in which case it consists of two or more disjoint pieces that are
themselves connected. Any graph15 consists of a finite number of connected
14
For this approach to work in practice, it turns out to be useful to truncate φ(J) as a
power series in J, the truncation order increasing by one with each iteration. If you don’t
do this, each iteration triples the highest power in J, leading to very unwieldy expressions
with only the first few terms being actually correct.
15
For us, the terms ‘diagram’ and ‘graph’ are interchangeable.
38 August 31, 2019
Note that the precise shape of the lines and the precise position of the ver-
tices are irrelevant. The important thing is the way in which the lines are
connected to the vertices17 .
16
Sophistry : it has no points between which one might wish to move.
17
As you will discover, I have endeavoured in these notes to avoid drawing straight
lines, or to draw blobs or closed loops as circles. Many texts do employ only straight lines
and circles. This not only leads to awfully unæsthetic-looking pictures, but is also deeply
misleading. There is a (natural) tendency to look at Feynman diagrams with the idea
that the lines represent ‘particles moving freely through space’ so that the lines ‘ought’
to be straight according to Newton’s first law. That this is completely wrong becomes
immediately clear if we realize that, in the zero-dimensional world we are dealing with for
now, there cannot be any notion of movement yet, let alone any Newton to pronounce on
it. In fact, Newton’s first law ought to be derived from our theory, and we shall do so in
due course.
August 31, 2019 39
↔ 1/µ
↔ −λ3
↔ −λ4
↔ +J
Feynman rules, version 1.1 (1.33)
A vertex at which a single line ends (and which carries a Feynman rule
factor +J) is called a source vertex. A disconnected diagram evaluates to
the product of the values of its disjunct connected pieces. Because of this
multiplicative rule, the value of the empty diagram E is taken to be unity.
In addition, we assign to every Feynman diagram a symmetry factor. The
symmetry factor is the single most nontrivial ingredient of the diagrammatic
approach. We shall therefore devote a separate section to this issue.
• for every set of k lines that may be permuted without changing the
diagram, there will be a factor 1/k! ;
• for every set of m vertices that may be permuted without changing the
diagram, there will be a factor 1/m! ;
carries a symmetry factor of 1/3! because the three internal lines are inter-
changeable. The graph
1 λ4 3
= −
4 µ7
carries a symmetry factor (1/2!)(1/2!) since there are now only two inter-
changeable internal lines, and a single ‘leaf’. Finally, the diagram
1 λ4 2
=
48 µ4
has a symmetry factor (1/4!)(1/2!) since there are 4 equivalent internal lines,
and moreover the diagram can be ‘flipped over’ without changing it. E7
H0 = E + + + + + ··· (1.36)
to be the set of all connected diagrams with precisely one external line, and
any number of source vertices. The shading indicates that all the diagrams
in the blob must be connected . Clearly, then, we have
X 1
Ψ(J) = J n Cn+1 , (1.39)
n≥0
n!
August 31, 2019 43
On the other hand, it may happen that the two branches end up in disjunct
connected pieces of the diagram, which then looks like
00000
11111
11111
00000
00000
11111
00000
11111
00000
11111
00000
11111
00000
11111
00000
11111
111111
000000
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
n≥0
n! ∂J
and 1111111
0000000
0000000
1111111
0000000
1111111
0000000
1111111
X 1 ∂2
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111 = J n Cn+3 = Ψ(J) , (1.43)
0000000
1111111
0000000
1111111
0000000
1111111
n≥0
n! (∂J)2
so that we can translate the diagrammatic equation (1.41) into an algebraic
equation for Ψ(J) by carefully implementing the correct Feynman rules, in-
cluding the symmetry factors for equivalent blobs and lines:
J λ3 1 2 1 ∂
Ψ(J) = − Ψ(J) + Ψ(J)
µ µ 2 2 ∂J
1 ∂2
λ4 1 3 1 ∂
− Ψ(J) + Ψ(J) Ψ(J) + Ψ(J) . (1.44)
µ 6 2 ∂J 6 (∂J)2
Now Eq.(1.44), obtained from the Feynman diagrams via the Feynman rules,
has exactly the same form as Eq.(1.32), valid for the field function φ(J) – note
the importance of the symmetry factors ! Moreover, the iterative solution
for φ(J) starts with φ(J) = J/µ, also identical to the diagrammatic starting
E8 point . We therefore conclude that Ψ(J) = φ(J), in other words Cn = Cn
(n ≥ 1). This proves that connected Green’s functions can be obtained
E9
by the following recipe: to obtain Cn (n ≥ 1), write out all connected
E10 Feynman diagrams with no source vertices and precisely n external
lines. Evaluate the diagrams using the Feynman rules, and sum
them.
other external lines), with the constraint that each connected piece of the
semi-connected graph is attached to at least one of the lines indicated on the
left. This may sound more intimidating that is actually is : an example is
1 1 1 1
2 = 2 + 2 + 3
3 3 3 2
1
2
+ 3 + 2 . (1.45)
1
3
A single semi-connected graph with n indicated lines stands for B(n) dia-
grams with explicit connected graphs, where B(n) is the so-called Bell num-
ber : the number of ways to divide n distinct objects into non-empty groups22 .
For ϕ3/4 theory, the SDe then becomes simply
= + + . (1.46)
22
For small n we have B(0) = 1, B(1) = 1, B(2) = 2, B(3) = 5, B(4) = 15, and
B(5) = 52 ; more general values can be obtained from the identity
X xn
B(n) = exp (ex − 1) .
n!
n≥0
where the ellipsis contains connected contributions with more source vertices.
Vacuum bubbles do not contribute to W (J). By taking careful account of
the symmetry factor assigned to identical connected parts of a disconnected
diagram, we can see that
111111
000000
000000
111111
111111
000000 000000
111111 000000
111111
000000
111111 111111
000000 000000
111111
000000
111111 000000
111111 000000
111111
000000
111111 000000
111111 000000
111111
000000
111111 000000
111111 000000
111111
000000
111111 000000
111111
1 000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
W (J)2
000000
111111
= 000000
111111
111111
000000
000000
111111
000000
111111
+ 111111
000000
000000
111111
000000
111111
+ 11111
00000
00000
11111
00000
11111
00000
11111
2! 000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
00000
11111
00000
11111
00000
11111
00000
11111
00000
11111
00000
11111
000000
111111 00000
11111
111111
000000 11111
00000
00000
11111
000000
111111 000000
111111 00000
11111
111111
000000
000000
111111 000000
111111 00000
11111
000000
111111 000000
111111 00000
11111
000000
111111 000000
111111
000000
111111 00000
11111
000000
111111 000000
111111 00000
11111
00000
11111
111111
000000
000000
111111 000000
111111 00000
11111
000000
111111
+ 000000
111111
111111
000000
000000
111111
000000
111111
+ 00000
11111
11111
00000
00000
11111
00000
11111
00000
11111
+ 00000
11111
00000
11111
11111
00000
000000
111111 00000
11111 00000
11111
000000
111111
000000
111111 00000
11111 00000
11111
000000
111111 00000
11111 00000
11111
00000
11111
000000
111111 00000
11111
00000
11111 00000
11111
00000
11111
00000
11111
corresponding to
∂2
J λ2 λ4 3 ∂
φ(J) = − φ(J) − φ(J) + 3φ(J) φ(J) + φ(J) . (1.51)
µ µ 6µ ∂J (∂J)2
Multiplying the equation by µ and transposing the λ2 term to the left, we
obtain
∂2
J λ4 3 ∂
φ(J) = − φ(J) + 3φ(J) φ(J) + φ(J) ,
µ + λ2 6(µ + λ2 ) ∂J (∂J)2
(1.52)
precisely what we woud have obtained by taking the combination (µ + λ2 ) as
the kinetic part from the start. This procedure, by which the effect of two-
point (effective) vertices is subsumed in a redefinition of the kinetic part,
is called Dyson summation. In the present example, the summation is of
course trivial ; but we shall see that two-point interactions can also arise
from more complicated Feynman diagrams corresponding to higher orders
in perturbation theory. The manner in which Dyson summation is usually
treated is by explicitly writing out the propagator, ‘dressed’ with two-point
vertices in all possible ways :
+ + + + ···
1 1 1 1 1 1 1 1 1 1
= − λ2 + λ2 λ2 − λ2 λ2 λ2 + ···
µ µ µ µ µ µ µ µ µ µ
k
1 X λ2
= −
µ k≥0 µ
1 1 1
= = , (1.53)
µ 1 + λ2 /µ µ + λ2
48 August 31, 2019
where it should come as no surprise that we cheerfully ignore all issues about
convergence, in the spirit of perturbation theory. Every propagator line can
(and must ! ) be dressed in this way once any two-point vertex (elementary
of effective, that is, as the result of a collection of closed loops with two legs
sticking out) occurs.
~
↔
µ
λ3
↔ −
~
λ4
↔ −
~
J
↔ +
~
Feynman rules, version 1.2 (1.59)
and
50 August 31, 2019
Since in the Feynman rules ~ appears all over the place, it is advisable to
check that the ~-behaviour of the Feynman graphs is indeed as desired. For
an arbitrary given diagram let us define the characteristics E : the number
of external lines, I : number of internal lines, Vq : the number of vertices
of q-point type, L : the number of closed loops, and P : the number of
disjunct connected pieces. There are precisely two ‘sum rules’ that involve
linear combinations of these quantities. They are derived in section 15.3 and
read X X
Vq = I + P − L , qVq = 2I + E . (1.60)
q q
We are now able to read off the power of ~ associated with an arbitrary
connected diagram (with P = 1). From the Feynman rules, we infer that
every line contributes a factor ~ and every vertex a factor 1/~. The total
power of ~ is, therefore
X
E+I − Vq = E + L − 1 . (1.61)
q
SDe will, for the ϕ3/4 theory, then take the form
11111
00000
00000
11111 11111
00000
00000
11111 00000
11111
00000
11111 00000
11111
00000
11111
11111
00000
00000
11111 00000
11111
00000
11111 00000
11111
00000
11111
00000
11111 00000
11111 00000
11111
00000
11111 00000
11111
00000
11111
00000
11111
00000
11111
00000
11111
00000
11111
= + 00000
11111
111111
000000
+ 0000
1111
0000
1111
0000
1111
0000
1111
0000
1111
. (1.62)
00000
11111 000000
111111
000000
111111 00001111
1111
00000000
1111
00000
11111 0000
1111
000000
111111
000000
111111
0000
1111
00001111
11110000
000000
111111 0000
1111
0000
1111
Let us now look at the path-integral picture of the classical limit. When ~
becomes small, the fluctuations in the path integrand
1
exp − S(ϕ) − Jϕ
~
become extremely exaggerated. The main contribution to hϕi therefore
comes from that value ϕc of ϕ for whichthe probability distribution attains
its maximum, that is,
hϕiJ ≈ ϕc , where S 0 (ϕc ) = J , S 00 (ϕc ) > 0 . (1.66)
Also in the classical limit, we therefore have φc (J) = ϕc .
25
see Appendix 15.15.12.
52 August 31, 2019
fast. Such subdominant solutions to the classical field equations are called
instantons. Their contribution to Green’s functions do not have a series
expansion around ~ = 0. Such effects are therefore not accessible using
Feynman diagrams : they are called nonpertubative. This is not to say that
they are irrelevant. Indeed, we usually have a finite value for ~ ; more
(1)
dramatically, if we let J vary as a parameter, ϕc , say, may for some value
(0)
of J take over from ϕc as the true maximum position of the probability
(0) (1)
density, causing a sudden shift in the value of φc (J) from ϕc to ϕc .
(1.72)
Let us denote the set of all 1PI graphs with precisely n external lines by
−γn /~, where the convention is that the Feynman factors for the external
lines are not included. Consider, now, what happens if we enter the field
function by way of its single external leg, as in the SDe. If we encounter
a vertex, that vertex is part of a 1PI subdiagram (possibly consisting of
only the vertex itself). Indicating the 1PI property with cross-hatches, we
therefore obtain the diagrammatic equation
11111
00000
00000
11111
00000
11111
111111
000000
000000
111111
00000
11111
00000
11111
00000
11111 = + 000000
111111
000000
111111
00000
11111
00000
11111
00000
11111
00000
11111
000000
111111
000000
111111
00000
11111
00000
11111
11111
00000
000000
111111 00000
11111
00000
11111
111111
000000 00000
11111
000000
111111
000000
111111 00000
11111
0000000
1111111
000000
111111 11111111
00000000
00000
11111
00000000
11111111
00000
11111
00000
11111
111111
000000
000000 111111
111111 000000
111111
000000 1111111
0000000
000000
111111
000000
111111
000000
111111
000000
111111 00000000
11111111
00000
11111
00000
11111
00000
11111
00000
11111
00000
11111
00000
11111
+ 000000
111111 000000
111111
000000
111111
+ 0000000
1111111
0000000
1111111
000000
111111
+ 00000000
11111111
00000000
11111111
00000
11111
00000
11111
00000
11111
· · · . (1.73)
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111 0000000
1111111
0000000
1111111 00000000
11111111
00000
11111
00000
11111
00000
11111
000000 111111
111111 000000
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111 0000000011111
11111111 00000
11111
00000
11111
00000
11111
00000
1111111
0000000 00000
11111
0000000
1111111
0000000
1111111 00000
11111
0000000
1111111 11111
00000
00000
11111
0000000
1111111 00000
11111
00000
11111
00000
11111
00000
11111
00000
11111
Algebraically, it reads
J 1 1 2 1 3
φ(J) = − γ1 + γ2 φ(J) + γ3 φ(J) + γ4 φ(J) + · · · , (1.74)
µ µ 2! 3!
30
Including them would be silly, since any diagram falls apart if we chop through an
external line.
56 August 31, 2019
in other words
Γ0 (φ) = J , (1.75)
where
1 1 1
Γ(ϕ) = γ1 ϕ + (γ2 + µ)ϕ2 + γ3 ϕ3 + γ4 ϕ4 + · · · . (1.76)
2! 3! 4!
We conclude that the vertices of the effective action are determined
by the 1PI diagrams. It must be noted that, in general, the effective
action contains vertices with an arbitrarily large number of legs, even if the
E12 original action S goes up only to ϕ3 or ϕ4 , say.
Such an object has, of course, its own SDe. Taking careful account of all
possibilities to attach external lines, we can write it as
n n n−1
= θ(n = 0) +
1
n n−2 n n−3
+ + + · · ·(1.79)
2 3
We define the generating function for such dressed propagators as
X zn n
P (z) = = , (1.80)
n≥0
n!
August 31, 2019 57
L(z) = + + + ···
z 2 λ5 S 000 (z)
1 λ3 λ4
= − P (z) − z P (z) − P (z) − · · · = − 00 (1.83).
2 ~ ~ 2! ~ 2S (z)
The symmetry factor 1/2 arises from the fact that the propagator is not
oriented and thus we have to avoid double-counting. Considering that a
propagator with n external legs leads to a closed loop with at least n + 1
external legs, we see that the one-loop effective action is given by
Z
~
Γ1 (φ) = −~ dφ L(φ) = log(S 00 (φ)) . (1.84)
2
A few remarks are in order here. In the first place, we see that the effective
action obtained in this way is only well-defined where the action itself is
concave, in agreement with the discussion in 1.5. In the second place, the
trick of closing the loop with an extra vertex, rather than just trying to
‘glue’ the endpoints of P (z) together, is technically useful since it avoids
potential problems with the symmetry factors. In the last place, the above
calculation is possible since all external lines are, so to speak, identical. In
more dimensions, where external lines can carry momentum, this is no longer
true. However, the effective potential, that is the effective action at zero
momentum, does lend itself to such a calculation in higher dimensions31 .
31
For simple scalar theories. Of course external lines may carry more than just momen-
tum information, that is, they can also carry spin/charge/colour· · · information. Then the
calculation is again more difficult.
58 August 31, 2019
We can extend this treatment to higher loop orders as well. Let us denote
a vertex where at least n + 1 lines come together by
+ + +
1.6 Exercises
Excercise 1 Green’s functions and connected Green’s functions
We have X Jn X Jn
Z(J) = Gn , W (J) = Cn
n≥0
n! n≥0
n!
and
W (J) = log(Z(J)) .
Using that G0 = 1, and the expansion
X xn x2 x3 x4 x5
− log(1 − x) = =x+ + + + − ··· ,
n≥1
n 2 3 4 5
to derive
s Z∞
3~ 3/(4g) −3/4 3 1
H= e dψ ψ exp − ψ+
4µg 8g ψ
0
5. The field function has only one-loop corrections ! Verify that the two-
loop correction to the propagator vanishes.
6. Show that the effective action is
µ µ
Γ(φ) = − 2 + ~ log(1 − aφ) − φ
a a
It is also free from higher-order corrections beyond one loop !
August 31, 2019 63
2. Show that the solution for the path integral has the form
aJ/~
a2 ~
µ + aJ µ
Z(J) = Γ /Γ
µ a2 ~ a2 ~
a2k−1 Bk
1 aJ a~ X
φ(J) = log 1 + − − ~k
a µ 2(µ + aJ) k≥2 k(µ + aJ)k
5. Show that all odd loop orders beyond the first one vanish.
On renormalization
µ , λ3 , λ 4 −→ C1 , C 2 , C 3 , C 4 , . . .
In this set-up, the parameters are supplied from outside the computational
and experimental context. Since, however, as (I hope) we are physicists
the situation is somewhat different : we first have to measure the values
of the parameters from inside the experimental context, using some of the
connected Green’s functions as measurement processes, and then predict some
other connected Green’s functions, which we shall call prediction processes.
This rather different situation may be depicted by the scheme
Ek = Ck , k = 1 . . . 4 −→ µ , λ3 , λ 4 −→ C5 , C 6 , C 7 , . . .
Here, the quantities E1,2,3,... stand for the experimentally observed values of
the connected Green’s functions : barring experimental errors or refinements,
these numerical values do not change under any improvement of the theory.
65
66 August 31, 2019
Now consider the fact that we are doing perturbation theory. That is, both
the measurement and the prediction processes are known only as truncated
series in ~. Let us suppose that by stolidity and perseverance a next higher
order in perturbation theory for the prediction processes has become avail-
able. Is this any good ? Obviously not, unless a similar increased level of
precision has been attained for the measurement processes. Only in that case
a new ‘fit’ of the parameters of the action can be made, and improved values
of the prediction processes can usefully be obtained. This order-by-order
improvement is called renormalisation. Let us denote by a superscript the
order to which the connected Green’s functions have been computed. The
‘physicist’s scheme’ can then be envisaged as follows :
Order by order, the parameters keep getting updated, but in the overall pic-
ture they are just bookkeeping devices that allow one to go from measurements
to predictions of the more physically interesting connected Green’s func-
tions. It should not come as a surprise that in the measurement-parameter-
prediction protocol, a higher-order correction in the parameters due to an
improved measurement expression is cancelled again, to some extent, in the
prediction. In fact, for certain classes of theories, which are called renormal-
isable, these cancellations may be quite extreme.
~3 λ
7 149 2 197 3
C4 = − 4 1− u+ u − u + ··· ,
µ 2 12 4
~5 λ2
1535 2 6405 3
C6 = 10 − 80u + u − u + ··· ,
µ7 3 2
~7 λ3
111755 2 3
C8 = − 10 280 − 3815u + u − 330925u + · · · ,
µ 3
~9 λ4
12672800 2 3
C10 = 15400 − 310940u + u − 49859600u + · · · ,
µ13 3
(2.1)
where u is the ubiquitous combination ~λ/µ2 . Now suppose that the two
measurements give C2 = ~/m and C4 = −~3 g/m3 . In lowest order of per-
turbation theory, this would imply that the action parameters are µ = m
and λ = g. However, in higher orders we would no longer have C2 = E2 and
C4 = E4 . Instead, we must choose
1 7 2 1 3
µ = m 1 − y − y − y + ··· ,
2 12 8
3 3 2 11 3
λ = g 1 + y + y + y + ··· , (2.2)
2 4 8
with y = ~g/m2 . The properly renormalised Green’s functions then become
~
C2 = ,
m
~3 g
C4 = − 4 ,
m
5 2
~g 3 375 3
C6 = 10 − 15y + 45y − y + ··· ,
m7 2
~7 g 3
2 67725 3
C8 = − 10 280 − 1155y + 5775y − y + ··· ,
m 2
~9 g 4 2 3
C10 = 15400 − 118440y + 882000y − 6963075y + · · · . (2.3)
m13
The difference between the ‘naive’ and the renormalised connected Green’s
functions is quite evident. In particular C2 and C4 are completely free of
higher-order corrections. For the other connected Green’s functions the co-
efficients in the perturbation expansion are smaller in absolute value than in
the ‘naive’ expressions (see also below, section 2.4).
68 August 31, 2019
1
This insight is, even at present, not as endemic as it ought to be.
2
This rule accords with ‘naive power counting’ for four-dimensional scalar theories
without derivative couplings, the most direct four-dimensional extension of the zero-
dimensional theories we are discussing in this chapter.
August 31, 2019 69
= + , ≡ c1 × ,
= + , ≡ c2 × . (2.4)
For example, under this rule the following two-loop diagrams are modified
accordingly :
→ + + +
= (1 + c1 )(1 + c2 ) ,
→ +
= (1 + c2 ) . (2.5)
≡ + + + ··· , (2.6)
≡ + + + ···
+ + + + ···
+ , (2.7)
≡ + + + ··· , (2.8)
≡ + + + ··· . (2.9)
3
With 5 or more legs our rule does not allow for diagrams with only dotted loops.
70 August 31, 2019
The only diagram that does not carry a ‘tower’ of loops on its back is the
last diagram in the two-point dotted series. Using these artefacts, we can
now rewrite the appropriate SDe for our ϕ3/4 theory with the added dotting
rule :
00000
11111 000000
111111
11111
00000 111111
000000
00000
11111 000000
111111
00000
11111 000000
111111
00000
11111
00000
11111
00000
11111
00000
11111
00000
11111
= + + 000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
+
00000
11111
00000
11111 000000
111111
000000
111111
000000
111111
11111
00000
00000
11111
00000
11111
00000
11111
00000
11111
00000
11111 11111
00000
00000
11111 00000
11111
00000
11111
00000
11111
000000
111111
+ 00000
11111
00000
11111
00000
11111
00000
11111
+
000000
111111
000000
111111 00000
11111
000000
111111 00000
11111
00000
11111
000000
111111
000000
111111
000000
111111
000000
111111 000000
111111
000000
111111
111111
000000
000000
111111 111111
000000
000000
111111 000000
111111
000000
111111 000000
111111 111111
000000 000000
111111
000000
111111 000000
111111 000000
111111 111111
000000
000000
111111 000000
111111 000000
111111 000000
111111
000000
111111 000000
111111 000000
111111 000000
111111
000000
111111 000000
111111 000000
111111 000000
111111
000000
111111 000000
111111
000000
111111
000000
111111
111111
000000
000000
111111
+ 000000
111111
000000
111111
111111
000000
000000
111111
+ 000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
+ 000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
+
000000
111111
000000
111111 000000
111111
000000
111111 000000
111111 000000
111111
000000
111111 000000
111111 000000
111111 000000
111111
000000
111111 000000
111111 000000
111111
000000
111111
000000
111111 000000
111111
000000
111111
11111
00000
00000
11111
00000
11111
00000
11111 000000
111111 000000
111111
111111
000000
00000
11111 111111
000000 000000
111111
00000
11111 000000
111111
000000
111111 000000
111111
00000
11111 000000
111111 000000
111111
00000
11111
0000
1111
0000
1111
0000
1111
0000
1111
0000
1111
+ 000000
111111
000000
111111
000000
111111
000000
111111
+ 000000
111111
000000
111111
000000
111111 +
0000
1111 000000
111111
0000
1111
00001111
11110000
0000
1111 0000
1111
0000
1111 0000
1111
00001111
11110000 0000
1111
0000
1111
0000
1111 0000
1111
0000
1111
0000
1111
0000
1111
00000
11111 00000
11111
11111
00000 00000
11111
00000
11111 11111
00000
00000
11111 00000
11111
00000
11111
00000
11111 00000
11111
00000
11111
00000
11111 00000
11111
00000
11111 00000
11111
00000
11111 00000
11111
00000000
11111111 00000000
11111111
00000000
11111111
00000000
11111111
00000000
11111111
00000000
11111111
00000000
11111111
00000000
11111111
00000000
11111111
+ 00000000
11111111
11111111
00000000
00000000
11111111
00000000
11111111
00000000
11111111
00000000
11111111
00000000
11111111
+
000000
111111
111111
000000
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
00000
11111
00000
11111
11111
00000
00000
11111
00000
11111
00000
11111
00000
11111
00000
11111
00000
11111
11111111
00000000
00000000
11111111
00000000
11111111
00000000
11111111
+ 11111111
00000000
00000000
11111111
00000000
11111111
00000000
11111111
00000000
11111111
00000000
11111111
00000000
11111111
. (2.10)
00000000
11111111
00000000
11111111 00000000
11111111
00000000
11111111
00000000
11111111
We can readily translate this SDe into algebraic form. If we take out the
external propagators from the ‘black box’ graphs, we can write
= B1 , = B2 , = B3 , = B4 . (2.11)
We shall leave the actual evaluation of these sets of graphs for later (cf
section 2.3.5): at this point, we shall simply treat them as effective vertices.
The ‘dotted-loop’-modified SDe then reads, when we work out the graphs
one after the other, in the order in which they are displayed above :
J B1 B2 λ3 2 ~λ3 0
φ = − − φ− φ − φ
µ µ µ 2µ 2µ
B3 B3 2 ~B3 0 ~B3 0
− φ2 − φ − φ − φ
2µ µ 2µ µ
λ4 ~λ4 0 ~2 λ4 00
− φ3 − φφ − φ
6µ 2µ 6µ
August 31, 2019 71
B4 3 ~B4 0 ~B4 0 ~2 B4 00
− φ − φφ − φφ − φ . (2.12)
2µ 2µ µ 2µ
We can simply rewrite this SDe as
(µ + B2 )φ = (J − B1 ) − (λ3 + 3B3 )(φ2 + ~φ0 )
−(λ4 + 3B4 )(φ3 + 3~φφ0 + ~2 φ00 ) . (2.13)
But hold it ! This is just the SDe equation belonging to the action
1 1 1
S(ϕ) = B1 ϕ + (µ + B2 )ϕ2 + (λ3 + 3B3 )ϕ3 + (λ4 + 3B4 )ϕ4 . (2.14)
2 6 24
Therefore, the spirit of renormalisation tells us that in every application the
bare parameters µ, λ3 and λ4 will never occur on their own, but always only
in the renormalised combinations µR = µ + B2 , λR R
3 = λ3 + 3B3 , and λ4 =
λ4 + 3B4 ; and that therefore, whatever the values of B2,3,4 , the combination E17
will automatically be finite if the experimental quantities in which they enter
are finite. We can therefore choose the action’s parameters such that all
Green’s functions come out finite ; and the remaining B1 can always be
completely compensated by an extra ϕ1 in the action4 . Indeed, this is the way
in which the notorious ‘loop divergences’ are absorbed into the bare action :
infinite loop corrections are compensated for by infinite bare parameters5 .
consider a Feynman rule in which a loop with three vertices acquires a dotted
counterpart : that is, we would have a (potentially infinite) contribution of
the form
shall lump all these effects together into a quantity s, which we shall call the
scale. It must be stressed that the scale also contains the (regularized) loop
divergences, and may be expected to become infinite at some stage.
Let us consider a theory with only one parameter : an example of such
a theory is massless QCD, that is the theory of massless quarks and gluons
and their interactions. The single parameter is then the coupling constant.
Let the bare parameter, as it occurs in the action, be denoted by v. The
renormalised parameter, extracted from experiment, will be denoted by w.
The renormalised coupling is then given by the bare coupling and the exper-
imental context, embodied by the scale s :
w = F (v, s) . (2.15)
v = G(w, s) . (2.16)
Obviously we have
This expression now contains the finite number w and the divergent scale s :
it can only make sense if s actually drops out. Note that this is not a proof,
it is a requirement that we make of the theory ! We therefore demand that
F (v, s) be such that
∂s F (G(w, s), s) = 0 . (2.21)
In other words,
∂s2 F
∂s F ∂s ∂v F ∂s F
− = ∂v =0 : (2.23)
∂v F (∂v F )2 ∂v F
All reference to the bare coupling has been removed : we see that the renor-
malised coupling has a definite, predictable dependence on the energy scale
of the measuring experiment14 . Note, also, that whereas we introduced h as a
function of the bare parameter, it enters in the beta function as a function of
the renormalised parameter. Finally, as we shall see in the next section, the
h functions usually starts at second order (O (v 2 )) in perturbation theory, so
that the two first terms in the beta function are independent of the form of
s0 (v) as long as this is non-singular for v = 0.
β(v) = β0 v 2 + β1 v 3 + β2 v 4 + · · · (2.31)
The requirement that the beta function depend not on s governs the form of
the functions αj (s) : to low order in v we have from Eq.(2.30)
Having computed the beta function for w, we can now simply obtain it for
w̃ :
dw̃ dw̃ dw
β(w̃) = =
ds
dw ds
2 3 2 3 4
= 1 + 2t1 w + 3t2 w + 4t3 w + · · · β0 w + β1 w + β2 w + · · ·
The two beta functions can be transformed from one scheme to another ; for
any scheme dependence for which Eq.(2.37) holds, the first two coefficients,
β0 and β1 , are seen to be independent of the actual scheme, as was to be
expected (cf Eq.(2.28)) : the two schemes correspond to different functions
s0 (v) as defined in section 2.3.
17
In practice, this difference can be quite small, as between the so-called MS and MS
schemes. With ‘different measurement processes’, we here mean two different, complete
operational schemes that both lead to a well-defined value for coupling constants.
18
This rules out possible but, for a practicing physicist useless and/or irrelevant, differ-
ences such as for instance obtained by defining w̃ = 2w. Get a life !
78 August 31, 2019
where ∂j stands for ∂/∂v j , and we adhere to the Einstein summation con-
vention. Note that the pullback function
0 = Pkj (v, s) ∂s2 F k (v, s) − Pkj (v, s) ∂s ∂` F k (v, s) Pj` (v, s) ∂s F j (v, s)
= Pkj (v, s) ∂s2 F k (v, s) + ∂s Pkj (v, s) ∂` F k (v, s) Pj` (v, s) ∂s F j (v, s)
of the unrenormalised ones. We can investigate this for the ϕ4 theory sys-
tematically as follows. The bare connected Green’s functions are given by
n
~ ~λ X (k)
C2n = (−u)n−1 t2n (u) , u = 2 , t2n (u) = t2n uk . (2.52)
µ µ k≥0
All these manipulations are well suited to computer algebra. We can then
(k) (k)
compute the ‘improvement factor’ t2n /τ2n . The results for C6,8,10,12 are de-
picted in figure 2.1 as a function of k, the loop order, up to 155 loops. The
improvement is less for higher connected Green’s functions but at high order
the factors approach an asympotic value. We see that the asymptotic charac-
ter of perturbation theory is completely unchanged by renormalisation, since
neither the superexponential (n!) nor the purely exponential (cn ) or even
the polynomial (na ) behaviour of the coefficients is affected. The asymptotic
value of all improvement factors is exp(15/4) = 42.521 · · ·, which is explained
in appendix 15.15.3.
August 31, 2019 81
again decide to have C2 and C4 free of loop corrections. Obviously it will be E16
sufficient to restrict ourselves to the 1PI diagrams. Denoting the counterterm
interactions (and their formal loop order) by dots, we start by computing the
one-loop correction to C2 :
1 1 1
+ = − − δ1 ⇒ δ1 = − . (2.60)
2 2
1 3 3
+ = − η1 ⇒ η1 = . (2.61)
2 2
1 2
1
+ + + +
1 1 δ1 η1 7
= + + − − δ2 ⇒ δ1 = − ,
4 6 2 2 12
1 1 2
+ + + + +
3 3 3
= − − − 3 − 3δ1 − 3η1 − η2 ⇒ η2 = . (2.62)
4 2 4
E18 The resulting values for δ1,2 and η1,2 are of course precisely those we already
encountered in Eq.(2.2), only now we can compute them in a systematic,
E19
diagrammatic way.
E20
2.6 Exercises
Excercise 15 Some reflection
Verify the claim made below Eq.(2.5) by showing that every diagram contain-
ing dotted loops can be assigned its place in the modified SDe of Eq.(2.10).
Excercise 17 Dotting in ϕ3
Show that for pure ϕ3 theory (never mind its ill-defined character) there
are only two dotted 1PI diagrams possible. There are therefore no ‘infinite
towers’ of dotted diagrams. Write the appropriate SDe in this dotted model.
Such theories, where only a finite number of diagrams diverge, are called
super-renormalizable. Show that, in this case, we can afford to also dot loops
with 3 propagators, and end up with a renormalizable theory19 .
85
86 August 31, 2019
The precise expression of the G’s in terms of the C’s is of course now some-
what more involved : for instance, for K = 3 we have
G1,0,0 = C1,0,0 ,
G1,1,0 = C1,0,0 C0,1,0 + C1,1,0 ,
G1,1,1 = C1,0,0 C0,1,0 C0,0,1 + C1,1,0 C0,0,1
+ C1,0,1 C0,1,0 + C0,1,1 C1,0,0 + C1,1,1 . (3.6)
2
Where possible, we denote multiple integral by a single integration sign. This usually
does not lead to confusion.
August 31, 2019 87
We now have K field functions, one for each field ; they are given by
∂
φj (J1 , . . . , JK ) = ~ W (J1 , . . . , JK ) . (3.7)
∂Jj
An important thing to note is that, since the field functions are derivatives,
∂ ∂
φj (J1 , . . . , JK ) = φk (J1 , . . . , JK ) . (3.8)
∂Jk ∂Jj
as can easily be verified. For the field functions, the SDe is best illustrated
with an example. Suppose that we have the following action for K = 2 :
1 1 λ
S(ϕ1 , ϕ2 ) = µ1 ϕ1 2 + µ2 ϕ2 2 + ϕ1 2 ϕ2 2 . (3.10)
2 2 4
This time, the coupling constant λ carries a factor 1/(2!)/(2!) since there
are not four identical fields ‘meeting’ at the vertex, but rather two pairs of
identical fields, as mentioned above. We indicate the field type with either
‘1’ or ‘2’. The Feynman rules for this case are
~ ~ 1 2 −λ
1
↔ , 2
↔ , ↔ ,
µ1 µ2 1 2 ~
1 J1 2
J2
↔ , ↔ . (3.11)
~ ~
There are two coupled Schwinger-Dyson equations, one for each field :
1111
0000
0000
1111
0000
1111
0000
1111
0000
1111
1111
0000
0000
1111
1 0000
1111
0000
1111
0000
1111 0000
1111
0000
1111
1 0000
1111
0000
1111 1 1 2
0000
1111
0000
1111
0000
1111
0000
1111
0000
1111
= + 00000
11111
00000
11111
00000
11111
00000
11111
0000
1111 00000
11111
0000
1111 0000
1111
00000
11111
2 0000
1111
00000
11111
0000
1111
0000
1111
0000
1111
0000
1111
0000
1111
0000
1111
0000
1111
000000
111111
000000
111111 000000
111111
111111
000000 000000
111111
1 111111
1 1 111111
000000
000000
111111
000000
111111 1 2 000000
111111
000000
111111
000000
111111
000000
000000
111111
111111
000000
000000
111111
111111
000000 000000
111111
000000
111111
000000
111111
000000
111111
000000
111111 000000
111111 1 2 000000
111111
000000
111111
+ 2
00000
11111
+ 1
00000
11111
+ 2
000000
111111
000000
111111
000000
111111
000000
111111
,
11111
00000
00000
11111 11111
00000
00000
11111
2 00000
11111
00000
11111
00000
11111
2 00000
11111
00000
11111
00000
11111
00000
11111
00000
11111 00000
11111
00000
11111
00000
11111
00000
11111 00000
11111
00000
11111
88 August 31, 2019
0000
1111
1111
0000
0000
1111
0000
1111
2 0000
1111
1111
0000 0000
1111
0000
1111
0000
1111 0000
1111
0000
1111 0000
1111
0000
1111
0000
1111 2 2
2 0000
1111
0000
1111
0000
1111
0000
1111
0000
1111
= + 00000
11111
00000
11111
1
00000
11111
00000
11111
0000
1111 00000
11111
0000
1111 1 0000
1111
00000
11111
1111
0000
00000
11111
0000
1111
0000
1111
0000
1111
0000
1111
0000
1111
0000
1111
0000
1111
000000
111111 000000
111111
111111
000000 000000
111111
2 000000
111111
000000
111111
000000
111111
1 111111
000000 1 111111
000000
000000
111111
000000
111111
000000
111111 000000
111111 000000
111111
2 000000
111111
000000
111111
000000
111111
2 000000
111111
111111
000000
000000
111111
000000
111111
1 111111
000000
000000
111111
2 000000
111111
+ 1
00000
11111
+ 2
00000
11111
+ 000000
111111
2111111
000000
000000
111111
000000
111111
,
11111
00000 11111
00000
00000
11111 00000
11111
00000
11111
00000
11111 00000
11111
00000
11111
1 00000
11111
00000
11111
1 00000
11111
00000
11111
00000
11111 00000
11111
00000
11111 00000
11111
00000
11111 00000
11111
(3.12)
E21 with the following analytical representation for the field functions φj =
φj (J1 , J2 ) (j = 1, 2) :
2
J1 λ 2 ∂ ∂ 2 ∂
φ1 = − φ1 φ2 + ~φ1 φ2 + 2~φ2 φ1 + ~ φ1 ,
µ1 2µ1 ∂J2 ∂J2 (∂J2 )2
2
J2 λ 2 ∂ ∂ 2 ∂
φ2 = − φ2 φ1 + ~φ2 φ1 + 2~φ1 φ2 + ~ φ2 .
µ2 2µ2 ∂J1 ∂J1 (∂J1 )2
(3.13)
As before, these sum rules are valid in any nonzero dimension as well.
A few things are of interest here. In the first place, we have here the first
instance of an important concept : that of oriented lines. The propagator is
oriented, it runs from ϕ to ϕ̄. In the second place, in the action we find the
two terms Jϕ ¯ and ϕ̄J, which would suggest that J is the source in the SDe
¯
of ψ̄, and J is the source in the SDe for ψ ; but it is actually the other way
E24 around ! What is the source for a given field function is seen by taking the
derivative of the action, and inspecting which field then occurs as a linear
term, and which source term is left by itself after the differentiation.
August 31, 2019 91
+ + + ··· (3.27)
We see that the logarithmic term in S(B) simply embodies the closed ϕ
loops ! As we shall see in section 10.2.7, in ‘grown-up’ QED we have Furry’s
92 August 31, 2019
theorem that states that ϕ loops with an odd number of B legs must vanish
when summed correctly. We can implement this by using counterterms for E26
only these odd-n loops :
2k+1
X ~ eB ~ eB eB
− = − log 1 + − log 1 − . (3.29)
k≥0
2k + 1 m 2 m m
3.6 Exercises
Excercise 21 Two-field action
Consider the two-field action, now with the sources included :
µ λ
S(ϕ1 , ϕ2 ) = ϕ1 2 + ϕ2 2 + ϕ1 2 ϕ2 2 − J1 ϕ1 − J2 ϕ2
2 4
1. Determine the 2 SDe’s for the path integral.
2. From these, determine the 2 SDe’d for the field functions φ1 and φ2 .
3. Verify that this is agreement with the diagrammatically obtained result.
2. There are now of course also 3 field functions φ1,2,3 . Prove that
∂ ∂
φj = φi , i, j ∈ (1, 2, 3)
∂Ji ∂Jj
Cij = 0 , i 6= j ,
4.1 Introduction
The main characteristic of a space(-time) of more than zero dimensions is
the fact that the quantum field is defined at more than one point ; in fact,
at an infinity of points. The possibility of sending signals from one point to
another one requires the existence of correlations between the field values
at different points. The nature of this correlation, and its reflection in the
appropriate Feynman rules, is our subject now.
where the field labels run from −∞ to +∞. The collection of all their
corresponding sources is denoted by {J}. We shall, as a working example,
consider a theory where the interaction consists of four fields with the same
label meeting at one point. Moreover, we shall assume the kinetic terms to
95
96 August 31, 2019
X 1 λ4 4
2
S({ϕ}, {J}) = µϕn − γϕn ϕn+1 + ϕn − Jn ϕn , (4.1)
n
2 4!
where we include the sources in the action2 . If γ were zero, the action would
be separable and the theory would be an uninteresting series of replicas of the
zero-dimensional action for a single field. We shall consider positive values of
γ ; in that case, the action tends to minimize if ϕn and ϕn+1 carry the same
sign : a positive correlation between ‘neighbour’ fields is the result. Note,
moreover, that the action is, by choice, invariant under the relabelling of n
by n + K with any fixed K : this is called translation invariance, in this case
translation by a fixed increment in labelling3 . The model is also invariant
under the relabelling of n by −n : this is called parity invariance.
The Feynman rules are easily derived from the action of Eq.(4.1) :
n ~
↔
µ
n m γ
↔ + δm,n+1 + δm,n−1
~
n n λ4
n n ↔ −
~
Jn
n ↔ +
~
Feynman rules, version 4.1 (4.2)
1
If not indicated explicitly otherwise, sums will run from −∞ to +∞.
2
Both µ and γ are independent of n simply because I choose them so. I might choose
differently, only I don’t want to.
3
This will lead to momentum conservation later on. Note however that, as indicated
above, momentum conservation is a consequence of our choice, or in practice of our belief
in the translation invariance of our physical laws. Other models are possible and not a
priori wrong : they are simply much more complicated.
August 31, 2019 97
The identity of the field is indicated by its label. Alternatively, the four-
vertex and the source vertex may be labelled. The SDe now takes the fol-
lowing form, for any n :
111111
000000
000000
111111 111111
000000 111111
000000
000000
111111 000000
111111
000000
111111 000000
111111
000000
111111
000000
111111
000000
111111 000000
111111 000000
111111
n 000000
111111 n 000000
111111
000000
111111 000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
= + n n+1
000000
111111
000000
111111
000000
111111
000000
111111
+ n n−1
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111 000000
111111
000000
111111
000000
111111
000000
111111
11111
00000 111111
000000
000000
111111 1111
0000
0000
1111
00000
11111 000000
111111 0000
1111
00000
11111
00000
11111 000000
111111 0000
1111
n 00000
11111 n 000000
111111
000000
111111 0000
1111
0000
1111
+ 0000
1111
00000
11111
+ 000000
111111
000
111
000
111
+ n
0000
1111
0000
1111
0000
1111
, (4.3)
1111
0000 000
111 0000
1111
0000
1111
0000
1111
0000
1111
000
111
000
111
0000
1111
0000
1111
0000
1111 000
111
0000
1111
0000
1111 000
111
000
111
0000
1111
0000
1111
0000
1111 000
111
0000
1111
0000
1111 000
111
000
111
or, in terms of the field functions φn ({J}), that depend on all sources :
Jn γ
φn = + (φn−1 + φn+1 )
µ µ
∂2
λ4 3 ∂ 2
− φn + 3~φn φn + ~ φn . (4.4)
6µ ∂Jn (∂Jn )2
n m
Πm,n ≡ (4.5)
the total set of diagrams that contain only two-point vertices (or no vertices),
and have fields n and m at its external legs4 . The SDe can then be rewritten
as follows :
000000
111111
000000
111111
000000
111111
000000
111111 n
n 000000
111111
000000
111111
X k k
000000
111111
000000
111111
000000
111111
= +
000000
111111 n
000000
111111
000000
111111
k
n k
+ + n ,
(4.6)
k
4
To go from ϕn to ϕm one needs, of course, at least |n − m| vertices, but more vertices
are also possible.
98 August 31, 2019
where we must sum over the label of the field exiting the Π. Therefore, we
have
X
φn = Πn,k ×
k
∂2
Jk λ 3 ∂ 2
− φk + 3~φk φk + ~ φk . (4.7)
µ 6µ ∂Jk (∂Jk )2
The object Πn,k describes to what extent the fields ϕn and ϕk influence one
another : it will be called the propagator from now on.
0 n = 0 n + n + n , (4.8)
0 1 0 −1
or
~ γ
Π(n) = δ0,n + Π(n + 1) + Π(n − 1) . (4.9)
µ µ
The easiest way to solve this set of equations is by Fourier transform. We
define5 X
R(z) = Π(n) e−inz , (4.10)
n
6
from which the propagator may be recovered using
Z+π
1
Π(n) = e+inz R(z) dz . (4.11)
2π
−π
5
We are forced to use exp(−inz), with absolute value unity ; any other absolute value
would make the generating function divergent either as n → ∞ or n → −∞.
6
We choose e−inz rather than e+inz in Eq.(4.10) by convention. Although this may
not be completely, glaringly obvious at this point, this convention is ultimately related to
the fact that, in nonrelativistic quantum mechanics, the Schrödinger equation has been
chosen to read i~∂|ψi/∂t = Ĥ|ψi rather than −i~∂|ψi/∂t = Ĥ|ψi.
August 31, 2019 99
where we have introduced u = eiz . This allows us to write the integral (4.11)
as
un
I
~
Πn = − du , (4.13)
2iπγ (u − u+ )(u − u− )
|u|=1
Unsurprisingly, the propagator falls off exponentially with |n|. Note that if E27
γ were negative, then u− would also be negative, and the propagator would
oscillate between positive and negative correlations. Moreover if µ were 2γ
or smaller the poles of the integrand would lie on the unit circle |u| = 1,
making the integral ill-defined.
With this explicit form of the propagator, we can now switch to a new
set of Feynman rules :
7
This derivation is valid for n ≥ 0. For negative n, Cauchy’s theorem on which it is
based does not hold immediately : but in that case we can perform the variable transform
from u to 1/u and obtain the result.
100 August 31, 2019
m
n ↔ Π(m − n)
n n λ4
n n ↔ −
~
n Jn
↔ +
~
Feynman rules, version 4.2 (4.17)
The difference with the previous set of rules is that now the line denotes a
propagator running between n and m. The SDe is now very similar to that
of the zero-dimensional ϕ4 theory :
1111
0000
0000
1111
0000
1111
000000
111111 0000
1111
0000
1111
111111
000000
000000
111111
000000
111111
X n m n m 0000
1111
0000
1111
000000
111111 000
111
0000
1111
= +
n 000000
111111
000000
111111
000
111
000000
111111
000000
111111
000
111
000
111
000000
111111
000000
111111
000000
111111 000
111
000
111
0000
1111
m
000
111
0000
1111
0000
1111
0000
1111
0000
1111
0000
1111
1111
0000
0000
1111 1111
0000
n 0000
1111 n m 0000
1111
0000
1111
+ 0000
1111 + 0000
1111
. (4.18)
m 0000
1111 0000
1111
0000
1111
0000
1111 0000
1111
0000
1111
000
111 0000
1111
000
111
000
111 0000
1111
000
111 0000
1111
0000
1111
000
111 0000
1111
000
111
000
111 0000
1111
000
111
000
111
ϕ ϕ ϕ ϕ ϕ3
−1 0 1 2
the further those points are deemed to be apart ; in the language of these
notes, the smaller Π(m − n), the larger the ‘distance’ between n and m. We
can therefore dress up our picture by introducing a fundamental distance ∆,
subsequent field locations being separated by this distance,
∆ ∆ ∆ ∆ ∆ ∆
}
}
}
}
}
}
ϕ ϕ ϕ ϕ ϕ3
−1 0 1 2
where the distances between the successive points are all equal since the
couplings γ are all equal. We have, as it were, constructed a one-dimensional
universe. It may come as a surprise that the concept of space is here presented
as a visualization device. If we reflect, however, on how someone who (like
a new-born infant) has no a priori concept of spacelike separations would
have to envisage the workings of the physical world, we shall conclude that
that person had better invent space in order not to go insane pretty quickly.
In its essence, space, like so much else in the world around us, is simply a
mental construction that allows us to come to grips with, and control, our
environment8 .
After all this has been said, we must acknowledge the empirical fact
that to our knowledge space seems not to be made up from single points9 .
Therefore we have to assume that ∆ must be much smaller than the smallest
distances that can, at present, be resolved10 . We therefore introduce the
continuum limit : we assume that the theories we consider are such that the
limit ∆ → 0 can be taken in a sensible manner, yielding sensible results.
This sidesteps the interesting question of whether ∆ is really zero or not.
Indeed, we do not know. Any theoretical result that depends sensitively on
whether ∆ = 0 or ∆ 6= 0 would be extremely important since experimental
information about it would allow us a look at the fundamental structure of
space ; but for us it is safer to construct theories the predictions of which
do not hinge on this unknown. As we shall see, this can be made to work ;
8
See also Peter L. Berger and Thomas Luckmann, The Social Construction of Reality :
A Treatise in the Sociology of Knowledge (Garden City, New York: Anchor Books, 1966).
Additionally, the thought of all of us sharing a single point with every particle of our
bodies clashes with my sense of personal space to say the least.
9
Nor does it appear to be one-dimensional – but that is easily repaired, as we shall see.
10
About 10−19 meter.
102 August 31, 2019
as an added bonus, we can feel free from misgivings about the mathematical
rigour of taking the continuum limit — we may not be at the limit anyway.
This assignment is called the Weyl ordering. Its converse reads, of course,
∆ 0 ∆ 0
ϕn = ϕ(x) − ϕ (x) , ϕn+1 = ϕ(x) + ϕ (x) . (4.28)
2 2
In a sense, the field value ϕ(x) is sitting ‘in between’ the points ϕn and
ϕn+1 . Other assignments can be proposed, for instance ϕn = ϕ(x). However,
these are less attractive13 . Upon careful application of Weyl ordering and the
assumed continuum limits for µ and γ, the kinetic part of the action (4.1)
has the following continuum limit :
i X 1 ∆2
X hµ
2 2 0 2
ϕn − γϕn ϕn+1 = (µ − 2γ)ϕ(x) + (µ + 2γ)ϕ (x) =
n
2 n
2 8
X 1 1 0 2
Z
1 2 1 0 2
2 2 2
m ϕ(x) + ϕ (x) ∆ = m ϕ(x) + ϕ (x) dx . (4.29)
n
2 2 2 2
The interaction and source terms in the path integral do not have a factor
∆ coming out naturally, but we may simply define the continuum limits by
redefining the objects in the action :
λ4 → ∆λ4 , Jn → ∆J(x) , (4.30)
so that the continuum limit of the full action, including this time also the
sources, becomes14
Z
1 2 2 1 0 2 λ4 4
S[ϕ, J] = m ϕ(x) + ϕ (x) + ϕ(x) − J(x)ϕ(x) dx . (4.31)
2 2 4!
Note the notation with square brackets: the action is now no longer a num-
ber depending on (a countably infinite set of) numbers, but rather on the
functions ϕ(x) and J(x) ; this is called a functional.
13
For example, consider a function ϕ(x) that vanishes for x → ±∞. The integral
2ϕ(x)ϕ0 (x) dx then vanishes upon partial integration. Weyl ordering tells us that
R
where the sum also vanishes explicitly after relabelling. For the alternative assignment
ϕn = ϕ(x) the vanishing cannot be proven.
14
Strictly speaking, the Weyl ordering requires the replacement of Jn not by ∆J(x) but
by ∆J(x) + ∆2 J 0 (x)/2. The additional term, however, vanishes in the continuum limit as
∆ → 0, as do the higher powers of ∆ involved in the ϕn 4 term.
August 31, 2019 105
This version of the SDe is not meant to be solved, rather it tells us the
Feynman rules in the continuum limit :
↔ Π(x − y)
x y
λ4
↔ −
~
J(x)
x ↔ +
Z ~
dx for a vertex at position x
This comes with the understanding that the positions of all vertices are to
be integrated over, and that the field function φ is now a functional of the
source J. For a free theory there are no interactions, and we find
Z
φ(x) = dy Π(x − y) J(y) . (4.40)
We see that the free field is the sum of its responses to the source, weighted
by the correlation between the position where the field is measured and that
of the strength of the source at all points. It is this property that establishes
the propagator as the ‘differential-equation’ Green’s function; but this cor-
respondence is only valid for non-interacting theories (see also footnote 3 in
Chapter 1).
√
in field values are of order ∆, as mentioned above. Now imagine ‘zoom-
ing out’, that is, disregarding the value of ϕ1 , and inspecting only ϕ0 and
ϕ2 , which
R are now separated by 2∆. This is obtained by integrating over
ϕ1 : dϕ1 K∆ (ϕ0 , ϕ1 )K∆ (ϕ1 , ϕ2 ) = K2∆ (ϕ0 , ϕ2 ) , where the proportionality
constant is absorbed in the normalization
√ of the path integral. The typical
jump from ϕ0 to ϕ2 is now of order 2∆. We conclude that, if we resolve
the continuum path down to√a scale ∆, the typical fluctuations over this
scale will always be of order ∆. The typical path has a fractal structure.
Such behaviour, with zigs and zags at every length scale, is encountered in
Brownian motion – and in the behaviour of the stock market16 . In figure
4.1 we plot a typical fractal path running over 10,000 points separated by a
distance of 0.01, with ∆ = 1. The first plot shows all points ; in the second,
only every 10th point is used, and in the third plot only every 100th point is
used. The qualitative form of the three paths remains the same, as expected
for a fractal path. The average absolute value of the point-to-point jumps
are 0.80, 2.49, and 6.97, √
respectively : the ratios between these numbers are
indeed roughly equal to 10.
16
Note that this qualitative picture holds only for one-dimensional theories (and, luckily,
the price of stocks, bonds, futures etc is expressed in one-dimensional currency). In more
dimensions, the paths’ behaviour is even more wild.
August 31, 2019 109
We now have to figure out what the correct Feynman rules are in this
new language. To do so we use (what else ? ) the SDe. It suffices to restrict
ourselves to ϕ3 theory at the tree level, where it reads
Z
1 λ3 2
φ(x) = dy Π(x − y) J(y) − φ(y) . (4.43)
~ 2~
17
Recall the discussion on loose terminology in Chapter 0.
18
Indeed, the more-dimensional theories have been constructed expressly to make fields
at different points correlate to one another!
19
We use the same notation for the position-dependent quantities and their momentum-
dependent Fourier transforms. This will not lead to confusion since we shall soon drop
the position-dependent ones anyway.
110 August 31, 2019
~
k ↔
k 2 + m2
k1 !
k2 λ X
↔ − (2π) δ kj
k4 ~ j
k3
k q
1
↔ + J(q) (2π) δ (q + k)
~
Z∞
dk
for every momentum k
2π
−∞
single momentum mode. In that case the external legs in a diagram carry a
single momentum ; but all momenta of the internal lines have to be integrated
over. Also note that the vertices now carry Dirac deltas imposing momentum
conservation. This is a direct consequence of our choosing the vertices of the
theory to be position-independent20 . In addition, it has becomes necessary
to indicate how the momenta involved in the vertices are to be counted. It
is usual to count all the momenta either incoming or outgoing. The precise
convention is unimportant, but it is important that you use it consistently.
q k
(4.47)
this gives us
1 1
J(q1 ) 2 J(q2 ) (2π) δ(q1 + q2 ) . (4.50)
~ q1 + m2
we see here an important fact : every connected diagram contains one Dirac
delta informing us that overall momentum must be conserved. The diagram
q1
q2 q3
(4.51)
20
This means that the homogeneity of space(-time) can be investigated by very carefully
checking momentum(-energy) conservation in interactions. Of course, if vertices take on
different values very far away in space or time these effects may be undetectable.
112 August 31, 2019
k4
k3 (4.53)
ϕn → ϕ~n , ~n = (n1 , n2 , . . . , nD ) .
The obvious visualization for this choice is that of a space rather than
a line, covered with a regular square grid of fields, each connected to 2D
nearest neighbors: the corresponding continuum picture, therefore, is that of
a theory in D equivalent dimensions. The propagator of this theory obeys,
of course, the SDe
D D
~Y γX
Π(~n) = δn ,0 + Π(~n + k) + Π(~n − k) , (4.57)
µ k=1 k µ k=1
so that, now, µ must exceed 2Dγ. The continuum limit takes a different
form than in the one-dimensional case. We define
~x = (x1 , x2 , . . . , xD ) , xj = nj ∆ ,
~k = (k 1 , k 2 , . . . , k D ) , k j = zj /∆ , (4.59)
114 August 31, 2019
γ → ∆D−2 , µ → 2Dγ + m2 ∆D , λ4 → ∆D λ4 ,
ϕn1 ,n2 ,...,nD → ϕ(~x) , Jn1 ,n2 ,...,nD → ∆D J(~x) . (4.60)
↔ Π(~x − ~y )
x y
λ4
↔ −
~
J(~x)
↔ +
x ~
Feynman rules, version 4.5 (4.63)
δ2
λ4 3 δ 2
− φ(~y ) + 3~φ(~y ) φ(~y ) + ~ φ(~y ) . (4.64)
6 δJ(~y ) (δJ(~y ))2
21
The propagator is still a smooth function ; but, as you can easily check by comparing
to the discussion in section 4.3.5, for D = 2 the field function ϕ is no longer continuous.
For D ≥ 3, the discontinuities are not even guaranteed to be finite, and only the interaction
terms can keep the field from jumping infinitely wildly !
August 31, 2019 115
can be obtained directly from the continuum action by the functional Euler-
Lagrange equation
!
δ ~ · δ
S[ϕ, J] − ∇ S[ϕ, J] =0 . (4.66)
δϕ(~x) ~ x)
δ ∇ϕ(~
D
X
2 2
ds(d~x) = |d~x| = (dxj )2 , (4.67)
j=1
the Euclidean distance between the points ; this type of quantum field theory
is therefore said to be Euclidean.
22
The increase in symmetry depends on an interplay between the lattice action and the
form of the continuum limit ; it is possible to construct actions in which the continuum
symmetry is not larger than that of the lattice theory. But I prefer to choose actions that
(will come to) look like the physics of the real world.s
116 August 31, 2019
~
k ↔
~
|k| + m2
2
k1 !
k2 λ X
↔ − (2π)D δ ~kj
k4 ~ j
k3
k q
1 D
~
↔ + J(q) (2π) δ ~q + k
~
Z∞
dD k
for every momentum k
(2π)D
−∞
The function K is the so-called modified Bessel function of the second kind,
discussed in appendix 15.15.8 ; here we have used its integral representation
August 31, 2019 117
4.7 Exercises
Excercise 27 Guessing the correlator
For the ‘one-dimensional discrete model’ we found for the correlator the
recursion relation
~ γ
Πn = δ0,n + Πn+1 + Πn−1
µ µ
Make the following Ansatz:
Πn = A B |n|
k
where the loop momentum is indicated.
(a) Show that the external line carries no momentum.
(b) Write down the expression for the tadpole diagram. By direct inte-
gration over k from −∞ to +∞, show that the result is −λ3 π/(2m).
(c) Redo the above calculation in another way : the integrand has
poles at k = ±im. Close the integration contour in the complex k
plane and contract it around a pole. Show that it does not matter
which pole you choose.
2. Consider the following diagram :
p−k
p
k
where the external momentum and the loop momenta are indicated.
Show, using the same contour technique as above, that it evaluates to
1/(2m)/(p2 + 4m2 ). Show that again the choice of the upper or the
lower half-plane is irrelevant.
3. Apply the same technique twice to show that
1
=
(8m2 )(p2 + 9m2 )
August 31, 2019 119
4. Show that
p2 p1 2 + p2 2 + p3 2 + 24m2
p1 =
p3 (p1 2 + 4m2 )(p2 2 + 4m2 )(p3 2 + 4m2 )
where all external momenta are counted ingoing, so that p1 +p2 +p3 = 0.
120 August 31, 2019
Chapter 5
121
122 August 31, 2019
and the real distance between two events with coordinates xµ and xµ + dxµ
is given by
3
X
2 0 2
s(dx) = (dx ) − (dxj )2 ≡ gµν dxµ dxν ≡ dxµ dxµ ≡ dx · dx , (5.2)
j=1
(summation over repeated indices implied), where gµν is the covariant metric
tensor4
1 if µ = µ = 0
gµν = diag(1, −1, −1, −1) ≡ -1 if µ = ν ∈ {1, 2, 3} (5.3)
0 otherwise
g µα gαν = δ µ ν , (5.4)
so that g µν is numerically equal5 to gµν . The metric tensors allow for the
raising or lowering of indices : for instance,
The special rôle of time in physics is evidenced by the relative minus sign in
the metric tensor.
Note the special treatment of the zeroeth component, which we again put in
by hand 6 We can now solve Eq.(5.7)
Z
1 D 0 0 1 1 D−1 D−1
Π(~n) = d z exp −in z + i(n z + · · · + n z R(z) ,
(2π)D
D−1
!−1
X
R(z) = i~ µ − 2γ0 cos(z 0 ) − 2γ cos(z k ) . (5.9)
k=1
e−ik·x
Z
i~ D
Π(x) = d k (5.12)
(2π)D k 2 − m2 + i
m2 − i
Z
D 1 µ 2 λ4 4
S[ϕ] = d x ∂ ϕ(x)∂µ ϕ(x) − ϕ(x) − ϕ(x) + J(x)ϕ(x) .
2 2 4!
(5.14)
As we have seen, the Minkowskian nature of the theory has been imposed,
rather than derived, at two separate moments : first, the distinguishing of
γ0 and γ in Eq.(5.6), and second, the definition (5.8). Of course, there is
no a priori reason why the universe would have Minkowskian rather than
Euclidean symmetry8 . So the difference between γ and γ0 can be understood
from a phenomenological point of view. It is somewhat surprising, then, that
a second ‘by hand’ intervention, having z 0 act differently from z 1,...,D−1 , is
necessary to have position x and momentum k have the right Minkowskian
product.
k i~
↔
k · k − m2 + i
k1 k2 i
↔ − λ4 (2π)4 δ 4 (k1 + k2 + k3 + k4 )
k4 k3 ~
k1 k2 i
↔ + J(k2 )(2π)4 δ 4 (k1 + k2 )
~
Z∞
dk
for every momentum k (5.15)
2π
−∞
The vertices also pick up an additional factor i, and all vectors from now on
are assumed to be Minkowskian four-vectors.
e−ik·x
Z
i~ 4
Π(x) = d k
(2π)4 k 2 − m2 + i
0 0 ~
e−ik x eik·~x
Z
i~ 0 3
= dk d k 0 2 , (5.16)
(2π)4 (k ) − ω(k)2
August 31, 2019 127
Im k0
Re k0
where q q
~ 2 2
ω(k) = |k| + m − i ≈ |~k|2 + m2 − i (5.17)
exp(−ik 0 s)
Z
i~
Π(x) = dk 0 d3 k . (5.18)
(2π)4 (k 0 − ω(k))(k 0 + ω(k))
The integrand has two simple poles, one at ω(k) and the other at −ω(k). For
s > 0 the contour must be closed in the lower half plane in order to make the
exponent exp(−ik 0 s) vanish at infinity (cf figure 5.1). We can now perform
the k 0 integral :
Z
~ 1
Π(x) = d3 k exp(−isω(k)) . (5.19)
(2π)3 2ω(k)
128 August 31, 2019
The integration over the solid angle of ~k is trivial, and we are left with a
single integral :
Z∞
~ k2
Π(x) = dk exp(−isω) , ω = ω(k) . (5.20)
(2π)2 ω
0
where we have used Eq.(15.296) and partial integration. For negative s the
same result is found, but with −s instead of s. So we can formulate the final
form as
~m
Π(x) = −i K1 (im|s|) . (5.22)
(2π)2 |s|
exp(i~k · ~x)
Z
i~ 0 3
Π(x) = dk d k
(2π)4 (k 0 )2 − ω(k)2
Z∞ Z1 Z2π
~ k2
= dk dc dϕ exp(iskc)
(2π)3 2ω(k)
0 −1 0
Z∞ Z∞
−i~ k ~ k
= dk sin(ks) = − 2 dk exp(iks) . (5.23)
(2π)2 s ω(k) 8π s ω(k)
0 −∞
The integral of k over the real axis can be deformed as indicated in figure
5.2. The integrand has two cuts along the imaginary axis in conformity with
the usuals choice where the cut of z 1/2 lies along the negative real axis. Since
s > 0 we have to move around the cut in the upper half plane. The result is
~m
Π(x) = K1 (ms) . (5.24)
(2π)2 s
August 31, 2019 129
im
−im
It is gratifying that, in spite of the very different way in which they are arrived
at, the Minkowskian propagator for spacelike separations and the Euclidean
propagator of Eq.(4.69) (where all separations are spacelike anyway) exactly
coincide !
For large values of ms, we can use the asymptotic expansion of Eq.(15.305)
to arrive at
m1/2
Π(x) ∼ exp(−im|s|) for x · x = s2 > 0 ,
|s|3/2
11
Incidentally, this can easily be obscured since K1 (iz) is sometimes written as
(2) (2)
(π/2)H1 (z), where H1 is a so-called Hankel function. The world of Bessel functions
can be bewildering...
130 August 31, 2019
m1/2
Π(x) ∼ exp(−m|s|) for x · x = −s2 < 0 . (5.26)
|s|3/2
In figure 5.3 we give the two functions |K1 (ix)| and K1 (x) for real x > 0. For
imaginary argument, (i.e. in the timelike region) it shows a power-damped
oscillatory behaviour while for real argument (i.e. the spacelike region) it falls
off exponentially ; similar to the behaviour of, say, exp(x). This practically
prohibits any faster-than-light signalling by particles obeying this propagator
over distances much larger than 1/m.
We have seen that for small x · x of either sign, the propagators coincide
and are independent of m. We can understand this as follows. If x · x ≈ 0,
then |x0 | ≈ |~x| can be made very small indeed using an appropriate Lorentz
transformation12 . In that case the k integral will be dominated by very large
values, for which ω(k) ≈ k and hence the value of m becomes unimportant.
12
If x · x = 0 then x0 can be brought arbitrarily close to zero.
August 31, 2019 131
13
Note that in Eq.(5.30), ∂ stands for a derivative to x, not to y !
132 August 31, 2019
Z µ 2 −ik·(x−y)
1 4 4 (∂ ∂µ + m ) e
= − d y d k J(y)
(2π)4 k 2 − m2
Z 2 2 −ik·(x−y)
1 4 4 (−k + m ) e
= − d y d k J(y)
(2π)4 k 2 − m2
Z Z
1 4 4 −ik·(x−y)
= dydke J(y) = d4 y δ 4 (x − y) J(y) ,(5.30)
(2π)4
since the integral over k has become trivial14 . The free-field function φ(x) is
therefore seen to obey the equation
∂ µ ∂µ + m2 φ(x) = J(x) ,
(5.31)
k 1
↔ i~
k · k − m2 + imΓ
The propagator for an unstable particle with mean life-
time 1/(Γc).
Feynman rules, version 5.1 (addendum) (5.37)
The i prescription is seen to just mean that we should treat stable particles
as the infinitely-long-lifetime limit of unstable particles.
Equation (5.35) describes particles at rest since there is no space depen-
dence in the wave function. The lifetime/width is therefore that of particles
at rest, which is indeed the usual definition. For particles in motion, time
dilatation indicates that the lifetime be increased by a factor p0 /m ; we shall
encounter this situation in the next chapter.
An issue that appears resolved is the direction of time flow. Whereas
Minkowski space itself, being essentially static, does not assign any preferred
direction associated with the time coordinate, the direction of time flow is
now defined to be that direction in which unstable particles disappear , rather
than appear19 .
Another point to be noted is the following. The unstable propagator by
itself is seen to lead to a decreasing overall probability : that contradicts
the normal unitary evolution of quantum mechanics. But it is not the whole
18
Remember that Γ has the dimension of inverse length, not kilograms ! In practice, it
is customary to give total decay widths, like particle masses, in GeV, and then τ = ~/Γ.
19
Attractive as the above argument appears, a drawback comes from the case x0 < 0.
In that case, the contour integral must be closed along the upper half plane, so that the
pole k 0 = −m + iγ/(2m) becomes the significant one. We find φ(x) ∝ exp(−|t|/τ ), which
is to be interpreted as a particle density that starts out as zero at t = −∞, and grows to a
crescendo at t = 0 ; this lacks an obvious interpretation. We ascribe this to the use of the
simple form (5.33). A better source, needed for a more rigorous treatment, can be simply
constructed. Notice that this really means that the direction of time is governed by the
sources !
August 31, 2019 135
Z∞
eikr e−ikr
1
= dk k −
4iπ 2 r k 2 + m2 k 2 + m2
Z0
1 k
= dk eikr . (5.40)
4iπ 2 r k2 + m2
For r > 0 we can close the integration contour in the upper half of the
complex-k plane, and we find
1 exp(−mr)
φ(x) = . (5.41)
4π r
This is the so-called Yukawa potential, introduced in the 1930’s as a model for
E34 the strong nucleon-nucleon force, with m the mass of the pion. The Compton
wavelength of the pion is, indeed, roughly the range of the nuclear forces. If
we take m → 0 we find the Coulomb potential of a static electric source ;
but the real propagator of the photon field, responsible for the Coulomb
interaction, is more complicated, so that the above derivation is more or less
just handwaving in the case of electromagnetism. A fuller treatment is given
in section 10.5.
shows that it emits particles with all kinds of wave vectors k µ = (k 0 , ~k),
centered around values pµ /~, with pµ = (p0 , p~). For a bridge to non-quantum
August 31, 2019 137
Im k0
0
Re k
physics to be built, both the position and wave representation of the source
should be adequately localized ; σ0 and σ should be neither too large nor too
small. For now, we do not assume any particular relation between p0 and p~.
We want to study the response of the field to this source for positive times.
We have
Z 0 0
exp −ik x + ik · ~x~
φ(x) ∼ d4 k J(k) . (5.44)
(k 0 )2 − |~k|2 − m2 + i
In figure 5.4 we exhibit the k 0 integral contour in the complex plane. For
x0 > 0, the contour is to be closed in the lower half complex-k 0 plane. The
integrand has simple poles at the loci
0 ~ 0 p0 i
k = ω(k) − i , k = − ,
~ σ0
p0 i
k 0 = −ω(~k) + i , k 0 = + ,
~ σ0
the latter two lying outside the contour.
The k 0 integral therefore leads to the following expression for φ(x) :
Z 2 !
p
~
φ(x) ∼ d3~k exp i~x · ~k − σ 2 ~k − A(k 0 , ~k) ,
~
138 August 31, 2019
exp −ix 0
ω(~k)
1
A(k 0 , ~k) = 2
2ω(~k) p0 /~ − ω(~k) + 1/σ0 2
The second term decays exponentially at the same rate as the source. Since
we are interested in the behaviour of the field when it is free, i.e. unaffected
by any interactions, we can only study that behaviour once the source has
died out, and then so has this term20 . The first term describes Fourier modes
of the field that obey the dispersion relation k 0 = ω(~k), together with the
resonance condition that tells us that the field can only be appreciable if
both p0 /~ ≈ ω(~k) and p~/~ ≈ ~k. We therefore expect any fruitful resonance
in the field, which can allow for the transmission of signals over macroscopic
distances, only if
p0
p~
≈ω . (5.46)
~ ~
If we relate the zero component p0 (with dimension kg m/s) to an energy E
by writing
p0 = E/c , (5.47)
we find that the only particle modes emitted by the source that have a chance
of propagating over distances much further than σ must satisfy
Mc
q
E≈ |~p|2 c2 + M 2 c4 , m= . (5.48)
~
This is the mass shell condition, which prescribes the relation between the
energy E (in Joule), momentum p~ (in kg m/s), and mechanical mass M (in
kg) of a particle moving freely through spacetime. We recognize the quantity
m that we have been using so far as the inverse Compton wavelength of the
particle. A particle is called on-shell if its momentum pµ satisfies Eq.(5.48) ;
if not, it is called off-shell. Off-shell particles are not exotic or improbable ;
they are just not visible as the result of an experiment since they cannot
20
This is comparable with what you would do classically: studying the trajectory of a
thrown ball to see whether Newton’s laws are obeyed only makes sense once the ball has
definitively left your hand.
August 31, 2019 139
∂
∂
~k
x k − ~x · ~k =
0 0
x ω(~k) − ~x · ~k =
0
x0 − ~x = 0 . (5.49)
∂k~ ~
∂k ω(k)~
5.2.6 Antimatter
We again consider the free SDe :
dk 0 d3~k exp(−ik · x)
Z Z
0
φ(x , ~x) ∼ J(k 0 , ~k) ,
2π 3 ~
(2π) (k ) − ω(k) + i
0 2 2
q
ω(~k) = |~k|2 + m2 . (5.51)
21
In popular literature, off-shell particles are often dicussed with a lot of mumbling
about ‘uncertainty relations’, ‘borrowing energy from the vacuum’, and so on. Do not
allow yourself to be misled ! When a theorist starts invoking the uncertainty principle as
a reason for something, keep your hand on your wallet. The ‘uncertainty principle’ is not
a principle, not a reason, but a result.
140 August 31, 2019
x x
B B
E>0 E<0
A A
t t
If x0 > 0, the integration contour can be closed through the lower half of the
complex k 0 plane :
d3~k
Z
0
φ(x , ~x) ∼ exp −i(x0 ω(~k) − ~x · ~k) J(ω(~k), ~k) . (5.52)
(2π)3 2ω(~k)
If, on the other hand, x0 < 0, the closure must be over the upper half of the
plane, and then
d3~k
Z
0
φ(x , ~x) ∼ exp −i(−x ω(k) − ~x · k) J(−ω(~k), ~k) . (5.53)
0 ~ ~
(2π)3 2ω(~k)
We see that the propagator essentially describes plane waves, with the fol-
lowing characteristic: positive energies travel towards the future, and negative
energies travel towards the past.
While the concept of particles with positive kinetic energy, moving from
past to future, conforms to our everyday experience, the idea of negative
(kinetic) energies and movement backwards in time is not only æsthetically
repellent but may lead to splitting headaches in the verbal description of
physical processes. When, however, we consider more closely how such a sit-
uation will appear, it becomes clear that negative energies moving backwards
in time are indistinguishable from positive energies moving forward. Some
simple bookkeeping with the help of the two diagrams in figure 5.5 comes
in handy here. Consider two loci in space, denoted by A and B. In the
August 31, 2019 141
first diagram a particle moves forward in time, with positive energy, from
A to B. As a result the energy at A decreases, and that at B increases.
In the second diagram, a particle with negative kinetic energy starts at B,
and moves backwards in time to A. The net effect on the energies at A and
B is exactly the same ! The two situations are indistinguishable from the
point of view of the energy balance22 . There may still be a difference, of
course ; if the particles have additional properties such as electric charge,
the backwards-moving particles will appear with the opposite charge. For
instance, a negatively charged electron moving backwards will appear as a
positively charged positron moving forward, as can be seen from the two di-
agrams above. Such re-interpreted time-reversed particles are called
antiparticles 23 . Every particulate object whose propagator contains the
denominator of Eq.(5.51) is seen to contain both the regular particles and
their antiparticles. Moreover, we find the fundamental result that particles
and their antiparticles must have exactly the same mass and lifetime. The
antiparticle interpretation is just the way we surrender to a prejudice about
motion in time24 . Particles and their antiparticles may be identical, the pho-
ton being an example. Such particles must, of course, be electrically neutral.
On the other hand, not all neutral particles are their own antiparticles ; neu-
trons and antineutrons are distinct from one another25 . We have thus found
the following result for free particles : if we (a) replace all particles by their
antiparticles and vice versa, the so-called charge conjugation operation C,
(b) inverse all space directions26 , the so-called parity transformation P, and
(c) invert the direction of time, the so-called time reversal operation T, then
the world will look exactly the same ! This is (a restricted form of) the CPT
theorem, valid for the propagation of free particles. The more interesting,
22
At this point you may wonder about lots of the particles that constitute you and
the world actually moving backwards in time. Try to formulate, say, your own digestive
processes in this light — it doesn’t bear thinking about.
23
A commonly made error is to say that antimatter moves backwards in time. This is
wrong. Antimatter is invented precisely to make everything move forward in time !
24
Physicists from some alien civilization might have less problems with the other inter-
pretation.
25
Once the neutron is seen to be a collection of charged quarks, the distinction be-
comes obvious. On the other hand, neutrinos, while electrically neutral, are not equal to
antineutrinos, and are yet believed to be elementary.
26
Since, as can be seen from our diagrams, inverting the direction of the motion through
time will simultaneously change motion towards the left (say) into motion towards the
right, and so on.
142 August 31, 2019
A
t
t0
real CPT theorem, valid also for interacting particles, needs more tools than
we have at our disposal right now : its proof is referred to Appendix 15.14.
Let us now consider the (classically depicted) path a particle tracks out
in spacetime, as given by the space-time diagram given in figure 5.6. In one
description, the particle starts at A and moves to B, where at time t0 it
reverses its time direction, and moves backwards in time to C. In the alter-
native description, a particle starts at A and its antiparticle starts at C, and
the pair collides at B at time t0 . For times later than t0 , the particle and/or
its antiparticle have disappeared ; but because of momentum conservation
their combined energy has to be transferred onto one or more other parti-
cles (not depicted). The two descriptions are completely equivalent, but the
second one conforms much better to the way we tend to view the world. At
the ‘collision/reversal-point’ B the particle coming from A must dump its
energy, and even an additional amount since its energy must become nega-
tive for it to start moving backwards to C. Therefore, particle-antiparticle
annihilation releases energy, often in the form of photons27 . For instance,
27
It is sometimes stated that particles can only annihilate with their own antiparticle.
This is a somewhat restricted point of view, since for instance electrons can annihilate
with anti-neutrinos into W particles, as we shall see. It may be more appropriate to
say that it needs particles with their own antiparticles to annihilate into something that
has quantum numbers (electric charge, fermion number, etcetera) equal to those of the
vacuum. Neutrinos and their antineutrinos cannot easily annihilate into photons, being
electrically neutral : but they can annihilate into one or more Z bosons.
August 31, 2019 143
d3~k
q
1
, ω(k) = ~k 2 + m2 .
~
(2π)3 2ω(~k)
Note that if k 0 is positive for an on-shell particle in any given inertial frame,
it is positive in all intertial frames that can be reached by Lorentz transfor-
mations from the first one. This ensures that the step function θ(k 0 ) always
has the same value, irrespective of any Lorentz boosts we may care to make.
Lorentz covariance of the phase space integration element is thus guaran-
teed. We shall use the density of states (5.54) for all on-shell particles in the
calculation of cross sections and lifetimes.
If, for a given scattering process, the final state contains N particles
with masses mj , j = 1, 2, . . . , N , and momenta pµ1 , pµ2 , . . . , pµN , the combined
phase-space integration element is
dV (P ; p1 , p2 , . . . , pN ) ≡
N
! N
!
Y 1 X
3
d4 pj δ(pj 2 − mj 2 ) (2π)4 δ 4 P − pj , (5.55)
j=1
(2π) j=1
28
Note that the simpler-seeming process e− e+ → γ is kinematically impossible if the
resulting photon is to be on its mass shell. On the other hand, a single off-shell photon
can be produced, but such a photon must immediately decay again, in for instance a
particle-antiparticle pair of some kind.
144 August 31, 2019
5.3 Exercises
Excercise 30 Dangerous
Explain why, in Eq.(5.11), it is important that > 0.
m2 − i 2 λ 4
S(ϕ) = ϕ − ϕ − Jϕ
2 4!
1. Using the i prescription, show that
Z∞ r
m2 2
2π~ −iπ/4
dϕ exp −i ϕ = e
2~ m2
−∞
2. Find the Feynman rules for the propagator and the vertices.
Excercise 32 Useful
Show that the integral Z
I0 = dx exp(ixy)
where x runs over the whole real axis and y is an arbitrary real number, is
ill-defined. Show that the integral
Z
I = dx exp ixy − x2
2
29
Conservation of total energy and momentum.
August 31, 2019 145
1. Show that the path integral over ψ(x) can be performed independently
at every point in Minkowski space.
2. Compute the result of the ψ integration for the path integral. This is
called integrating out the field ψ.
3. Show that after having integrated out ψ we are left with the ‘original’
ϕ4 theory, provided we choose λ to be a combination of g and µ. Find
this combination.
4. Write out the coupled SDe’s for the ϕ−ψ theory in configuration space,
using the formalism of functional derivatives. Insert the SDe for ψ into
that for ϕ and show that the correct ϕ4 -SDe for ϕ is again obtained.
6. Show that the method of auxiliary fields does not work in Euclidean
space since there the path integral is not defined (it is easiest to do this
in zero dimensions).
Chapter 6
Scattering processes
6.1 Introduction
In this chapter we turn our attention to the bread-and-butter subject of parti-
cle phenomenology : the description of scattering processes. We shall discuss
the way in which Feynman diagrams and their evaluation are postulated to
predict the probability for finding specified final states given specified initial
states. We also investigate the consequences of the claim that our approach
describes quantum physics and is therefore of a probabilistic nature : that
is, we can only compute probabilities, which are necessarily bounded1 . This
leads to the notion of unitarity and the use (and usefulness) of cutting rules.
147
148 August 31, 2019
which the initial-state particles approach one another and meet, hopefully2 ,
in the interaction point, where the dynamics takes place. The final state is
observed by the detector operated by the particle physicists.
Since not only the scattering itself but also the initial-state preparation
and the final-state observation are quantum processes, all these parts of the
process must, according to our assumptions, be described by Feynman di-
agrams in a manner still to be established. The diagrammatic form of the
complete process will then look as follows :
and here (and in the following) we adopt the convention that the initial state
appears on the left-hand side of the diagrams, and the final state on the right-
hand side3 . This does not imply any spatial or timelike relation between any
of the vertices in the diagram: indeed, they are supposed to be integrated
over all of spacetime4 . Another observation on the above diagram is also
relevant : the initial-state preparation and the final-state observation should
contain physics that is better understood than the scattering part, and there
should be a clear notion of precisely which particles constitute the initial
and final states. This is indicated by the identifiable propagators connecting
the various ingredients of the process. We therefore adopt the idealization
that the only relevant part of the scattering should reside in the central, or
2
In the idealised setting in which particles with perfectly well-defined momenta form
plane waves of infinite spatial extent, they can hardly avoid meeting. In practice, the
momenta and spatial extensions of the particles’ wave packets are of course more limited.
3
This convention is based on not much more than the fact that in Latin writing you
move from left to right. Indeed a not uncommon alternative is to move upwards, so that
the incoming state is at the bottom. I do not know of any script that moves like that. It
could be worse : fortunately no-one employs boustrophedonic diagrams !
4
Of course, if there is any justice the contribution from paths in which a vertex is very
far out ought to be small.
August 31, 2019 149
We now have to confront the two following questions. In the first place,
which Feynman diagrams should occur in the scattering part ? And sec-
ondly, in actual experiments the initial- and final-state particles travel over
many meters between preparation, scattering, and detection. These particles
should therefore be on their mass shell, but isn’t this precisely the case in
which their propagators blow up ? The situation obviously calls for some
reinterpretation and additional Feynman rules, to which we shall come.
Before finishing this section, let us remark that also initial states consist-
ing of only a single particle occur :
In this case, we simply study the decay properties of the particle, such as its
total or partial decay width.
where as before the shading indicates connected diagrams. Now, recall that
every vertex in any diagram contributes a Dirac delta imposing energy-
momentum conservation. Therefore, every connected diagram has an overall
150 August 31, 2019
asks for particles carrying positive energy to originate (by some interactions)
from the vacuum. Such contributions therefore vanish by energy conserva-
tion, and the only contributing diagrams are contained in the totally con-
nected blob. Next, consider two-particle scattering. Since we forbid satellite
diagrams the only possible contributions are given by
0000000
1111111
1111111
0000000
0000000
1111111
1111111
0000000 0000000
1111111
0000000
1111111 0000000
1111111
0000000
1111111
0000000
1111111
= 0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
+ 0000000
1111111
0000000
1111111
+ · · · (6.3)
0000000
1111111 1111111
0000000
0000000
1111111 0000000
1111111
0000000
1111111
0000000
1111111
Now, the second term here is in principle possible but only if a) the two
E38 incoming particles are inherently unstable5 and b) the outgoing particles ar-
range themselves in precisely two groups according to the indicated decay
patterns. Leaving aside such special cases, we conclude that the scattering
amplitude is given by the connected Feynman diagrams. Note that
the restriction to connected diagrams only arises here from simple energy
considerations, and not from any deep inherent superiority of connected di-
agrams over disconnected ones : in essentially all cases of interest, the result
of the disconnected diagrams vanish anyway6 .
We may imagine situations where particles can be created from the vac-
uum. This is the cases in ‘field theories at high temperature’ where processes
take place in a heat bath which can deliver energy to create particles, so en-
ergy conservation is relaxed. In such a picture the heat bath is the ‘vacuum’
of the theory, and diagrams such as that of Eq.(6.2) are not automatically
zero. Another delicate situation is that of more incoming particles : for
instance, we might consider four particles scattering into four, in which we
5
This makes the notion of particles ‘coming in from infinity’ conceptually dubious in
this scattering.
6
There are plans on the table for a muon collider. Muons are unstable particles, but
their lifetime (especially at high energies) is extremely long as particle lifetimes go, so we
can treat them as essentially stable in most processes. However, it is possible that final
states are produced where one of the muons actually decays ; then we have to take more
care (see also exercise 38).
August 31, 2019 151
Here, M stands for the transition amplitude, which we still have to establish.
The symbol h|M|2 i indicates that in accordance with quantum-mechanical
practice we have to square the absolute value of M in order to arrive at
a probability, and the brackets indicate summation and/or averaging over
degrees of freedom other than the momenta : at present such degrees of
freedom are not in our theory yet, but they will arrive ! The symbol ΦΓ
denotes the collection of factors that must be included to account for the
density of states for the incoming particle, etcetera. That such a factor must
be present is clear from the fact that dΓ must have dimension 1/L. dV
denotes the phase-space integration element discussed in section 5.2.7. The
total decay width in this particular decay channel is obtained by integrating
over the available phase space7 . The momentum P µ is that of the incoming
particle at rest. The symmetry factor Fsymm is included to handle identical
particles in the final state. In quantum mechanics, the statement that two
particles are identical means that an interchange of these particles leads to
the physically identical final state, so that an unconstrained summation over
7
This can become very challenging !
152 August 31, 2019
with its own flux factor, also still to be determined. We see that, in order to
get the formulae (6.5) and (6.6) to actually work, we have to establish
• the algorithm to derive from the connected Green’s function the am-
plitude. In particular this calls for a special treatment of the external
lines.
together with the k’s. Nevertheless, the complete final state consists of both
the k’s and the q’s. The relevant diagrams thus look like
pa 111111
000000
000000
111111
000000
111111
000000
111111
}k 1,2,...
M=
000000
111111
000000
111111
A
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111 p
1111111
0000000
0000000
1111111
0000000
1111111
0000000
1111111
B
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
}q 1,2,...
(6.7)
pb
Note that the connected blobs may themselves contain many different in-
dividual diagrams. By separating the blobs A and B we indicate that the
unstable particles is actually quite long-lived so that the place where it is
produced and that where it decays tend to be clearly separated.
Now, we shall assume that we have, somehow solved the problem of how
to go from connected Green’s function to amplitude, and that we have applied
this procedure to the above process, in particular to the external lines. We
then have for the amplitude the form
i~
M=A B , (6.8)
p2 − m2 + imΓ
where p = q1 + · · · + qn is the momentum of the (internal !) line correspond-
ing to the unstable particle, and p2 = p · p. The unstable particle’s mass
is m, and its total decay width is Γ. The symbols A and B stand for the
‘processed’ connected Green’s functions for the ‘production’ process A and
the ‘decay’ process B, but with the Feynman factors for the unstable inter-
mediate particle removed. Assuming, for simplicity, that Fsymm = 1, we then
have for the differential cross section the form
~2
dσ = Φσ |A|2 |B|2 dV (P ; k1 , . . . , kj , q1 , . . . , qn ) , (6.9)
(p2 − m2 )2 + m2 Γ2
where P = pa + pb . In order to emphasize that p is the sum of the q’s, we
may write this also as
dσ = Φσ |A|2 |B|2 dV (P ; k1 , . . . , kj , q1 , . . . , qn )
~2 d4 p
(2π)4 δ 4 (p − Σq) , (6.10)
(p2 − m2 )2 + m2 Γ2 (2π)4
with obvious notation for the sum over the wavevectors q.
154 August 31, 2019
Now, we let the unstable particle approach stability, so that the location
where it decays becomes widely separated from that where it is produced.
That is, we examine the case that Γ becomes very, very small, and we may
E39 approximate
1 π
→ δ(p2 − m2 ) . (6.11)
(p2 − m2 )2 + m2 Γ2 mΓ
dσ = ~ |A|2 dV (P ; k1 , . . . , kj , p)
1 1
~ |B|2 dV (p; q1 , . . . , qn ) .
(6.14)
Γ 2m
Now, step back and consider what it is we are actually computing here : it
is the cross section for producing an almost-stable particle p, together with
the k’s in a specified configuration, followed by the decay of the particle p
into a specified configuration of q’s. Under the usual ideas of conditional
probability, this is the same as first computing the cross section for the
production of p and the k’s, followed by the conditional probability that,
given p, we see it decay into the q’s. This conditional probability, called the
(differential) branching ratio, is the partial decay width for p to go into the
q’s (computed in the p rest frame), divided by the total decay width, in this
case Γ (also defined in the rest frame). We conclude that
• ~ |A|2 is h|M|2 i for the process pa + pb → k1 + · · · + kj + p ;
• ΦΓ is given by 1/(2m).
August 31, 2019 155
In a sense, we have managed to cut through the p line, and interpret the
process rather as it would be given by the diagrams
pa 000000
111111
111111
000000
000000
111111
000000
111111
}k 1,2,...
M →
000000
111111
000000
111111
A
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
p p 1111111
0000000
0000000
1111111
0000000
1111111
0000000
1111111
B
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
}q 1,2,...
(6.15)
pb
A point to be noted here has been somewhat hidden so far. The connected
Green’s functions contain factors (2π)4 δ 4 (· · ·) for overall momentum conser-
vation. This conservation has been imposed already, however, in our choice
of the phase space integration elements dV . We therefore have to remove
these factors as well in the transition from connected Green’s function to M.
What about the treatment of the external lines ? In the above discussion
we started with p as an internally ocurring unstable particle, carrying its
own propagator. As we let it become stable, the propagator has disappeared
into the phase space counting, leaving only a residue of a factor ~2 . At the
end of the story the particle p has become a stable particle occuring as an
external line in the blob A. This, therefore, must be the prescription for
the external lines ! This is called truncation or amputation of external lines.
An external line√must apparently carry, instead of its undefined propagator,
simply a factor ~. We arrive at the following, expanded set of rules for the
calculation of scattering amplitudes M (as opposed to Green’s functions) :
156 August 31, 2019
k i~
↔ internal
k · k − m2 + i
111111111
000000000
000000000
111111111
000000000
111111111
000000000
111111111
000000000
111111111
000000000
111111111 k √
000000000
111111111
000000000
111111111
000000000
111111111
000000000
111111111
000000000
111111111
000000000
111111111
↔ ~ external
000000000
111111111
000000000
111111111
000000000
111111111
k1 k2 i
↔ − λ4 (2π)4 δ 4 (k1 + k2 + k3 + k4 )
k4 k3 ~
k1 k2 i
↔ + J(k2 )(2π)4 δ 4 (k1 + k2 )
~
Z∞
dk
for every momentum k (6.16)
2π
−∞
Let us now turn to the two-particle initial state. The flux factor ΦΓ for
particle decay has been found to be 1/(2m). It is related to how we count
the density of states of the incoming particle9 . We can directly translate this
to the case of two-particle scattering. Let us work in the Lorentz frame in
which particle b is at rest while particle a impinges upon it10 . Keeping in
mind the effect of Lorentz transformations on the density of states we see
that whereas mb remains, ma has to be replaced by p0a , in accordance with
the discussion in section 5.2.3. The density-of-states factor for the two-body
initial state is therefore 1/4p0a mb . Since, however, we are asking for a cross
section rather than a transition rate, we have to divide this by the velocity
of particle a in b’s rest frame, that is, by a factor |~pa |/p0a . The flux factor
therefore becomes
Φσ = (4mb |~pa |)−1 .
This expression, being given in a specific Lorentz frame, is not very attractive.
9
If we fix the particle to be at rest, we must count the states as d3 (~k) d3~k/2ω(~k) =
R
1/2m.
10
Not the most practical frame, to be sure : at the LHC it is a frame in which the whole
ATLAS detector, say, moves at 99.99999882% of the speed of light.
August 31, 2019 157
It often happens that the colliding particles have masses that are negligible
compared to their combined invariant mass, the square of which is commonly
denoted by the Mandelstam variable s. In that case, we may write
1
Φσ ≈ , s ≡ (pa + pb )2 . (6.19)
2s
This finishes our bootstrap treatment of the relation between connected
Green’s functions and scattering amplitudes, or matrix elements.
as required. Similarly, for the cross section of two particles going into n
particles we have
again as required. Note that the above analysis is kept simple because we
have restricted ourselves to the use of wavevectors rather than mechanical
momenta, which would introduce additional factors of ~ in the calculation.
The other natural constant, c, need not enter here.
where we have indicated the momenta of the particles. Let us write the
corresponding amplitude as M(p1 , p2 , q1 , q2 ). By moving particles from the
initial to the final state13 , or vice versa, we can then find the amplitudes for
13
You can visualize this by taking an outgoing particle, say, and dragging its external
leg from the final to the initial state.
August 31, 2019 159
where S is the matrix describing the time evolution of the incoming state from
t = −∞ to t = +∞. An important consideration here is the conservation of
probability. That is, any initial state |ii must go to some final state |f i with
100% probability15 ; of course |f i = |ii may also be one of the possibilities.
Writing this using the S matrix we have
X X
i S † f f S i = i S † S i .
1= | hf |S| ii |2 = (6.28)
f f
Conversely, any final state |f i must have come from some initial state |ii
with 100% probability, so that
X X
f S i i S † f = f SS † f .
1= | hf |S| ii |2 = (6.29)
i i
E40 Since this must hold for all states |i, f i we have
S † S = SS † = 1 , (6.30)
and the S matrix is unitary. Note that we have had to assume that the set of
initial states |ii and final states |f i are complete16 . The free-particle states
are natural choices for complete orthonormal bases, and we see that M is
simply a matrix element of the S matrix. We shall investigate this in some
more detail.
For simplicity, let us assume that we can label the initial states with a
discrete label i, and the final states by a similar discrete label f . We can
then write the S matrix element as
S f i = δ f i + Mf i , (6.31)
where the Kronecker delta embodies what would happen if there were no in-
teractions : the only possible observed final state would in that case be iden-
tical to the initial state (two particles, say, continuing on their way without
having interacted). The remainder Mf i is the object described by Eq.(6.27) ;
it is the result of the interactions of the theory, and is described by the Feyn-
man diagrams. Note that Mii 6= 0 is quite possible ; it corresponds to the
case where the final state happens to reproduce the initial state, so to speak
15
Try to imagine a world in which this does not hold ! Conservation of probability is a
dogma — but a reasonable one.
16
Again, try to imagine what things would look like if that were not the case. . .
August 31, 2019 161
or, in terms of M :
X
Mf i + M∗if + M∗kf Mki = 0 . (6.33)
k
which immediately shows that the forward scattering amplitude must have
negative real part18 . In Eq.(6.34), the first term is called the dispersion term,
and the second one the absorptive term, terms which originate in the study
of waves moving through a medium. Also, in our ‘discrete’ formulation,
which implies that Mf i can not be arbitrarily large ; we shall employ this
idea extensively later on. In appendix 15.10 we derive a more precise bound,
called partial-wave unitarity, but that calls for Feynman rules we still have
to develop.
A an elementary illustration of the optical theorem we consider the fol-
lowing physical process. We start with an empty initial state i (that is, a
state containing no particles). At some moment a source kicks in, producing
an unstable particle with wavevector p, mass m and total width Γ. Sometime
later, the same source absorbs the particle, and at the end the final state f
is empty again. The simple Feynman diagram describing this is
J
p J
17
Since S may be an infinite matrix, both conditions are necessary, whereas for a finite
matrix one would suffice.
18
A word of caution : in much of the literature, the statement reads that the amplitude
must have positive imaginary part. This is simply due to the fact that in those texts, the
S matrix element is written not δ + M but δ + iM. I do not see any particular virtue in
this.
162 August 31, 2019
and since the initial and final state coincide (f = i) this is a forward scattering
amplitude ; it must obey the optical theorem. We shall now verify this. The
matrix element is given by19
J i~ J
Mii = i 2 2
i , (6.36)
~ p − m + imΓ ~
so that
J2 mΓ
Re (Mii ) = − , (6.37)
~ (p2 − m2 )2 + m2 Γ2
which is indeed negative. Now, we consider the matrix elements Mki as used
in Eq.(6.34). These describe the initial state i going over in any final state
k, that is, they describe the decay of the particle after it has been produced
by the source :
111
000
J 000
111
000
111
000
111
000
111
000
111
p 000
111
000
111
000
111
But this is, of course, precisely the prescription for the decay width of the
particle, if we realize that the final state k indicates not only all possible
final states, but also that the summation over k should include the phase-
space integration. This short excercise illustrates both the optical theorem
and the computational prescriptions arrived at before. Note that the factor
~ corresponds √ precisely with the Feynman rule that an external line should
carry a factor ~.
19
You might opt for the ‘absorbing’ source to be J ∗ rather than J. The conclusions are
the same.
August 31, 2019 163
If the diagram contains oriented lines, the time-reversal also inverts the ori-
entation of those lines (if the orientation is indicated by an arrow, we reverse
the arrow). We can write Eq.(6.33) diagrammatically as
000000
111111
111111
000000
000000
111111
111111
000000
000000
111111 111111
000000 X 111111
000000 000000
111111
000000
111111 000000
111111 000000
111111 000000
111111
000000
111111 000000
111111 000000
111111 000000
111111
i 000000
111111
000000
111111
000000
111111 f + i 000000
111111
000000
111111
000000
111111
f + i 000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111 =0 . (6.41)
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
k k 000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
f
000000
111111 k 000000
111111 000000
111111
000000
111111
(6.42)
20
If Eq.(6.41) were not to hold order-by-order, this would imply subtle relations between
coupling constants, ~, and the like. We would then be in a position to actually compute
coupling constants from first principles, which would be good — too good to be true, in
fact, if we believe that we are allowed some freedom in choosing the values of the coupling
constants while retaining a consistent theory.
164 August 31, 2019
+ +
21
The secret resides in the fact that in V the external fields 1,6 and 8 occur precisely
once, and the other fields precisely twice.
22
Note that, for instance, the choice k = {5, 7, 8} would result in the right-hand half of
the diagram being disconnected ; the choice k = {2, 4, 7} is inconsistent since both 6 and
8 are in the final state.
August 31, 2019 165
+ + + = 0 , (6.46)
where we have omitted the line labellings : indeed, the same identity must
hold for the original diagram (6.42). This establishes the so-called cutting
rules (also called the Cutkosky rules), which can be most simply expressed in
words : take a diagram and move the cutting line through it from right to left
in all possible manners, making sure that the two halves in which the diagram
is cut remain connected and that neither the inital state or the final state
is dissected. The particles described by internal lines through which the cut
runs must be assumed to be on their mass shell23 . The sum of all the possible
contributions then vanishes24 . It must be stressed, however, that we have not
proven the Cutkosky rules in this section : rather we have established that a
theory must satisfy them if it is to be a good, probability-conserving theory.
where the oriented lines stand for electrons and positrons, and the wavy
line denotes the photon. Let us consider the 1PI two-loop corrections to the
photon propagator. These are given by
111111111
000000000
000000000
111111111
000000000
111111111
000000000
111111111
000000000
111111111
000000000
111111111
000000000
111111111
000000000
111111111
000000000
111111111
000000000
111111111
000000000
111111111
000000000
111111111
000000000
111111111
000000000
111111111
= + +
000000000
111111111
000000000
111111111
000000000
111111111
23
This may mean that the situation thus described fails to meet the restrictions of
momentum/energy conservation ; then, that contribution vanishes.
24
You might object that in a theory with many different fields the symmetry factors of
the diagrams will, in general, be different from those of a theory with only a single field,
and this is true : however, in the summation over the ‘intermediate states’ k we must of
course also include the ‘indentical-particle’ symmetry factor Fsymm , which precisely repairs
the correspondence — another illustration of the crucial rôle of the symmetry factors !
166 August 31, 2019
By applying the cutting rules we can investigate the real part of this two-loop
contribution:
000000000
111111111
111111111
000000000
000000000
111111111
000000000
111111111
000000000
111111111
∗
000000000
111111111
111111111
000000000
000000000
111111111
000000000
111111111
000000000
111111111
000000000
111111111 000000000
111111111
+
000000000
111111111
000000000
111111111
000000000
111111111
000000000
111111111
000000000
111111111
000000000
111111111
000000000
111111111
000000000
111111111
↔
000000000
111111111
000000000
111111111
000000000
111111111
000000000
111111111
000000000
111111111
000000000
111111111
000000000
111111111
000000000
111111111
000000000
111111111 000000000
111111111
000000000
111111111 000000000
111111111
000000000
111111111 000000000
111111111
↔ + + + +
+ + +
+ + . (6.47)
where integration over the final state is implied. As we shall see, the presence
of a photon in the final state leads to a so-called infrared (IR) divergence
arising from the fact that the probability of emitting an on-shell photon goes
to infinity as the photon energy goes to zero. The process described by the
last two diagrams has therefore an infrared divergence. This divergence is
neatly cancelled by a compensating divergence in the diagrams with a virtual
photon in the first line. This is a well-known fact26 ; but it is instructive to
see that the statement about the cancellation of the infrared divergences can
be replaced by the simpler statement that the photon propagator is free from
infrared divergences27 . This is one example of a useful rule of thumb : when
25
Recall the remark about oriented lines just above Eq.(6.41).
26
And a fortunate one.
27
Two remarks are in order here. In the first place, the virtual-photon diagrams do
contain divergences related to the loop momentum going to infinity : these are ultraviolet
(UV) divergences. ß The photon propagator is therefore still ultraviolet divergent, and
this is cured in the usual manner by renormalization. In the second place, the cancellation
of IR divergences takes place even when we restrict the phase space for the outgoing
particles, provided that zero-energy photons are admitted. This last point is, of course,
not proven by our argument ; but if it were not true you would see really infinite physical
cross sections by making suitable cuts on the final state ! Nature, fortunately, is more
friendly than that.
August 31, 2019 167
you encounter loop diagrams, try to envisage the physics that is described by
cutting them. In fact, the cancellation can be pinpointed further ; the single
statement that the single diagram
1 µ m2 2
L = (∂ ϕE ) (∂µ ϕE ) − ϕE
2 2
1 M 2 2 mλ
+ (∂ µ ϕF ) (∂µ ϕF ) − ϕF − ϕF ϕE 2 . (6.49)
2 2 2
There exists a single coupling between two E’s and one F . Note that the
Feynman rule for the vertex is given28 by −imλ/~ ; we have introduced a
factor m in order to ensure that
P µ = q1 µ + q2 µ , s = P µ Pµ . (6.50)
The phase space (and with it widths and cross sections) is often most easily
evaluated in the rest frame of P µ , in which ~q1 = −~q2 . The phase space
integration element is given by29
1
dV (P ; q1 , q2 ) = d4 q1 δ(q1 2 − m1 2 ) d4 q2 δ(q2 2 − m2 2 ) δ 4 (P − q1 − q2 ) .
(2π)2
(6.51)
As a first step, we cancel d4 q2 against the four-dimensional Dirac delta, and
write the q1 integration in its not-explicitly-covariant form :
1 d3 ~q1 2 2
dV (P ; q1 , q2 ) = δ (P − q 1 ) − m2 . (6.52)
(2π)2 2q1 0
Now, the q1 integration element can be expressed in polar coordinates as
d3 ~q1 |~q1 |2 d|~q1 | dΩ 1
0
= 0
= |~q1 | dq1 0 dΩ , (6.53)
2q1 2q1 2
where we denote the ~q1 solid angle by
dΩ = d cos θ dφ (6.54)
The Dirac delta imposing the mass shell condition on q2 can be written as
√
δ (P − q1 )2 − m2 2 = δ s + m1 2 − m2 2 − 2q1 0 s
s + m1 2 − m2 2
1 0
= √ δ √ − q1 , (6.56)
2 s 2 s
29
It is usual not to include the step functions that require the energies to be positive.
August 31, 2019 169
where the rest frame of P has been used. We immediately find that
s + m1 2 − m2 2 s + m2 2 − m1 2
q1 0 = √ , q2 0 = √ , (6.57)
2 s 2 s
and
1
|~q1 | = |~q2 | = √ λ(s, m1 2 , m2 2 )1/2 , (6.58)
2 s
where the Källén function crops up again. In the P µ rest frame, the phase
space integration element is therefore given by
1/2
m1 2 m2 2
1
dV (P ; q1 , q2 ) = λ 1, , dΩ . (6.59)
32π 2 s s
F (P ) → E(q1 ) E(q2 ) .
q1
P q2
mλ √ 3 √
M = −i ~ = −imλ ~ , (6.60)
~
so that it has dimensionality dim[1/L] as it should. The decay width is
therefore
1 1
dΓ(F → EE) = |M|2 dV (P ; q1 , q2 )
2M r 2!
2 2 2
mλ~ 4m
= 2
1− dΩ . (6.61)
128π M M2
170 August 31, 2019
Note the occurrence of the symmetry factor 1/2 arising from the fact that the
two final-state E particles are indistinguishable. The angular integration is
of course trivial in this simple case, and we immediately find the total width
r
m2 λ2 ~ 4m2
Γ(F → EE) = 1− , (6.62)
32πM M2
with the correct dimensionality dim[Γ] = dim[1/L].
p1 q2 p1 q2
M= q1 + q1 (6.63)
p2 p2
pµ1 = (p, q~ep ) , pµ2 = (q, −q~ep ) , q1µ = (p, q~eq ) , q2µ = (q, −q~eq ) . (6.64)
flattens out. We can also see that in the strictly forward direction, θ = 0, the
amplitude vanishes. The flux factor and the phase space integration element
read
1 1 s − m2
Φσ = , dV (p 1 + p ; q
2 1 2, q ) = d cos(θ) dϕ , (6.66)
2(s − m2 ) 32π 2 s
with ϕ the azimuthal angle, on which M does not depend. The differential
cross section thus becomes
2
~2 λ4 m4
1 − cos(θ)
dσ = d cos(θ) dϕ . (6.67)
64π 2 s (s + m2 ) + (s − m2 ) cos(θ)
The angular integral readily performed, and we find the total cross section
~2 λ4 m4
2s s s
σ(s) = 1− log + 2 . (6.68)
16πs(s − m2 )2 s − m2 m2 m
~2 λ4 m4
σ(m2 ) = , (6.69)
48πm2
which can be considered the cross section for static scattering. Let us consider
this expression more closely. In the first place, the factor λ4 and consequently
the factor ~2 could have been foreseen from the start. The fact that the cross
section must have dim[σ] = dim[L2 ] implies that at the threshold, where
m is the only length scale in the problem, there must also be an overall
factor 1/m2 . Moreover, n body phase space contains a power π 4−3n from its
definition ; and also it contains n − 1 solid angles to be integrated over, each
giving rise to30 a factor π. This means that the total cross section for an E43
n-body final state will contain a factor π 3−2n . In this way, almost the whole
E44
cross section formula is determined, and all the calculational effort is only
used to find the numerical factor 1/48. E45
30
This is to say that the angular integral does not necessarily evaluate to π, but rather
that a factor π invariably arises in the result of a solid-angle integral.
172 August 31, 2019
k p
2
p
L(p ) = (6.70)
p+k
We have dropped the coupling constants, but included the symmetry factor
1/2. We start by combining the two denominators by using the so-called
Feynman trick, which we shall now discuss.
In this integral, we may define σ as the sum of the z’s, and define xj as zj /σ,
as follows:
n Z∞ Y n
!
dzj zj νj −1
Y 1
νj
= exp − Σj zj aj
j=1
a j j=1
Γ(ν j )
0
Y !
zj
dσ δ σ − Σj zj dxj δ xj − . (6.73)
j
σ
August 31, 2019 173
We can now eliminate the z’s in favor of the x’s, and integrate out σ :
n Z∞ Yn
dxj xj νj −1
Y 1
= δ 1 − Σj xj
a νj
j=1 j j=1
Γ(νj )
0
Z∞ P
−1+ j νj
× dσ σ exp − σ Σj xj aj
0
(dxj xj νj −1 ) δ(1 −
P Q P
Γ Z1 xj )
j νj
j
j
= Q Pj νj . (6.74)
j Γ(νj )
0 x 1 a1 + x 2 a2 + · · · + x n an
In our example we have n = 2 and ν1 = ν2 = 1 :
Z1
1 dx
= 2 . (6.75)
a1 a2 xa1 + (1 − x)a2
0
by simply putting a1 = a2 = · · · = an = 1.
where we have switched the notation for the infinitesimal width from to η.
Replacing the loop integral momentum k by k − xp we obtain
Z1 Z
1 −2
L(s) = dx d4 k k 2 + x(1 − x)s − m2 + iη , s = p2 ,
2(2π)4
0
(6.78)
and we see explicitly that the loop integral in this case depends only on s
(and m).
174 August 31, 2019
Dimensional regularization
Z1
µ2
Z
−2
L(s) = dx dD k k 2 + x(1 − x)s − m2 + iη (6.79)
2(2π)4−2
0
Now we let the iη come into its own. The k integral can be written as
dD k = dk 0 dD−1~k, and in terms of k 0 the integrand skirts along a singularity
if (k 0 )2 ∼ ~k 2 − sx(1 − x) + m2 . we see that we can change the integration
contour from s(−∞, +∞) to (−i∞, +i∞) by a counterclockwise rotation.
That is, we may write dk 0 = idk 4 and let k 4 run over the real axis. This
(a) gives an overall factor of +i and (b) changes every occurrence of k 2 into
31
This is mainly due to the fact that the dimensionality of a theory usually has only a
small influence on properties such as causality or current conservation.
32
That is, inverse length. Loosely speaking, an energy scale.
August 31, 2019 175
0
Z∞
= t D/2−1
dy1 · · · dyD y1 −1/2 · · · yD −1/2 δ(y1 + · · · + yD − 1)
0
Γ(1/2)D
= tD/2−1 . (6.82)
Γ(D/2)
√
To get this result we have written k j = yj t and used Euler’s formula,
Eq.(6.76). Our loop integral thus becomes, with D = 4 − 2 :
Z1 Z∞
iµ2 t1−
L(s) = dx dt . (6.83)
2(4π)2− Γ(2 − ) (t + m2 − x(1 − x)s − iη)2
0 0
The t integral
By writing t = a(1 − y)/y and using again Eq.(6.76) we can compute the t
integral :
Z∞ Z1
tα Γ(β − α − 1)Γ(α + 1)
dt = a1+α−β dy y β−α−2 (1−y)α = a1+α−β .
(t + a)β Γ(β)
0 0
(6.84)
176 August 31, 2019
The expansion
Keeping in mind that at some point we want to take → 0 it is sensible to
expand in relevant powers of . First, we use Eq.(15.274),
1
Γ() = (1 − γE + O 2 ) ,
(6.87)
where γE ≈ 0.5772 is Euler’s constant33 , to see that our integral diverges as
−1 whern vanishes. This implies that in all other factors we must keep the
first-order terms but may discard higher orders34 . For the other factors we
may use
A = exp( log(A)) = 1 + log(A) + O 2 ,
(6.88)
and putting everything together we arrive at the final, intermediate result :
i 2
L(s) = R − R(m , s) + O () ,
2(4π)2
1
R = − γE + log(4π) + log(µ2 ) ,
Z1
R(m2 , s) = dx log(m2 − x(1 − x)s − iη) . (6.89)
0
33
Lots and lots of mathematical things are called after Euler. To spread the credit
somewhat, it is also called the Euler-Mascheroni constant. See also appendix 15.15.4
34
Higher-loop integrals typically diverge with higher powers of 1/, and figuring out
what to keep becomes accordingly more complicated.
August 31, 2019 177
So far the computation has been quite straightforward, in that many of its
steps are the same for all one-loop calculations. In particular we can see that
1/ always occurs with −γE + log(4π) + log(µ2 ) in the combination R .
3. s > 4m2 . This is the more tricky case since the argument crosses zero
twice as xpmoves between 0 and 1. The two roots are given as x0,1 ,
with β = 1 − 4m2 /s (but now 0 < β < 1) :
x1 = (1 + β)/2 + iη , x0 = 1 − x1 = (1 − β)/2 − iη . (6.92)
178 August 31, 2019
Z1
R(m2 , s) = log(s) +
dx log(x − x0 ) + log(x − x1 )
0
1
= log(s) + (x − x0 ) log(x − x0 ) + (x − x1 ) log(x − x1 ) − 2x
0
= log(s) + x1 log(x1 ) + log(−x1 ) + x0 log(x0 ) + log(−x0 ) − 2
= log(s) + 2x1 log(x1 ) + 2x0 log(x0 ) + iπ(x0 − x1 )
2 1+β
= log(m ) − 2 + β log − iπ . (6.94)
1−β
The form of R(m2 , s) in the three different regions can be seen as analytic
continuations of each other, but of course precisely how to analytically con-
tinue is the tricky point here ! In figure 6.1 we plot the function R(m2 , s) for
August 31, 2019 179
Mf = (6.95)
The second term is obviously divergent. But we should realize that what we
are trying to do here is to relate the imaginary part of an amplitude to its real
part ; and that real part is indeed divergent, as the R shows. The divergent
part of the v integral is therefore just a reflection of the loop divergence, and
since we have already taken that into account we may write
Z1
2 β2
<R(m , s) = 2 P dv +C , (6.100)
β 2 − v2
0
~λ2
Mee = −i : (6.104)
s − M2
its absolute value squared, summed over all final states (that is, the cross
section without the flux factor) is given by
~2 λ4
Z
X 2 1 β 1
|Mee | = |Mee |2 dV (p; q1 , q2 ) Fsymm = (4π) ,
(s − M 2 )2 (2π)2 8 2
(6.105)
and we find precisely
X
=Mf + |Mee |2 = 0 . (6.106)
6.7 Exercises
Excercise 38 Stability implies safety
Consider a process in which two stable particles collide and produce a number
of final-state particles. One possible diagram is
i~
p2 − m2 + i
and this would blow up if p2 = m2 , making the cross section infinite. Show
that this cannot happen.
182 August 31, 2019
This proves that f (x) approaches the Dirac delta function δ(x) as → 0.
1. By considering |ci = |ai − |bi hb| ai, prove that | ha| bi |2 ≤ 1, and that
| ha| bi | = 1 implies that |bi = eiφ |ai for some complex phase eiφ .
Excercise 41 Find’em !
In section 6.4.2, it is mentioned that in ϕ3 theory the three-point amplitude
contains 58 diagrams at two loops. Find these diagrams, and show that their
symmetry factors add up to 175/8. Hint : also have a look at appendix
15.2.1.
where Ω is the solid angle in the rest frame of P . There is one vertex :
E F mλ
↔ −i
~
E
2. Give the three tree-level diagrams for E(q1 )E(q2 ) → E(p1 )E(p2 ).
2. Conider he same process but now with the additional emission of also
an F of momentum k. Consider the diagram where the F comes from
the external E line, and draw it.
6. Show that the total cross section for this process of F emission must
be divergent.
f (x) = f0 + f1 x + f2 x2 + f3 x3 + · · · , f0 6= 0
and has no singularities in an interval a < 0 < b. Show that computing the
principal-value integral
Zb
f (x)
P dx
x
a
Dirac particles
So far we have been studying particles that can carry only a limited amount of
information : such a particle is completely specified once we have determined
its identity and its momentum. In this chapter we shall start increasing the
number of properties that particles can carry, by examining how the Feynman
propagator can be modified. Since the pole structure of the propagator is
closely related to the causality1 of the theory, and must be used to derive
Newton’s first law in the approximation of propagation over macroscopic
distances, we will not mess with the denominator of the propagator. The
generalisations we shall propose therefore involve the numerator, and are of
the form
1 T
i~ → i~ , (7.1)
p2 − m2 + i p2 − m2 + i
where T is some object that informs us that the particle propagating is not
as simple as we have seen so far, but has additional properties. What those
properties are depends, of course, on the choice of T , and in the following
chapters we shall investigate some of the possible choices.
1
Positive energies moving to the future, and all that.
185
186 August 31, 2019
In such a case, the row ‘vector’ could be assigned to one of the vertices, and
the column ‘vector’ to the other vertex, and the remaining propagator would
again be trivial. We therefore assume that the numerator of the propagator
has a matrix structure that is not a simple dyad, but rather a sum of dyads4 :
X
U (n) W (n) .
T = (7.2)
n
where a, b are some indices living in some linear space. They may be Lorentz
indices5 , but not necessarily. The sum over n must contain at least two terms.
The vertices of the theory must, of course, contain corresponding indices a, b
2
This also holds true also in the case of external lines, if we keep in mind that they are
defined in the square of the matrix element.
3
A number, or a function of p2 .
4
Any matrix r × s can be written as the sum of at most rs dyads.
5
As in the case of spin-1 particles, see later on.
August 31, 2019 187
with which those of the propagator are contracted, otherwise the matrix ele-
ment could not be a scalar. The objects | Ui are ‘columns’ indicated by upper
indices, and the hW| are ‘rows’ indicated by lower indices. It is tempting to
think of the hW| as ‘hermitean conjugates’ of the | Ui but this is not neces-
sarily true6 .
with K some constant (that is, the external-line factors are (multiples of)
the elements of an orthonormal set). This implies that
T2 ∝ T , (7.5)
6
For Dirac particles, the hW| are the Dirac conjugates of the | Ui. More about this
later.
7
Where the U and the W are assigned depends, of course, on which vertex their re-
spective indices are coupled to.
188 August 31, 2019
We have noted that the sum over dyads is actually a sum over distinct
additional properties of the particle that propagates. Now, we should like the
different ‘versions’ of the particle to propagate in the same manner, otherwise
by just sitting and waiting we would see the ‘additional’ properties of the
particle change10 . We shall therefore require that T depends only on the
momentum (and possibly on the mass) of the particle : T = T (p, m). The
particles we have studied in the previous chapter, whose propagator has the
trivial numerator T = 1, of course transform trivially (i.e. not at all) under
rotations : such particles are therefore scalars, or spin-0 particles.
but, as is conventional, we shall not explicitly write out the Dirac indices
unless it is unavoidable. Note also that Eq.(7.8) immediately confirms that
the Dirac objects γ cannot be simple numbers13 . Dirac matrices with different
indices anti-commute, while
(γ 0 )2 = 1 , (γ k )2 = −1 (k = 1, 2, 3) . (7.9)
11
Another argument against the γ’s being simple numbers is that, in that case, they
would define a preferential vector γ µ . This would destroy the assumed isotropy of
Minkowski space, and a frame in which ~γ vanishes would deserve to be equated with
Newton’s absolute reference frame.
12
The anticommutation is necessary because of the symmetry of g µν in its indices.
Another possibility might read something like γ µ γ ν γ α γ β = µναβ 1 but this woud not
allow us to remove fewer than 4 Dirac matrices in any matrix element. The factor 2 in
Eq.(7.8) is simply conventional.
13
Because in that case the fact that g 01 = 0 would imply that γ 0 or γ 1 , or both, vanish:
and that would clash with g 00 = −g 11 = 1.
190 August 31, 2019
γ µ γµ = 4 . (7.10)
(γ 0 )† = γ 0 , (γ k )† = −γ k (k = 1, 2, 3) . (7.11)
For the rest of these notes, the eigenvalues of the Dirac matrices are ac-
tually unimportant. Any choice of Dirac matrices satisfying Eqs.(7.8) is
acceptable ; many have been proposed in the literature15 . That none of
them possesses a physical advantage over the others follows from the ‘funda-
mental theorem of Dirac matrices’ which shows that any two representations
of the Dirac algebra (7.8) can be transformed into each other16 . This again
strengthens our conviction that any result involving Dirac particles should
be derivable without any reference whatsoever to the particular form of the
Dirac matrices, and we shall endeavour to adhere to this. Note that, at this
point, we have not specified the dimensionality of the Dirac matrices. In or-
der to avoid confusion with Lorentz indices, the Dirac indices will be called
spinor indices, and the objects | Ui and hW| for this propagator will be called
spinors. Spinors carry a single spinor index.
Before finishing this section, let us introduce the Feynman slash notation :
if aµ is a Lorentz vector, we shall mean by a / its contraction with Dirac
matrices:
/ ≡ aµ γµ .
a (7.12)
14
At least in precisely four spacetime dimensions. In so-called dimensional regularization
this may change.
15
You may argue that we must show that solutions to Eq.(7.8) actually exist ! That is
correct, and so I am forced to exhibit at least one solution. Here goes, the Pauli-Dirac
representation in block notation :
1 0 0 σk 0 1
γ0 = , γk = , γ5 =
0 -1 −σk 0 1 0
where σk are the three Pauli matrices (γ 5 is defined later on). Can we continue now ?
16
We defer the proof of this theorem to Appendix 15.11.
August 31, 2019 191
a / = 2 (a · b) for all aµ , bν , a
/b/ + b/a /a/ = a2 . (7.13)
γµ a / , γµ a
/ γµ = −2 a / b/ γµ = 4(a · b) , γ µ a
/ b/ c/ γµ = −2 c/ b/ a
/ . (7.14)
γ 5 γ µ = −γ µ γ 5 , (γ 5 )2 = 1 . (7.16)
Γ = S 1 + Vµ γ µ + Tµν σ µν + Aµ γ 5 γ µ + P γ 5 , (7.19)
17
It is customary to leave the unit matrix 1 out of the notation. Its presence can always
be inferred where necessary.
18
Again, in four spacetime dimensions.
19
In some texts the definition of γ 5 is slightly different, for instance it may lack the
factor i, therefore take care when comparing different texts. The reason why it is called
γ 5 and not γ 4 is that in some older treatments the Minkowski indices were assumed to
run from 1 to 4, with the 4th index playing the rôle of our 0th one.
192 August 31, 2019
and these objects form the Clifford algebra. We see that T (p) must be an ele-
ment of the Clifford algebra. The various coefficients are called, respectively,
the scalar (S), vector (Vµ ), tensor (Tµν ), axial-vector (Aµ ) and pseudo-scalar
(P ) coefficients. Since the tensor coefficient may be taken antisymmetric,
there are in total 1+4+6+4+1=16 coefficients. This suggests (but does not
prove) that the Dirac matrices are 4 × 4 matrices. Given an element Γ in
the Clifford algebra, we can recover its coefficients using the trace identities
that we shall discuss below. Finally, we can define the two handy Clifford
elements
1
1 ± γ5 .
ω± = (7.20)
2
They are mutually exclusive complete projection operators, that is
2
ω± = ω± , ω+ ω− = ω− ω+ = 0 , ω+ + ω− = 1 . (7.21)
so that the trace of a single Dirac matrix vanishes ; and by the same method
we see that the trace of a product of an odd number of Dirac matrices is also
zero, in particular
Tr γ 5 γ µ = 0 .
(7.24)
For two Dirac matrices we have
1
Tr (γ µ γ ν ) = Tr (γ µ γ ν + γ ν γ µ ) = N g µν , (7.25)
2
from which we see that the trace of a σ matrix must vanish21 . To continue,
Tr γ 5 = Tr γ 5 γ 0 γ 0 = Tr γ 0 γ 5 γ 0 = −Tr γ 5 γ 0 γ 0 = −Tr γ 5 , (7.26)
20
We shall prove later on that N = 4.
21
Generally, all commutators are always traceless, by the cyclicity property of traces.
August 31, 2019 193
so that also this trace evaluates to zero. The trace of 4 Dirac matrices requires
a bit more work, anticommuting γ m u to the right results in
Tr γ µ γ ν γ α γ β = Tr 2g µν γ α γ β − 2g µα γ ν γ β + 2g µβ γ ν γ α − γ ν γ α γ β γ µ ,
(7.27)
so that, by cyclicity,
Tr γ µ γ ν γ α γ β = N g µν g αβ − g µα g νβ + g µβ g να ;
(7.28)
and the same method may be used to arrive at the 15 terms for a trace
of 6 Dirac matices, the 105 terms for a trace of 8 matrices, and so on22 .
Furthermore, since the anticommutation operations used in Eq.(7.27) might
as well have moved to the left inside the trace instead of to the right, we
immediately find that23
Since γ 5 is the product of all four different Dirac matrices, the product γ 5 γ µ γ ν
(with µ 6= ν) is actually a product of two different Dirac matrices, and
therefore
Tr γ 5 γ µ γ ν = 0 .
(7.30)
Finally, it is immediately24 seen that
Tr γ 5 γ µ γ ν γ α γ β = iN µναβ .
(7.31)
Tr (Γ) = N S , Tr (Γ γ µ ) = N V µ , Tr (Γ σ µν ) = 2N T µν ,
Tr Γ γ 5 γ µ = −N Aµ , Tr Γ γ 5 = N P ,
(7.32)
This shows that we can indeed recover all coefficients from a given Γ25 . It
also leads to the following useful insight : if all the above five traces vanish,
then Γ itself must be identically zero. The above method of computing the
Clifford coefficients from the algebra element is also called Fierzing.
A final, important remark : we have shown that the trace identities,
which have been obtained using only Eq.(7.8), evaluate to expressions con-
taining only the metric and the Levi-Civita symbol, which are Lorentz ten-
sors. Therefore, if we can show that all matrix elements (or, at a pinch, their
absolute squares) can be written as traces, we have realized our goal : the
particular representation of the Dirac matrices is irrelevant, and all possible
E51 choices will lead unambiguously to a unique result.
Below, we shall prove that we may take N = 4 : for this choice, we here
summarize the more important trace identities.
Tr (1) = 4
Tr (γ α1 γ α2 · · · γ α2n+1 ) = 0
Tr γ α γ β = 4 g αβ
Tr γ α γ β γ µ γ ν = 4 g αβ g µν − g αµ g βν + g αν g βµ
Tr γ 5 = 0
Tr γ 5 γ α γ β = 0
Tr γ 5 γ α γ β γ µ γ ν = 4i αβµν
Tr (γ α1 γ α2 · · · γ αn ) = Tr (γ αn · · · γ α2 γ α1 ) (7.34)
Γ = Vµ γ µ + Aµ γ 5 γ µ . (7.37)
Let us define the reverse ΓR as the result of writing all the Dirac matrices
involved in the reverse order26 . By the reflection property of Eq.(7.29), this
means that
Tr ΓR = Tr (Γ) ,
(7.38)
for all elements of the Clifford algebra. In the present case, we have
ΓR = Vµ γ µ − Aµ γ 5 γ µ . (7.39)
Therefore,
ΓR + Γ = 2Vµ γ µ . (7.40)
We immediately arrive at the so-called Chisholm identity :
N
γµ Tr (Γ γ µ ) = Γ + ΓR .
(7.41)
2
In any computation it is always wise to try and get rid of repeated Lorentz
indices, and the Chisholm identity is one of the tools to do so.
γ µ = γ µ , µ = 0, 1, 2, 3 . (7.42)
Γ = Ω Γ† Ω−1 (7.43)
26
Note that, fortunately, (γ 5 )R = γ 5 , so that (γ 5 γ µ )R = −γ 5 γ µ .
196 August 31, 2019
for any Clifford element Γ ; such a form ensures the reasonable property
Γ1 Γ2 = Γ2 Γ1 (7.44)
for two Clifford elements. Double conjugation should be equal to the identity :
†
Γ = Γ = Ω Ω−1 Γ Ω† Ω−1 = B −1 Γ B , B = Ω† Ω−1 .
(7.45)
The element B must therefore commute with all Clifford elements, which
implies that B is a multiple of the unit element27 . Without loss of generality
we may therefore take B = 1, so that Ω is Hermitean. The straightforward
choice is therefore Ω = γ 0 , and the Dirac conjugate is defined as
Γ = γ 0 Γ† γ 0 . (7.46)
For a spinor ξ (which carries an upper spinor index) we have
ξ = ξ† γ0 , (7.47)
which is seen to carry a lower spinor index. A conjugate spinor η, which
carries a lower index, obeys
(η) = η , (7.48)
which has an upper index. A spinor sandwich 28 is an object of the form
ηΓξ ,
and it carries no spinor indices as can be seen ; reasonably, we have
∗
η Γ ξ = ξ Γ η = ηΓξ . (7.49)
Further conjugacy properties follow immediately from Eq.(7.42) :
σ µν = σ µν , γ 5 γ µ = γ 5 γ µ , γ 5 = −γ 5 , ω± = ω∓ . (7.50)
In order for a general Clifford element of the form (7.19) to be self-conjugate,
the coefficients S, V µ , T µν and Aµ must be real, and P imaginary.
The standard Dirac spinors which we shall investigate are defined such
that W = U, although as we have already mentioned this is not an unavoid-
able choice to make. Note that the Dirac choice implies that
T (p) = T (p) . (7.51)
27 5 5
From Bγ = γ B, B must consist of an even number of Dirac matrices. The require-
ment Tr (Bγ α − γ α B)γ β ∼ Tr Bσ αβ = 0 then shows that B can only contain 1 and
γ 5 , and γ 5 does not commute with γ µ .
28
Named after John Montagu, 4th Earl of Sandwich, PC, FRS (13 November 1718 - 30
April 1792).
August 31, 2019 197
Once we realise that the individual terms in this double sum are, in fact,
simple numbers, it is clear that we may also write
X
ηΓξ= ξ b (η)a (Γ)a b = Tr (ξ η Γ) , (7.53)
a,b
On the left-hand side we have two sandwiches, on the right-hand side only
one. This is in particular useful if the dyad ξ1 η̄2 has some simple form.
198 August 31, 2019
F (1, 2, 3, 4) = ξ 1 ω+ γ µ ξ2 ξ 3 ω+ γµ ξ4 , (7.55)
where the ξ’s are arbitrary spinors. Obviously, F (1, 2, 3, 4) = F (3, 4, 1, 2).
Now, as F stands denoted above, it appears to be the (Minkowski) product
of two spinor sandwiches, but we may also (by the ‘mental flip’ mentioned
above) see it as the single sandwich
F (1, 2, 3, 4) = ξ 1 ω+ γ µ ξ2 ξ 3 ω+ γµ ξ4 .
(7.56)
ξ2 ξ 3 = S + Vα γ α + Tαβ σ αβ + Aa γ 5 γ α + P γ 5 . (7.57)
ω+ γ µ ξ2 ξ 3 ω+ γ µ = ω+ γ µ Vα γ α + Aα γ 5 γ α ω+ γµ
= ω+ γ µ Vα γ α + Aα γ 5 γ α γµ = −2 ω+ (Vα γ α − Aα γ α ) , (7.58)
where we have used the fact that ω+ Γω+ = 0 if Γ contains an odd number of
Dirac matrices29 . We can therefore write
u+ = ω+ u+ , u− = ω− u− (7.63)
u− u− = ω− u− u− ω+
= ω− S + Vµ γ µ + Tµν σ µν + Aµ γ 5 γ µ + P γ 5 ω+
= ω− Vµ γ µ + Aµ γ 5 γ µ ω+ = ω− (Vµ − Aµ ) γ µ ,
(7.64)
and similar for u+ . We see that there must exist some four-vector q, called
the spinor’s momentum, such that32
u− u− = ω− q/ . (7.65)
30
This is very suggestive, once we are convinced that the Dirac system describes
fermions. However, the Fierz identity holds only for this particular sandwich, and re-
lies heavily on the presence of the ω± . On the other hand again, it is eminently suited to
resolve a potential problem in the Fermi model of muon decay, that we shall discuss later
on.
31
Considerably, in fact.
32
In this sense, spinors are ‘square roots of vectors’.
200 August 31, 2019
We shall therefore denote the chirality spinors by u− (q) and u+ (q). The
momentum q can be recovered from the spinor components once we know N ,
since
N µ
u± (q) γ µ u± (q) = Tr (u± (q)u± (q)γ µ ) = Tr (ω± q/γ µ ) = q . (7.66)
2
This also tells us immediately that q µ is real-valued and finite, and that
ua± (q)2 (7.67)
X
2q 0 = Tr ω− q/γ 0 = u± (q) γ 0 u± (q) = u†± (q)u± (q) =
so that q 0 is positive 33 .
(7.69)
for n ≥ 1. This allows us to build a generating function, defined34 for suffi-
ciently small z :
X
f (z) = z n Θn = z Θ − z 2 q 2 ω− + 2q 0 z f (z) − z 2 q 2 f (z)
n≥1
z Θ − z 2 q 2 ω−
= . (7.70)
1 − 2q 0 z + q 2 z 2
Now, Tr (Θ) = N q 0 /2 and Tr (ω− ) = N/2 so that
!
X zN −2q 0 + 2zq 2
Tr (f (z)) = Tr z n Θn = − 0z + q2z2
, (7.71)
n≥1
4 1 − 2q
33
Note that in the sum 7.67 we still have not specified how many indices are to be
summed over.
34
This is always possible since both q 0 and q 2 are finite.
August 31, 2019 201
where the product runs over the N eigenvalues λj of Θ. Now, a dyad can
have at most one nonzero eigenvalue35 . The determinant in Eq.(7.74) must
therefore have the form 1 − xλ, where λ is the single nonzero eigenvalue.
Two conclusions fall into our lap : N , the dimension of Dirac space, must
be equal to 4 (and we shall use that everywhere from now on), and q 2 = 0.
We conclude that the momentum of any chirality spinor is finite,
real-valued and massless36 , with positive time component37 .
1 1
k= (q1 + q2 ) , s= (q1 − q2 ) , (7.79)
2w 2w
so that
k·k =1 , s · s = −1 , k·s=0 . (7.80)
All this allows us to write
1
φ = arctanh(sin(θ)) , (7.84)
2
so that
1 sin(θ)
cosh(2φ) = , sinh(2φ) = , (7.85)
| cos(θ)| | cos(θ)|
and finally arrive at
1 + σγ 5 s/
Σ η η̄ Σ = w| cos(θ)| k/ + σ , σ = sign(cos(θ)) . (7.86)
204 August 31, 2019
where we have introduced the Dirac spinor and Dirac antispinor by the
definition of their dyads :
1
p/ + m 1 + γ 5 s/ ,
u(p, s) u(p, s) = U(p, s) =
2
1
p/ − m 1 + γ 5 s/ .
v(p, s) v(p, s) = V(p, s) = (7.88)
2
Note that the overall complex phase of the spinors is not determined40 . It
E52 can easily be checked that the U’s and V’s are mutually exclusive projecting
operators, and span the whole Clifford space :
Noticing that the k/ in Σ can be absorbed into the spinors41 , we finally arrive
at
(cosh(φ) + i sinh(φ) γ 5 ) u(p, s) for σ = +1
η= . (7.90)
5
(cosh(φ) − i sinh(φ) γ ) v(p, −s) for σ = −1
The rewriting that we have done is only sensible if we can actually find
E53 the various vectors and phases ! Fortunately that is possible : from Eq.(7.81)
we see that
η̄ g µ η = 2k µ , η̄ γ 5 γ µ η = −2a5 k µ , (7.101)
p
which gives us k µ and a5 , and also a⊥ = 1 − a25 : we can always take
a⊥ > 0 without loss of generality. To find out about sµ⊥ , we pick any vector
rµ (except r = k) and construct
1 (s⊥ · r) α
yα = η̄ γ 5 r/γ α η = sα⊥ − k . (7.102)
2a⊥ (k · r) (k · r)
y0 µ
sµ⊥ = y µ − k . (7.103)
k0
45
This is a curious-looking expression, that you might not think to be dyad at all at
first sight. The secret is that a25 + a2⊥ = 1.
August 31, 2019 207
η η̄ = w k/ + γ 5 s/ + iλ γ 5 − k/ s/
, λ = sin(θ) = ±1 . (7.104)
We can actually choose the one we like best ! The phases z1,2 of Eq.(7.77)
and the spinor product s+ (q1 , q2 ) have conspired to make this possible.
It has to be noted (see sect.7.4.6) that the forms (7.100) and (7.107) are
also dyads of Dirac spinors, taken in the limit m → 0. The conclusion of
all these exercises is the following : a general spinor can always be brought
in the form of a Dirac spinor, using similarity transformations that depend
only on k. This will help us in choosing a propagator later on.
where sµ and s0µ are spin vectors. If the spinors refer to orthogonal quantum
states, then the absolute square of their product must vanish. We shall now
compute this exactly, by turning the product into a trace using the so-called
Casimir trick, an example of the ‘mental flip’ :
X a b
|u(p, s)u(p, s0 )|2 = u(p, s) a u(p, s0 ) u(p, s0 ) b u(p, s)
a,b
X b a
u(p, s) a u(p, s0 ) u(p, s0 ) b
= u(p, s)
a,b
X b a
= u(p, s)u(p, s) a u(p, s0 )u(p, s0 ) b
a,b
1
Tr (/p + m)(1 + γ 5 s/)(/p + m)(1 + γ 5 s/0 )
=
4
1
Tr (/p + m)2 (1 + γ 5 s/)(1 + γ 5 s/0 )
=
4
m
Tr (/p + m)(1 + γ 5 s/)(1 + γ 5 s/0 )
=
2
m
Tr p/ + m + p/γ 5 s/ + mγ 5 s/ + p/γ 5 s/0 + mγ 5 s/0 + p/γ 5 s/γ 5 s/0 + mγ 5 s/γ 5 s/0
=
2
m m
Tr m + mγ 5 s/γ 5 s/0 = Tr (m − m/ss/0 ) = 2m2 1 − (s · s0 ) (. 7.109)
=
2 2
46
Note that there is a price: the length of the expressions is doubled by the squaring, and
if the amplitude contains many diagrams the algebra can become very cumbersome indeed.
A lot of computational shortcuts have been proposed, the most useful of which appears to
be not to bother with squaring at all but rather to evaluate the spinor products themselves
directly as complex numbers, by so-called spinor techniques. On the other hand, the
existence of the Casimir trick ensures that, as required, one can completely get rid of
the Dirac matrices in the prediction of cross sections using only their anticommutation
properties.
August 31, 2019 209
Note that only two out of the eight terms contain combinations of Dirac
matrices that survive the trace. Since we can work in the pµ rest frame,
where the spin vectors must be spatial unit vectors, we conclude that, in
that frame
|u(p, s)u(p, s0 )|2 = 2m2 1 + ~s · ~s0 .
(7.110)
The states are only strictly orthogonal if ~s0 = −~s. E54
We can use the Casimir trick also to simply prove some properties of
E55
Dirac (anti)spinors :
E56
u(p, s)u(p, −s) = 0 , u(p, s)v(p, s0 ) = 0 ,
u(p, s)u(p, s) = 2m , v(p, s)v(p, s) = −2m , E57
u(p, s)γ µ u(p, s) = 2pµ , v(p, s)γ µ v(p, s) = 2pµ ,
u(p, s)γ 5 γ µ u(p, s) = −2msµ , v(p, s)γ 5 γ µ v(p, s) = 2msµ ,
u(p, s)γ 5 u(p, s) = 0 , v(p, s)γ 5 v(p, s) = 0 ,
u(p, s)γ 5 v(p, s) = 0 , |u(p, s)γ 5 v(p, −s)| = 2m . (7.111)
T (p) = p/ + m , (7.113)
47
Here it is important that the similarity transformations depend only on k, i.e. the
momentum, discussed in sect.7.3.4 Vertices are perfectly allowed to depend on the momen-
tum of a leg (these are called derivative couplings), but we would surely run into trouble
if they would also depend on, say, s.
48
Note that this implies that a funny, non-Dirac propagator can always be mimicked by
a Dirac propagator, where the particle has peculiar interactions !
210 August 31, 2019
and adopt this choice also off the mass shell (where it is actually used). The
Dirac propagator therefore takes the form
p/ + m
i~ .
p2 − m2 + i
The fact that the numerator is linear in p means that the propagator is ori-
ented, in contrast to what we have used so far. To indicate this we define the
orientation with an arrow, and adhere to the convention that the momentum
is counted in the direction of the arrow, irrespective of the sign of the energy
component. The first Dirac Feynman rule therefore becomes
k/ + m
↔ i~
k k · k − m2 + i
Feynman rules, version 7.1
(7.114)
In writing out Feynman diagrams containing Dirac particles, we of course
have to keep track of the Dirac indices resident in propagator and vertices.
This may lead to incredibly cumbersome notation, that may however be
greatly simplified if we adopt the following writing convention : write out
the Dirac-index carrying factors in order, moving against the ori-
entation of the line. Then, all these factors are contracted together using
the usual rules for matrix multiplication, and one hardly ever needs to write
the Dirac indices explicitly. This convention is really to be urged on anyone
contemplating any calculation involving Dirac particles49 !
A final word on notation : since
p/ + m p/ − m = p2 − m2 ,
(7.115)
pa }k 1,2,...
M= A
p B }q 1,2,...
(7.117)
pb
According to the convention described above we then have for the amplitude
i~(/p + m)
M=B A . (7.118)
p2 − m2 + imΓ
Note that, in this amplitude, the factor A must carry the upper Dirac index
of a spinor, and B the lower index of a conjugate spinor. Obviously, pµ
carries positive energy. As we let Γ vanish and pµ approaches the mass shell,
we may then write X
p/ + m = u(p, s) u(p, s) , (7.119)
s
where the sum over s runs over two values, sµ and −sµ . Following the
truncation argument, we readily see that the spinor u(p, s) must then be
included in the decay amplitude, and u(p, s) in the production amplitude.
In the alternative case, where the line is oriented against the flow of
energy, the amplitude is given by
pa }k 1,2,...
M= A
p B }q 1,2,...
(7.120)
pb
and reads (again with our convention !)
i~(−/p + m)
M=A B . (7.121)
p2 − m2 + imΓ
212 August 31, 2019
Note that it is now A that is the conjugate spinor, and B the regular one.
Of course, they describe a physical process different from the first case ! We
are now forced by the negativity of the energy to write
X
−/p + m = − v(p, s) v(p, s) . (7.122)
s
The sign flip in the projecting operator is of course precisely that which
turns a particle description (with negative energy, moving backwards in time
along the orientation of the propagator) into the antiparticle description, with
positive energy. The truncation argument then tells us that v(p, s) must be
the factor associated with the production, and v(p, s) must be associated with
the annihilation, of the antiparticle. There remains the question of where to
put the left-over Fermi minus sign. Consistently, we may decide to keep it
with the v, in which case we arrive at the following Dirac Feynman rules :
k/ + m
↔ i~
k k · k − m2 + i
√
↔ ~ u(p, s)
p,s
√
↔ ~ u(p, s)
p,s
√
↔ ~ v(p, s)
p,s
√
↔ − ~ v(p, s)
p,s
Feynman rules, version 7.2
(7.123)
The awkward-looking minus sign is usually subjected to the argument that
any matrix element containing an incoming antiparticle will have the factor
−v in each of its diagrams, and since we are interested in absolute values
squared anyway, there would appear to be little harm in deleting this overall
minus sign from the Feynman rules : and this is what is commonly done. A
little reflexion, though, will remind us that the sign of the amplitude’s real
part is fixed by unitarity, and now we have changed it ! Clearly, the minus
sign will be back to haunt us later on50 .
50
The minus sign becomes expecially urgent when we use amplitudes to describe inter-
action potentials, see section 10.5.
August 31, 2019 213
Λ(p; q) : p/ → Σ1 p/ Σ2 . (7.124)
Since we must ensure that Dirac conjugation commutes with Lorentz trans-
formation, we must have Σ2 = Σ1 ; and in order to have matrix multiplication
commute with Lorentz transformations as well52 we must have Σ2 Σ1 = 1. We
conlude that
Λ(p; q) : p/ → Σ p/ Σ , Σ Σ = 1 . (7.125)
The explicit form of Σ, the minimal Lorentz transformation reads53
p2
q/p/
Σ=C 1+ 2 , |C|2 = . (7.126)
p (p + q)2
2(pq) p2 q 2
2 q/p/ + p/q/ q/p/p/q/ 2
Σ Σ = |C| 1 + + 4 = |C| 1 + 2 + 4 =1 ,
p2 p p p
(7.127)
51
Not to be confused with an infinitesimal Lorentz transformation. A product of two
minimal transformations is, in general, not itself minimal.
52
So that we can either first mutiply /p1 and /p2 , and then Lorentz-transform them, or
do the Lorentz transform first and the multiplication afterwards.
53
This form tacitly assumes that under minimal Lorentz transforms the sign of p2 and
(p + q)2 are the same. If p2 < 0 this is not always true ; however, for boosts and spatial
rotations it does hold.
214 August 31, 2019
and
2 q/p/p/ + p/p/q/ q/p/p/p/q/ 2 q/p/q/
Σ p/Σ = |C| p/ + + = |C| p/ + 2/q + 2 = q/ ,
p2 p4 p
(7.128)
where we have used the anticommutation result q/p/q/ = 2(pq)/q − p/q 2 . The
other requirements, Σ/qΣ = p/ and Σ/rΣ = r/, are proven trivially. For general
Clifford elements Γ, we have now also ensured that
Γ → ΣΓΣ (7.129)
ξ → Σξ , ξ → ξΣ . (7.130)
Tx = β/yz/ , Ty = β/zx
/ , Tz = β/xy/ , (7.133)
54
Here the confusing active-passive distinction rears its ugly head. We shall not worry
about it since the rotation algebra is the same in each case.
August 31, 2019 215
where we have used cyclicity, but not specified the constant β. This constant
can be determined from the rotation group algebra requirement:
~2
Tz 2 = β 2 x / y/ = −β 2 x2 y 2 =
/ y/x = Tx 2 = Ty 2 , (7.136)
4
we conclude that the total-spin operator comes to
2 2 2 1 1
Tx + Ty + Tz = + 1 ~2 . (7.137)
2 2
As m → 0, the spin vector diverges, and the massless limit is not so obvious.
We may, however, write for this case
1 0 p0 − p 1
sµk = 1, −~e = pµ + O m/p0 ,
p , p ~e + (7.139)
m m m
since (p0 −p)/m = m/(p0 +p). The projecting operator can then be evaluated
by
1
1 + γ 5 s/k (/p + m)
u(p, s)u(p, s) =
2
1 1 5 0
= 1 + γ p/ + O m/p (/p + m)
2 m
1
= (1 + γ 5 )(/p + m) + O (m)
2
≈ ω+ p/ , (7.140)
But states without pure helicity are also possible : we can consider the case
where p~ and ~s make a fixed angle θ. In that case the spin vector reads
m cos θ sµk + p0 sin θ sµ⊥
µ
s = p , sµ⊥ = (0, ~e⊥ ) , (7.143)
(p0 )2 − p2 cos2 θ
with ~e · ~e⊥ = 0. If we now let m → 0 so that p → p0 , while keeping φ fixed,
then the limit of the projecting operator becomes
1
u(p, s)u(p, s) ≈ (1 + γ 5 s/⊥ )/p , (7.144)
2
and we see that this limit is indistinguishable from a masless, transversely
polarised Dirac particle. We may also decide to let θ also approach zero with
θ/m fixed, and then we will end up with
1
1 + PL γ 5 + P⊥ γ 5 s/⊥ p/ ,
u(p, s)u(p, s) ≈ (7.145)
2
August 31, 2019 217
with PL2 + P⊥2 = 1 : precisely the special case discussed in sect.7.3.6 ! The
message is that the massless limit is always defined, but must be taken with
some care57 .
Here, the vectors with and without hats are related as follows :
q q q
+ + = 0 (7.150)
p p p
where Γ1,2 represent the rest of the diagrams. The momenta p and q are
assumed to run from left to right. We have not indicated the spins since
anyway we have to sum over them. Therefore we would have to evaluate the
trace
Tr (/p + mp )Γ1 (/q − mq )Γ2 ,
where we have indicated that the two Dirac particles are not necessarily of
the same type. Let us now shift our attention to the first diagram, say. A
closed loop of Dirac particles is automatically also a trace: this diagram,
then, requires the analogous trace
Tr (/p + mp )Γ1 (−/q + mq )Γ2 ,
61
The requirement that amplitudes do not contain uncontracted indices essentially forces
us to use Feynman rules in which the orientation of Dirac lines is conserved at every vertex.
For so-called Majorana fermions this is not true : Majorana fermions, therefore, have no
distinction between particle and antiparticle.
August 31, 2019 219
k/ + m
↔ i~
k k · k − m2 + i
√
↔ ~ u(p, s)
p,s
√
↔ ~ u(p, s)
p,s
√
↔ ~ v(p, s)
p,s
√
↔ ~ v(p, s)
p,s
-1 for every closed Dirac loop
Feynman rules, version 7.3
(7.151)
where the lines without arrows have no Dirac but scalar propagators. The
cut crosses two Dirac lines, and we might conclude that a minus sign is called
62
We disregard the denominators of the Dirac propagators since they do not influence
our argument.
220 August 31, 2019
where now the cut crosses one Dirac line and one line without an arrow.
Since the propagator in that line is even in its momentum, we can always
choose the loop momentum to run in the ‘correct’ direction for the Dirac
line, and no minus sign is needed. Therefore, the first diagram also takes no
extra minus sign, since crossing symmetry forbids an amplitude to suddenly
pick up an extra minus sign under crossing. It is only when a closed loop
consists of only Dirac particles that no crossing can be found for which the
loop momentum can be chosen to run in the ‘correct’ direction. Therefore,
only for such loops is a minus sign unavoidable63 .
where the first diagram contains a fermion loop and hence carries an overall
minus sign ; the second one does not. Now consider the cut versions of these
diagrams :
and notice that the left-hand sides of the cut-through diagrams are identi-
cal. The right-hand sides differ in the way that the in-going fermions are
connected to the out-going ones ; the ingoing ones are interchanged in the
second diagram with respect to the first one. This, then, must correspond to
63
This holds true later on, where we also introduce vector particles, the propagator of
which is also even in the momentum.
August 31, 2019 221
k/ + m
↔ i~
k k · k − m2 + i
√
↔ ~ u(p, s)
p,s
√
↔ ~ u(p, s)
p,s
√
↔ ~ v(p, s)
p,s
√
↔ ~ v(p, s)
p,s
-1 for every closed Dirac loop
-1 for every Dirac particle interchange
Feynman rules, version 7.4
(7.152)
Note that the interchange rule only determines the relative sign between two
Feynman diagrams. How the interchange sign can be determined is best
illustrated by an example. Consider, for instance, a process with 6 external
fermions. Three of them must then be oriented outward from the diagram,
carrying a u or v, and the other three must be oriented inward and carry a u
or a v. Let us assume that there are three Feynman diagrams, schematically
given by64
diagram 1 : u1 Γ1 u2 v 3 Γ2 u4 u5 Γ3 v6
diagram 2 : u1 Γ4 u2 v 3 Γ5 v6 u5 Γ6 u4
diagram 3 : u1 Γ7 u4 v 3 Γ8 v6 u5 Γ9 u2 ,
Clearly, we have left out an enormous amount of detail here, and the Γ’s
can be anything. Note that we have written the three diagrams in such a
way that the conjugate spinors u1 , v 3 and u5 are in the same order in each
diagram : this is always possible. Now, we see that to go from diagram 1 to
diagram 2, the positions of u4 and v6 must be interhanged, whereas one can
64
The process e− e− e+ → e− e− e+ is an example.
222 August 31, 2019
x
x = . (7.155)
With the field function of the Dirac field denoted by ψ(x), this SDe reads
Z
1 k/ + m
ψ(x) = 4
d4 y d4 k e−ik·(x−y) 2 J(y) , (7.156)
(2π) k − m2
where matrix multiplication is implied as usual. We can now apply a well-
chosen differential operator :
Z
1 k/ + m
(i/∂ − m) ψ(x) = 4
d4 y d4 k e−ik·(x−y) (/k − m) 2 J(y)
(2π) k − m2
of matter. Its only connection to the rest of the universe would arise through some
non-diagrammatic process, involving possibly gravity since that appears not to be very
amenable to diagrammatics. Of course, classical quantum mechanics finds that the com-
bined wave function for identical-state electrons vanishes identically, but again quantum
and gravity do not see completely eye to eye.
224 August 31, 2019
Z
1
= 4
d4 y d4 k e−ik·(x−y) J(y)
(2π)
Z
= d4 y δ 4 (x − y) J(y) = J(x) , (7.157)
x
x = , (7.159)
which is written as
Z
1 ¯ k/ + m −ik·(y−x)
ψ̄(x) = d4 y d4 k J(y) e . (7.160)
(2π)4 k 2 − m2
By the same simple manipulation as above, we can then show that the con-
jugate Dirac equation reads
←
ψ̄(x)(−i ∂ ¯
/ −m) = J(x) , (7.161)
where the leftward arrow indicates that the derivative must be taken towards
the left 67 .
7.7 Exercises
Excercise 48 Lorentz-contracted gamma matrices
Prove the results given in Eq.(7.14).
Excercise 49 Five spacetime dimensions
Let γ 4 be defined as γ 4 = iγ 5 . Show that this γ 4 has just the right properties
so that the relations
γ µ γ ν + γ ν γ µ = 2 g µν , γ µ = γ µ
are valid in a five-dimensional spacetime (µ, ν = 0, 1, 2, 3, 4) with metric
γµν = diag(+1, −1, −1, −1, −1).
Excercise 50 Cyclic sigmas
Prove, by explicit calculation, that
γ 5 γ 0 γ j = σ kn , γ 5 γ j γ k = σ n0
where (j, k, n) is a cyclic permutation of (1, 2, 3).
Excercise 51 Another relation between σ’s
Use trace identities and Fierzing to prove that
i
γ 5 σ αβ = − αβµν σµν
2
226 August 31, 2019
p µ = xµ , q µ = cos(θ) xµ + sin(θ) y µ
Compute Σ(p → q) for θ = π/3, and from there, by taking powers, for
θ = 2π/3 and finally for θ = 2π.
means that we have not only E particles but also Ē particles. There is one
vertex :
E
i
E F ↔ ~λ
1. There are now 3 diagrammatic SDe’s for this model. Write them out.
2. Assume M > 2m. Compute the decay width Γ(F → EE) at the tree
level.
(a) Write the tree diagrams for the process E(p1 ) Ē(p2 ) → E(q1 ) Ē(q2 )
by F exchange. We have indicated the momenta of the E parti-
cles.
(b) Compute h|M|2 i for this process, where we sum over the final-
state spins and average over the initial-state spins of the E’s.
(c) Compute the total cross section σ(E Ē → E Ē), as a function of
the total invariant mass-squared s ≡ (p1 + p2 )2 . This is best done
in the centre-of-mass frame.
(a) Write down the 2 diagrams for E(p1 ) F (k1 ) → E(p2 ) F (k2 ) at
the tree level.
(b) Work out h|M|2 i for this process.
(c) Compute the total cross section as a function of s = (p1 + k1 )2 .
This is best done in the centre-of-mass frame.
(d) Compute the total cross section in the limit s → m2 .
6. Make no assumption on m or M .
228 August 31, 2019
E F F
D
where both vertices have Feynman rule iλ/~ as before. We shall assume
mE > mD .
2. Compute the cross section σ(F F → E Ē) at the tree level, as a function
of the total invariant mass-squared s.
Chapter 8
1
Or at least faster !
2
I think that this really is because ‘chirality’ pronounces less fluenty than ‘helicity’.
3
At the LHC, electrons are almost perfectly massless.
229
230 August 31, 2019
factor)
u± (p) ∼ p/ ξ∓ , ξ∓ = ω∓ ξ∓ , (8.1)
where the helicity spinors ξ∓ can be chosen almost4 arbitrarily. Note that
this is a phase convention ! Then, the spinor product can be written in the
form
s+ (p, q) = u+ (p) u− (q) ∼ ξ¯− p/q/ ξ+ . (8.2)
You might try to find a spinor product that is symmetric in p ↔ q, but that
is useless since it would give
1 1
s+ (p, q) = (s+ (p, q) + s+ (q, p)) ∼ ξ¯− (/pq/ + q/p/) ξ+ = (p · q) ξ¯− ξ+ , (8.3)
2 2
and such a spinor product is simply the vector product. The alternative is
to choose the spinor product antisymmetric, so that s+ (p, q) + s+ (q, p) = 0,
and we see that this requires ξ¯− ξ+ = 0 : the helicity spinors ξ± must be
orthogonal. Then we shall have
and easily verify that u− (k0 )u+ (k0 ) = 0. Reversal ‘flips’ the helicity :
(u− (k0 )u− (k0 ))R = (ω− k/0 )R = k/0 ω− = ω+ k/0 = u+ (k0 )u+ (k0 ) . (8.8)
Using the basis spinor, we now define all other massless spinors by
1
u± (p) = √ p/ u∓ (k0 ) . (8.9)
2p · k0
This is by construction a valid form since
1 1
u± (p)u± (p) = p/ ω∓ k/0 p/ = ω± p/k/0 p/ = ω± p/ . (8.10)
2p · k0 2p · k0
This choice is at the basis of the so-called spinor techniques, or helicity meth-
ods. we shall use Eq.(8.9) extensively to good effect.
u+ (p1 )/p2 p/3 u− (p4 ) + u+ (p1 )/p3 p/2 u− (p4 ) = 2(p2 · p3 ) u+ (p1 )u− (p4 ) . (8.15)
232 August 31, 2019
s+ (p1 , p2 )s− (p2 , p3 )s+ (p3 , p4 ) + s+ (p1 , p3 )s− (p3 , p2 )s+ (p2 , p4 )
= s+ (p2 , p3 )s− (p3 , p2 )s+ (p1 , p4 ) . (8.16)
Using the antisymmetry property of s, and dividing out the factor s− (p2 , p3 ),
we obtain the so-called Schouten identity :
s+ (p1 , p2 )s+ (p3 , p4 ) + s+ (p1 , p3 )s+ (p4 , p2 ) + s+ (p1 , p4 )s+ (p2 , p3 ) = 0 . (8.17)
E61 Note the cyclicity in p2,3,4 . Obviously, the identity holds for s− as well. At
the end of this chapter we give a summary table of all these spinor techniques.
You can easily check the spinor products’ properties by explicit inspection.
From |sλ (p, q)|2 = 2(pq) we can see that spinor products are square roots
of vector products. Actually, in practical applications it pays to compute
the spinor products of massless momenta in your problem anyway. As an
example, if two massless momenta pµ and q µ make a small angle θ, their
vector product reads (pq) = pp q 0 (1 − cos θ) ∼ θ2 . For their spinor product,
on the other hand, we have |s+ (p, q)| ∼ θ so the numerical cancellations are
much less severe in te spinor product than in the vector one. The vector
product can then be obtained as a simple square.
µ
We may consider an explicit choice for the vectors k0,1 :
which is very useful for actual numerical applications. Note that this choice E62
presupposes that none of the light-like vectors in the problem is oriented
E63
exactly along the x-axis. Since the ‘special’ direction in many problems is
traditionally chosen to be the z-axis, this is usually safe. E64
E65
8.1.5 The standard form for massive particles
The standard form for Dirac spinors given in Eq.(8.9) can be generalised to
the case of massive particles. Let pµ be the momentum of such a particle,
and let m be its mass. We then simply define
1
u± (p) = √ (/p + m)u∓ (k0 ) ,
2p · k0
1
v± (p) = √ (/p − m)u∓ (k0 ) . (8.21)
2p · k0
We can find out the spin vector for these two cases : writing u± (p) =
u(p, ±s0 ) we obtain
1
s0 µ = − u+ (p) γ 5 γ µ u+ (p)
2m
1
Tr ω− k/0 (/p + m)γ 5 γ µ (/p + m)
= −
4m p · k0
1 µ m
= p − k0 µ , (8.22)
m (pk0 )
which is indeed the only vector built from p and k0 that can have the right
properties s0 2 = −1 and (ps0 ) = 0. Note that for small(ish) m and generally
positioned k0 , ~s0 points in the general direction of p~. Therefore we call u+ (p)
a right-handed spinor, and u− (p) a left-handed spinor. In addition, from the
fact that, for the antispinor v± (p),
1
v + (p) γ 5 γ µ v+ (p) = −s0 µ , (8.23)
2m
234 August 31, 2019
7
Remember the ‘not too small’ insect.8.1.2 ? In a frame where ~k0 and p~ are parallel the
dot product (k0 p) is of order O m2 and the ‘subleading’ term with k0 in Eq.(8.22) is no
longer subleading. Whereas normally the spinor u+ (p) looks a lot like a positive-chirality
spinor u+ for small m, it then looks more like u− : a continuous source of worry especially
in large computer programs that deal with many fermions numerically.
8
In cases like this, you will usually hear your lecturer going ‘mumble mumble analytic
continuation mumble mumble’ but that is unhelpful unless you know how to analytically
continue. This section, in fact, explains how.
9
I have never encountered this need. Of course, the complex conjugate amplitude occurs
in the cross section, but that can be accounted for by replacing every u by w and vice
versa, if you really insist on computing the complex conjugate amplitude separately (why
would you ?).
August 31, 2019 235
We have now found the dyadic forms we wanted. In what way do they differ
from the Dirac case ? The spinor products
and, in fact, it is easily checked that all the other relations such as the reversal
formula, the Chisholm identity, and the Schouten identity still hold.
• Dirac equations :
• projecting operator :
uλ (p) uλ (p) = ωλ p/ (8.29)
• Spinor products :
uλ (p) uλ (q) = 0
sλ (p, q) = uλ (p) u−λ (q) = −sλ (q, p) = −s−λ (p, q)∗
|sλ (p, q)|2 = sλ (p, q) s−λ (q, p) = 2(p · q) (8.30)
• Chisholm identity :
• Reversal :
uλ1 (p) Γ uλ2 (q) = λ1 λ2 u−λ2 (q) ΓR u−λ1 (p) (8.33)
• Schouten identity :
p k1
M= q (8.35)
k2
2
= 64 G2F ~2 (q · k1 ) (p · k2 ) . (8.40)
Here, the sum is over the electron and muon spins. It is practical to evaluate
this in the muon rest frame. We shall write E1,2 for k1,2 0 in this frame. Then
(p · k2 ) is equal to mµ E2 , and by momentum conservation we find
1 1
(q + k1 )2 − me 2 = (P − k2 )2 − me 2 = mµ (K − E2 ) ,
(q · k1 ) =
2 2
(8.41)
12
Unless lepton flavour number is invoked.
13
In the standard form of spinors, the helicity for antispinors is reversed. The antineu-
trino therefore actually has positive chirality.
238 August 31, 2019
where
mµ 2 − me 2
K= . (8.42)
2mµ
The transition rate then takes the form
|M|2 = 64 GF 2 ~2 mµ 2 E2 (K − E2 ) ,
(8.43)
and for the partial decay width we find
dΓ(µ → eνµ ν e ) = 32 GF 2 ~2 mµ E2 (K − E2 ) dV (p; q, k1 , k2 ) . (8.44)
where θ is the angle between the neutrino momenta. Hence we can integrate
trivially over the other polar, and the two azimuthal, angles (leading to a
factor 8π 2 ), and the integral over θ is resolved by the delta function. The
result is
π2
dV (p; q, k1 , k2 ) = dE1 dE2 . (8.47)
(2π)5
In terms of these variables, the phase space is perfectly flat14 . Since |cos θ|
cannot exceed unity, we also have the restrictions
mµ 2 − me 2 − 2mµ E1 − 2mµ E2 ≤ 0 ,
14
This flatness does not depend on the masslessness of the neutrinos. For massive
neutrinos the same phase space density is found, only the boundaries of the phase space
become (horribly) complicated.
August 31, 2019 239
GF 2 ~2 mµ 5
Γ(µ → eνµ ν e ) = F (me 2 /mµ 2 ) ,
192 π 3
F (x) = 1 − 8x + 8x3 − x4 − 12x2 log(x) . (8.51)
The function F (x), shown in figure 8.1, is strictly decreasing since with
increasing me /mµ the available phase space shrinks. For the realistic values
240 August 31, 2019
15
Since the spin sum of uu contains /p.
August 31, 2019 241
∼ (q · k1 ) (k2 · p) − mµ (k2 · s)
∼ k20 (mµ − 2k20 ) (1 + cos(θ) cos(θ2 ) + sin(θ) sin(θ2 ) cos(φ2 )) ,
(8.53)
where θ is the angle between ~q and ~s, and θ2 , φ2 are the polar and azimuthal
angles of ~k2 with respect to ~q. As we have seen above, the phase space
integration element is, up to overall factors,
mµ (q 0 + k20 − mµ /2)
cos(θ2 ) = 1 − . (8.54)
q 0 k20
mµ (q 0 + k20 − mµ /2)
0 0
dΓ ∝ k2 (mµ −2k2 ) 1 + cos(θ) 1 − dq 0 dk20 d cos(θ) .
q 0 k20
(8.55)
The integral over k20 from mµ /2 − q 0 to mµ /2 is straightforward, leading to
the following, properly normalised distribution :
1 d2 Γ 2q 0
= y 2 3 − 2y + cos(θ)(1 − 2y) ,
y= . (8.56)
Γ dy d cos(θ) mµ
1 dΓ 1
= 3 − cos(θ) , (8.57)
Γ d cos(θ) 6
1 dΓ
= 2y 2 (3 − 2y) . (8.58)
Γ dy
Let us now consider the parity properties of the decay. Under the parity
transform, Eq.(7.149) tells us that ~q → −~q and ~s → ~s, so that effectively
cos(θ) → − cos(θ). We see that the muon decay amplitude, which is not
242 August 31, 2019
to simplify
8.4 Exercises
Excercise 61 Helicity operator for massless spinors
Show that, for a general timelike or lightlike momentum pµ , the operator
1 5
Σp = cos(θ/2) + i sin(θ/2) γ [/p, γ 0 ]
2|~p|
244 August 31, 2019
where (as usual for massless fermions) we have disregarded the distinction
between spinors and antispinors. The quantity Qe is a fixed coupling con-
stant. The helicities λ1,2 and ρ1,2 are explicitly indicated. There are, in total,
16 helicity combinations.
1. Using spinor techniques, compute M for the sixteen helicity cases.
2. Assume that we are in the centre-of-mass frame, where p~1 + p~2 = 0.
Let θ be the angle between p~1 and ~q1 , and E = p1 0 . Give the simplest
form you can find for h|M|2 i.
3. Show that h|M|2 i is really divergent for θ = 0, and that this has nothing
to do with neglecting the electron mass.
Vector particles
247
248 August 31, 2019
The first Feynman rule for these particles, that we call vector particles since
they carry a Lorentz index, is therefore established :
which means that, for the objects U, W three mutually orthogonal choices
can be made, for instance U (1) = x, U (2) = y, and U (3) = z. Of course,
complex linear combinations of these are also possible : in general, we can
say that there can be found three polarisation vectors µλ , with λ = −1, 0, 1,
such that
1
µ µν
X ν
(λ ) ( )µ = −δ
λ0 λ,λ0 , T (p) = (λ )µ (λ ) . (9.6)
λ=−1
E69 If pµ is not in the rest frame, we simply boost both p and the various to
2
That is, its spacetime part is unoriented. There may of course be other properties
such as charge that do impose a distinction between production and decay of the particle.
August 31, 2019 249
the actual Lorentz frame. We can now go once more through the truncation
argument of chapter 6, with the obvious result that the polarisation vectors
are to be assigned to the external lines, and we immediately arrive at the full
set of Feynman rules for massive vector particles :
000000
111111
111111
000000
000000
111111
000000
111111
000000
111111
000000
111111
√ outgoing lines
000000
111111
000000
111111
000000
111111
000000
111111
000000
111111
↔ ~ λ µ
000000
111111
with Σ as defined in section 7.4.5 ; since this must hold for arbitrary a, we
have the relation
Λ(p; q)µ ν γµ = Σ γν Σ , (9.10)
By multiplying with γ α on both sides and taking the trace, we immediately
E70 find the form of Λ(p; q) in Minkowski space :
1 1
Λ(p; q)α ν = Tr Λ(p; q)µ ν γµ γ α = Tr Σ γν Σ γ α
4 4
2
p q/p/ p/q/
= Tr 1 + 2 γν 1 + 2 γ α
4(p + q)2 p p
2 2
= δαν − 2
(p + q)α (p + q)ν + 2 q α pν . (9.11)
(p + q) p
so that the generators of the rotation group must in this case have the form
we conclude that the spin is indeed unity. The total spin operator contains,
as it must, the projection of all vectors on the spatial subspace. In words:
to be a good polarisation vector, µ must be properly normalised and satisfy
the Lorenz condition4 :
· ∗ = −1 , ·p=0 . (9.18)
µ i
↔ − J µ (x) (9.23)
~
The SDe for a free vector particle’s field function V µ is then again very
simple : 000000
111111 111111
000000
000000
111111
000000
111111
000000
111111 x
x 000000
111111
000000
111111
000000
111111
000000
111111
= , (9.24)
000000
111111
000000
111111
000000
111111
e−ik·(x−y)
Z
µ 1 4 4 µν 1
V (x) = dydk 2 −g + 2 kµkν Jν (y) . (9.25)
(2π)4 k − m2 m
The remaining integrals over y and k now lead immediately to the so-called
Proca equation for V µ :
∂ · ∂ V µ − ∂ µ ∂ · V + m2 V µ = J . (9.28)
F µν = ∂ µ V ν − ∂ ν V µ . (9.30)
spinors. Let the momentum of the vector particle be pµ and its mass m. We
can choose two massless momenta k1µ and k2µ that sum to pµ :
with ~q. Similarly, ~0 is only parallel to p~ if ~k1,2 are antiparallel, therefore in
general 0 is not a purely longitudinal state. Nevertheless, the representations
(9.32) form a perfectly acceptable orthonormal set of polarisations, and in
µ
particular (as we shall see) the freedom in choosing k1,2 can often be used to
good effect.
In a frame in which ~k1,2 are antiparallel we can find the helicities as
follows. Using the result of exercise 72, we can write the rotation over an
angle θ around the common direction as
uλ (k1 ) γ µ uλ (k1 ) →
µν 1 − cos θ µ ν µ ν
sin θ µν
cos θg + k1 k2 + k2 k1 − (k1 , k2 ) uλ (k1 ) γν uλ (k2 )
k1 · k2 k1 · k2
(9.37)
and since we can write
i
µν (k1 , k2 )uλ (k1 ) γν uλ (k2 ) = − Tr γ 5 γ µ γ ν k/1 k/2 uλ (k1 ) γν uλ (k2 )
4
i
= − Tr γ γ u−λ (k1 )u−λ (k2 ) k/1 k/2 = −i(k1 k2 ) u−λ (k2 )γ 5 γ µ u−λ (k1 )
5 µ
2
= −iλ (k1 k2 ) uλ (k1 )γ µ uλ (k2 ) , (9.38)
this rotation results in just a phase factor exp(iλθ). E74
Before finishing this section we point out that also the (trivial) external-
E75
line Feynman factor for scalar particles can be written in terms of spinors.
µ
For a massive scalar with momentum q µ , the same choice of k1,2 is of course
possible. We simply note that √ we can always find a complex phase eiϕ such
that the external-line factor h can be cast in a form containing two spinors :
√ √ eiϕ
~ → ~ √ s+ (k1 , k2 ) . (9.39)
2 p1 · p2
It should not come as a surprise that an external integer-spin particle can
conventiently be represented by a spinor-antispinor pair. After all, this is
precisely the way in which particles like the W and Z (and, to a lesser extent,
the Higgs as well) are most often seen in experiment : namely, through their
decay into a fermion-antifermion pair.
7
Traditionally, the spin-statistics theorem, like the CPT theorem, is considered to be
very deep and difficult. Make up your mind.
8
This can also be proven for particles of higher spin. There is a single exception, the
so-called Kalb-Ramond state. See Appendix 15.12.
August 31, 2019 257
0 = 0 , ~k · ~ = 0 . (9.40)
However, a problem immediately arises: for the above equations are not
invariant under Lorentz boosts. If we boost k µ and µ to a generically other
frame, they no longer hold. Let us assume that we are in such a frame ; there
we have the Lorentz-invariant conditions
Inserting this into the last equation of Eq.(9.41), we find immediately that
0 = |~k |, and the second equation then gives |~⊥ | = 1. We see that, whatever
the value of µ , we can always write
0
µ µ
= ⊥ + kµ , (9.43)
k0
where ⊥ µ does satisfy Eq.(9.40). We can therefore have a consistent and uni-
tary theory of massless vector particles, provided that the k µ term decouples
from the physics. Now, any matrix element involving an external massless
vector particle with momentum k µ and polarisation vector µ will be of the
form
M = J (k)µ µ , (9.44)
where J µ (k) stands for the rest of the amplitude. Note that J µ does not
carry any information about µ , but it does know what k µ is, by momentum
conservation. Our requirement then is that the interactions of the theory be
such that
J µ (k) kµ = 0 . (9.45)
9
The current limit is about 10−36 GeV/c2 .
258 August 31, 2019
That is, if we replace the polarisation vector by the momentum, the ampli-
tude must vanish.
We shall use the convention that the momentum under the handlebar is
counted outgoing. The requirement for strictly massless external vector par-
ticles then becomes 0000000
1111111
1111111
0000000
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
=0 . (9.47)
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
A similar message can be gotten from the propagator. After all, the
massive-vector propagator
−g µν + k µ k ν /m2
i~
k 2 − m2
clearly becomes horribly singular at m = 0. The solution, as before, is to
require that in our theory the k µ k ν term should drop out. There is a catch,
10
The term ‘handlebar’ is my own, biker’s privilege. You might use the term ‘taking
the divergence’ but that obscures the very mechanical way by which we investigate the
suitability of amplitudes.
11
One may for instance have the source J represent a charge whose momentum changes,
thereby emitting radiation.
12
Whether they exist is another question ; at any rate we cannot produce them, nor
observe them.
August 31, 2019 259
however: whereas external vector particles must be on the mass shell, the
momentum of internal lines is off the mass shell. We therefore arrive at the
sharper requirement that Eq.(9.47) must hold even if the particle is off-shell.
Using the fact that the spinors of massless particles are homogeneous of
degree 1/2 in the argument :
√
u± (αp2 ) = α u± (p2 ) , (9.51)
we see that (for instance) the polarisation vector + can be written, in analogy
to Eq.(9.32), as
1
+ µ = √ u+ (p1 ) γ µ u+ (p2 ) . (9.52)
2 p1 · p2
Since α does not occur in the polarisation vector, we may consider the limit
α → 0. In that case, q = p1 is massless, and the only condition on the
massless vector p2 is that (p1 · p2 ) must not vanish. By a judicious choice of
overall complex phase, this leads us to propose, for a massless vector particle
with momentum k µ , states of definite helicity as follows, where the spinors
are again in the standard form :
uλ (k)γ µ uλ (r)
λ µ = √ , λ=± . (9.53)
λ 2 s−λ (k, r)
and √
k/ /λ = λ 2 u−λ (k) uλ (k) , (9.57)
and this object is explicitly gauge-invariant.
August 31, 2019 261
µν µ ν µ ν
µ k ν ↔ i~ −g + (k r + r k )/(k · r) massless internal lines
k 2 + i
Note the appearance of the arbitrary vector r. This way of writing the
propagator is called the axial gauge. The propagator is explicitly orthogonal
to rµ whatever the value of k : The vector r acts as an ‘axis’ with respect
to which the field is always orthogonal, hence the name. The fact that
the vector r is arbitrary is of course bothersome, in the same way that the
arbitrariness of the representation chosen for the Dirac matrices in the case
of Dirac particles is bothersome. We solve it in the same way, by insisting
that we ought to be able to remove r from the final expressions for matrix
elements. This can of course not be by virtue of any property of r itself, but
must come from the handlebar condition, since every term containing r also
contains k. Two things are worthy of remark here. In the first place, the
propagator is homogeneous of degree zero in r, so any result cannot depend
on the length of r anyway. In the second place, in contrast to the propagator
proposed before, with pµ pν /m2 , the propagator in the axial gauge does not
diverge.
262 August 31, 2019
λ µ (k, r1 ) − λ µ (k, r2 ) =
λ uλ (k)γ µ uλ (r1 ) uλ (k)γ µ uλ (r2 )
√ −
2 s−λ (k, r1 ) s−λ (k, r2 )
λ u−λ (r1 )γ u−λ (k)u−λ (k)uλ (r2 ) + u−λ (r1 )uλ (k)uλ (k)γ µ uλ (r2 )
µ
= −√
2 s−λ (r1 , k)s−λ (k, r2 )
µ µ
λ u−λ (r1 )(γ k/ + k/ γ )uλ (r2 ) √ s−λ (r1 , r2 )
= √ =λ 2 k µ . (9.60)
2 s−λ (k, r1 ) s−λ (k, r2 ) s−λ (k, r1 ) s−λ (k, r2 )
we see that the two states differ only by the vector particle’s momentum. In
any current-conserving set of diagrams we may therefore choose the gauge
vector at will ; there is no risk of picking up a phase difference if two different
gauge vectors are used for two different current-conserving sets of diagrams.
As an illustration of how the gauge vector can disappear from a current-
conserving object, let us consider
p q
λ · − ,
2k · p 2k · q
with p and q two massless momenta. The form of section 9.3.4 turns this
into
λ sλ (k, p)s−λ (p, r) sλ (k, q)s−λ (q, r)
√ −
2 s−λ (k, r) 2k · p 2k · q
λ s−λ (p, r) s−λ (q, r)
= √ −
2 s−λ (k, r) s−λ (p, k) s−λ (q, k)
λ s−λ (p, r)s−λ (q, k) − s−λ (q, r)s−λ (p, k)
= √ (9.61)
2 s−λ (k, r)s−λ (p, k)s−λ (q, k)
Now, the Schouten identity tells us that
s−λ (p, r)s−λ (q, k) + s−λ (p, k)s−λ (r, q) = −s−λ (p, q)s−λ (k, r) (9.62)
August 31, 2019 263
9.4 Exercises
Excercise 69 Polarisation averages
If, for vector particles, we take not a spin sum but a spin average, we find
3
pµ p ν
1X µ ν 1 µν
j ¯j = −g + 2
3 j=1 3 m
δ µ ν = tµ tν − xµ xν − y µ yν − z µ zν
sin(α/2) 5 0
Σ = cos(α/2) + i γ [γ , p/]
2|~p|
and
Σu+ (k2 ) = exp(+iα/2) u+ (k2 )
4. Show that, therefore, the object u+ (k1 )γ µ u+ (k2 ) acquires a phase exp(iα)
under this rotation.
1 1
µ± = √ u± (k1 ) γ µ u± (k2 ) , µ0 = (k1µ − k2µ )
m 2 m
August 31, 2019 265
2. Use this to show that the vectors pµ and ±,0 form an orthogonal set
Quantum Electrodynamics
10.1 Introduction
In this chapter we shall start to work our way to realistic theories about the
actual elementary particles encountered in nature1 . Almost all elementary
particles seen so far have nonzero spin2 . We shall defer the discussion of
charged spin-1 particles to a later chapter ; at this point we shall only dis-
cuss how to set up a consistent theory of spin-1/2 particles (charged leptons
and/or quarks) and photons. This is the theory of quantum electro-dynamics,
or QED.
267
268 August 31, 2019
Dirac index3 : and since the photon is involved, it must also carry a Lorentz
index. The simplest, and – as we shall see – indeed the correct form of the
vertex is that of a Dirac matrix. We therefore propose the following Feynman
rule :
Q QED vertex
µ ↔ i γµ
~
Feynman rules, version 10.1 (10.1)
q p2
Of course the photon has to be off-shell here, but that is no problem since also
off-shell photons must obey current conservation. The part of the amplitude
3
Since we cannot form constant (i.e. momentum-independent) objects with one Dirac
index, or with, say, two upper Dirac indices, we are essentially forced to have conservation
of the orientation of Dirac lines in each vertex. This is modified to some extent if Majorana
particles are present in the theory.
4
Or, rather, it is related to the charge. The precise form of this relation must, of course,
be established by investigating the coupling in a well-defined physical situation.
August 31, 2019 269
q p2
With the convention, to which we shall try to adhere, that the momentum
assigned in the handlebar must be counted outgoing from the vertex, so in
this case should read −q, the handlebarred M becomes
where m is the mass of the fermion. Now, we know that the spinors u(p1 )
and v(p2 ) satisfy the Dirac equations
for on-shell momenta, so that half of the expression 10.6 ‘cancels to the left’
and the other half ‘cancels to the right’, and we end up with
Mc = 0 . (10.8)
We shall see that this is the general mechanism by which unitarity and cur-
rent conservation are ensured.
process at once, but already we can learn a few useful things. In the first
place, a possible alternative coupling, with γ 5 γ µ instead of γ µ , is ruled out
since we cannot obtain two Dirac equations :
so that either the cancellation to the left would be spoiled, or that to the
right. In the second place, it is necessary that both fermions have pre-
cisely the same mass. Since all known different fermion types have different
masses, this means that the QED interaction must conserve fermion type, or
‘flavour’. Electromagnetic muon decay, µ → eγ, is therefore forbidden, not
by conservation of the electric charge (which is indeed the same for muons
E78 and electrons) but by conservation of the whole electromagnetic current6 .
where the fermion momenta p and q are indicated and for the photon mo-
mentum k we have k µ = (p − q)µ . The momenta p and q may be on-shell
(in which case the corresponding blob isn’t actually there), but any of them
may be off-shell, and hence leads into a further piece of Feynman diagram. In
that case the blobs stand for the other vertices, where the fermion is created
and absorbed7 . The part of the diagram between the blobs is
Qγ µ
q/ + m p/ + m
i~ 2 i i~ 2 ,
q − m2 ~ p − m2
6
Fortunately, the decay µ → eγ has never been observed, and the branching ratio is
smaller than about 10−11 .
7
Actually, the p and q lines are attached to a semi-connected graph rather than two
separate connected ones, but here the distinction is irrelevant.
August 31, 2019 271
p q
and see what happens. Algebraically, we must multiply the above expression
by kµ , and then
iQ 2 q/ + m µ p/ + m
(i~) γ 2 kµ =
~ q 2 − m2 p − m2
iQ q/ + m p/ + m
= (i~)2 2 2
(/p − q/) 2
~ q −m p − m2
iQ 2 q /+m p/ + m
= (i~) 2 (/
p − m) − (/
q − m)
~ q − m2 p2 − m 2
iQ q/ + m p/ + m
= (i~)2 2 2
− 2 . (10.10)
~ q −m p − m2
We see that under the handlebar the double propagator splits up into two
single ones. Note that, for this to be possible, it is again essential that
the mass of the fermion does not change at the vertex. We may write this
operation diagrammatically as
p q = − , (10.11)
↔ i~ , (10.12)
272 August 31, 2019
where the external line may belong to the initial or final state, and the arrow
orientation may be also reversed. An important result follows immediately
from the triviality of our new Feynman-rule tools :
= . (10.15)
≡ (10.16)
k
The fermions’ momenta are reckoned along their respective arrows, and the
photon momentum k is counted outgoing. Note that, since we are considering
August 31, 2019 273
here a Green’s function and not an amplitude, any of the external lines may
be off-shell. It is this object that we shall submit to a handlebar operation.
The SDe’s of QED are, in our notation :
= + ,
= + ,
= + . (10.17)
We therefore have
= , (10.18)
= − . (10.19)
Each term on the right-hand side can be subjected to its own SDe, to give
qj
X pj
X
= −
j j
By virtue of Eq.(10.15) the last two terms cancel precisely, and we are left
with the Ward-Takahashi identity :
qj
X pj
X
= − . (10.21)
j j
274 August 31, 2019
We can conveniently express this in a more analytic form. Let us denote our
Green’s function by
Gµ (p1 , . . . , pr ; q1 , . . . , qs ; k) ≡ , (10.22)
where the explicit Lorentz index is that of the photon line, and the same
Green’s function, only with the special photon removed, by
G0 (p1 , . . . , pr ; q1 , . . . , qs ) ≡ . (10.23)
Taking into account the flow of the momenta and the fact that the two trivial
Feynman rules (10.12) and (10.13) together yield a factor of (iQ/~)(i~) =
−Q, we can write the diagrammatic Ward-Takahashi identity (10.21) as fol-
lows8 :
Gµ (p1 , . . . , pr ; q1 , . . . , qs ; k) kµ =
X s
Q G0 (p1 , . . . , pr ; q1 , . . . , qj + k, . . . , qs )
j=1
Xr
−Q G0 (p1 , . . . , pj − k, . . . , pr ; q1 , . . . , qs ) . (10.24)
j=1
It is this result that proves that, indeed, the choice of the vertex (10.2.1)
leads to an acceptable theory.
So far we have considered the general case, with no constraint on the
external momenta. If we now specialize to amplitudes, in which all external
momenta except perhaps for k µ , must be on their mass shell, the rule (10.14)
applies, and we find the even more attractive Ward identity :
= 0 . (10.25)
8
In many, or even most, cases of interest fermion number is conserved, which means that
in every (fundamental or effective) vertex the number of incoming and outgoing fermions
is the same ; in that case we have r = s. But since we have nowhere used this, the Ward-
Takahashi identity may be expected to hold also for processes in which fermion number
is not conserved, for example in supersymmetry where Majorana fermions occur. Note,
however, that Majorana fermions are necessarily neutral and themselves do not couple to
photons.
August 31, 2019 275
in other words
Z Z
1 −ik·(x−y) k/ + m Q
ψ(x) = d4 y 4
dke i~ 2 i γµ ψ(y)Aµ (y) ,
(2π)4 k − m2 + i ~
(10.27)
whence
i/∂ − m + QA(x)
/ ψ(x) = 0 , (10.28)
which is the Dirac equation in the presence of an electromagnetic field. Let us
work this expression towards classical physics. In the first place, the deriva-
tive is, by the standard assignment rules for quantum mechanics, related to
the momentum operator :
pµ = i~ ∂ µ , (10.29)
and the mass m to the mechanical mass M by (as we have seen)
Mc
m= . (10.30)
~
The Dirac equation can therefore be written as
dim[E] = dim[q] /m2 . Since this is the gradient of the classical e.m. vector
potential Ac we have dim[Ac ] = dim[q] /m, and because the photon field
A has dimensionality dim[A2 ] = kg/sec, it follows that the correct relation
between the photon field and the classical e.m. field must read
Ac 2 = c A2 . (10.33)
From this it follows that the coupling Q and the charge q are related by
√
Q = −q/(~ c) , (10.34)
√
which implies the correct dimensionality dim[Q] = dim[1/ ~] ; moreover,
we find immediately that, for particles with unit electric charge,
4π
Q2 = α , (10.35)
~
where α stands for the electromagnetic fine structure constant :
α ≈ 1 / 137.036 . (10.36)
Since in QED every next loop order contains two extra powers of Q and
one (effective) power of ~, the loop expansion is in QED equivalent to an
expansion in powers of α.
What we have√done here is a pure √ dimensional analysis, it does not decide
between Q = q/~ c and Q = 2q/~ c. Later on, the discussion of Thomson
scattering will assure us that we have made the right choice. Of course, any
extra numerical factor in Q can be compensated for by a rescaling of A.
Since
q/γ µ = q µ − iqα σ αµ , γ µ p/ = pµ + ipα σ αµ , (10.39)
the current takes the form
ie
Jµ = u(q) (p + q)µ + i(p − q)α σ αµ u(p) .
(10.40)
2m~
This is called the Gordon decomposition : the vertex is split up into a piece
that we shall later recognize as the vertex by which a scalar particle interacts
with photons, which is called the convection term, and a tensorial part, called
the spin term. Both terms vanish individually under the handlebar operation.
The Gordon decomposition is especially useful in nonrelativistic situations,
where q µ and pµ are both dominated by their time components. In that
case u(q)u(p) ≈ 2m, and the vertex becomes simply (p + q)µ . Note that for
antispinors we find exactly the same result, namely v(p)γ µ v(q) ≈ (p + q)µ in
the nonrelativistic situation9 .
Here, we have indicated the Lorentz indices on the photon lines, and the
momenta across the photon lines are considered incoming into the loop. In
addition to this diagram, there is also a similar diagram in which the orien-
tation of the loop is reversed :
k ν
p2
D+ = µ p3
p1
λ
9
q v(q) = −mv(q), say, which gives us a minus sign :
Since the Dirac equation reads /
on the other hand, v(p)v(q) now evaluates to approximately −2m, thereby cancelling the
minus sign.
278 August 31, 2019
Note that these graphs cannot be twisted into one another. For loops with
only one or two vertices they can be so twisted, and then do not count as
separate diagrams ; for three or more vertices, there are two distinct ones.
Without pretending to evaluate the whole loop, let us concentrate on the
Dirac structure of their numerators. The first diagram contains the trace10
= − T− , (10.44)
since no terms with an odd power of m survives the trace. We see that
the two loops cancel each other precisely ! This can obviously be extended
to loops with more vertices, and we find Furry’s theorem : fermion loops
with an odd number of vector vertices11 and opposite orientation
cancel each other ; with an even number of vector vertices, they
are identical12 . Furry’s theorem does not hold if one or more of the ver-
tices are of axial-vector type, and so it is not generally valid for the weak
interactions. For QCD, in which the quark-gluon couplings have the Dirac-
matrix form as in QED, Furry’s theorem holds in a more restricted form :
the spacetime part of the two quark loops with even(odd) number of vertices
are equal(opposite), but the additional colour structures of the diagrams are
different. This implies, for instance, that the two quark loops with three
gluon vertices do not cancel completely. We shall come back to that case
later on.
10
By the rules of Dirac particles, closed loops automatically evaluate to traces.
11
That is, vertices consisting of a single Dirac matrix, such as in QED.
12
Furry’s theorem is usually proved by invoking the charge-conjugation matrix, discussed
in section 15.11.2. However, as we see this is not necessary.
August 31, 2019 279
Now consider a diagram (or set of diagrams) N with only one single
external line which is a photon :
p
N ≡ , (10.45)
N µ = pµ f (p2 ) . (10.46)
M = (10.47)
q2
p
2
Both the electron and muon are Dirac particles. We shall denote the electron
charge by Qe , and the muon charge by Qµ , and their masses by me and mµ ,
respectively. The total invariant mass squared is conventionally denoted by
s, and of course momentum is conserved :
~Qe Qµ
M=i v(p2 ) γ α u(p1 ) u(q1 ) γα v(q2 ) , (10.49)
s
and is strictly dimensionless: dim[M] = dim[1], as it ought to be for a
2 → 2 process at tree order.
~2 Qe 2 Qµ 2
Tr (/p2 − me )γ α (/p1 + me )γ β
= 2
4s
Tr (/q1 + mµ )γα (/q2 − mµ )γβ
4~2 Qe 2 Qµ 2
p2 α p1 β + p1 α p2 β − (p1 · p2 )g αβ − me 2 g αβ
=
s2
q1 α q2 β + q2 α q2 β − (q1 · q2 )gαβ − mµ 2 gαβ
4~2 Qe 2 Qµ 2
= 2(p1 · q1 )(p2 · q2 ) + 2(p1 · q2 )(p2 · q1 )
s2
2 2
+ s(me + mµ ) (10.50)
4π α2 mµ 4
σ≈ 1 − 6 2 + ··· . (10.59)
3s s
p p
q k2
M = + (10.60)
k1 k2 k1 q
The amplitude is given by
M = M1 + M2 ,
A1
M1 = −i~Qe 2 , A1 = u(q) /2 (/p + k/1 + m) /1 u(p) ,
2(p · k1 )
A2
M2 = +i~Qe 2 , A2 = u(q) /1 (/q − k/1 + m) /2 u(p) ,
2(q · k1 )
(10.61)
where 1,2 are the polarisation vectors of the respective photons. Taking into
account the averaging factor 1/4, we find17 (with m for me )
1
|A1 |2 Tr (/q + m) γ α (/p + k/1 + m) γ β (/p + m) γβ (/p + k/1 + m) γα
=
4
= 16m4 − 8(pq)m2 + 8(pk1 )(qk1 ) + 16(pk1 )m2 − 8(qk1 )m2 ,
1
|A2 |2 = Tr (/q + m) γ β (/q − k/1 + m) γ α (/p + m) γα (/q − k/1 + m) γβ
4
= 16m4 − 8(pq)m2 + 8(pk1 )(qk1 ) + 8(pk1 )m2 − 16(qk1 )m2 ,
hA1 A∗2 i = hA2 A∗1 i
1
Tr (/q + m) γ α (/p + k/1 + m) γ β (/p + m) γα (/q − k/1 + m) γβ
=
4
= 8(pq)(pk1 ) − 8(pq)(qk1 ) + 16(pq)m2 − 8(pq)2
−4(pk1 )m2 + 4(qk1 )m2 . (10.62)
1 1 K
dV (p + k1 ; q, k2 ) = dΩ , (10.65)
(2π)2 8 s
where Ω is the solid angle of the emitted electron. The flux factor is
1 1
2 1/2
= . (10.66)
2λ(s, m , 0) 2K
K(s + m2 )
Z
1
dΩ (qk1 ) = ,
4π 4s
Z
1 1 2s K
dΩ = log 1 + 2 ,
4π (qk1 ) K2 m
Z
1 1 4s
dΩ = . (10.68)
4π (qk1 )2 m2 K 2
We therefore have for the transition rate, now also averaged over the scat-
tering angle :
m2 m2 s
2 2 2
|M| = 16π α 1 + + 16 2
s K
2 4
ms ms s K
+ −8 2 − 16 3 + 2 log 1 + 2 (10.69)
K K K m
286 August 31, 2019
|M|2
σ= . (10.70)
16πs
It is interesting19 to note that the ‘static’ limit K → 0 is well-defined :
8π α2
lim σ = . (10.71)
K→0 3 m2
This is called the Thomson cross section. It may serve as the ‘measurement’
prediction by which the electric charge of the electron is defined.
Note that, just as in the case of muon pair production, the cross section
does not depend on ~. This means that this cross section had better coincide
with the prediction from classical electromagnetism — as, indeed, it does.
Returning to the arguments that led us to make the identifications
q
Acl 2 = c A2 , Q= √ , (10.72)
~ c
r+ = p2 , r− = p1 (10.77)
M(ν, λ, λ) = 0 (10.78)
which puts paid to 4 out of 8 helicity configurations20 . Of course the fact that E79
these amplitudes vanish cannot depend on our choice of r± , but the proof
that they vanish is now very simple ! The next case is that of λ1 = −λ2 = +.
We see that A1 still vanishes, and we can compute A2 as follows (using
(p2 k2 ) = (p1 k1 )) :
√ √
1 2 − 2
A2 =
2(p1 k1 ) s− (k1 , p2 ) s+ (k2 , p1 )
20
For ν = 0 it suffices to interchange s+ and s− , and include a minus sign for each
polarization vector.
288 August 31, 2019
× u+ (p1 )u− (k1 )u− (p2 ) (/p2 − k/2 ) u− (p1 )u− (k2 )u+ (p2 )
u+ (p1 )u− (k1 )u− (p2 ) k/2 u− (p1 )u− (k2 )u+ (p2 )
=
(p1 k1 ) s− (k1 , p2 )s− (k2 , p1 )
s+ (p1 , k1 )s− (p2 , k2 )s+ (k2 , p1 )s− (k2 , p2 )
=
(p1 k1 ) s− (k1 , p2 )s− (k2 , p1 )
s− (k2 , p2 )2
= −2 (10.79)
s− (k1 , p1 )s− (k1 , p2 )
the singularity is not due to our neglecting the electron mass ; indeed, for
nonzero mass we have
t = (p1 − q1 )2 = 2m2 − 2(p1 0 )2 + 2|~p1 |2 cos θ
= −2|~p1 |2 (1 − cos θ) . (10.88)
E81 To this order in perturbation theory, the total cross section for Bhabha
scattering is therefore indeed divergent.
E82
Note the Fermi minus sign between M1,2 and M3,4 . The four pairs of dia-
grams are separately current-conserving, i.e.
Mi c→k = 0 , i = 1, 2, 3, 4 . (10.92)
q
M0 = A (10.93)
where A denotes the rest of the diagram(s). The corresponding radiative pro-
cess will (amongst others) contain diagrams in which the photon is emitted
by this particular fermion :
q
Ms = A (10.95)
k
292 August 31, 2019
which evaluates to
√ q/ + k/ + m
Ms ≡ −(Qe ~) u(q) / A(q + k) . (10.96)
2q · k
Notice that the denominator q · k goes to zero as the photon energy vanishes,
and hence the diagram diverges in the limit of soft Bremsstrahlung . In
the soft-photon approximation ( and assuming that the object A does not
depend on q in too drastic a manner23 ) we have
√ q/ + m
Ms ≈ −(Qe ~) u(q) / A(q) . (10.97)
2q · k
Anticommuting / and q/, and using the property of the Dirac spinor, which
tells us that u(q)/q = mu(q), we then find
√ q·
Ms ≈ −(Qe ~) u(q)A(q) , (10.98)
q·k
that is, the diagram factorizes into the nonradiative result and an ‘infrared
factor’24 . We can repeat this procedure for those diagrams in which the
photon is emitted by the other external particles. There are, of course, also
(possibly) diagrams in which the photon is emitted from internal lines ; but,
as can easily be checked, such diagrams do not diverge as k 0 → 0. In the
soft-photon approximation, they do therefore not contribute. For radiative
Mœller scattering, we therefore have the nicely factorized form
√
M = −(Qe ~) (VIR · ) M0 ,
µ q1µ q2µ pµ1 pµ1
VIR = + − − , (10.99)
q1 · k q2 · k p1 · k p1 · k
where M0 is the amplitude for the nonradiative process ; and, using the
polarisation sum rule Σµ ν = −g µν , we find
s4 + t4 + u4
|M|2 = −2Qe 6 ~3
(VIR · VIR ) ,
t2 u2
m2e m2e m2e m2e
−VIR · VIR = − − − −
(p1 k)2 (p2 k)2 (q1 k)2 (q2 k)2
23
This assumption fails, for instance, close to a resonance. However, since every reso-
nance has a finite width, the soft-photon approximation is formally correct for infinitesimal
photon energies.
24
Since infrared light has low energy compared to visible light.
August 31, 2019 293
As has already been intimated, the double poles are indeed suppressed by a
factor me 2 .
with µ1,2 , ν1,2 , λ = ±. The amplitude is then a function of the helicities, and
we write M(µ1 , µ2 ; ν1 , ν2 ; λ). We first consider M1 (+, +; +, +; +). Using
Eq.(9.56) this can be written as
√ √
(Qe ~)3 2
M1 (+, +; +, +; +) = i
2(p2 · q2 )s− (k, r)
q/1 + k/ α /1 − k/
αp
× u+ (q1 ) u− (k)u− (r) γ −γ u− (k)u− (r) u+ (p1 )
2k · q1 2k · p1
× u+ (q2 )γα u+ (p2 ) , (10.101)
M1 (+, +; +, +; +) =
√ √ s+ (q1 , k)u− (p1 )(/q1 + k/ )u− (q2 )s− (p2 , p1 )
i(Qe ~)3 8 . (10.102)
(2p2 · q2 )(2k · q1 )s− (k, p1 )
so that
√ √ s− (p1 , p2 )2
M(+, +; +, +; +) = i(Qe ~)3 8 . (10.104)
s− (p2 , q2 )s− (k, p1 )s− (k, q1 )
Finally, we can make use of the identity of Eq.(9.63) to arrive at the form
√ s− (p1 , p2 )2
3 + · p1 + · q1
M1 (+, +; +, +; +) = −2i(Qe ~) − .
k · p1 k · q1
s− (p1 , q1 )s− (p2 , q2 )
(10.105)
The infrared factor also appears in this case ! Performing the appropriate
subtitutions we can write the complete amplitude as
√
M(+, +; +, +; +) = −2i(Qe ~)3 s− (p1 , p2 )2 (VIR · + )
1 1
× − . (10.106)
s− (p1 , q1 )s− (p2 , q2 ) s− (p1 , q2 )s− (p2 , q1 )
The minus sign in the last term is the Fermi sign ; it helps us to simplify our
expression even further using the Schouten identity, and the final form for
the amplitude is
M(µ1 , µ2 ; ν1 , ν2 ; λ) =
(VIR · λ ) σ s−λ (a, b)3 s−λ (c, d)
2i~3/2 Q3 , (10.108)
s−λ (p1 , q1 )s−λ (p2 , q2 )s−λ (p1 , q2 )s−λ (p2 , q1 )
where the sign σ and the identities of the four vectors a, b, c and d are given
in the following table :
August 31, 2019 295
ν1 µ1 ν2 µ2 λ σ a b c d
+ + + + + + p1 p2 q1 q2
+ + + + – + q1 q2 p1 p2
+ + – – + + q2 p1 q1 p2
+ + – – – – q1 p2 q2 p1
+ – – + + – q2 p2 q1 p1
+ – – + – + q1 p1 q2 p2
– – – – + – q1 q2 p1 p2
– – – – – – p1 p2 q1 q2
– – + + + + q1 p2 q2 p1
– – + + – – q2 p1 q1 p2
– + + – + – q1 p1 q2 p2
– + + – – + q2 p2 q1 p1
A few things may be noted, apart from the surprising simplicity of the am-
plitudes. In each case, only s+ or s− occur. Moreover, when the electrons
have equal spins the amplitudes are explicitly antisymmetric in p1 ↔ p2 or
q1 ↔ q2 , as required by Fermi statistics25 . The spin-averaged matrix element
squared therefore has the following form in the strictly massless case :
|M|2 me =0 = −2Qe 6 ~3 (VIR · VIR )
where, as before, A stands for the rest of the diagram(s). We shall not
assume the soft-photon limit, however. Let us assume that the photon is
emitted as small angle θ with respect to the fermion momentum. We then
find, assuming the fermion energy to be large compared to its mass m :
(k · q) = k 0 q 0 − |~q| cos θ
2 !
1 me
≈ k 0 (q 0 − |~q|) + |~q|θ2 /2 ≈ k 0 q 0 θ2 +
(10.111)
,
2 q0
where we have used the fact that q 0 − |~q| = me 2 /(q 0 + |~q|) ≈ me 2 /(2q 0 ).
we conclude that as soon as θ is of order me /q 0 or smaller26 , the product
(k · q) becomes of order m2e ; and this means that in that case the ‘single
pole’ (k · q)−1 and the ‘double pole’ me 2 (k · q)−2 are of the same order27 . The
squared matrix element (summed over fermion and photon spins) contains
of course
Qe 2 ~
|Mc |2 = −
4(k · q)2
× A(q + k)(/q + k/ + m)γ α (/q + m)γα (/q + k/ + m)A(q + k)(10.112)
The second term in this expression enters into the ‘massless’ result since it
will give rise only to single-pole terms, whereas the first term tells us that
the double-pole term coming from this Mc must read
2
2 m2
|Mc | = −Qe ~ A(q + k)(/q + k/ )A(q + k) , (10.114)
(k · q)2
gives us the double-pole terms : keeping all four collinear situations in sight,
we can write the transition rate including the double-pole terms as
2 6 2
|M| = Qe ~
1 1 2
= 0 0 ∼ , (10.116)
p·k k (p − |~p| cos θ) k 0 p0 (m/p0 )2 + θ2
where we have assumed that m/p0 and θ are small. This has been the reason
for statements such as ‘the radiation is essentially confined to a cone of angle
m/p0 around the fermion direction’, the so-called radiation cone. This is E84
quite wrong since the polar angular integration variable is d cos θ ∼ d(θ2 )/2
and not dθ.
Suppose that the photon is strictly parallel to the fermion, ~k//~p. Since
the photon polarisation is strictly orthogonal, ~ · ~k = 0, we must also have
~ · p~ = 0, and the soft-photon approximation 10.99 then tells us that the
diagram with denominator (p · k) vanishes in the soft-photon limit : the
298 August 31, 2019
angular distribution shows a sharp dip known as Yennie’s crater. For photons E85
that are not infinitesimally soft, the crater is less deep but still there.
where the charge flow is indicated by the arrow. The photon index is µ. The
momenta p and q are counted along the arrow. Note that the propagator
of scalar particles may be unoriented, but the vertices do not have to be so
if there is a quantum number, such as charge, that distinguishes between
particle and antiparticle. Since no Dirac matrices occur the only quantities
in this vertex that carry a Lorentz index are the momenta p and q (and
of course the photon’s own momentum, but that is fixed by p and q). We
therefore propose a Feynman rule of the form
q Q
µ ↔ i (c1 pµ + c2 q µ ) ,
p ~
with constants c1,2 to be determined. This is simple, since we can study the
annihilation of the charged scalar-antiscalar pair into an off-shell photon :
under the handlebar operation, the amplitude becomes
p1 √
= iQ ~ (c2 p2 µ − c1 p1 µ ) kµ
p2 k √
= iQ ~ (c2 p2 µ − c1 p1 µ ) p1 µ + p2 µ
√ (c2 − c1 )
= iQ ~(p1 + p2 )2 . (10.117)
2
28
Or ultra-collinear.
29
Elementary charged scalar particles have to date not been observed, although they
are predicted in extensions of the standard model. We include them here since they will
provide indications on how to treat charged vector particles.
August 31, 2019 299
We see that c1 = c2 is required, and therefore the first Feynman rule for
scalar electrodynamics (sQED) reads
Let us now consider the more complicated process of annihilation into two
on-shell photons. With the above vertex two diagrams are involved :
p1 k1 p1 k2
M = k2 + k1 (10.119)
p2 p2
The solution is to introduce a four-point vertex into the Feynman rules. For
reasons lost in the mists of time, such a vertex is called a sea-gull vertex30 .
The improved Feynman rules are now :
30
To me it does not look very gully nor even particularly birdy.
300 August 31, 2019
µ Q2 µν sQED 4-vertex
↔ 2i g
ν ~
sQED Feynman rules, version 10.2 (10.122)
+ + = 0 . (10.123)
You might worry that annihiliation into three photons will necessitate a five-
point vertex, and so on. Fortunately, the above two vertices are sufficient to
guarantee current conservation in all sQED processes, as we shall now show
using some more handlebar diagrammatics.
As in our proof for regular QED, none of these lines is necessarily on-shell. .
Momentum conservation again fixes the photon momentum to be k = p − q.
In analogy to regular QED we can now invent some handlebar diagrammatics
as follows :
(p − q) · (p + q)
= −iQ~
(p2 − m2 )(q 2 − m2 )
i~ Q Q i~
= 2 2
i (i~) + (i~) i
q −m ~ ~ p − m2
2
= − , (10.124)
August 31, 2019 301
Q
= i~ , =i . (10.125)
~
These rules are very similar to those we adopted in regular QED : however,
in general we have
6= (10.126)
in other words,
= − . (10.128)
The proof of current conservation again relies on the SDe’s for this model :
= + + ,
= + + ,
= + + ,
(10.129)
= +
302 August 31, 2019
= −
+ −
(10.130)
If we now iterate the SDe cleverly for the first two of these four diagrams,
we obtain
= +
− −
+ −
= 0 , (10.131)
since we do have
= , (10.132)
owing to the simple, momentum-independent structure of the seagull vertex.
Comparing the lines of the proof for sQED with that of regular QED, the
general proof strategy becomes clear : if in a diagram a slashed propagator
occurs as one of the indicated lines of a (semi-)connected graph, we must
iterate de SDe for that line, and then we can collect the various canceling
contributions.
to derive (up to an overall factor) the Yukawa potential. But we used a single
source and looked at the field’s response to it : now, we have to consider two
particles, say, scattering ever so slightly. Let us start with two scalars31 with
masses M1,2 and charge Qe . Using the sQED Feynman rules we find for the
tree amplitude :
p q i~Q2
1 1
M= k = (p1 + q1 · p2 + q2 ) . (10.134)
p2 q2 k 2 + i
Since p01 = q10 and p02 = q20 in the centre-of-mass frame we have k 0 = 0 and
we can write
i~Q2e (4M1 M2 )
M = M(~k) = − , (10.135)
|~k|2
where we have assumed the static limit (SL) in the numerator ; and this we
have to Fourier-transform to find the interaction energy V (r). The correct
convention (as we shall see) is to use
Z
i 1
V (r) = d3~k M(~k) exp(i~k · ~x) , (10.136)
4M1 M2 (2π)3
so that we find, for this scattering, V (r) = α/r, a repulsive interaction32 .
If we now change the direction of the charge flow in the (p2 , q2 ) line in
Eq.(10.134), the amplitude changes sign by the sQED Feynman rules, and
the interaction energy becomes attractive. Therefore, like charges repel, and
opposite charges attract33 . We may also consider the ‘original’ Yukawa inter-
action, mediated by a scalar particle. Its propagator has numerator i~ rather
than the −i~g µν for the photon, with no sign change in the interaction vertex
if we replace particle by antiparticle. We conclude that in the exchange of a
scalar particle rather than a vector one, the interaction is always attractive.
Let us now turn to the case of regular QED, and consider e− µ− scattering :
p q −i~Q2e
1 1
M(~k) = k = u(q1 )γ µ u(p1 ) u(q2 )γµ u(p2 ) . (10.137)
p2 q2 ~
|k|2
31
We shall assume that they are not identical so as to avoid trouble with having to deal
with two diagrams rather than one.
32
What about the factor 4M1 M2 ? It is just the wave function normalization factor for
two nonrelativistic particles.
33
If we had chosen the other convention in Eq.(10.136), with −i rather than i, we would
have found that like charges attract, contrary to experience. The point here is that the
interaction can change sign between like-charge and opposite-charge.
304 August 31, 2019
and find, from the Gordon decomposition, that opposite-sign fermions also
repel ! But we have forgotten the Fermi minus sign lurking backstage. The
correct Feynman rules are those of Eq.(7.123) rather than the ‘prettified’
ones of Eq.(7.151), and we obtain an extra sign that tells us that oppositely-
charged fermions attract each other just like oppositely-charged scalars34 .
This is the Klein-Gordon equation for charged scalar fields. We see that the
same ‘minimal substitution rule’ pµ → pµ − eAµ as in the Dirac case is
employed to account for the presence of the e.m. field ; and we see that the
charge coupling constant e is defined in the same way for both scalar and
Dirac particles.
We find, for spin-1/2 electrons, the relativistic Pauli equation, the Klein-
Gordon equation with an extra spin term, the so-called Stern-Gerlach term
added on : E86
2
i∂ − eA(x) − m2 − eσ µν ∂µ Aν (x) φ(x) = 0 .
(10.146)
306 August 31, 2019
which represents the coupling between the magnetic field and the angular
momentum of the moving charge. For the Stern-Gerlach term, we have
e µν
−e σ µν ∂µ Aν (x) = σ εµναβ tα B β
2
= ieγ 5 σαβ tα B β = −eB
/ γ 5 t/ . (10.151)
We shall again alculate this for massless electrons. At the tree level there are
6 Feynman diagrams, so that the amplitude reads
M(λ; λ1 , λ2 , λ3 ) = i~3/2 e3
(/k3 − p/1 ) (/p2 − k/1 )
× uλ (p1 )/λ3 (k3 ) /λ2 (k2 ) /λ (k1 )uλ (p2 )
2(k3 p1 ) 2(k1 p2 ) 1
+ (perm) . (10.157)
Here we have explicitly indicated the various helicities, and ‘perm’ stands
for the other 5 permutations of the photons. For λ = + we take the photon
polarisations as in section 10.3.3, and that immediately tells us that the
amplitude vanishes if λ1 = λ2 = λ3 (as was already anticipated in exercise
79). We see that the only helicity amplitude that we have to work hard on
is, say, M(+; − + +). Neglecting, for now, the overall factor i(2~e2 )3/2 we
can write
M(+; − + +) =
u− (k3 )u− (p2 ) (/k3 − p/1 ) u− (k2 )u− (p2 ) (/p2 − k/1 ) u− (p2 )u− (k1 )
u+ (p1 ) u+ (p2 )
s− (k3 , p2 ) 2(k3 p1 ) s− (k2 , p2 ) 2(k1 p2 ) s+ (k1 , p2 )
+ (perm) . (10.158)
Actually, only two diagrams contribute here, namely the one written down
and the one where k2 and k3 are interchanged. With
and
u− (p2 )(/p2 − k/1 )u− (p2 )
= −1 (10.160)
2(k1 p2 )
we can rewrite
u− (p2 )(/k3 − p/1 )u− (k2 ) s− (k1 , p2 )
M(+; − + +) = − + (k2 ↔ k3 ) .
s− (k3 , p1 ) s− (k3 , p2 ) s− (k2 , p2 ) s+ (k1 , p2 )
(10.161)
We can simplify further :
u− (p2 )(/k3 − p/1 )u− (k2 ) = u− (p2 )(/k2 + k/3 − p/1 − p/2 )u− (k2 )
= −u− (p2 )/k1 u− (k2 ) , (10.162)
August 31, 2019 309
We can easily infer the other helicity amplitudes. The final answer is
3
(kj p1 )(kj p2 ) [(kj p1 )2 + (kj p2 )2 ]
P
j=1
|M|2 = 2~2 e4 (p1 p2 )
3
. (10.167)
Q
(kj p1 )(kj p2 )
j=1
Looking at this result, we notice two interesting things. In the first place,
it is very simple, something you would not have guessed right off, and cer-
tainly would have had to work on very hard using the classical Casimir-trick
approach. In the second place, Eq.(10.166) contains only s− and no s+ ,
as in the two-photon annihiliation case of section 10.3.3. This is a general
feature : conceding that M(+; +++) = 0 is the simplest possible amplitude,
the ‘next-simplest’, in our case M(+; − + +), is both simple and holomor-
phic in the spinor products37 . Such amplitudes are called maximal helicity
violating (MHV) and are an object of research in their own right, occurring
in many theories with massless particles such as QCD with massless quarks
and gluons, and even in gravity.
37
In the sense that s+ is the complex conjugate of s− up to a sign.
310 August 31, 2019
In the static limit the kinetic energies are negligible with respect to the
rest energies of massive particles. We shall, for our process
p q p
k2 p q
M = + + (10.170)
k1 k2 k1 k2
k1 q
Here e is the electron charge. The photon helicities are denoted by λ1,2 , and
we shall use the representation
uλ (k1 )γ α uλ (k2 ) uλ (k2 )γ α uλ (k1 )
1 (λ)α = p , 2 (λ)α = p , (10.172)
4(k1 · k2 ) 4(k1 · k2 )
Since the amplitude must be dimensionless in energy units and m is the only
scale in the SL, it is not surprising that m drops out altogether. The second
amplitude is given by
2 1 1 2
M(+, −) = −2ie ~ − p · 1 (+)
(pk1 ) (qk1 )
(k1 k2 )
= 2ie2 ~ eiφ |p · 1 (+)|2
(pk1 )(qk1 )
2(pk1 )(qk1 ) − m2 (k1 k2 )
= ie2 ~ eiφ
(pk1 )(qk1 )
(E + k)(1 + cos θ)
= ie2 ~ eiφ . (10.178)
E + k cos θ
312 August 31, 2019
A++
1 = u(q) 2(qk2 )ω+ k/1 + 2(pk1 )ω− k/2 − mω+ k/1 k/2 − mω− k/2 k/1 u(p)
= u(q) 2(pk1 )/k2 − 2m(k1 k2 )ω+ u(p) ,
A++
2 = u(q) 2(qk1 )ω + k
/ 2 + 2(pk 2 )ω − k
/ 1 − mω − k
/ 1 k
/ 2 − mω + k
/ 2 k
/ 1 u(p)
= u(q) 2(qk1 )/k1 − 2m(k1 k2 )ω+ u(p) . (10.183)
Since
u(q) k/2 − k/1 u(p) = u(q) p/ − q/ u(p) = 0 , (10.185)
we arrive at the exact result
m(k1 k2 )
M(+, +) = −ie2 ~ u(q, s0 ) 1 + γ 5 u(p, s) ,
(10.186)
2(pk1 )(qk1 )
where we have indicated the spins. In the SL, we may in this expression
approximate q by p. The γ 5 term then drops out, and the amplitude vanishes
if s0 = −s. We conclude that M++ SL takes again the form (10.177), and that
the electron spin is not influenced in the SL. The second amplitude is given
by
+−
−ie2 ~ A+−
A1 2
M(+, −) = − ,
(k1 k2 ) 2(pk1 ) 2(qk1 )
A+−
1 = u(q) u− (k1 )u − (k 2 ) + u + (k2 )u + (k1 ) (/p + k/1 + m) ×
u+ (k2 )u+ (k1 ) + u− (k1 )u− (k2 ) u(p) ,
A+−
2 = u(q) u+ (k2 )u+ (k1 ) + u− (k1 )u− (k2 ) (/q − k/1 + m) ×
u− (k1 )u− (k2 ) + u+ (k2 )u+ (k1 ) u(p) . (10.187)
Again using the various standard-form techniques we can establish that
A+−
1 = u(q)u− (k1 )u− (k2 )/pu− (k1 )u− (k2 )u(p)
+ u(q)u+ (k2 )u+ (k1 )/pu+ (k2 )u+ (k1 )u(p)
= eiφ u(q) ω− k/1 p/k/2 + ω+ k/2 p/k/1 u(p) ,
(10.188)
where eiφ is the phase factor of Eq.(10.179). For A+−
2 we find the exact same
result, and thus
ie2 ~ eiφ
M(+, −) = u(q) ω− k/1 p/k/2 + ω+ k/2 p/k/1 u(p) . (10.189)
2(pk1 )(qk1 )
In this expression, the denominator is of order O (k 2 ) which is already com-
pensated by the occurrence of k/1 and k/2 in the numerator. In the SL we can
therefore again replace q by p since q = p + O (k), and use
ω− k/1 p/k/2 + ω+ k/2 p/k/1 = (pk1 )/k2 + (pk2 )/k1 − (k1 k2 )/p − iγ α α (k1 , k2 , p) (10.190)
which yields 4(pk1 )(pk2 ) − 2m2 (k1 k2 ) when sandwiched between u(p, s) and
u(p, s), but zero when sandwiched between u(p, −s) and u(p, s). We see that
314 August 31, 2019
also M(+, −)SL is the same as in the scalar case, while the spin again remains
unaffected by the Thomson scattering.
We have thus established that Thomson scattering will not distinguish
between scalar or Dirac electrons. It is interesting to note, however, that
the amplitudes depend on θ even in the SL, while of course in that precise
limit (zero photon energy) the value of θ becomes undetermined ! We see
that Thomson scattering (and consequently the determination of the elec-
tron charge by this process) is only meaningful if there is some momentum
transfer, no matter how small39 .
M = A1 (q ·0 )(η1 ·η2 )+A2 ε(P, 0 , η1 , η2 )+A3 ε(P, q, η1 , η2 )(q ·0 ) . (10.194)
The coefficients A are of course undetermined, but they can only depend on
P 2 , q 2 and (P · q). This last product is zero, and P 2 = −4q 2 = M 2 where M
is the positronium mass, so the A’s are effectively just constants. We now
come to the main observation : under interchange of the two photons we
40
Looking really closely would in this case mean splitting it up, and then the positronium
is gone.
41
You might be tempted to write down a term like ε(q, 0 , η1 , η2 ) but since all these
vectors have vanishing zeroth component, this Levi-Civita product is simply zero.
316 August 31, 2019
10.8 Exercises
Excercise 77 Compton Current Conservation
Consider the process
e− (p1 ) γ(k1 , 1 ) → e− (p2 ) γ(k2 , 2 )
where the momenta and polarisations are indicated, and write the two dia-
grams that describe it at tree level. Then, substitute 1 → k1 and show that
the amplitude, so treated, vanishes.
Excercise 78 Rare or impossible ? The process µ → e γ
The process
µ− (p) → e− (q) γ(k)
is not allowed in standard QED since it violates current conservation. Nev-
ertheless, it could be possible with a different interaction vertex coming from
some ‘new physics’. The amplitude would then read
√
M = i ~g u(q) cos θ + sin θγ 5 k/ / u(p)
42
In the words of Feynman, ‘everything that is not explicitly forbidden is allowed’.
August 31, 2019 317
where g is the coupling in the new physics, and the angle θ simply parametrizes
the relative weight of the two alternative couplings.
3. Compute h|M|2 i using the Casimir trick and the trace identities.
Γ(µ → eγ)
≤B , B ≈ 10−11 .
Γ(µ → all)
We shall assume that the decay µ → eνµ ν¯e is by far the dominant one.
Show that we can relate this to B as follows :
K
Λ≥ √ .
B
Compute K, and find the current lower limit on Λ.
3. If all photon helicities except one are equal, only (n − 1)! diagrams
contribute if we choose the gauge vectors cleverly.
318 August 31, 2019
p 1 + p2 → q1 + q2
We take the incoming particles to have mass m and the outgoing ones to
have mass M . From the conservation of energy and momentum,
1. Show that the ‘canonical’ radiation cone angle derived from Eq.(10.116)
is about 5 × 10−6 radians.
August 31, 2019 319
3. Show that α ∼ 3 × 10−3 radians for f = 0.5, i.e. half the radiation is
emitted at larger angles than this.
1. Show that the partial decay width contains the soft-photon factor
2(p1 p2 ) m2 m2
S= − −
(p1 k)(p2 k) (p1 k)2 (p2 k)2
2. Let the angle between p~1 and ~k in the centre-of-mass frame have cosine
equal to c. Show that
2(E 2 + p2 ) 2m2 (E 2 + p2 c2 )
S= −
E 2 − p 2 c2 (E 2 − p2 c2 )2
where the polarisations and spins are indicated along with the mo-
menta. Write out the amplitude at the tree level, keeping the electron
mass m nonzero.
2. Choose the gauge vectors such that not only (k1 1 ) = (k2 2 ) = 0 but
also (k1 2 ) = (k2 1 ) = 0. Show that in that case 1,2 0 = 0 in the
centre-of-mass frame.
3. Take the static limit, in which p1 and p2 become equal. Show that in
this limit the amplitude is proportional to
= − ,
= = = . (11.2)
321
322 August 31, 2019
F µν (m2 , p) = pµ pν − s g µν K(m2 , s) , s = p2 .
(11.3)
where θ is the scattering angle between p~1 and ~q1 in the centre-of-mass
frame. The full sum over the final state of course includes the phase-space
integration :
~2 Q2 Q02
Z X
|Ma |2 dV (p1 + p2 ; q1 , q2 ) = β(3 − β 2 )θ(s > 4m2 ) . (11.6)
spins
12π
The forward scattering amplitude is that for s(p1 ) s̄(p2 ) → s(p1 ) s̄(p2 ) with
the fermion loop inserted :
p2 p2 ~2 Q02
p1 = (p1 − p2 )µ F µν (m2 , p1 + p2 )(p1 − p2 )ν
p1 s2
= ~2 Q02 K(m2 , s) . (11.7)
Note that the pµ pν term in Eq.(11.3) drops out here because of current con-
servation. The optical theorem 6.34 then tells us that
α Q2 ~
< K(m2 , s) = − β(3 − β 2 ) θ(s > 4m2 ) , α= . (11.8)
6~ 4π
August 31, 2019 323
Now,
= + ,
−i~ g µν −i~ g µα
Πµν (s) = −s gαβ K(s) Παν (s)
+
s s
−i~ g µν −i~ g µν
= = . (11.16)
s 1 − i~K(s) s (1 − Π(s))
β(3 − β 2 )
α 2 β+1 2 8
Π(s) = −R + log m + log +β − , (11.17)
3π 2 β−1 3
3
This saves us a lot of unnecessary work : the ‘other terms’ contain k µ k ν , g µν k 2 and
suchlike, with integrals more cumbersome by far.
August 31, 2019 325
Note that we have defined β with the mass m2 augmented by its ‘natural’
extension into m2 −iη (η ↓ 0); the iη takes care of the correct imaginary part,
as it should. Plot 11.1 shows the behaviour of Π(s) − Π(0) as a function of
s/m2 . It is continuous both at s = 0 and at s = 4m2 , and we see the opening
of a branch cut for s ≥ 4m2 . I used the value η = 10−8 and so obviated the
need to distinguish between the three different cases s < 0, 0 < s < 4m2 and
s > 4m2 .
order effects from strong interactions4 . But all is not lost ! Let us write the
hadronic self-energy of the photon as5
µ ν
= (pµ pν − s g µν ) H(s) , s = p2 . (11.19)
p
p1 p1
Mf =
p2 p2
Q2 ~2
= 2
uλ (p1 )γµ uλ (p2 ) (pµ pν − s g µν ) H(s)uλ (p2 )γµ uλ (p1 )
s
= 2Q2 ~2 H(s) , (11.20)
+ −
where σhad (s) stands for the total
√ cross section for the process e e →
hadrons at total collision energy s, and dV includes a sum over all hadronic
final states6 . The optical theorem (6.34) then tells us that
s σhad (s)
<H(s) = − : (11.22)
4πα~
4
After all, free quarks are never observed, but only their more-or-less stable bound
states. The interactions responsible must be drastic, to say the least !
5
Here we again assume current conservation in the photon-quark interactions. If that
does not hold, then the vacuum polarization is very far down the list of our problems. . .
6
The prefactor 4s arises from excising the flux factor 1/2s, realising that λ can take
two values, and that there is a factor 1/4 for the spin average. Additionally, we need the
cross section with the radiative corrections to the initial state somehow removed ; but this
is doable.
August 31, 2019 327
I have taken the external fermions to be on their mass shell for simplicity,
otherwise we would need even more diagrams10 . That these cannot easily
7
The first actually useful phenomenological application of this idea is in F.A. Berends
and G.J. Komen, Radiative corrections to Bhabha scattering and µ pair production from
the hadronic vacuum polarization, Phys.Lett. 63B (1976) 432. But the references therein
show that the idea is a bit older.
8
Around the turn of the century, the hadronic vacuum polarization was the dominant
uncertainty in the precision analysis of the LEP data (helping to establish bounds on the
Higgs boson mass, for instance). Not surprisingly, a small industry arose of competing
computations of Πh (s). If you wish, you can also try to extract ‘effective’ quark masses by
fitting them to the observed polarization. I have always found that of dubious relevance.
9
In order not to let the diagrams become too cumbersome I indicate the fermion flow
with just a single arrow.
10
In that case, a fermion would enter into another ‘blob’ and we would have to let the
photon also enter in there.
328 August 31, 2019
= − , = − , = .
(11.24)
By using Eq.(10.15) we see that the total set of graphs vanish. This is,
of course, just an illustration of the Ward identity of section 10.2.4. The
situation is slightly more subtle if we consider the internal photon. If we
single out the to-be-handlebarred vertex we see that we. have to consider
not 3 but the 6 diagrams
+ + + + +
E91 and then the whole lot cancels11 . The upshot is that we are allowed to use
the simple photon propagator −g µν /(k 2 + i) for the internal photon line.
1 − a(m2 )
∼ m 1 − y(m2 ) , y(m2 ) = a(m2 ) + b(m2 ) . (11.28)
m0 = m 2
1 + b(m )
Note that we keep only the one-loop corrections, so that both a and b are
formally infinitesimal. Implementing Eq.(11.28) in the summed propagator
we may Taylor-expand to find that, to one-loop accuracy, the inclusion of
the self-energy leads to
i~ i~ 1
→ + (terms with no poles) ,
p/ − m p/ − m 1 − z(m2 )
z(m2 ) = 1 − a(m2 ) − 2m2 a0 (m2 ) − 2m2 b0 (m2 ) . (11.29)
We shall restrict ourselves to on-shell (external) fermions only, so the non- E92
pole terms can be neglected.
We will have to compute A(m2 ), B(m2 ), m2 A0 (m2 ) and m2 B 0 (m2 ), and this
simplifies our lives since then12 n(x) = m2 x2 + λ2 . In computing A, B(m2 )
we can let λ vanish with impunity :
Z1
1
A(m2 ) ≈ dx(1 − x) (1 − log(m2 x2 )) = 1 − (log m2 − 3) ,
2
0
Z1
B(m2 ) ≈ dx (1 − log(m2 x2 )) = 1 − log m2 − 2 .
(11.34)
0
This means that a(m2 ) and b(m2 ) have an UV divergence residing in Γ().
The other two are more interesting. We may decide to keep λ nonzero or let
it vanish. In the first case we find (taking to zero where possible) :
Z1
m2 x(1 − x)2
2
2 0 2 m
m A (m )λ ≈ dx 2 2 2
≈ log −3 ,
(m x + λ ) 2 λ2
0
12
Since λ only becomes relevant when x is very close to zero, we can replace λ2 (1 − x)
by just λ2 .
August 31, 2019 331
Z1
m2 x(1 − x)
2
2 0 2 m
m B (m )λ ≈ dx 2 2 2
≈ log − 2 . (11.35)
m x +λ 2 λ2
0
We see that the overall factor will remove the singularity in Γ() since
Γ() ≈ 1. The UV divergence is now gone in these integrals, and instead we
find an IR divergence regularised by λ. The alternative, taking λ = 0, leads
to
Z1
0 −1
2 2
dx x−1−2 (1 − x)2 = 1 − (log m2 − 3) .
m A (m )0 =
(m2 ) 2
0
Z1
−1
m2 B 0 (m2 )0 = dx x−1−2 (1 − x) = 1 − (log m2 − 2) .
(m2 ) 2
0
(11.36)
The IR divergence is now represented by the 1/ in Γ() with the opposite
sign provided by the −1 in Eq.(11.36). Putting everything together we finally
arrive at
α
y(m2 ) = 3R − 3 log m2 + 4 ,
4π
2 α
−3R + 3 log m2 − 4 ,
z(m )0 =
4π
2 α
−R − 2 log λ2 + 3 log m2 − 4 .
z(m )λ = (11.37)
4π
In the same spirit, we can say that the photon self-energy Eq.(11.15) has
an IR divergence regularized by the fermion mass in log m2 . Indeed, if we
compute K(0, 0) directly we find again a vanishing result.
where p1,2 are two external momenta, and k is the photon’s momentum. This
we have to integrate over, with some (small !) maximum value E0 for the
energy. The integral that we have to compute is then, after the Feynman
trick,
0
(p1 · p2 )
Z
4πα 3~ θ(k < E0 )
B(p1 , p2 ) = d k
(2π) 3
2|~k| (p1 · k)(p2 · k)
Z1 0
3~ θ(k < E0 ) (p1 · p2 )
Z
4πα
= dy d k ,
(2π)3 2|~k| (py · k)2
0
py µ = y p1 µ + (1 − y)p2 µ . (11.39)
If we again give the photon a very tiny mass λ then λ < k 0 < E0 , and λ
will regulate the IR divergence15 ; but we can also let k 0 start at zero and
use dimensional regularization16 . Let us again work in 4 − 2 dimensions17 .
Using the t-shell formula (6.82) we then can cast our integration element into
the following form :
µ2 ~
3−2~ θ(0 < |k| < E0 )
Z
d k
(2π)3−2 2|~k|
2
2 ZE0 Z1
µ d cos θ
= 3/2−
dt t− , (11.40)
2(4π) Γ(3/2 − ) 2
0 −1
we find that
Moreover18 ,
2
Z1 ZE0
d cos θ 1 1 1
0 2
= , dt t−1− = − E0 −2 . (11.42)
2 (py − |~py | cos θ) py · p y
−1 0
Z1
α 1 p 1 · p2
B(p1 , p2 )0 = log(2E0 ) − R − 1 dy . (11.43)
π 2 py · py
0
Z1 Z1
1 + β2 1 + β2
p1 · p2 1+β
dy = dy = log . (11.45)
py · py 1 − β 2 (2y − 1)2 2β 1−β
0 0
In the limit of very small mass it approaches log(s/m2 ). The final result, in
dimensional regularization, is therefore
1 + β2
α 1 1+β
B(p1 , p2 )0 = log(2E0 ) − R − 1 log ,
π 2 2β 1−β
α 1
B(p1 , p1 )0 = log(2E0 ) − R − 1 . (11.46)
π 2
18
Here is ‘negative’, remember ? Therefore t− = 0 at t = 0.
19
The case of unequal masses gives us a result that is more complicated but not more
enlightening : see exercise 95.
August 31, 2019 335
ZE0 Z1
α (1 + β 2 ) 0 |~k|
B(p1 , p2 )λ = dk d cos θ
2π (k 0 )2 − β 2 |~k|2 cos θ2
λ −1
ZE0 !
α 1 + β2 1 0 k 0 + β|~k|
= J(β) , J(β) = dk 0 log . (11.47)
π 2β k k 0 − β|~k|
λ
4E0 2 2
ZE0 Z /λ
∂ 2|~k| dk 0 1 (v − 1)2
J(β) = = dv
∂β (k 0 )2 − β 2 |~k|2 1 − β2 v(v + r)(v + 1/r)
λ 1
2
1 4E0 1 1−β
≈ log + log(r) , r = . (11.48)
1 − β2 λ2 β 1+β
The term containing log(E0 ) comes out the same, but we pick up extra terms.
For small mass, these approach −(log(s/m2 ))2 /4 − π 2 /6.
20
For the dilogarithm function Li2 , see appendix 15.15.10.
336 August 31, 2019
Note that we have taken out a factor iQ/~ so that V µ equals γ µ at the tree
level. Q is the fermion charge which we shall take equal to the electron
charge. In addition we shall only discuss the case where the two fermions are
on-shell, p2 = q 2 = m2 . The emitted photon is therefore, generally, off-shell.
Following the Feynman rules we find, again in the Feynman gauge, and using
a photon mass λ for possible later use :
−iQ2 ~ µ
Z
µ 4 Ω
V (p, q) = dk ,
(2π)4 N
1/N = (k 2 + 2(pk) + iη)(k 2 + 2(qk) + iη)(k 2 − λ2 + iη) ,
Ωµ = γ α (/q + k/ + m) γ µ (/p + k/ + m) γα . (11.51)
Here we have dropped all terms linear in k̂, and for the quadratic term we
used the fact that k̂ ρ k̂ ν can only integrate to something proportional to g ρν
and therefore we can replace
1
k̂ ρ k̂ ν → g ρσ k̂ 2 . (11.53)
4 − 2
The vertex correction thus becomes
(4πα) 2 µ2 2 µ µ
4−2 A0 k̂ γ − A1 γ − m[/ p − q/, γ µ ]B1
Z Z
µ
V (p, q) = δxy d k̂ ,
(2π)4−2 (k̂ 2 + D)3
D = m2 x2 + m2 y 2 + 2σxy + λ2 . (11.54)
the photon mass is only important if x = y = 0. The symbol δxy stands for
dx dy θ(0 < x < 1) θ(0 < y < 1 − x). After the t-shell formula-cum-integral
we are then left with
Z
α
µ
V (p, q) = 2
(4πµ ) δxy A0 γ µ Γ()(2 − ) D−
4π
µ µ −1−
− A1 γ + m[/p − q/, γ ]B1 Γ(1 + ) D . (11.55)
11.6 Exercises
Excercise 89 Doing it with fermions only
Prove that Eq.(11.8) is also obtained if we start with an initial state consisting
of a fermion-antifermion pair instead of scalars. Take care not to average
over the incoming spins, as that could not possibly correspond to a forward
scattering amplitude (why ?).
N = p/ − m0 − a(s)/p − m0 b(s)
where
3. Show that
Z1
s2 + κ2 − δ 2 (s + κ)2 − δ 2
p1 · p2
dy = log
p2y 4κ s (s − κ)2 − δ 2
0
4. Show that we obtain the correct result if δ → 0 ; and that the integral
diverges if one or both masses vanish.
340 August 31, 2019
Chapter 12
Quantum Chromodynamics
341
342 August 31, 2019
2. All colours are equal and none are ‘more equal than others’, which
means that particles that only differ by their colours propagate through
spacetime in the same way 3 .
pα nλ nλ pβ
αβ −1 α λβ
Π (p) = i~ 2 g λ− g −
p + i (p · n) (p · n)
α β α β 2 α β
i~ αβ p n +n p n p p
= 2 −g + − . (12.3)
p + i (p · n) (p · n)2
The first line of Eq.(12.3) shows that, indeed, Παβ (p) nα = Παβ (p) nβ = 0.
Note that we have extended the definition of the axial-gauge propagator to
nonzero values of n2 . However, since Πα α (p) ∝ −2+n2 p2 /(p·n)2 , the number
of degrees of freedom is still equal to 2 on the mass shell, where p2 = 0.
b a
c
j
j b
k
a
contains the colour factor
X
(T j )a b (T k )b a = Tr T j T k
,
a,b
we must have b = a since the colour is conserved and the photon is colourless ;
therefore the T matrices must be traceless :
Tr T j = 0 for all gluon colours j .
(12.7)
Finally, we consider the following two-loop self-energy diagram of the photon :
Here, the fermions are quarks and the internal line labelled j is a gluon of
colour type j : of course, we have to sum over all j values. If we compare this
diagram to the corresponding QED one, we see that apart from the overall
charges (g 2 instead of Q2 ) the only difference is the colour factor, in this case
X
Tr T j T j
j
Now, if our theory is to be unitary, it must obey the Cutkosky rules, and
therefore we demand that
+ j
+ j
j
+ j
+ j
=0 . (12.8)
For the QED diagram, this indeeds holds. In the coloured case, however, the
colour structures of the diagram cut in the various ways are no longer the
same : the three lines in Eq.(12.8) are proportional to, respectively,
X X † †
j j†
X
j j
Tr T j T j
T = Tr T T , Tc = Tr T T , and Tcc = .
j j j
Unitarity can therefore only be safe if these three different traces are, in fact,
equal to one another. We may therefore write
†
X
Tr Aj Aj = 0 , Aj = i T j − T j
2Tc − T − Tcc = . (12.9)
j
346 August 31, 2019
hence all Aj are actually identically zero, and the matrices T j must be Her-
mitean. The number of different gluon colours type is therefore N 2 − 1, and
the constant k of Eq.(12.5) is equal to (N 2 − 1)/2N .
Tr T j T k T l Tr T j T k T l
1 k l k l
1 k l
k l
= Tr T T T T − Tr T T Tr T T
2 N
1 l
l
2 l l
1 l
l
= Tr T Tr T − Tr T T + 2 Tr T Tr T
4 N N
2
1 N −1
Tr T l T l = −
= − , (12.17)
2N 4N
and
Tr T j T k T l Tr T j T l T k
1 k l l k
1 k l
l k
= Tr T T T T − Tr T T Tr T T
2 N
1 k k
2 k k
1 k
k
= Tr T T Tr (1) − Tr T T + 2 Tr T Tr T
4 N N
2 2 2
N −2 (N − 1)(N − 2)
Tr T k T k =
= . (12.18)
4N 8N
8
Empirically.
348 August 31, 2019
M= q2 + , (12.19)
kq
p2 k p2 2
b b
They read
q/1 − p/1 + m
M1 = −i~g 2 v(p1 )/1 /2 u(p2 ) (T j T k )a b ,
−2(q1 · p1 )
p/2 − q/1 + m
M2 = −i~g 2 v(p1 )/2 /1 u(p2 ) (T k T j )a b . (12.20)
−2(q1 · p2 )
Let us now put the handlebar on gluon 1, so that we replace µ1 by q1µ . By
the same reasoning as in chapter 10, we arrive at the handlebar rule
= − , (12.21)
j j
M1+2 c1 →q1 = +
k
k
j j
= − + j
− j
k k
k k
j
= − j
k
k
where the square brackets denote, of course, the commutator of the matrices
T j and T k . Because of the colour structure we have a non-vanishing result,
and current conservation is in trouble ! The remedy must be to introduce a
third diagram, with a nontrivial ggg vertex,
a j
p1
q1
n
q2
p2
b k
and it is now our job to determine the form of the new three-gluon vertex.
We shall do this by investigating loop diagrams.
10
In calculations such as this one it is often sufficient to simply label the gluons by their
colour, thus simplifying the typography.
350 August 31, 2019
p −q 2 ρ
l
ν k q3
q2
K ∼ (12.24)
p p + q1
j q1
µ
Here three gluons are effectively coupled by a quark loop. We have explicitly
indicated the momentum flows. Note especially that the gluon momenta are
all counted flowing out of the vertex, so that we have
q1 + q2 + q3 = 0 . (12.25)
Apart from overall coupling constants and the like, the loop diagram is given
by
K ∼ Y (q1 , µ; q2 , ν; q3 , ρ) Tr T j [T k , T l ] ,
(12.27)
with
for some numbers a1 , . . . , a6 . For large p, each of the three propagators goes as
1/p, and the loop integral is therefore divergent. We see that indeed a three-
gluon counterterm must be allowed, and therefore a three-gluon coupling
must appear in the action, otherwise the theory would not be renormalizable ;
and the form of the three-gluon vertex must be that of Eq.(12.28).
Without evaluating the loop integral completely, we can glean all the
information we need. Consider the following transformation on K :
q1 ↔ −q2 , q3 → −q3 , µ ↔ ν . (12.29)
This transformation leaves the momentum conservation law (12.25) intact,
and also preserves the value of T (by the reversal property (7.29) of Dirac
traces). The same holds, of course, for the transformations
q1 ↔ −q3 , q2 → −q2 , µ ↔ ρ ,
q2 ↔ −q3 , q1 → −q1 , ν ↔ ρ . (12.30)
The function Y must therefore satisfy
Y (q1 , µ; q2 , ν; q3 , ρ) = Y (−q2 , ν; −q1 , µ; −q3 , ρ) =
= Y (−q3 , ρ; −q2 , ν; −q1 , µ) = Y (−q1 , µ; −q3 , ρ; −q2 , ν) ; (12.31)
and by inspection we then find that c1 = c3 = c5 = −c2 = −c4 = −c6 . We
shall therefore from now on use the Yang-Mills three-boson vertex
Y (q1 , µ; q2 , ν; q3 , ρ)
≡ (q1 − q2 )ρ g µν + (q2 − q3 )µ g νρ + (q3 − q1 )ν g ρµ . (12.32)
Note that this form is antisymmetric in the interchange of any two gluons,
and therefore invariant under a cyclic permutation11 . E100
Note that the gluon momenta are counted outgoing from the vertex. The
value of g3 must be determined, as well as the colour factor hjkl . The Y
function’s total antisymmetry strongly suggests that we take the h symbols
antisymmetric as well, and later on we shall show that this is indeed the case.
The following object will turn out to be useful :
∆(q)αβ ≡ q α q β − q 2 g αβ , (12.34)
for which
∆(q)αβ = ∆(q)βα , ∆(q)αβ qβ = 0 . (12.35)
Also,
∆(q)αβ β = q α (q · ) − q 2 α = 0 (12.36)
if is the polarisation vector of an on-shell gluon with momentum q.
We now come to an important result. Let us consider the vertex of
Eq.(12.33), and let us put a handlebar on gluon q3 . We find, using momentum
conservation in the form q3 = −q1 − q2 ,
Y (q1 , µ; q2 , ν; q3 , q3 ) ≡ Y (q1 , µ; q2 , ν; q3 , ρ) q3 ρ
= (q1 − q2 · q3 )g µν + (q2 − q3 )µ q3 ν + (q3 − q1 )ν q3 µ
= (q2 − q1 · q2 + q1 )g µν − q2 µ (q1 + q2 )ν + q1 ν (q1 + q2 )µ
= ∆(q1 )µν − ∆(q2 )µν . (12.37)
Πµα (q1 ) ∆αβ (q1 ) Πβν (q2 ) = (i~g µα ) (gαβ ) Πβν (q2 ) , (12.38)
= m + , (12.39)
m m
k k
k
j k ↔ i~ g µν δ jk (12.40)
µ ν
August 31, 2019 353
and
m g3 αβ jkm
↔ i g h . (12.41)
β α
~
k j
It is now time to return to the q q̄ → gg process. The new available
Feynman diagram, given by
a
p q1
1
n j
p2 q2 k
b
reads
n j n j n j
= + (12.43)
k k k
so that
M3 c = −i~gg3 v(p1 )/2 u(p2 ) hnkj (T n )a b . (12.44)
The total handlebarred amplitude now reads
a
M1+2+3 c = i~g v(p1 )/2 u(p2 ) g3 hjkn T n − g[T j , T k ] b (12.45)
g3 = g (12.46)
and
[T j , T k ] = hjkn T n . (12.47)
354 August 31, 2019
Note that since the matrices T are Hermitean, the constants h must be purely
imaginary12 . Moreover, we can compute them, using Eq.(12.6), as
hjkl = 2 Tr T j T k T l − T l T k T j ,
(12.48)
from which we see that the h symbols must be totally antisymmetric. Since
they are related to commutators, we can use the Jacobi identity to find
relations between them13 :
hmnj hmnk = 8 Tr T m T n T j Tr T m T n T k − Tr T m T k T n
, (12.52)
12
It is customary to write [T j , T k ] = i f jk n T n . The f ’s are then called the structure
constants, and the set of T matrices are then the generators of the Lie algebra of the group
SU (N ). The i is then combined with the overall i of the vertex to give a Feynman rule
without any i. This is of course a matter of taste.
13
Here and in the following, raising or lowering colour labels has no physical meaning ;
I do it only for typographical reasons.
August 31, 2019 355
1 jk
Tr T m T n T j Tr T m T n T k = −
δ ,
4N
m n j m k n N 1
δ jk ,
Tr T T T Tr T T T = − (12.53)
8 4N
we arrive at
hmnj hmnk = −N δ j k . (12.54)
j l j l j l
M = n + n
+ n (12.62)
k m k k m
m
15
These are precisely those of the Riemann-Christoffel tensor, familiar from the theory
of general relativity.
16
Since in this section we only consider gluons anyway, I shall here denote them by
smooth rather than wriggly lines ; this makes them somewhat easier to read, and certainly
easier to draw.
August 31, 2019 357
j l j l j l
Mc1 →q1 = n + n + n (12.63)
k m k m k m
and this combination does not obviously vanish. The three-gluon vertex
is somewhat cumbersome, but we can streamline our calculations a bit by
introducing a ‘partial’ three-gluon vertex :
λ
n g
j ↔ i (q1 + q2 )λ g αβ hjkn . (12.64)
~
k q1 α
β q2
Here, the momenta are counted in the direction of the arrows. Note that
reversing the arrows leaves the vertex unchanged owing to the antisymmetry
of the h symbol. There is therefore no ambiguity. We may write
= + + . (12.65)
This makes it easier to single out particular terms in Eq.(12.63). For instance,
the terms proportional to (2 · 4 ) are given by the diagrams
j l j l j l
n + n + n
k m k m m
k
(3 · q2 + q4 ) hmkn hnlj + (3 · q1 + q2 + q4 ) hmnl hnkj + (3 · 2q2 − q3 ) hnkl hnmj
= (3 · q2 + q4 ) [mklj] − (3 · q1 + q2 + q4 ) [mlkj] + (3 · 2q2 − q3 ) [klmj]
= (3 · q2 + q4 ) −[mljk] − [mjkl]
+(3 · q1 + q2 + q4 ) [mljk] − (3 · 2q2 − q3 ) [mjlk]
= [mljk] (3 · q1 + q2 + q4 − q2 − q4 ) + [mjlk] (3 · q2 + q4 − 2q2 + q3 )
= (q1 · 3 ) [mjlk] + [mljk] . (12.66)
358 August 31, 2019
+ g αµ g βν [jknm] + [jnkm]
+ g αν g µβ [jkmn] + [jmkn]
. (12.69)
= − − − . (12.70)
bad luck : with the three- and four-gluon vertices all amplitudes will vanish
under the handlebar. We shall prove this in the same manner as for QED,
only the proof will be obviously somewhat more involved. Before we turn to
our reliable working horse, the SDe, we first need one more result.
and let us now attach a fifth gluon in all possible ways, and slash the inter-
mediate propagators. This leads to the following expression :
j j j
n j
k n k
k k
n
l l
V = + l + l
+ n
. (12.71)
m
m m m
We now call upon the various symmetry properties and Jacobi identities :
So we find that V vanishes completely, and we shall employ that fact in E96
what follows next.
360 August 31, 2019
= + + (12.74)
Let us now apply the handlebar rules we have developed in this chapter :
= −
+ − . (12.75)
It is important to realize that the internal lines in the SDe take on all pos-
sible identities, and therefore the third diagram stands for both graphs in
Eq.(12.39), and the fourth stands for all three diagrams in Eq.(12.70)17 . We
now iterate the SDe for the slashed propagators in the first three diagrams,
and this gives us
= −
+ + +
− = 0 , (12.76)
since the first diagram on the second line cancels against those of the first line,
and the last diagram on the second line vanishes all by itself, as we have seen
in the previous section. This establishes the proof of current conservation in
QCD. It is important to realise that the above proof relies on our use of the
axial gauge for the gluon propagator. Indeed, that choice is what makes the
identity (12.39) possible.
17
Are you worried about possible double-counting here ? Don’t worry, be happy : the
symmetry factors are there for precisely that reason.
August 31, 2019 361
with the summation convention implied. Here the R are unitary 3×3 matrices
with unit determinant : they constitute the group SU(3). The quarks are
said to be triplets (3) under colour, and the antiquarks to be anti-triplets
(3̄). It pays to examine what kind of colour combined states of quarks and
antiquarks can have18 . In the first place, the combination q̄a q a transforms as
this combination doesn’t change colour at all, it is the ‘white’ singlet. The
other eight possible (orthogonal) q q̄ colour combinations are the colour octet
(8). In group-theory speak we say that 3 ⊗ 3̄ = 1 ⊕ 8. A typical octet state is,
for instance, q̄1 q 2 . The fact that the matrices T j are traceless combinations
of a colour and and anticolour index tells us that gluons are colour octets
as well. We can also investigate how the special quark-quark combination
abc q b q c transforms, where abc is the three-dimensional Levi-Civita symbol :
abc q b q c → abc Rm
b
Rnc q m q n . (12.79)
these quarks are also in the antitriplet, 3̄ state ! The remaining orthogonal
quark-quark colour combinations are the sextet (6), so that we can say that
3 ⊗ 3 = 3̄ ⊕ 6. A typical sextet state is, for instance, q 3 q 3 . Finally we
can observe that the colour combination abc q a q b q c is also a colour singlet,
a ‘white’ state. The empirical fact that we see only mesons (q q̄), baryons
(qqq), and antibaryons (q̄ q̄ q̄) occurring as more-or-less stable, more-or-less
free hadrons can therefore be nicely summed up by the postulate that only
E97 ‘white’ (singlet) colour states can occur freely.
We conclude that a q q̄ pair in a colour singlet state attracts, but in the octet
state repels. And indeed, colour octet states are never observed21 . We can
also consider qq scattering :
b a
p q
1 1 −i~g 2
M(~k) = p k j = u(q1 )γ µ u(p1 ) u(q2 )γµ u(p2 ) Cbd
ac
. (12.84)
2 q2 ~
|k|2
d c
Two quarks cannot form a singlet state, but they can form an antitriplet
state, an antisymmetric combination of two colours, for which we may take
1 and 3, say. The colour factor now evaluates to22
2
1 13 13 31 31
N +1
√ C13 − C31 − C13 + C31 =− (12.85)
2 2N
and the interaction is attractive. The remaining six colour combinations are
the symmetric sextet ones, for which we may simply choose a = b = c = d =
33
3, say, and then find C33 = (N − 1)/2N . A quark pair in such a state feels a
23
repulsive interaction .
to establish the three-gluon vertex, but what interests us here is not how
the unwanted terms vanish, but rather what are the acceptable terms that
are left. As we have seen, at the tree level the amplitude is described by 3
Feynman diagrams :
p1 q1 p1 q2 p1 q1
M= (12.86)
p2 q2 p2 q1 p q2
2
The polarisations
The first (and perhaps most crucial) step24 is to make a clever choice for the
polarisation vectors. It turns out to be smart indeed to take
u+ (qi )γ µ u+ (p2 ) u− (qi )γ µ u− (p1 )
i (+)µ = √ , i (−)µ = √ , (12.88)
2 s− (qi , p2 ) 2 s+ (qi , p1 )
just like we did in QED, since then
√ √
2 u− (qi )u− (p2 ) 2 u− (p1 )u− (qi )
ω− /i (+) = , ω− /i (−) = ; (12.89)
s− (qi , p2 ) s+ (qi , p1 )
24
Again, you must realise that the choice of polarisation vector cannot change the final
result for the amplitude : but making the intermediate steps easy or even trivial can save
a lot of sweat and tears (if not blood).
August 31, 2019 365
Spinorial workout
It is easy to see that out of the twelve cases, eight vanish straight away25 :
An (±, ±) = A1 (−, +) = 0 = A2 (+, −) = 0 (n = 1, 2, 3) , (12.92)
The four remaining cases are readily evaluated in terms of spinor products26 :
s− (p2 , q2 )
A1 (+, −) = 2s+ (p1 , q1 )2 ,
s+ (p1 , q2 )
A3 (+, −) = −A1 (+, −) ,
s− (p2 , q1 )
A2 (−, +) = 2s+ (p1 , q2 )2 ,
s+ (p1 , q1 )
A3 (−, +) = A2 (−, +) . (12.93)
Perhaps surprisingly, the A3 ends up looking just like the A1,2 ! If we neglect E98
overall complex phases, we have
A1 (+, −) ∼ 2t t/u , A2 (−, +) ∼
p p
= 2u u/t , (12.94)
and so we find
p 2 1 j k 1 j k
M(+, −) ∼ 2g ~ t t/u − T T − [T , T ]
t s
2 p
2g ~
t/u u T j T k + t T k T j .
= (12.95)
s
25
Note that, since in this calculation, u(p1 )/i u(p2 ) is always zero, we here have
Y (q1 , 1 ; q2 , 2 , −q1 − q2 , α) ∼ (q1 − q2 )α (1 2 ).
26
In the A3 results, it may be useful to note that there are alternative forms : u+ (p1 )(/q1 −
q2 )u+ (p2 ) equals 2s+ (p1 , q1 )s− (q1 , p2 ) as well as −2s+ (p1 , q2 )s− (q2 , p2 ).
/
366 August 31, 2019
Here we have used −1/t − 1/s = u/ts. In exactly the same way we find
2g 2 ~ p
u/t u T j T k + t T k T j .
M(−, +) ∼ (12.96)
s
In these expressions, replacing T j,k by unity gives us the QED result for
e+ e− → γγ if we remember that u + t = −s.
E99 The colour sums are most easily handled using the Fierz identities (12.16) :
(N 2 − 1)2
Tr T j T k T k T j
= ,
4N
(N 2 − 1)
Tr T j T k T j T k = −
. (12.98)
4N
The final result is
g 4 ~2 (N 2 − 1) 2
2
(N − 1) 2N 2
2 2
|M| = (t + u ) − 2 . (12.99)
2N 3 ut s
The differential cross section is
dσ 1
2
= |M| : (12.100)
dΩ 128π 2 s
note the symmetry factor Fsymm = 1/2 for the 2-gluon final state !
where the quark and the antiquark, of mass m, have the same momentum, i.e.
they annihilate at rest. As before, we shall employ the modified polarisation
vectors
(q2 1 ) µ
η1µ = µ1 − q (12.101)
(q1 q2 ) 1
(and vice versa), and q µ = (q1µ − q2µ )/2. With the colour labels of gluon 1
and 2 denoted by j and k, respectively, the amplitude is again given by the
diagrams (12.86) but now it reads
ig 2 ~
M = − 2 2
v(p)/η1 (/q1 − p/) η/2 u(p) T j T k
(p − q1 ) − m
ig 2 ~
− 2 2
v(p)/η2 (/q2 − p/) η/1 u(p) T k T j
(p − q2 ) − m
ig 2 ~
+ v(p)γα u(p) Y (q1 , η1 ; q2 , η2 ; −2p, α) [T j , T k ]
(2p)2
ig 2 ~
= + 2 v(p)/η1 q/η/2 u(p) {T j , T k } + [T j , T k ]
4m
ig 2 ~
− 2 v(p)/η2 q/η/1 u(p) {T j , T k } − [T j , T k ]
4m
ig 2 ~
+ 2 v(p)/qu(p) (η1 η2 ) [T j , T k ] . (12.102)
2m
The first two lines corresponds to the ‘QED-like’ diagrams with quark ex-
change, and the third line derives from the s-channel diagram containing the
368 August 31, 2019
g2~
M=− (p, q, η1 , η2 ) {T j , Y k } v(p)γ 5 u(p) . (12.105)
m3
We now consider the total spin of the q q̄ system. If it has spin-1, the spin
vectors of the quark and antiquark must be the same, and then we find
1
v(p, s)γ 5 u(p, s) = v(p, s)γ 5 1 + γ 5 s/ u(p, s)
2
1
v(p, s) 1 − γ 5 s/ γ 5 u(p, s) = 0 .
= (12.106)
2
For opposite spin vectors, the amplitude is nonzero : |v(p, s)γ 5 u(p, −s)|2 =
4m2 . Thus, a quark-antiquark pair at relative rest in a spin-1 state cannot
decay into two gluons, and the Landau-Yang theorem appears to hold here
as well - at the tree level. We might turn QCD into QED by replacing the T
matrices by unity, and so find the proof for orthopositronium.
27
This last diagram is discussed in section 12.3.2.
August 31, 2019 369
and that the t- and u-channel diagrams do not pick up similar quark loops.
Since the number (and certainly the masses) of the other quarks are not fixed
by any principle there is, therefore, no possibility of a similar cancellation
holding at the one-loop level and beyond ; and we conclude that the Landau-
Yang theorem does not hold for gluons28 .
12.7 Exercises
Excercise 96 Prove the Antkaz !
Prove the result (12.73) by using the properties of the [jklmn] symbols.
and show that this is completely transverse to any external polarisation vec-
tor.
28
Note, however, that this violating would show up only at second loop order, since the
interference of the tree-level result with the one-loop result would of course still vanish.
Experimentally, therefore, the violation is very modest.
370 August 31, 2019
7. By choosing the gauge vectors cleverly, show that the amplitude van-
ishes if all n gluons have the same helicity.
8. By choosing the gauge vectors cleverly, show that the amplitude van-
ishes if n−1 gluons have the same helicity and one the opposite helicity.
9. Show that, when two gluons have helicity opposite from the other n − 2
gluons we cannot choose all gauge vectors such that the amplitude
vanishes.
These amplitudes, with 2 gluons having polarization opposite from the other
ones, are called MHV amplitudes, see also section 10.7.1.
August 31, 2019 371
We conider the case where all gluons have the same helicity.
1. Let us first concentrate on a single ‘blob’ where an off-shell gluon com-
ing from the q − q̄ line decays into n gluons:
µ
Electroweak theory
373
374 August 31, 2019
We can therefore derive the ‘energy scale’ of the interaction responsible for
muon decay1 : s
~c2
ΛW = ≈ 292.5 GeV . (13.6)
GF
where θ is the angle between the muon and electron momenta. By taking
also the angular average we obtain
16
|M|2 GF 2 ~2 s2 .
= (13.11)
3
The total cross section is therefore given by
− − GF 2 ~2
σ(µ ν µ → e ν e ) = s (13.12)
3π
As we have seen before, only the factor 1/3 is not immediately obvious in
this expression but has to be computed from the Feynman diagrams.
The scattering cross section rises linearly with s, and will therefore violate
the unitarity bound at sufficiently high energy. Since the the muon and its
antineutrino couple with a single γ µ , they must be in a J = 1 state. The
unitarity bound on this cross section (cf appendix 15.10) is therefore
1 16π 24π
σ(µ− ν µ → e− ν e ) ≤ (2J + 1) = , (13.13)
2 s s
which leads to a fundamental failure
√ of the Fermi model (at least, at the tree
level) at a scattering energy of s ≈ 1.5 TeV.
with f (x) decreasing at least as fast as 1/x for large arguments, and Λ
some constant energy. Such energy-dependent couplings, called form factors,
376 August 31, 2019
The more elegant, and (as it turns out) the correct way to go is to make
the Fermi model look more ‘QED-like’: instead of using a contact interaction
between four fermions, we postulate the existence of a new particle, the so-
called W boson. This couples to fermion-antifermion pairs in a way reminis-
cent of the photon. The four-fermion interaction then resolves into two f f¯W
interactions, with the W boson mediating between the two vertices ; the cor-
responding Feynman diagram for the process µ− (p) to e− (q) νµ (k1 ) ν e (k2 )
is therefore given by
p k1
M = Q (13.15)
q
k2
At the time this model was first seriously discussed, it went under the name
of Intermediate Vector-Boson (IVB) hypothesis. We take the W to couple
to the fermion pairs eνe and µνµ , so that (as we shall check!) the W must be
electrically charged, and assume that the coupling is in both cases of equal
strength4 (for now). We therefore postulate the following Feynman rules :
3
In the present case, GF , being dimensionful, sets such a scale by itself.
4
At this point, these are of course just assumptions. Since 1983, when the W boson was
first freely produced, they have been tested with great accuracy. The alternative scenario
of the charge retention form, with an electrically neutral W , is completely ruled out by
the fact that the decay W → e+ µ− is never seen. The equality of the couplings is verified
by the fact that the branching ratios for W → eν e and W → µν µ are the same up to
computable mass effects.
August 31, 2019 377
i
gW 1 + γ 5 γ µ f f 0 W vertices
↔
µ ~
However, from Eq.(13.22) we see that for such large values the dimensionless
coupling constant is so large that the tree-level approximation for the cross
section is questionable.
5 −7
In fact, for the actual values of the
√ masses the suppression factor is about 10 .
6
This value is close to the value of s at which unitairy breaks down in the unmodified
Fermi model, see Eq.(13.13). This is not a coincidence. Whatever we do to the electroweak
interactions, 1.5 TeV appears to be the energy régime where things get tricky.
August 31, 2019 379
2~2 gW 4 s
σ(µ− ν µ → e− ν e ) = . (13.25)
3π (s − mW ) + mW 2 ΓW 2
2 2
This is well below the unitarity limit. The IVB hypothesis therefore indeed E103
cures the unitarity problem in this process.
E104
Because of these successes, we shall adopt the notion of an existing W
particle of spin 1 (and hence obeying the lines laid out in chapter 9), coupling
to pairs of fermions separated by one unit of charge7 .
M = + +
q2 k2 q2 k1 q2 k1
(13.27)
The three diagrams correspond to the three partial matrix elements
k/1 − q/1 + mD
M1 = −i~gW QD v(q1 )/ (1 + γ 5 ) /W u(q2 ) ,
(k1 − q1 )2 − mD 2
q/2 − k/1 + mU
M2 = −i~gW QU v(q1 ) (1 + γ 5 ) /W / u(q2 ) ,
(q2 − k1 )2 − mU 2
M3 = +i~gW QW v(q1 ) (1 + γ 5 ) γα u(q2 )
g αβ − P α P β /mW 2
W β (2k2 + k1 · ) , (13.28)
s − mW 2
where s = P 2 , P = q1 + q2 = k1 + k2 .
Since this process involves a produced photon, the handlebar identity
must hold : if we replace µ by k1 µ the amplitude must vanish. We shall
investigate this in some detail, by explicit computation rather than by the
diagrammatic manipulations of the previous chapters. In the first place, we
perform some simple Dirac algebra to note that
v(q1 )/(/k1 − q/1 + mD ) →k1 = v(q1 )/k1 (/k1 − q/1 + mD )
= v(q1 ) k1 2 − 2(q1 · k1 ) + (/q1 + mD )/k1
= (k1 − q1 )2 − mD 2 v(q1 ) ,
(13.29)
where in the second line we have used anticommutation between k/1 and q/1 ,
and in the third line the Dirac equation for v(q1 ). We see that
M1 c→k1 = −i~gW QD v(q1 ) (1 + γ 5 ) /W u(q2 ) , (13.30)
August 31, 2019 381
and similarly
If we were allowed to consider only the first of the two terms of the result
(13.32), we could obtain the desired cancellation :
3
%
X
Mj = 0 ⇒ QW = QD − QU : (13.33)
j=1 →k1
but the second term in Eq.(13.32) spoils this idea by having a quite different
algebraic structure ; no tuning of coupling constants is going to ensure that
a W W γ vertex of the form (13.26) can do the job.
+(a5 k1 + a6 k2 )β (+ · )
= δ α β − P α Pβ /mW 2
+(a5 k1 + a6 k2 )β (+ · ) .
(13.38)
The replacement → k1 now leads, after some simple algebra (and use of
momentum conservation !) to the form
Z α c→k1 = (a1 − a2 )(k1 · k2 )+ α + T α ,
T α = (−a3 + a4 + a5 − a6 )k1 α
(k1 · k2 )
− (a1 − a2 − a3 + a4 + a5 + a6 )P α . (13.39)
mW 2
August 31, 2019 383
+ mU 2 − mD 2 v(q1 ) (1 + γ 5 ) /u(q2 ) .
(13.43)
Of these three lines, the second is suppressed with respect to the first one by
a factor (mass/energy), and the third line even by (mass/energy)2 . In the
high-energy limit, therefore, the second and third line will not contribute to
any unwanted high-energy behaviour of the amplitude : we shall call such
terms safe terms10 . We can therefore write
M1 c+ →k2 = +i~gW QD v(q1 ) (1 + γ 5 ) /u(q2 ) + · · · ,
(13.44)
9
Any common factor in the a’s is always absorbed in the value of QW so this is no loss
of generality.
10
Which is not to say that they are negligible ! The point here is that they do not
contribute to any condition on the coupling constants. At the end of a cross section
calculation it is the safe terms that we want !
384 August 31, 2019
where the ellipsis denotes safe terms. For the second diagram, we find in a
similar way :
M2 c+ →k2 = −i~gW QU v(q1 ) (1 + γ 5 ) /u(q2 ) + · · · ,
(13.45)
For the third graph we find, after some algebra,
Z α c+ →k2 = (a4 − a3 )(k1 · k2 )α − a3 mW 2 α
(k2 · )(k1 · k2 )
− (a1 − a2 − a3 + a4 + a5 + a6 )P α
mW 2
+ (k2 · )(−a1 + a2 + a5 − a6 )k1 α . (13.46)
Requiring M3 to cancel against M1 + M2 up to safe terms therefore leads
to yet more relations between the a’s :
a3 − a4 = 2 , a1 − a 2 = a 5 − a 6 . (13.47)
Combining the requirements (13.37), (13.40), (13.41) and (13.47) we find the
unique solution
a1 = a3 = a5 = 1 , a2 = a4 = a6 = −1 . (13.48)
We find that the spacetime part of the vertex has, after all, exactly the same
form as that of the three-gluon vertex, Eq.(12.33) ! We have thus established
the W W γ vertex to be
µ
+
W (p1 ) γ (p3 ) i
↔ QW Y (p1 , µ; p2 , ν; p3 , ρ)
_ ρ ~
W (p2 )
ν
All particles and momenta counted outgoing
EW Feynman rules, part 13.2 (13.49)
We shall now pursue the same strategy for different processes. Since we
shall be interested in the high-energy behaviour of amplitudes we shall allow
ourselves to neglect particle masses wherever possible.
Let us consider the process
U (p1 ) U (p2 ) → W + (q+ , + ) W − (q− , − )
With the vertices available so far, we have the following two Feynman dia-
grams
_ _
U _ U +
W W
D
γ _
U W+ U W
which evaluates to
~ gWWZ 5
M3 = i v(p 1 ) vU + a U γ γµ u(p2 )
(q+ + q− )2 − mZ 2
Y (q+ , + ; q− , − ; −q+ − q− , µ) . (13.55)
Note that nothing has been neglected in this expression ; the second term in
the massive-boson propagator drops out when we multiply it into the Yang-
Mills vertex. Since this diagram is so similar to M2 it is easy to perform the
handlebar operation :
M3 c+ →q+ ≈ i~ gWWZ v(p1 ) vU + aU γ 5 /− u(p2 ) ,
(13.56)
388 August 31, 2019
0 = vU gWWZ + 2gW 2 + QU QW ,
0 = aU gWWZ + 2gW 2 . (13.57)
0 = vD gWWZ − 2gW 2 + QD QW ,
0 = aD gWWZ − 2gW 2 . (13.58)
In the following we shall use the notation sW = sin θW and cW = cos θW . This
angle is called the weak mixing angle13 , and it parametrizes essentially all of
the minimal model of electroweak interactions we are constructing here. In
13
The subscript W was originally used to refer to S. Weinberg, one of the early propo-
nents of the electroweak model, but nowadays it appears more fair to take it to mean just
‘weak’.
August 31, 2019 389
the first place, we know that the charge of the W must be equal to the charge
of the electron (since neutrinos are neutral) and therefore we might prefer to
write
QW
gW = √ (13.63)
8 sW
which leads to a parametrization of the W mass itself14 :
πα 1
(~ c mW )2 = √ 2
GeV2 , (13.64)
2 1.16 10 sW
−5
or
37.3
~ c mW = GeV . (13.65)
sW
As we see, the assumption of the existence of a single, neutral Z boson
immediately implies that the W has a mass of at least 37.3 GeV. Notice that
no prediction for the mass of the Z is obtained, however.
The other unknowns in our treatment can now be expressed in terms of
θW . Adopting the usual convention of denoting by e the positive unit charge,
we find by straightforward algebra
cW
QW = −e , gWWZ = −e ,
sW
e
aU = −aD = ,
4sW cW
QU
vU = aU 1 − 4sW 2 ,
e
2 QD
vD = aD 1 + 4sW . (13.66)
e
We note here that θW is defined at this stage as a relation between coupling E105
constants ; later on we shall encounter it in another guise !
much quarried all possible information15 about this sector, we now turn to
the 2 → 2 processes involving four bosons. First we consider the process
W + (p1 , 1 ) γ(p2 , 2 ) → W + (p3 , 3 ) γ(p4 , 4 )
With the available vertices we have two Feynman diagrams for this process :
+
W+
γ W+ W
M = + γ (13.67)
γ W+ γ
× Y (p1 − p4 , µ; −p1 , 1 ; p4 , 4 )
~QW 2
= −i − mW 2 (2 · 3 )(1 · 4 )
2(p2 · p3 )
+Y (p3 , 3 ; p2 − p3 , µ; −p2 , 2 )Y (p1 − p4 , µ; −p1 , 1 ; p4 , 4 ) ,
~QW 2
M2 = i Y (p3 , 3 ; −p3 − p4 , ν; p4 , 4 )
(p3 + p4 )2 − mW 2
× g µν + (p1 + p2 )µ (−p3 − p4 )ν /mW 2
so that
We might also have chosen choose to put the handlebar on 2 instead ; the
result would then have been
2
%
X
Mj =
j=1 2 →p2
2
= i~QW (2(1 · 3 )(p2 · 4 ) − (1 · p2 )(3 · 4 ) − (p2 · 3 )(1 · 4 )) .
(13.73)
Going to the limit of large energies, we can also envisage putting a handlebar
on 1 or 3 . Neglecting safe terms leads to
2
%
X
Mj =
j=1 1 →p1
2
= i~QW (2(p1 · 3 )(2 · 4 ) − (p1 · 2 )(3 · 4 ) − (2 · 3 )(p1 · 4 )) ,
(13.74)
and
2
%
X
Mj =
j=1 3 →p3
2
= i~QW (2(1 · p3 )(2 · 4 ) − (1 · 2 )(p3 · 4 ) − (2 · p3 )(1 · 4 )) .
(13.75)
392 August 31, 2019
We can repair the high-energy behaviour of the amplitude, for all these cases
at once, by introducing a four-boson vertex :
µ ν _
W+ W i
↔ − QW 2 X µναβ (13.76)
γ ~
γ
α β
where
X µναβ = 2 g µν g αβ − g µα g νβ − g µβ g να . (13.77)
The occurrence of such a four-point vertex should not surprise us, with our
experience of a similar vertex in sQED and QCD. Its precise algebraic struc-
ture can, of course, not be inferred from that example16 .
W+ W+ W+ Z,γ W+
M = Ζ,γ + . (13.79)
_ _ _ _
W W W W
16
Except, perhaps, the idea that it contains only the metric tensor, and not any of the
momenta : but momenta would only worsen the high-energy behaviour
August 31, 2019 393
It will turn out to be useful to take the γ and Z exchanges together so that
we have two contributions :
M1 = i~QW 2 Y (p3 , 3 , −p1 , 1 , p1 − p3 , µ)
g µν cW 2 g µν − (p1 − p3 )µ (p1 − p3 )ν /mZ 2
+
(p1 − p3 )2 sW 2 (p1 − p3 )2 − mZ 2
Y (−p2 , 2 , p4 , 4 , p2 − p4 , ν) ,
M2 = i~QW 2 Y (−p2 , 2 , −p1 , 1 , p1 + p2 , µ)
g µν cW 2 g µν − (p1 + p2 )µ (p1 + p2 )ν /mZ 2
+
(p1 − p3 )2 sW 2 (p1 + p2 )2 − mZ 2
Y (p3 , 3 , p4 , 4 , p2 − p4 , ν) . (13.80)
Because the masses of the external particles are all equal, the second term in
the Z propagator can be seen to drop out exactly. We can therefore afford
to take the limit s mZ 2 without more ado, and combine the γ and Z
propagators to arrive at the following high-energy form of the contributions :
~QW 2 1
M1 = i Y (p3 , 3 , −p1 , 1 , p1 − p3 , µ)
sW (p1 − p3 )2
2
Y (−p2 , 2 , p4 , 4 , p2 − p4 , µ) ,
~QW 2 1
M2 = i Y (−p2 , 2 , −p1 , 1 , p1 + p2 , µ)
2
sW (p1 + p2 )2
Y (p3 , 3 , p4 , 4 , p2 − p4 , µ) . (13.81)
Let us now take the outgoing W − longitudinal, i.e apply the handlebar on
4 , and drop safe terms :
Y (p3 , 3 , −p1 , 1 , p1 − p3 , µ)Y (−p2 , 2 , p4 , 4 , p2 − p4 , µ)c4 →p4
= Y (p3 , 3 , −p1 , 1 , p1 − p3 , µ)
× (p1 − p3 )µ ((p1 − p3 ) · 2 ) − ((p1 − p3 )2 − mW 2 )2 µ
M1 + M2 c4 →p4 =
~QW 2
−i (2(p4 · 1 )(2 · 3 ) − (p4 · 2 )(1 · 3 ) − (p4 · 3 )(1 · 2 )) :
sW 2
(13.85)
is, the external vector particles are assumed to be purely longitudinal in the
centre-of-mass frame of the scattering17 . The longitudinal polarisation of an
on-shell vector particle with momentum pµ and mass m is then given by
m2 µ m2
NL
µ
L = µ
p − c , NL −2 = 1 − , (13.87)
m c·p (c · p)2
which expression is well-defined as long as p~ 6= 0. We see that, as before,
L = p/m + O (m/p0 ). In the cases studied so far, the subleading terms in
L have only led to safe terms so that they could be neglected18 ; now, this
is no longer automatically the case.
Before continuing we may note that another possible choice for longitu-
dinal polarisation is to take a lightlike vector tµ and define
1 m µ
µL = pµ − t . (13.88)
m p·t
In many applications this is useful. However, the boson is purely longitudinal
only in a frame where ~t k p~. In scattering processes, not all bosons can then
be strictly longitudinal at the same time. In this section we therefore opt for
cµ rather than tµ
W W → ZZ
The first Gedanken process19 is
W + (p1 , 1 ) W − (p2 , 2 ) → Z 0 (p3 , 3 )Z 0 (p4 , 4 )
So far, we have the following three Feynman graphs available at the tree
level :
1 W Z 3 1 W Z 4
1 W Z 4
M = W + W + , (13.89)
Z Z W
2 W 4 2 W 3 2 Z 3
17
That this is not a trivial point becomes clear when we realise that in ‘W W scattering’
at the LHC, say, the centre-of-mass frame of the scattering does not coincide with the
laboratory frame, in which the detector is at rest, and in which the polarisation analysis
of the produced bosons is presumably performed.
18
From the point of view of restoring unitarity, not that of actually getting the cross
section right!
19
As usual, with improving technology and the commissioning of machines like the LHC,
Gedanken processes are gradually being turned into actual ones.
396 August 31, 2019
Nj
Mj = −i~gWWZ 2 , j = 1, 2 ,
∆j
N1 = Y (p1 − p3 , µ; −p1 , 1 ; p3 , 3 )
µν 1 µ ν
× −g + (p1 − p3 ) (p1 − p3 )
mW 2
× Y (−p2 , 2 ; p2 − p4 , ν; p4 , 4 ) ,
∆1 = (p1 − p3 )2 − mW 2 = mZ 2 − 2(p1 · p3 ) ,
N2 = Y (p1 − p4 , µ; −p1 , 1 ; p4 , 4 )
µν 1 µ ν
× −g + (p1 − p4 ) (p1 − p4 )
mW 2
× Y (−p2 , 2 ; p2 − p3 , ν; p3 , 3 ) ,
∆2 = (p1 − p4 )2 − mW 2 = mZ 2 − 2(p1 · p4 ) ,
M3 = −i~gWWZ 2 N3 ,
N3 = X(1 , 2 , 3 , 4 ) . (13.90)
Owing to the work we have done so far, we may already anticipate some
cancellations between the diagrams when we make all bosons longitudinal
and the safe terms are therefore not the subleading ones, but rather the sub-
subleading ones. We have to proceed carefully20 . Denoting by the subscript
L the ‘fully longitudinal’ case, it appears best to write the result as
3
%
X N123
Mj = −i~gWWZ 2 ,
j=1
∆ 12
L
mZ 2
N123 = N1 ∆2 + N2 ∆1 + ∆12 N3 = −4E 6 (sin θ)2 + · · · ,
mW 4
∆12 = ∆1 ∆2 = 4E 4 (sin θ)2 + · · · , (13.91)
can never play a rôle in any dynamical cancellation, and their subleading
terms are therefore always safe. The non-safe contribution from our three
Feynman graphs is therefore
3
%
X mZ 2
Mj = i~ gWWZ 2 E 2 4
+ ··· , (13.92)
j=1
mW
L
Before we proceed to the next Gedanken process, a few remarks are in order.
In the first place, the choice for a scalar Higgs particle is almost unavoidable.
It certainly cannot be a fermion ; if it were a vector particle, its propagator
would contain unwanted higher powers of the energy E, the W W H and
ZZH would presumably be of Yang-Mills type hence also E-dependent. The
vertices given above are essentially the only ones possible for the interactions
between two vectors and a scalar if we want them to be energy-independent.
Note that gWWH and gZZH may √ both be expected to contain a mass, that is,
they are of dimension L−1 / ~. The assumption that there is just one type
of neutral scalar involved is, of course, based on nothing but a prejudice in
favour of simplicity. Finally, at high energy all contributions from mH end up
in safe terms, and we do not expect to glean any information on the Higgs
mass from our considerations.
W W → W W scattering
Another four-boson scattering process of interest is
W + (p1 , 1 )W + (p2 , 2 ) → W + (p3 , 3 )W + (p4 , 4 )
for which we have five purely vector-boson diagrams :
1 1
3 4 1 3
M = γ,Ζ + γ,Ζ + (13.97)
4 2 3 4
2 2
whose contributions can be conviently written as
M1 = −i~ Y (p3 , 3 ; −p1 , 1 ; p1 − p3 , µ)
µν
+ (p1 − p3 )µ (p1 − p3 )ν /mW 2
2 −gµν 2 −g
× QW + gWWZ
(p1 − p3 )2 (p1 − p3 )2 − mZ 2
× Y (p4 , 4 ; −p2 , 2 ; p2 − p4 , ν) ,
M2 = M1 cp3 ,3 ↔ p4 ,4 ,
QW 2
M3 = i~ X(3 , 4 , 1 , 2 ) . (13.98)
sW 2
By the same methods as used in the previous section we arrive at
3
%
X ~ E 2 QW 2
−4mW 2 + 3mZ 2 cW 2 + · · ·
Mj =i 4 2
(13.99)
j=1
mW sW
L
August 31, 2019 399
1 1
3 4
M4,5 = H , H (13.100)
4 3
2 2
so that %
5
X gWWH 2
Mj = i~ E 2 + ··· (13.102)
j=4
mW 4
L
In this process, then, good high-energy behaviour is obtained under the con-
dition
QW 2
gWWH 2 = 4mW 2 − 3mZ 2 cW 2 .
2
(13.103)
sW
Again, no restrictions on mH occur.
HZ → W W scattering
We have now run out of four-vector Gedanken processes. ZZ → ZZ scat-
tering has no Yang-Mills contributions21 , and any four-vector process in-
volving photons will have vanishing amplitudes under a handlebar on any
photon. However, in the same spirit by which we boldly proposed the pro-
cess U D → W Z as soon as the Z was hypothesised, we can consider the
process
H(p1 )Z 0 (p2 , 2 ) → W + (p3 , 3 )W − (p4 , 4 )
Since only three out of four particles can become longitudinal here, the uni-
tarity violations are not so bad, and the safe terms are of sub- rather than
21
Under the Higgs hypothesis ZZ → ZZ scattering is described by three diagrams each
containing Higgs exchange : no Higgs means no scattering at all ! The amplitude is safe
by itself and hence does not lead to any constraints.
400 August 31, 2019
that contribute as
M1 = −i~ gWWZ gWWH Y (p3 , 3 ; p2 − p3 , µ; −p2 , 2 )
−g µν + (p2 − p3 )µ (p2 − p3 )ν
× (4 )ν ,
(p2 − p3 )2 − mW 2
M2 = −i~ gWWZ gWWH Y (p2 − p4 , µ; p4 , 4 ; −p2 , 2 )
−g µν + (p2 − p4 )µ (p2 − p4 )ν
× (3 )ν ,
(p2 − p4 )2 − mW 2
M3 = −i~ gWWZ gZZH Y (p3 , 3 ; p4 , 4 ; −p3 − p4 , µ)
−g µν + (p1 + p2 )µ (p1 + p2 )ν
× (2 )ν . (13.105)
(p1 + p2 )2 − mZ 2
The kinematics of this process is a little different from that of the two previous
ones, since mH and mZ cannot be assumed to be equal. Still, at high energy
we may apply massless kinematics since we only have to cancel the leading
non-safe terms. Neglecting, therefore, mW , mZ and mH in the kinematics22
we find
3
%
X
2 mZ 1
Mj = i~ E cos θ gWWZ gWWH − gZZH +· · · (13.106)
j=1
mW 4 mZ mW 2
L
Q W mW QW mZ
gWWH = , gZZH = , (13.108)
sW s W cW
mW = mZ cW . (13.109)
It is apposite to dwell on this last result. The weak mixing angle θW was E106
introduced to parametrize the system of coupling constants, as discussed in
section 13.3.2 : we now see it come back here as a relation between masses
instead ! From the treatment of the Electroweak Standard Model presented
in these notes, it also becomes clear that the mixing angle as a description
of coupling constants is, in a logical sense, prior to that as a description of
masses. The assumption of a single Z 0 particle determines the couplings as
described in section 13.3.2 : but it takes the supposition of a single, neutral
Higgs particle to obtain Eq.(13.109). If the Higgs sector of the Standard
Model turns out to be different, with more Higgs-like particles, say, the W
and Z mass become uncorrelated ; but the couplings of W and Z with the
fermions and each other remain unaffected. In the usual textbook deriva-
tion of the model this distinction tends to be obscured by the simultaneous
obtention of all couplings at once after symmetry breaking.
As a final comment we remark that, if unitarity is restored by any Higgs-
like phenomenon whatsoever, the weak mixing angle must always obey the
bound
4 mW 2
cW 2 < (13.110)
3 mZ 2
as can be seen from Eq.(13.103)23 .
with
gWWH 2 QW 2
gWWHH = = . (13.118)
2mW 2 2sW 2
M4 = (13.128)
24
In fact, the observation that the nonsafe part in this process is proportional to vu is
the strongest argument in favour of a scalar Higgs.
August 31, 2019 405
f H i e mf
↔ 1 (13.133)
f ~ 2sW mW
M1,2,3 = , , (13.134)
25
Note that for D-type fermions, aD has opposite sign ; but also the W + and W − are
interchanged in the first diagram.
406 August 31, 2019
where as usual the dotted lines denotes Z’s and the solid lines stand for H
particles, and we have to take into account the appropriate permutations of
the external Z particles. The amplitude is given by the three corresponding
contributions :
M1 = A1 (1, 2, 3, 4, 5) + A1 (2, 1, 3, 4, 5) + A1 (3, 4, 1, 2, 5)
+ A1 (4, 3, 1, 2, 5) + A1 (1, 3, 2, 4, 5) + A1 (3, 1, 2, 4, 5)
+ A1 (2, 4, 1, 3, 5) + A1 (4, 2, 1, 3, 5) + A1 (1, 4, 3, 2, 5)
+ A1 (4, 1, 3, 2, 5) + A1 (3, 2, 1, 4, 5) + A1 (2, 3, 1, 4, 5) ,
M2 = A2 (1, 2, 3, 4, 5) + A2 (3, 4, 1, 2, 5) + A2 (1, 3, 2, 4, 5)
+ A2 (2, 4, 1, 3, 5) + A2 (1, 4, 3, 2, 5) + A2 (3, 2, 1, 4, 5) .
M3 = A3 (1, 2, 3, 4, 5) + A3 (1, 3, 2, 4, 5) + A3 (1, 4, 2, 3, 5) ,
A1 (i1 , i2 , i3 , i4 , i5 ) = i~3/2 gZZH 3 i1 µ Πµν (pi1 + pi5 ) i2 ν (i3 · i4 )
× ∆Z (pi1 + pi3 ) ∆H (pi3 + pi4 ) ,
A2 (i1 , i2 , i3 , i4 , i5 ) = −i~3/2 gZZH gZZHH (i1 · i2 )(i3 · i4 )
× ∆H (pi3 + pi4 ) ,
A3 (i1 , i2 , i3 , i4 , i5 ) = i~3/2 gZZH 2 gHHH (i1 · i2 )(i3 · i4 )
× ∆H (pi1 + pi2 ) ∆H (pi3 + pi4 ) ,
1
Πµν (q) = −gµν + qµ qν ,
mZ 2
−1 −1
∆Z (q) = q 2 − mZ 2 , ∆H (q) = q 2 − mH 2 . (13.135)
Here we have, for once, taken all momenta outgoing, which means that the
momenta of the incoming Z’s have negative zeroth component. For this 5-
particle process the phase space is of course more complicated, and here we
demonstrate a numerical method to investigate cancellations. This can be
quite sensitive if done right26 . Although naı̈vely each diagram A1 and A2
grow quadratically with the energy in the fully longitudinal case, both M1
and M2 actualy become energy-independent at sufficiently high energy E.
But this is not safe : a 2 → 3 amplitude must go at most as E −1 , and
therefore cancellations between (M1 + M2 ) and M3 are still necessary. We
find that the required HHH coupling is given by
gHHH 3 QW mH 2
↔ i , gHHH = − (13.136)
~ 2 mW sW
26
An algorithm for sampling the relevant phase space is given in appendix 15.13
August 31, 2019 407
M1,...,6 = , , , , , .
(13.137)
where we have already anticipated a quartic Higgs coupling in the last dia-
gram. The contributions to the amplitude are
M1 = B1 (1, 2, 3, 4, 5) + B1 (1, 2, 4, 5, 3) + B1 (1, 2, 5, 3, 4)
+ B1 (1, 2, 5, 4, 3) + B1 (1, 2, 3, 5, 4) + B1 (1, 2, 4, 3, 5) ,
M2 = B2 (1, 2, 3, 4, 5) + B2 (1, 2, 4, 5, 3) + B2 (1, 2, 5, 3, 4)
+ B2 (2, 1, 3, 4, 5) + B2 (2, 1, 4, 5, 3) + B2 (2, 1, 5, 3, 4) ,
M3 = B3 (1, 2, 3, 4, 5) + B3 (1, 2, 4, 5, 3) + B3 (1, 2, 5, 3, 4) ,
M4 = B4 (1, 2, 3, 4, 5) + B4 (1, 2, 4, 5, 3) + B4 (1, 2, 5, 3, 4)
+ B4 (2, 1, 3, 4, 5) + B4 (2, 1, 4, 5, 3) + B4 (2, 1, 5, 3, 4) ,
M5 = B5 (1, 2, 3, 4, 5) + B5 (1, 2, 4, 5, 3) + B5 (1, 2, 5, 3, 4) ,
M6 = −i~3/2 gZZH gHHHH (1 · 2 ) ∆H (p1 + p2 ) ,
B1 (i1 , i2 , i3 , i4 , i5 ) = i~3/2 gZZH 3 1 µ Πµ λ (pi1 + pi3 ) Πλν (pi2 + pi5 ) 2 ν
× ∆Z (pi1 + pi3 ) ∆Z (pi2 + pi5 ) ,
B2 (i1 , i2 , i3 , i4 , i5 ) = i~3/2 gZZH 2 gHHH i1 µ Πµν (pi2 + pi5 ) i2 ν
× ∆Z (pi2 + pi5 ) ∆H (pi3 + pi4 ) ,
B3 (i1 , i2 , i3 , i4 , i5 ) = i~3/2 gZZH gHHH 2 (i1 · i2 ) ∆H (p1 + p2 ) ∆H (pi4 + pi5 ) ,
B4 (i1 , i2 , i3 , i4 , i5 ) = −i~3/2 gZZHH gZZH i1 µ Πµν (pi2 + pi5 ) i2 ν
× ∆Z (pi2 + pi5 ) ,
B5 (i1 , i2 , i3 , i4 , i5 ) = −i~3/2 gZZHH gHHH (i1 · i2 ) ∆H (pi3 + pi4 ) , . (13.138)
A treatment analogous to that of the previous paragraph leads to the follow-
ing, final Feynman rule :
i 3 QW 2 mH 2
↔ gHHHH , gHHHH = − (13.139)
~ 4 mW 2 sW 2
In figure 13.2 we plot the ratio − M6 cL / M1+···+5 cL obtained in the same
manner as in the previous paragraph. Again, the choice of the factor 3/4
in gHHHH is justified by the fact that the ratio goes to 1 with great accuracy
as the total energy w increases. At high energy there are no outliers : M6
depends only on w.
August 31, 2019 409
V1 V3
V2
γ, Z, and Z a γZZ coupling in the bare action implies that the electrically
neutral Z’s are, actually, not neutral at all. In that case, what does electric
charge mean ?
If all three fermion-fermion-vector couplings are of vector type, Furry’s
theorem (cf section 10.2.7) shows how the two diagrams have opposite sign,
all other things being equal. Therefore, in QED, where the three vector par-
ticles are photons, (V1 , V2 , V3 ) = (γ, γ, γ), such diagram pairs will completely
cancel. In QCD, with three gluons, (V1 , V2 , V3 ) = (g, g, g), the two diagrams
form a colour-antisymmetric contribution to the one-loop correction of the
three-gluon vertex. Two mixed cases are (V1 , V2 , V3 ) = (γ, γ, g), where the
diagrams vanish individually because Tr (T j ) = 0 for gluonic colour matrices
T j , and (V1 , V2, V3 ) = (γ, g, g)
where they again cancel against one another
since Tr T j T k = Tr T k T j . We shall now investigate how things stand
if we include the electroweak sector. In particular we have to worry about
diagrams containing one or three γ 5 ’s, since in such cases the two diagrams
with oppositely-oriented fermion loops have the same sign. Since we shall be
interested only in the possible ultraviolet behaviour of the theory, we may
assume that all fermions have the same (negligible) mass. Also, we consider
the contribution of a single fermion family, that is a charged lepton (with
charge QL ), a neutrino (with charge Qν , which we alllow to be nonzero for
the moment), and an up-type quark (with charge QU ) and a down-type quark
August 31, 2019 411
Nc QU 2 − QD 2 = QL 2 − Qν 2 ,
Nc QU + QD = − QL − Qν . (13.140)
QU − QD = Qν − QL (= −QW ) , (13.141)
with P = p1 + p2 = p3 + p4 . Using
and
Y (p3 , + , −P, −P ; p4 , ) = (mW 2 − mZ 2 )(+ · ) (13.146)
we can rewrite M3 as
~ gW gWWZ
M3 = i v(p2 )(1 + γ 5 )γα u(p1 ) Y (p3 , + ; −P, α; p4 , )
s − mW 2
mZ 2
5 5
+ 1− (+ · ) mU v(p2 )(1 + γ )u(p1 ) − mD v(p2 )(1 − γ )u(p1 ) .
mW 2
(13.147)
In the fully longitudinal case all three diagrams actually grow as E 2 at most,
at high energy E. From this point on, therefore, we take the bosons to be
massless since their masses can only lead to safe terms. The fully longitu-
dinal expressions, in which we systematically ignore safe terms, can then be
evaluated as follows.
~gW
M1 c L = i v(p2 ) Σ u(p1 ) ,
2(p2 · p4 )mW mZ
Σ = (vD + aD γ 5 )/p4 (−/p2 + mD )(1 + γ 5 )/q3
= −2(p2 · p4 )(vD + aD γ 5 )(1 + γ 5 )/p3
+(vD + aD γ 5 )(/p2 + mD )/p4 p/3 (1 − γ 5 )
= −2(p2 · p4 )(vD + aD )(1 + γ 5 )/p3
+2mD aD aD γ 5 p/4 p/3 (1 − γ 5 )
= −2(p2 · p4 )(vD + aD )(1 + γ 5 )/p3
−4(p2 · p4 )mD aD aD (1 − γ 5 ) , (13.148)
where we have used the fact that, up to safe terms, we can write p/4 p/3 =
p/4 (/p3 + p/4 − p/1 ) = p/4 p/2 ; so that for the first diagram we have
~gW
M1 c L = i − (vD + aD ) v(p2 )(1 + γ 5 )/p3 u(p1 )
mW mZ
5
− 2aD mD v(p2 )(1 − γ )u(p1 ) . (13.149)
414 August 31, 2019
These diagrams each contain a term that goes as E 2 , and these contributions
cancel one another under the condition
vU + aU − vD − aD + gWWZ = 0 (13.152)
mZ 2 mZ 2 mZ 2 cW 2
e
gWWZ + 4aU = gWWZ − 4aD = 1− , (13.154)
mW 2 mW 2 cW s W mW 2
so that unitarity is respected provided that mW = cW mZ . What if the observed
W and Z masses have a different ratio? In the first place, we must then
27
In the computation of this diagram, it is important to notice that the unsafe terms
going as E 1 come from terms in the brackets, and not just from the second term.
August 31, 2019 415
conclude that there moust be more than just one single type of neutral
Higgs boson, for it was this assumption that gave us the relation 13.109, and
it would now be seen to be incorrect. In the second place, there must be
a Higgs-like object that compensates the unsafe behaviour in U D → W + Z,
and it must be charged: a charged Higgs particle! We would need to introduce
couplings of this object to fermions and vector bosons as follows:
D
H+ i
↔ gUDH (mU ω+ − mD ω− )
~
U
+
ν W
_ i
H ↔ gWZH g µν
µ Z
~
At this point we shall assume that there is just a single type of charged
Higgs H ± in the model, but we shall allow for more than one type of neutral
Higgs, denoted therefore by Hj0 , where j can run over several values.
Let us now find some more couplings. In the first place, since the H ± is
a charged scalar, we know immediately what the structure of its couplings
to photons must be:
q i
µ ↔ Qc (p − q)µ ,
p ~
where the momenta are counted along the arrows, and of course
µ
i
↔ (2Qc 2 ) g µν
ν ~
As discussed in section 10.4. However, we must also allow for the possibility
of charged bosons coupling to H 0 H ± pairs. Simple analogy then motivates
the choices of two additional vertices:
q i
µ ↔ gWcj (p − q)µ ,
p ~
µ
i
↔ gWcjγ g µν .
~
ν
Here the momenta are counted along the arrow as before. The coupling
constants can be restricted by considering the process
for which we now have three Feynman diagrams at the tree level:
August 31, 2019 417
The handlebar operation on the photon, which tests for current consaerva-
tion, gives
M123 c4 →p4 = i~ (p1 · 3 ) QW gWcj + gWcjγ
+ (p2 · ) − QW gWcj + gWcjγ − 2gWcj Qc ,(13.159)
M2 = i~gWWcc (3 · 4 ) ,
QW Qc gWWZ gZcc
M3 = i~ + Y (p3 , 3 ; p4 , 4 ; −P, p1 − p2 ) (13.161)
.
s s − mZ 2
Putting a handlebar on 3 results in
M123 c3 →p3 = i~ (p1 · 4 ) gWWcc + QW Qc + gWWZ gZcc
" #)
X
+(p2 · 4 ) − 2 gWcj 2 + gWWcc − QW Qc − gWWZ gZcc (.13.162)
j
13.8 Exercises
Excercise 103 The width of the W
Assume that the weak coupling constant gW is universal, i.e. independent of
the fermion’s flavour.
1. Compute (at the tree level) the decay width for the decay
W − → e− ν̄e
2. Assuming that all quarks and leptons are essentially massless compared
to the W mass mW , with of course the exception of the top quark.
Determine the total W decay with ΓW . Hint : quarks have colour !
3. Insert your result into Eq.(13.25), and verify that unitarity is not vio-
lated in this process. For the unitarity limit, see Appendix 15.10.
W ud : gW cos(θc ) , W us : gW sin(θc ) ,
W cd : −gW sin(θc ) , W cs : gW cos(θc ) .
August 31, 2019 419
Here, θc ≈ 13o is the so-called Cabibbo angle. Show that this refinement does
not change the total W decay width.
1. Compute the total cross section for F1 F̄1 → F2 F̄2 at energies large
compared to M .
2. Compare your result with the unitarity limit and find a restriction on
M.
Example computations
421
422 August 31, 2019
allowed in the massless case we can write the two diagrams as follows, where
we indicate their dependence on the helicity:
i~ aν
M1 (λ) = uλ (p2 )(ve + ae γ 5 )γ α uλ (p1 ) u− (q1 )(1 + γ 5 )γα u− (q2 ) ,
s − mZ 2
i~ gW 2
M2 (λ) = u− (q1 )(1 + γ 5 )γ α uλ (p1 ) uλ (p2 )(1 + γ 5 )γα u− (q2 ) .
t − mW 2
(14.2)
4i~ aν (ve − ae )
M1 (+) = s+ (p2 , q2 ) s− (p1 , q1 ) ,
s − mZ 2
4i~ aν (ve + ae )
M1 (−) = s− (p2 , q1 ) s+ (q2 , p1 ) ,
s − mZ 2
M2 (+) = 0 ,
8i~ gW 2
M2 (−) = s− (q1 , p2 ) s+ (q2 , p1 ) . (14.5)
t − mW 2
Keeping in mind the antisymmetry of the spinor products, and the Fermi
minus sign, we find that up to an irrelevant overall complex phase the am-
plitudes are given by
aν (ve − ae )
M(+) = 4~ t ,
s − mZ 2
2gW 2
aν (ve + ae )
M(−) = 4~ + u . (14.6)
s − mZ 2 t − mW 2
August 31, 2019 423
We arrive at
2
2
2 aν (ve − ae )2 2 aν 2 (ve + ae )2 2
|M| = 4~ t + u
(s − mZ 2 )2 (s − mZ 2 )2
4gW 2 aν (ve + ae ) 4gW 4
2 2
+ u + u . (14.7)
(s − mZ 2 )(t − mW 2 ) (t − mW 2 )2
The centre-of-mass frame is the obvious choice to work in ; in this frame
s
t = − (1 − cos θ) , (14.8)
2
where θ is the angle between p~1 and ~q1 , and the cross section has no azimuthal-
angle dependence. We may therefore write the phase space as follows :
d cos θ dφ d cos θ 1
dV (p1 + p2 ; q1 , q2 ) = 2
→ = dt , (14.9)
32π 16π 8πs
with the integration interval being t ∈ [−s, 0]. The various integrals are
easily worked out ; we have
Z0
u2 mW 2
s s
=s +2−2 1+ log 1 + ,
(t − mW 2 )2 mW 2 s mW 2
−s
Z0 2 !
u2 3 mW 2 mW 2
s
2
= s2 + − 1+ log 1 + ,
t − mW 2 s s mW 2
−s
Z0 Z0
2 1
t dt = u2 dt = s3 . (14.10)
3
−s −s
Putting everything together2 we obtain for the total cross section the expres-
sion
σ(e+ e− → νe ν̄e ) =
( 2
~2
2 2 2
2 s
aν v e + ae
4πs 3 s − mZ 2
2
2 2
!
s 3 m m s
+ 4gW 2 aν (ve + ae )
W W
+ − 1+ log 1 +
s − mZ 2 2 s s mW 2
mW 2
4 s s
+ 4gW +2−2 1+ log 1 + . (14.11)
mW 2 s mW 2
2
And not forgetting the flux factor !
424 August 31, 2019
For muon (or tau) neutrinos only the first line remains :
2
~2 2 2
+ −
2 s
σ(e e → νµ ν̄µ ) = aν (ve + ae ) . (14.12)
6πs s − mZ 2
At the Z peak
As it stands, the cross section diverges for s = mZ 2 . This is of course due to
our neglecting ΓZ . To remedy this, we may replace mZ 2 by mZ 2 − imZ ΓZ , and
neglect the second (W -exchange) diagram3 . Close to the Z pole, the cross
section for each neutrino type is then given by
~2 aν 2 (ve 2 + ae 2 ) mZ 2
σ(e+ e− → ν ν̄) = (s ≈ mZ 2 ) ,
6π (s − mZ 2 )2 + mZ 2 ΓZ 2
(14.13)
while at the very peak we have4
~2 aν 2 (ve 2 + ae 2 )
σ(e+ e− → ν ν̄) = (s = mZ 2 ) , (14.14)
6πΓZ 2
This can be cast in an instructing form, using the fact that
~ (ve 2 + ae 2 ) mZ ~ aν 2 mZ
Γ(Z → e+ e− ) = , Γ(Z → ν ν̄) = . (14.15)
12π 6π
The cross section at the peak can therefore be written as
12π Γ(Z → e+ e− )
+ − Γ(Z → ν ν̄)
σ(e e → ν ν̄) = (s = mZ 2 ) ,
s ΓZ ΓZ
(14.16)
which is exactly the form demanded by unitarity for an intermediate state
(the Z) with unit spin (see section 15.10).
3
This is really justified ! In the sense in which ΓZ 6= 0 comes about by interactions, ΓZ
is formally of higher order in perturbation theory, and then a factor ΓZ −1 actually lowers
the order of such diagrams. Thus, around the Z pole, the W -exchange diagram is formally
of higher order in perturbation theory.
4
Remember, this is all strictly tree level...
August 31, 2019 425
where the index 1 refers to the νe -exchange diagram, 2 to the combined γ/Z
diagrams. The A’s collect the coupling constants and propagators :
1
A1 (λ) = −4gW 2δλ,− ,
∆
e2 gWWZ (ve − λae ) e2 mZ 2 4gW 2
A2 (λ) = + = − + δλ,− (14.20)
.
s s − mZ 2 s(s − mZ 2 ) s − mZ 2
pµ1 = (E, E~ep ) , pµ2 = (E, −E~ep ) , q1µ = (E, q~e) , q2µ = (E, −q~e) .
(14.22)
The beam energy is E so that s = 4E 2 , and the W velocities are β = q/E,
where q 2 = E 2 − mW 2 . The ~ep and ~e are unit vectors, with ~ep · ~e = c, the
August 31, 2019 427
1
(± )µ = √ u± (a) γ µ u± (b) . (14.24)
8
The two longitudinal polarisations are of different : we write
2mW 2 2mW 2
0 1 0 1
1 = q1 − P , 2 = q2 − P . (14.25)
mW β s mW β s
~e2 (1 + c)
M(−; +, +) ≈ (1 − c2 )1/2 ,
2sW 2 (1 − c)
~e2
M(−; −, −) ≈ 2
(1 − c2 )1/2 . (14.37)
2sW
The apparent singularity at c = 1 is of course due to our approximating ∆.
Note also that e2 ~ is actually the properly dimensionless quantity 4πα, see
sec. 10.2.5. Next we have
~e2 mZ 2
M(+; 0, 0) ≈ (1 − c2 )1/2 . (14.38)
2 mW 2
The most difficult one is
~e2 mZ 2
2 1/2 1 1
M(−; 0, 0) ≈ (1 − c ) − −1 . (14.40)
2 sW 2 mW 2 2sW 2
8
This is what we did all the hard work for, after all.
August 31, 2019 431
We have now proven that the amplitudes surviving the high-energy limit are
dimensionless and well behaved ; but they still depend on the ratio mZ /mW .
Since we have taken the fermions to be massless, the Higgs boson is absent at
the tree level here. Nevertheless (but this is, as we see, an additional input !)
we may take mW /mZ = cW from our discussion of the minimal Higgs sector
in sec. 13.4.2, and then we find the even more pleasing-looking results
1 1+c
M(−; +, +) ≈ 4πα (1 − c2 )1/2 ,
2sW 2 1 − c
1
M(−; −, −) ≈ 4πα (1 − c2 )1/2 ,
2sW 2
1
M(+; 0, 0) ≈ 4πα (1 − c2 )1/2 ,
2cW 2
1
M(−; 0, 0) ≈ 4πα (1 − c2 )1/2 . (14.41)
4sW 2 cW 2
MγγH = + , (14.42)
which differ in the orientation of the fermion line. If our theory is to be con-
sistent, this amplitude must be ultraviolet-finite (since otherwise we would
have to introduce a counterterm which would mean a direct photon-photon-
Higgs coupling after all) and it must obey current conservation. We shall
verify this point first, by putting a handlebar on one of the photons :
9
And it provided the first mechanism for observing Higgs particles at the LHC by
measurement of the invariant mass of photon pairs.
432 August 31, 2019
= − + −
(14.43)
= , = (14.44)
we see that the third and fourth diagram are actually equal to the first and
second one (flipped over), so that the sum vanishes and current conservation
(gauge invariance) is assured.
The process we investigate is, more explicitly, given as
γ(q1 , 1 ) + γ(q2 , 2 ) → H
where we have explicitly given the momenta and polarisations of the (on-
shell) photons. We denote the Higgs mass by mH and the fermion mass by
M . The fermion charge is denoted by Q, and its coupling to the Higgs by
gffH . Although we do not expect (or hope to see) divergences, we shall still
work in a general dimension 4 − 2 for reasons that will be come clear later
on : but we shall take → 0 wherever we can. One of our diagrams10 is
given by
Considering the first term, we notice that in our integral we can again, just
like in section 11.5, replace
α β 1 2 αβ 1
k̂ k̂ → k̂ g ∼ + k̂ 2 g αβ . (14.49)
4 − 2 4 8
Thus effective form of N is
2 2 2
N → M (η1 · η2 ) 4k̂ + 4M + mH (4xy − 2) , (14.50)
M1 = Q2 gffH M (η1 · η2 ) A ,
2µ2 2 2 2
4−2 4k̂ + 4M + mH (4xy − 2)
Z Z
A = δxy d k̂ 3 (14.51)
(2π)4−2
k̂ 2 − M 2 + xy mH 2 + iη
We see that in fact M2 = M1 , owing to our use of the η’s rather than the
’s11 . Proceeding with the one-loop cookbook of section 6.612 we can write
Z∞
4t − 4M 2 − mH 2 (4xy − 2)
Z
i
A= 2 δxy dt t1− 3 (14.52)
8π t + M 2 − xy mH 2 − iη
0
11
You might start to worry at this point. Weren’t the two diagrams necessary for the
cancellations proving current conservation ? How, then can they be equal ? The answer
is that by going from to η we have exchanged contributions between the diagrams, and
indeed in the η formulation they are separately current-conserving.
12
And taking → 0 wherever that is justified by the event.
434 August 31, 2019
Since
Z∞ Z∞
t2− 1 t1− 1
dt 3
≈ + O (1) , dt 3
≈ + O () , (14.53)
(t + a) (t + a) 2a
0 0
we find
1 − 4xy
Z
i
A = 2 δxy (14.54)
8π (M /mH 2 ) − xy − iη
2
14
Since it involves both a loop-induced production of the Higgs by gluon fusion and a
loop-induced decay of the Higgs into photon pairs, this process may also be considered as
a ‘proof’ of the actual reality of loop effects ; and as a true ‘quantum’ process since at the
tree level it is impossible.
436 August 31, 2019
15
At least, its version of August 31, 2019.
16
Indeed, if you present the Standard Model to a lay audience, you should always try
to elicit this question.
17
You might construct theories in which the extra generations are somehow masked by
some cancellation mechanism - but then they are no longer copies of the known ones.
Chapter 15
Appendices
437
438 August 31, 2019
simply because the coefficients do not grow as rapidly. Let us suppose that
this is indeed the case. We then may employ the formula
Z∞
dy exp(−y) (xy)n = n! xn , n = 0, 1, 2, . . . (15.4)
0
as <(x) < 1, thus immeasurably enlarging the region of x values where the
Borel-summed version makes sense. A more challenging example :
X
f (x) = k! (−x)k , x > 0 . (15.7)
k≥0
and this integral runs into problems around y = 1/x. One may of course
extend the integral to complex y values, and then skirt around the singularity.
But it is not clear whether we should pass the point y = 1/x above, or below,
the real axis. The ambiguity, that is, the difference between the results from
the alternative contours, is of course given by the number
e−y e−1/x ,
I
dy = 2πi
1 − xy x
y∼1/x
August 31, 2019 441
and, since during the integration we might decide to circle around the sin-
gularity any number of times, arbitrary multiples of the ambiguity may be
added. We see that the Borel integral becomes ambiguous : it may be some
consolation that the ambiguity is nonperturbative in nature, i.e. it has no
series expansion for infinitesimal but real and positive x. We conclude that
the function F (x) is given by
Z∞
e−y e−1/x
F (x) = P dy − (2n + 1)iπ ,
1 − xy x
0
(15.12)
Z∞ Z1/x
−y −1/x z −z
P dy
e
=
e −E1 (1/x) + dz e − e . (15.13)
1 − xy x z
0 0
(15.14)
To build the diagram, we first connect one of the external lines to the vertex :
there are 4 possible ways to do so since all legs of the vertex are indistin-
guishable. We then have
4
4!
4×3
4!
For the remaining operation of linking the two other legs of the vertex there is
of course only one possibility. The resulting symmetry factor is 4×3/24 = 1/2
as advertised.
4
If you have ever played with K’NEX this may look familiar
August 31, 2019 443
9
4! × (3!)4
444 August 31, 2019
9×6×3
4! × (3!)4
Now take one of the outside legs. It must be attached to another outside leg
that comes from another vertex, hence 4 possibilities :
9×6×3×4
4! × (3!)4
In the same way, for the next outside leg there are possibilities :
9×6×3×4×2
4! × (3!)4
and then the diagrams is closed uniquely. The symmetry factor is therefore
1/24, corresponding to the symmetry of permuting the vertices.
Note that the two crossing lines do not touch but form an ‘overpass’ : this
diagram is nonplanar in the sense that an overpass remains no matter how
you try to draw it. The toolkit is now even more impressive :
1
5! × (3!)5
We can attach the external lines to different 3-vertices in 15, 12, and 9 ways,
respectively :
15 × 12 × 9
5! × (3!)5
August 31, 2019 445
Now the leftmost object can be attached to the two remaining 3-vertices in
6 and then in 3 ways :
15 × 12 × 9 × 6 × 3
5! × (3!)5
The topmost leg of the object on the left has to be connected to both the
upper and the lower objects on the right, for which there are thus 4 and 2
options, respectively :
15 × 12 × 9 × 6 × 3 × 4 × 2
5! × (3!)5
By now, it ought to be obvious that there are 2 ways to finish off the dia-
gram, leading to a symmetry factor of (15.12.9.6.3.4.2.2)/(5!(3!)5 ), in simpler
terms 1/2. The actual symmetry is, once again, the permutation of the two
innermost vertices.
An example is
E = 2 , I = 6 , V1 = 1 , V3 = 3 ,
V4 = 1 , P = 1 , L = 2 .
We now look for linear combinations T of these numbers that are the same
for all diagrams. That is, whatever we do to a diagram, the value of T must
remain unchanged. It is easy to see that any diagram can be transformed into
446 August 31, 2019
(iii) cutting through a line such that the graph falls apart :
0000000
1111111
1111111
0000000 1111111
0000000 0000000
1111111
1111111
0000000 1111111
0000000
0000000
1111111 0000000
1111111
0000000
1111111 0000000
1111111 0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111 0000000
1111111 0000000
1111111
0000000
1111111 0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
↔ 0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111 0000000
1111111
0000000
1111111 0000000
1111111 0000000
1111111
0000000
1111111
where αE and β are undetermined. We see that we have precisely two di-
agrammatic sum rules. By inspection of an arbitrary5 diagram we see that
T = 0, so that the sum rules are
X
qVq = 2I + E , (15.21)
q
X
Vq = I + P − L . (15.22)
q
These two sum rules each have their own ‘physical’ interpretation. In the
first place, the sum rule 15.21 simply tells us that external lines have one
endpoint to be attached to a vertex, and internal lines have two endpoints
that must be accomodated. The second rule, 15.22, comes into its own when
we assign momenta to flow along the lines. The momenta in the external lines
are assumed to be given In every vertex the total momentum is conserved,
and each connected piece must contain an overall momentum conservation.
We see that in that picture L is simply the number of internal lines whose
momentum is not fixed by conservation. This derivation of the sum rules
given above is to convince you that there is not a third type of sum rule,
based on yet again different arguments, lurking about.
5
Arbitrary, except that it must contain at least one vertex. There are two connected
diagrams without vertices: the first one, , conforms to the sum rules by choosing
I = −1, and the second one, , fits in if we choose I = 0. But these choices are
obviously somewhat forced.
448 August 31, 2019
The requirement for the path integral to be defined is that at both endpoints
(still assumed to be at infinity in some complex direction) the real part of
the action goes to positive infinity. That is,
<(ϕm ) → +∞ ⇒
π 2π π 2π
− + k < arg(ϕ) < + k , k = 1, 2, . . . , m .(15.24)
2m m 2m m
The argument of the endpoints are restricted to certain intervals. Inside
each interval the precise value of the argument is irrelevant since the path
integral will be precisely the same: we may therefore say that for a theory
(m)
with highest interaction term ϕm the admissible endpoints are ∞n , with
n = 0, 1, 2, . . . , m − 1, where
(m) iφn 2π 1 2π 1
∞n = lim re , φn ∈ (n − ), (n + ) . (15.25)
r→∞ m 4 m 4
Since the path integrand is analytic, the theory is completely determined by
the endpoints. We see that for a theory with highest interaction of the form
ϕm there are precisely m − 1 independent solutions to the SDe, as necessary
since the SDe is a linear differential equation of order m − 1. We may take
(m)
these as given by a contour running between ∞0 and any of the m−1 other
(m)
∞n . By suitably combining several integrals we can of course also obtain
(m)
a theory based on a contour running between any two distinct ∞n .
August 31, 2019 449
If λ, µ and J are all real, the SDe can be iteratively solved starting from the
classical solution, and the perturbation series is completeley fixed as well as
real ; whereas the fact that the two integration contours are really distinct
from the real axis tells us that the path integral (and hence φ(J)) ought
to be complex, with the results from the two contours related by complex
conjugation. We conclude that the difference between the two alternative
path integrals must be non-perturbative in nature.
every diagram as being of the same ‘order of magnitude’ this gives an idea,
however crude, of the amplitude to be expected. Of particular interest is
the behaviour of the number of graphs under extreme circumstances such
as when the number of external lines becomes very large. In this section
we shall consider the simplest case, that of tree-level Green’s functions of a
single self-interacting field.
In order to count diagrams, we can simply consider the zero-dimensional
theory so that we are not bothered by summing diagrams over internal de-
grees of freedom. Secondly, we replace every vertex, and every propagator by
unity. This reduces every Feynman diagram to just its symmetry factor. For
tree diagrams, the symmetry factor is unity; for loop graphs, the symmetry
factors are nontrivial and getting rid of them is quite cumbersome8 . The
appropriate action reads
1 X k
S(ϕ) = ϕ2 − F (ϕ) , F (ϕ) = ϕk , (15.30)
2 k≥3
k!
where k is unity for every k-point interaction proposed in the theory, oth-
erwise zero. Since we only consider counting graphs, the fact that S may
become negative infinity for infinite ϕ does not bother us. The number-of-
diagrams generating function
X Nn
Φ(J) = Jn , (15.31)
n≥0
n!
This is useful in theories with only a single coupling, such as pure ϕ4 the-
ory. In more complicated theories, the best approach for n not too large is
simply to iterate Eq.(15.32) by computer algebra. For pure ϕp theories we
can explicitly work out the result of the Lagrange expansion. The counting
equation is
1 m
φ=J+ φ , m=p−1 , (15.34)
m!
so that Lagrange’s formula gives
n−1
X 1 ∂
φ = J+ J mn
n>0
n! (m!)n ∂J
X (mn)!
= J mn−n+1 . (15.35)
n≥0
n!(m!)n (mn − n + 1)!
(mn)!
Nn(m−1)+1 = , n = 0, 1, 2 . . . (15.36)
n!(m!)n
Asymptotic methods
For asymptotically large n, we can estimate the form of Nn by realizing that
these must be given by the behaviour of Φ(J) near that of its singularities
that lies closest to the origin in the complex-J plane. Now, if Φ(J) is singular,
then Φ0 (J) is divergent10 , so that dJ/dΦ must vanish. We therefore solve the
equation
∂
J = 1 − F 00 (Φ) = 0 (15.37)
∂Φ
for Φ. If the highest power of interaction in the theory is ϕm , this equation
has m − 2 complex roots Φ1 , Φ2 , . . . , Φm−2 , and
Jp = Φp − F 0 (Φp ) , p = 1, 2, . . . , m − 2 . (15.38)
10
The divergence might also show up in higher derivatives only, but in every actual case
that I have studied the divergence shows up in Φ0 .
August 31, 2019 453
Now, single out that Jp that has the smallest absolute value11 , which we shall
call J0 , and its corresponding Φp will be writtten Φ0 . For J and Φ very close
to the values J0 and Φ0 , respectively, we may use Taylor expansion to write
1
J ≈ J0 − F 000 (Φ0 )(Φ0 − Φ)2 , (15.39)
2
since the linear term vanishes by definition. Hence
1/2 s
J 2J0
Φ ≈ Φ0 − 1 − 000
(15.40)
J0 F (Φ0 )
This estimate grows roughly as n!, as ought to have been immediately obvious
from the fact that Φ(J) has a finite radius of convergence ; the above, more
careful, treatment gives an estimate that is quite good even for non-huge n.
As an application, we may consider purely gluonic QCD. In this theory, the
only interactions are between 3 or 4 gluons, and the theory is equivalent, as
far as counting is concerned, to the ϕ3/4 theory, with
1 3 1 4
F (ϕ) = ϕ + ϕ . (15.43)
3! 4!
The solutions of Eq.(15.37) and the corresponding J values are
√ 4 √ √ 4 √
Φ1 = −1 + 3 , J1 = − + 3 ; Φ2 = −1 − 3 , J2 = − − 3 ,
3 3
(15.44)
11
The case that there are several such values is discussed in the next paragraph.
12
This can be proven by applying the Lagrange expansion to the object u = y + u2 /2 =
√
1 − 1 − 2y, and putting y = x/2.
454 August 31, 2019
n Nn (exact) Nn (asymptotic)
1 1 0.85
2 1 1.07
3 4 4.01
4 25 25.17
5 220 220.94
6 2485 2493.60
7 34300 34397.35
8 559405 560754.85
9 10525900 10547973.57
Table 15.1: Exact and asymptotic number of tree diagrams in ϕ3/4 theory
√ √ √
so that J0 = 3 − 4/3, Φ0 = 3 − 1, and F 000 (Φ0 ) = 3. In table 15.5.1 we
give the exact number Nn , and its asymptotic estimate. The approximation
is better than one per cent for n ≥ 3. The non-polynomial (that is, n!) growth
of the number of diagrams with n can be seen as an immediate indication
of the failure of perturbation theory as a convergent series, as discussed in
Appendix 1.
Coarse-graining effects
In the above we have assumed that there is only a single J0 . This is indeed
usually the case ; for pure ϕp theories, however, Eq.(15.37) reads
1 q
ϕ =1 , q =p−2 , (15.45)
q!
and this has solutions
1/q n
φn = (q!) exp 2iπ , n = 1, 2, . . . , q ; (15.46)
q
the corresponding values for J are
1 q
Jn = 1 − φn q+1 = φn , n = 1, 2, . . . , q , (15.47)
(q + 1)! q+1
and these have all the same absolute value. The thing to do is therefore
to take the asymptotic contributions from all these q singular points into
August 31, 2019 455
The sum over the n values of φ will vanish completely, except when k − 1 is a
multiple of q, and then it evaluates to q/(q!)k−1 ; this is exactly the behaviour
we found using Lagrange expansion.
We might have proceeded otherwise, by simply taking the single real
solution φq = (q!)1/q as the only singular point. The number of diagrams
Nk will then be nonvanishing for every k value, while in the asymptotic
expression (15.48) the sum over n φ’s is replaced by φq −(k−1) , that is precisely
q times smaller than the nonvanishing sums of Eq.(15.48). We see that the
taking into account of only the single, real solution causes the asymptotic
values of Nk to be ‘smeared out’ ; Nk is then never zero anymore, but its
average value13 is still correct.
N0 (19) = 11081983532721088487500 ,
N1 (19) = 2900013601350201168582750 . (15.50)
13
For the correct definition of ‘average’.
456 August 31, 2019
The number N0 (19) is the actual number of diagrams since tree diagrams
always have unit symmetry factor ; but the number N1 (19) underestimates
the actual number of diagrams since the symmetry factors are not trivial. We
can see, however, that the only possible nonntrivial symmetry factor at the
one-loop level is 1/2, as evidenced by the factor ~/2 in Eq.(15.49). Inspection
tells us that in this theory the only elementary Feynman diagrams that have
symmetry factor 1/2 are
E1 = , E2 = , E3 = ,
E4 = , E5 = . (15.51)
All diagrams that contain one of these elementaries as a subgraph will have
a symmetry factor 1/2, and it will suffice to determine their number and
multiply it by two14 . Alternatively, we may get rid of all such diagrams,
and work with the difference. This is the more useful approach ; and it
illustrates how we may go about using counterterms to impose constraints
on the structure of Feynman diagrams. The procedure is best explained by
going through it step by step. In the first place, it will become necessary to
again distinguish betwee three- and four-point vertices. We therefore modify
Eq.(15.49) be reinserting labels for these couplings:
g3 2 g4 3 ~
Φ(J) = J + Φ + Φ + (g3 + g4 Φ) Φ0 (15.52)
2 6 2
Iterating this gives for the first N :
~
N (0) = g3 ,
2
1 2
N (1) = 1 + ~ g4 + g3 ,
2 7
N (2) = g3 + ~ 4g3 + g4 g3 ,
2
2 7 2 4 59 2
N (3) = g4 + 3g3 + ~ g4 + 24g3 + g4 g3 . (15.53)
2 2
14
This relies, of course, on the fact that there can be no diagrams containing two (or
more) of the elementaries, since that would be a two-loop diagram (or even higher).
August 31, 2019 457
We can now start to remove graphs. We shall get rid of all diagrams with a
tadpole by introducing a tadpole counterterm ~T in the SDe:
g3 2 g4 3 ~
Φ(J) = J + Φ + Φ + (g3 + g4 Φ) Φ0 − ~T (15.54)
2 6 2
We see that this amounts to replacing J by J − ~T , and the N ’s become
1
N (0) = ~ g3 − T ,
2
1 2
N (1) = 1 + ~ g4 + g3 − g3 T ,
2 7 2
N (2) = g3 + ~ 4g3 + g4 g3 − g4 T − 3g3 T ,
2
2 7 2 4 59 2 3
N (3) = g4 + 3g3 + ~ g4 + 24g3 + g4 g3 − 10g4 g3 T − 15g3 T .
2 2
(15.55)
, , and ,
16
Since Γmn is symmetric, so is Hmn although this is not obvious from the form it is
written here.
462 August 31, 2019
un−1
I
~
Π(n) = du ,
2iπ f (u)
|u|=1
1 2 1
f (u) = µ − γ1 u + − γ2 u + 2 . (15.90)
u u
In the continuum limit, the only relevant poles of the integrand are those at
values of u such that |u| = 1 − O (∆). Let uj (j = 1, 2, . . .) be these poles:
then
X uj |x|/∆
Π(x) = ~ . (15.91)
j
f 0 (uj )
Writing u = 1 − v∆, we can approximate
There are now two possible continuum limits. In the first case, we can assume
that γ1 + 4γ2 does not vanish. In that case, we can take γ1 + 4γ2 ∼ 1/∆, and
the resulting continuum limit is indistinguishable from the nearest-neighbour
case. For later reference we shall denote this propagator by
~
P1 (x) = exp(−m|x|) . (15.93)
2m
466 August 31, 2019
1 4 1
γ1 + 5γ2 ∼ → γ1 ∼ , γ2 ∼ − 3 , (15.94)
∆3 ∆3 ∆
and
6
µ = m4 ∆ + 2γ1 + 2γ2 ∼ m2 ∆ + . (15.95)
∆3
The poles of the integrand are therefore approximately given by
f (u) = ∆ m4 + v 4 + O ∆2 = 0 ,
(15.96)
Only u1 and u2 are inside the unit circle, and we obtain the propagator
~ −m|x| m|x| m|x|
Π(x) = √ exp √ cos √ + sin √ (15.98)
,
m3 8 2 2 2
which we shall denote by P2 (x): it has the interesting property that Π2 (x)
is negative for mx between 3π/4 and 7π/4, modulo 2π. An discrete action
such as the one belonging to this continuum limit, in which nearest-neighbour
and next-to-nearest-neighbour couplings have opposite sign, are called frus-
trated17 . The continuum limit of the propagator can also be written as
Z
~ exp(ikx)
P2 (x) = dk , (15.99)
2π k 4 + m4
We shall now arrange for the only the highest possible power of 1 − u to
survive in this expression. We first put u = exp(ik∆), so that the function
f (u) becomes
p
X X
f (u) = µ − 2γj cos(jk∆) = µ − (k∆)2r Br ,
j=1 r≥0
p
X 2(−)r
Br ≡ j 2r γj . (15.103)
j=1
(2r)!
with p
X 2(−)r
Q(j) = cr j 2r . (15.106)
r=1
(2r)!
The polynomial Q(j) is even and of degree 2p in j, and Q(0) = 0. We can
now, for any preassigned q with 1 ≤ q ≤ p, choose the numbers cr such that
Q(0) = · · · = Q(q − 1) = Q(q + 1) = · · · = Q(p) = 0 , Q(q) 6= 0 ,
(15.107)
468 August 31, 2019
upon which
γq = Bp /Q(q) . (15.108)
Obviously, the polynomial Q(j) is given by
2(−)p Y
2 2
Q(j) = j −n , (15.109)
(2p)!
0≤n≤p
n 6= q
x. The values of p are 1,3,10, and also the asymptotic form of Eq.(15.115)
is plotted. For large p the asymptotic form is approximated smoothly. The
higher the value of p, the more frustrated the lattice is, and the more diffi-
cult it becomes for momentum modes with high wave number to propagate
through the lattice, as is evident from Eq.(15.111). For the totally frustrated
lattice, all wave numbers smaller than m propagate equally, and all wave
nubers larger than m do not propagate at all.
For the time dependence of the source we examine three alternatives, which
may be called slow, fast, and abrupt, respectively : in position language,
|x0 |
(1) 0 1 i 0 0
Jt (x ) = exp − − px ,
(a1 )1/2 a1 ~
!
2
(2) 0 1 x0 i 0 0
Jt (x ) = exp − 2 − p x ,
(2πa22 )1/4 4a2 ~
(3) 0 1 0
i 0 0
Jt (x ) = θ −a3 < x < a3 exp − p x . (15.120)
(2a3 )1/2 ~
These three sources are all normalised to unit strength :
Z
(j)
dx0 |Jt (x0 )|2 = 1 , j = 1, 2, 3 . (15.121)
To compare the spread of the sources in the time domain we can use
D E Z
2 (j) a1 2 a3 2
x0 = dx0 (x0 )2 |Jt (x0 )|2 = = a2 2 = . (15.122)
2 3
In momentum language we have
(1) 2a1 1/2
Jt (k 0 ) = ,
∆2 + 1/a1 2
(2) 2 2
Jt (k 0 ) = (8πa22 )1/4 e−a2 ∆ ,
(3) −i −ia3 ∆
Jt (k 0 ) = ia3 ∆
e − e , (15.123)
(2a3 )1/2 ∆
August 31, 2019 471
with ∆ = k 0 − p0 /~. The response of the field to the timelike part of the
source is 0 0
e−ik x
Z
0 (j)
ψj ≡ dk 0 2 Jt (k 0 ) , (15.124)
k − ω 2 + i
and this is what we now investigate for positive times : x0 > 0.
For large times (τ → ∞) the two integral terms in the second both behave
as ∼ 1/τ so, as in the previous case, they give rise to a contribution that dies
out with the source. We therefore have
simply close the contour in the lower half complex k 0 plane to find
0 0 /~)
−π eia3 (ω−p /~) − e−ia3 (ω−p
ψ3 = √ exp(−iωx0 ) √ . (15.129)
ω 2 a3 (ω − p0 /~)
The numerator in the last factor remains bounded in absolute value. There-
√
fore, as long as ω 6= p0 /~, ψ3 goes to zero as 1/ a3 for large a3 , while at
√
ω = p0 /~ it approaches infinity as a3 : yet another situation in which the
on-shell condition is enforced.
where m and m0 are the masses of the two particles in 1. The symmetry
factor S1 equals 1 if the particles are distinguishable, and 1/2 if they are not.
Ω is of course the solid angle of one of the particles in the rest frame of X.
The angle- and spin-averaged transition rate is therefore
16πM Γ1 N M 2
Z
1 X dΩ
Ak · uj uj · Ak = , (15.133)
K j,k 4π S1 K λ1/2
18
This might be just a number, or a spinor, or a polarisation vector,. . . take your pick.
August 31, 2019 475
where l denotes the discrete quantum numbers in the state 2. The width for
the process is given by
Z
1 1 X
Γ2 = B l · uj uj · Bl dVn S2 , (15.135)
2M N j,l
where dVn is the n-particle phase space factor going with the state 2, and S2
is the appropriate symmetry factor.
where it must be realized that we have rewritten the integral over B-cum-A
by the integral over B times the average over A. Due to angular-momentum
conservation we can now write, using Eq.(15.133),
Γ1 Γ2 N 16π s
σcs=M 2 = . (15.143)
Γ Γ S1 K λ(s, m2 , m02 )
Now, the factor Γ2 /Γ is understandable since the X particle has only a frac-
tional probability to decay into state 2 (there may be other decay channels
available, in fact at least the decay X→1), and then symmetry between the
reactions 1→2 and 2→1 requires also the presence of the factor Γ1 /Γ. We
August 31, 2019 477
conclude that the cross section for the initial state 1 to go into any final state
with spin J is bounded by the unitarity limit
2J + 1 16π s
σUL = , (15.144)
S1 K λ(s, m2 , m02 )
γ µ γ ν + γ ν γ µ = 2 g µν , γ̂ µ γ̂ ν + γ̂ ν γ̂ µ = 2 g µν , (15.145)
γ̂ µ = S γ µ S −1 . (15.146)
Γ0 = 1 , Γ1 = γ 0 , Γ2 = iγ 1 , Γ3 = iγ 2 , Γ4 = iγ 3 ,
Γ5 = γ 0 γ 1 , Γ6 = γ 0 γ 2 , Γ7 = γ 0 γ 3 , Γ8 = iγ 1 γ 2 ,
Γ9 = iγ 1 γ 3 , Γ10 = iγ 2 γ 3 , Γ11 = iγ 0 γ 1 γ 2 , Γ12 = iγ 0 γ 1 γ 3 ,
Γ13 = iγ 0 γ 2 γ 3 , Γ14 = γ 1 γ 2 γ 3 , Γ15 = iγ 0 γ 1 γ 2 γ 3 , (15.147)
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
0 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
1 1 0 5 6 7 2 3 4 11 12 13 8 9 10 15 14
2 2 5 0 8 9 1 11 12 3 4 14 6 7 15 10 13
3 3 6 8 0 10 11 1 13 2 14 4 5 15 7 9 12
4 4 7 9 10 0 12 13 1 14 2 3 15 5 6 8 11
5 5 2 1 11 12 0 8 9 6 7 15 3 4 14 13 10
6 6 3 11 1 13 8 0 10 5 15 7 2 14 4 12 9
7 7 4 12 13 1 9 10 0 15 5 6 14 2 3 11 8
8 8 11 3 2 14 6 5 15 0 10 9 1 13 12 4 7
9 9 12 4 14 2 7 15 5 10 0 8 13 1 11 3 6
10 10 13 14 4 3 15 7 6 9 8 0 12 11 1 2 5
11 11 8 6 5 15 3 2 14 1 13 12 0 10 9 7 4
12 12 9 7 15 5 4 14 2 13 1 11 10 0 8 6 3
13 13 10 15 7 6 14 4 3 12 11 1 9 8 0 5 2
14 14 15 10 9 8 13 12 11 4 3 2 7 6 5 0 1
15 15 14 13 12 11 10 9 8 7 6 5 4 3 2 1 0
In the first place, Γk 2 = 1 for all k. Secondly, for every pair j and k there
are numbers n and cn such that
Γj Γk = cn Γn , cn = 1, −1, i or − i. (15.148)
From these properties it follows that simultaneously
1
Γk Γj = Γn (15.149)
cn
We can thus construct the multiplication table 15.11.120 , where the possible
values of j define the rows, and those for k the columns: the corresponding
entry is then the value of n. For instance,
Γ6 Γ4 = Γ13
20
Kids! Don’t do this at home, since constructing this multiplication table is extremely
tedious. The numbers cn are not given: they are anyhow only defined up to a sign, since
we can always replace Γj by −Γj (using γ 2 γ 0 instead of γ 0 γ 2 , say) without changing the
Dirac anticommutation relation.
August 31, 2019 479
(in this case c13 happens to be 1). Note that in this table every row and every
column contains each of the numbers from 0 to 15 precisely once. Hence, if
we keep j fixed and let k run from 0 to 15, the value of n will also take on
all values from 0 to 15 (although generally in a different order). Obviously,
for the set Γ̂ exactly the same multiplication table holds.
We are now ready to prove the theorem. Let A be an arbitrary matrix,
and define S by
X15
S≡ Γ̂k A Γk . (15.150)
k=0
in other words,
Γ̂j S = S Γj . (15.152)
It remains to ensure that the matrix S actually has an inverse. Since A can
be chosen at will (except A = 0) this is not a problem. Let us pick another
matrix B and construct
X 15
T = Γk B Γ̂k . (15.153)
k=0
Γj T = T Γ̂j . (15.154)
It is easy to see that the spin algebra is correctly constructed once we have
raising and lowering operators
with
[[S+ , S− ], S+ ] = 2~2 S+ . (15.159)
We can then find the other algebra elements via
1 1 1
Sx = S+ + S− , S y = S+ − S− , Sz = [S+ , S− ] ,
2 2i 2~
1
S2 = S+ , S− + (Sz )2 .
(15.160)
2
We will start with particles in their rest frame21 . The spin representations
are built using four unit vectors, with obvious notation, as tµ , xµ , y µ and z µ ,
which obey
t·t = 1 , x·x = y·y = z·z = −1 , t·x = t·y = t·z = x·y = x·z = y·z = 0 .
(15.161)
Things will become easier if we also define
1
x± µ = √ xµ ± i y µ
(15.162)
2
so that
x± · x± = 0 , x+ · x− = −1 , x± · z = x± · t = 0 . (15.163)
Since the spin of a particle informs us about its behaviour under rotations in
the three-dimensional spacelike part of Minkowski space, we always require,
for particles in their rest frame,
This means that the appropriate tensors in fact contain only the three vec-
tors x+ , x− , and z ; for instance the rank-4 tensor |s, miµ1 µ2 µ3 µ4 may contain
a term x+ µ1 x− µ2 z µ3 x+ µ4 . In general, the particle’s tensor is a linear combi-
nation of such terms : which precise linear combination it is depends on s
and m, and this is what we want to look into.
21
This implies that the particles are massive. For massless partices, see later on.
482 August 31, 2019
so that
Note that
∆µα ∆αν = −∆µ ν , ∆µ µ = −3 . (15.170)
We now proceed to set up the spin algebra. A general raising operator can
always be written in the form
√
S+ = 2~ a |+i h0| + b |0i h−| , (15.171)
From
2
2 2
S+ S− = −2~ |a| |+i h+| + |b| |0i h0| ,
2 2 2
S− S+ = −2~ |a| |0i h0| + |b| |−i h−| , (15.173)
August 31, 2019 483
we find that to get the correct form of Sz we have to take |a| = |b| = 1, since
only then22
Sz = −~ |+i h+| − |−i h−| ; (15.174)
which shows that we have here indeed a spin-one system. For reasons that
will become clear later on we shall choose a = −1 and b = 1. Thus,
√ √
S+ |+i = 0 , S+ |0i = 2~ |+i , S+ |−i = − 2~ |0i . (15.176)
Σj µν αβ = Sj µ α δ ν β + δ µ α Sj ν β , j = +, −, z , (15.178)
and it is easily checked that these also obey the correct commutation relations
There is precisely one rank-2 tensor with a spin 2~ along the z axis : it is
the tensor product
|2, 2i = |++i ,
√
|2, 1i = |+0i + |0+i / 2 ,
√
|2, 0i = − |+−i − |−+i + 2 |00i / 6 ,
√
|2, −1i = − |−0i − |0−i / 2 ,
These five objects are totally symmetric. They are also traceless in the sense
that |2, miµν gµν = 0 ; this is due to our choice for the constants a and b made
above. The one object made up from |+0i and |0+i that is orthonormal to
|2, 1i is |+0i − |0+i, which forms the basis of a spin-1 sector :
√
|1, 1i = |+0i − |0+i / 2 ,
√
|1, 0i = |−+i − |+−i / 2 ,
√
|1, −1i = |−0i − |0−i / 2 . (15.183)
which upon inspection is seen to have zero spin. The orthonormality of these
nine states is easily checked. Some simple algebra also tells us that
2
X 1 µ 1 1
|2, miµν h2, m|αβ = ∆ α ∆ν β + ∆µ β ∆ν α − ∆µν ∆αβ ,
m=−2
2 2 3
1
X 1 µ 1
|1, miµν h1, m|αβ = ∆ α ∆ν β − ∆µ β ∆ν α ,
m=−1
2 2
1 µν
|0, 0iµν h0, 0|αβ = ∆ ∆αβ , (15.185)
3
so that there is a completeness relation of the form
2
X s
X
|s, miµν hs, m|αβ = ∆µ α ∆ν β . (15.186)
s=0 m=−s
spin-3 :
|3, 3i = |+ + +i
|3, 2i = |+ + 0i + |+0+i + |0 + +i
|3, 1i = 2 |+00i + 2 |0 + 0i + 2 |00+i
− |+ + −i − |+ − +i − |− + +i
|3, 0i = 2 |000i − |+0−i − |0 − +i − |− + 0i
− |−0+i − |+ − 0i − |0 + −i
|3, −1i = |+ − −i + |− + −i + |− − +i
−2 |−00i − 2 |0 − 0i − 2 |00−i
|3, −2i = |− − 0i + |−0−i + |0 − −i
486 August 31, 2019
|3, −3i = − |− − −i
spin-2(1) :
|2, 2i = |+0+i + |0 + +i − 2 |+ + 0i
|2, 1i = 2 |00+i − |+ − +i − |− + +i
− |+00i − |0 + 0i + 2 |+ + −i
|2, 0i = |+0−i + |0 + −i − |−0+i − |0 − +i
|2, −1i = 2 |00−i − |+ − −i − |− + −i
− |0 − 0i − |−00i + 2 |− − +i
|2, −2i = 2 |− − 0i − |0 − −i − |−0−i
spin-2(2) :
|2, 2i = |+0+i − |0 + +i
|2, 1i = |+00i − |0 + 0i − |+ − +i + |− + +i
|2, 0i = − |0 − +i + |−0+i − |+0−i
+ |0 + −i − 2 |+ − 0i + 2 |− + 0i
|2, −1i = |+ − −i − |− + −i + |−00i − |0 − 0i
|2, −2i = |0 − −i − |−0−i
spin-1(1) :
|1, 1i = 6 |+ + −i + 3 |0 + 0i + 3 |+00i
+ |+ − +i + |− + +i − 2 |00+i
|1, 0i = 3 |0 + −i + 3 |+0−i + 3 |−0+i
+3 |0 − +i − 2 |+ − 0i − 2 |− + 0i + 4 |000i
|1, −1i = 2 |00−i − 3 |−00i − 3 |0 − 0i
− |+ − −i − |− + −i − 6 |− − +i
spin-1(2) :
|1, 1i = |+00i − |0 + 0i + |+ − +i − |− + +i
|1, 0i = |0 + −i − |+0−i + |0 − +i − |−0+i
|1, −1i = |+ − −i − |− + −i + |0 − 0i − |−00i
spin-1(3) :
|1, 1i = |+ − +i + |− + +i + |00+i
|1, 0i = |+ − 0i + |− + 0i + |000i
|1, −1i = |+ − −i + |− + −i + |00−i
August 31, 2019 487
spin-0 :
|0, 0i = |+ − 0i + |−0+i + |0 + −i
− |0 − +i − |+0−i − |− + 0i (15.187)
Note that the spin-0 state is totally antisymmetric : obviously, this is the
only possible such state in three space dimensions. We can also compute the
‘partial’ completeness relations pertaining to each spin sector. Some algebra
teaches us that these are the following set of mutually orthogonal projection
operators :
3
X
spin-3 : |3, miµνρ h3, m|αβγ =
m=−3
1
∆µ α ∆ν β ∆ρ γ + ∆µ β ∆ν γ ∆ρ α + ∆µ γ ∆ν α ∆ρ β
6
µ ν ρ µ ν ρ µ ν ρ
+ ∆ β ∆ α∆ γ + ∆ α∆ γ ∆ β + ∆ γ ∆ β ∆ α
1
∆µν ∆ρ α ∆βγ + ∆ρ β ∆γα + ∆ρ γ ∆αβ
−
15
+ ∆νρ ∆µ α ∆βγ + ∆µ β ∆γα + ∆µ γ ∆αβ
ρµ ν ν ν
+ ∆ ∆ α ∆βγ + ∆ β ∆γα + ∆ γ ∆αβ
2
X
spin-2(1) : |2, miµνρ h2, m|αβγ =
m=−2
1 µ ν
∆ α ∆ β + ∆ α ∆ β ∆ρ γ ν µ
3
1 µ ν ν µ
ρ µ ν ν µ
ρ
− ∆ β ∆ γ + ∆ β ∆ γ ∆ α + ∆ α∆ γ + ∆ α∆ γ ∆ β
6
1 µν ρ ρ
+ ∆ ∆ α ∆βγ + ∆ β ∆αγ
6
1 µρ ν νρ µ
+ ∆ ∆ γ + ∆ ∆ γ ∆αβ
6
1 µρ ν µρ ν νρ µ νρ µ
− ∆ ∆ α ∆βγ + ∆ ∆ β ∆αγ + ∆ ∆ α ∆βγ + ∆ ∆ β ∆αγ
12
488 August 31, 2019
1
− ∆µν ∆ρ γ ∆αβ
3
2
X
spin-2(2) : |2, miµνρ h2, m|αβγ =
m=−2
1
∆ α ∆ β − ∆ α ∆ β ∆ρ γ
µ ν ν µ
3
1 µ ν ρ µ ν ρ ν µ ρ ν µ ρ
+ ∆ γ ∆ β ∆ α − ∆ γ ∆ α∆ β − ∆ γ ∆ β ∆ α + ∆ γ ∆ α∆ β
6
1 µρ ν µρ ν νρ µ νρ µ
+ ∆ ∆ α ∆βγ − ∆ ∆ β ∆αγ − ∆ ∆ α ∆βγ + ∆ ∆ β ∆αγ
4
X 1
spin-1(1) : |1, miµνρ h1, m|αβγ =
m=−1
1 µν ρ
∆ ∆ γ ∆αβ
15
1 µν ρ ρ
− ∆ ∆ α ∆βγ + ∆ β ∆αγ
10
1 µρ ν νρ µ
− ∆ ∆ γ + ∆ ∆ γ ∆αβ
10
3 µρ ν µρ ν νρ µ νρ µ
+ ∆ ∆ α ∆βγ + ∆ ∆ β ∆αγ + ∆ ∆ α ∆βγ + ∆ ∆ β ∆αγ
20
X1
spin-1(2) : |1, miµνρ h1, m|αβγ =
m=−1
1 µρ ν µρ ν νρ µ νρ µ
− ∆ ∆ α ∆βγ − ∆ ∆ β ∆αγ − ∆ ∆ α ∆βγ + ∆ ∆ β ∆αγ
4
1
X
spin-1(3) : |1, miµνρ h1, m|αβγ =
m=−1
1 µν ρ
∆ ∆ γ ∆αβ
3
spin-0 : |0, 0iµνρ h0, 0|αβγ =
1
∆µ α ∆ν β ∆ρ γ + ∆µ β ∆ν γ ∆ρ α + ∆µ γ ∆ν α ∆ρ β
6
August 31, 2019 489
ν µ ρ ν µ ρ ν µ ρ
−∆ α ∆ β ∆ γ − ∆ β∆ γ∆ α − ∆ γ ∆ α∆ β . (15.188)
p0
µ |~p| µ
z → t + zµ . (15.191)
m m
T1 = |· · · + + − · · ·i , T2 = |· · · + − + · · ·i , T3 = |· · · − + + · · ·i .
(15.194)
The rest of the content of the kets (indicated by the ellipses, and consisting of
some sequences of +’s and −’s) is identical for the three kets. The candidate
state contains these T ’s in some linear combination :
ones :
T1 → |· · · + 00 · · ·i + |· · · 0 + 0 · · ·i + · · · . (15.195)
Similarly, we find for T2 and T3 :
T2 → |· · · 00 + · · ·i + |· · · + 00 · · ·i + · · · ,
T3 → |· · · 00 + · · ·i + |· · · 0 + 0 · · ·i + · · · . (15.196)
We now note a few things. In the first place, a resulting ket like |· · · 0 + 0 · · ·i
can only come from the T ’s (in this case, from T1 and T3 ). In the second
place, our candidate state cannot contain this ket by itself, since it must be
free of 0’s. In the third place, such unwanted kets must drop out because our
state is an eigenstate of Σ2 . We must therefore rely on cancellations between
the T ’s. In fact, we need simultaneously
• Only +’s, or only −’s, occur. These are precisely the rank-s, spin-s
states such as we have found, and this persists also for s > 3. Note that
these states are totally symmetric — not for some deep field-theoretical
reason, but because they can’t help it.
Compared to the massive case, some coefficients are different : -1/2 rather
than -1/3 in the spin-2 case, and -1/12 instead of -1/15 for spin-3. This is
due, of course, to the different traces of ∆ and ∇. The spin sum for the
massless vector particle (rank-1, spin-1) is in fact that of the axial gauge
discussed in Chapter 9, with the gauge vector r chosen to be p̄. Note that,
whatever rµ , we can always move to the centre-of-mass frame of pµ and rµ ,
and in that frame we have precisely rµ = p̄µ .
August 31, 2019 493
ψ µνρσ pρ (A + Bγ 5 ) γσ ψ
µνρσ pρ µν κλ pκ = 2 pσ pλ − s g σλ
(15.203)
we have
1
M = −2i~
s − M2
494 August 31, 2019
v(p1 ) A(m2 − m1 ) + B(m1 + m2 )γ 5 u(p2 )
0 0 5
× u(p3 ) A (m3 − m4 ) − B (m3 + m4 )γ v(p4 )
−s v(p1 ) A + Bγ 5 γ µ u(p2 )
0 0 5
× u(p3 ) A + B γ γµ v(p4 ) . (15.204)
So now we have constructed three spin-1 states using two spin-1/2 states.
There ought to be another, spin-0 state as well. Indeed there is, somewhat
trivially : it is the state
−1 1
v − γ 5 γ µ u+ = u+ γ µ u+ = pµ , (15.214)
2m 2m
496 August 31, 2019
the only object with one Lorentz index orthogonal to the eµ ’s. That it has
zero spin can be seen by lowering26 which turns v − γ 5 γ µ u+ into v − γ 5 γ µ u− =
−u+ γ µ u− = 0.
That this state has the properties of a spin-1/2 as well as of a spin-1 particle
can be seen from
(/p − m) uµ3/2 = 0 , pµ uµ3/2 = 0 . (15.216)
There is one more property : again using Eq.(15.209) we see that also
γµ uµ3/2 = 0 . (15.217)
The properties (15.216) and (15.217) are the so-called Rarita-Schwinger con-
ditions. Now, repeated application of the lowering operator (and normaliza-
tion) give us, successively :
1
uµ1/2 = √ u− v + γ µ u+ + u+ v − γ µ u+ + u+ v + γ µ u−
24m2
1 √
= √ u− eµ+ + 2u+ eµ0 ,
3
1
uµ−1/2 = u− v − γ µ u+ u− v + γ µ u− + u+ v − γ µ u−
√
24m2
1 √
= √ u+ eµ− + 2u− eµ0 ,
3
1
uµ−3/2 = √ u− v − γ µ u− = u− eµ− . (15.218)
8m 2
These three states also obey the Rarita-Schwinger conditions. Getting the
spin sum is now less trivial than before. It helps to introduce the shorthand
[αρβστ ] = uα uα γ 5 γ ρ uβ uβ γ 5 γ σ uτ uτ , α, β, τ = ±
26
And, conversely, also by raising.
August 31, 2019 497
µ 1 √ µ µ
1 √
µ µ µ
w1/2 =√ 2 u− e+ − u+ e0 , =√ 2 u+ e− − u− e0 ,
w−1/2
3 3
(15.221)
µ
are orthogonal to the four u ’s. Now, note that
√ 1
2 u− eµ+ − u+ eµ0 = (u− v + γ µ u+ − u+ v − γ µ u+ )
2m
1
u− u− γ 5 γ µ u+ + u+ u+ γ 5 γ µ u+
=
2m
1
= (/p + m)γ 5 γ µ u+ , (15.222)
2m
498 August 31, 2019
so, indeed
X 1
wjµ wνj = (/p + m)γ 5 γ µ (/p + m)γ 5 γ ν (/p + m) . (15.223)
12m2
j=±1/2
M− u+ = ~ u− , v + M− = ~ v − . (15.225)
This implies that, here, the lowering and raising operators must be
~ ~
M− = (u− u+ − v+ v − ) , M+ = M − = (u+ u− − v− v + ) .
2m 2m
(15.226)
The spin operator in the s direction is then
1 ~
Ms = (M+ M− − M− M+ ) = p/ γ 5 s/ . (15.227)
2~ 2m
Since Ms M+ = −M+ Ms = −~M+ /2 we also have [Ms , M+ ] = ~M+ as we
ought to. And the operator for total spin, finally, is
1 3~2
M2 = (M+ M− + M− M+ ) + Ms 2 = . (15.228)
2 4
August 31, 2019 499
where θ is the angle between ~q1 and ~q2 . The allowed phase space points
therefore obey
2w(q10 + q20 ) − w2 + 2q10 q20 + m23 − m21 − m22 ≤ 2|~q1 ||~q2 | . (15.231)
What we do know are the lower and upper limits on the energies :
1
mj ≤ qj0 ≤ w2 + m2j − (mk + m` )2 = qjmax , {j, k, `} = {1, 2, 3} .
2w
(15.232)
Armed with this knowledge we can use (pseudo-)random numbers ρ, uni-
formly and independently distributed in the interval [0, 1], to sample the
final-state momenta29 . First, we repeat qj0 ← mj + ρ qjmax − mj for j = 1, 2
until Eq.(15.231) is satisfied : typically, one has to try only a few times. This
gives us |~q1 | and |~q2 |, and cos(θ) from Eq.(15.230). We then put
~q1 ← (0, 0, |~q1 |) , ~q2 ← (0, |~q2 | sin(θ), |~q2 | cos(θ)) , ~q3 ← −~q1 − ~q2 .
This gives us the final-state momenta, sampled uniformly. For the initial-
state momenta we generate a uniformly distributed solid angle by
cos(η) ← −1 + 2ρ , φ ← 2π ρ
p~1 ← (|~p1 | sin(η) sin(φ), |~p1 | sin(η) cos(φ), |~p1 | cos(η)) , p~2 ← −~p1
The fact that ~q1 always points in the positive z direction is not a problem in
any theory that respects angular momentum conservation, since the incoming
momenta will take on all directions. You might want to have the direction
of p~1 fixed instead : in that case a global rotation on all the momenta will
do the trick.
This is, of course, only a phase convention, where the phase choice is not
explicit but implied by the choice of k0 , k1 and the complex phase of u− (k0 ).
Now, let us apply γ 5 to these states. It is easy to see that
helicity) :
+ → + , − → − −
+ → − + , − → − . (15.235)
As we can see, the effect of CPT on any diagram is not only to interchange
initial and final states, but also to reverse the arrows on intermediate fermion
lines.
At first sight, these CPT transforms look nothing like the original. Note,
however, that using the standard form we can write them as traces :
whereas
Keeping track of which terms in these traces actually survive34 , we see that,
appearances notwithstanding,
and so on.
D= (15.242)
q/ → − q/ , (15.245)
The arguments given in the previous section show that this evaluates again
to Aµν itself. Finally, for the polarisation vectors we have, for instance,
DCPT = , (15.248)
two vertices µ µ
_ and
e e+
µ
are both assigned the value iQγ /~. More poignantly, in the electroweak
sector we use the same vertex for
+ _
W W
and D
U D U
that the ‘vacuum’ state is itself simply not Lorentz invariant since there is a
‘preferred momentum’.
Z∞ Z2π Z∞
2 2 2
dr r exp −r2
G = dx dy exp − x + y = dφ
−∞ 0 0
Z∞ Z∞
dr r exp −r2
= 2π = π ds exp(−s) = π . (15.250)
0 0
43
As an example, we can use, for the kinetic part of a Lagrangian, the object f µν ∂µ ϕ ∂ν ϕ
rather than the usual g µν ∂µ ϕ ∂ν ϕ.
August 31, 2019 507
Thus we find
x2 √ x2 √
Z Z
dx exp − 2 = σ 2π , dx exp ±i 2 = σ π(1 ± i) ,
2σ 2σ
(15.251)
44
where the last result is obtained by analytic continuation .
where D`+1 /n` is the sum of the (` + 1)-loop vacuum diagrams for the above
action, including the disconnected ones. We can make life simpler by concen-
trating only on the sums of the connected vacuum diagrams, C`+1 /n` , and
then we write
1 X C`+1
log(n!) ≈ (n + 1/2) log(n) − n + log(2π) + `
. (15.256)
2 `≥1
n
44
And inspection of cos(x2 ) and sin(x2 ) which shows that these both integrate to positive
numbers.
508 August 31, 2019
and f (x) g(x) if g(x) ≺ f (x). Obviously, f (x)+g(x) ∼ f (x) if f (x) g(x).
Let us now consider the product f (x)g(x) :
X X n
f (x)g(x) = xn fk gn−k . (15.260)
n≥0 k=0
Because the coefficients grow so fast, the sum over k is (for n → ∞) com-
pletely dominated by gn + fn , and we conclude that
f (x)g(x) ∼ g(x) + f (x) . (15.261)
For instance, we find 1/f (x) ∼ −f (x). Also, for finite p, f (x)p ∼ pf (x).
Asymptotic pullbacks
Consider two asymptotic sums f (x) and g(x) as above, and assume that
f (x) g(x). We want to study the asymptotic form of the pullback f (x g(x)),
that is, an asymptotic sum of powers of an asymptotic sum ; at first sight
this would seem hopelessly tremendous. However, we can write45
X
f (x(g(x)) = fn xn g(x)n
n≥0
2
X
n n 2
= x fn + fn−1 (ng1 ) + fn−2 g1 + ng2
n≥0
2!
3
n 3 2
+fn−3 g1 + n g1 g2 + ng3 + · · · (15.262)
3!
Discarding the 1/n corrections, the leading behaviour of this sum is
n
X X nk
f (xg(x)) ∼ xn fn−k g1 k , (15.263)
n k=0
k!
Asymptotic inversion
The equation xf (x) = y can (at least formally) be inverted to give x as a
function of y. We can write
X fn+1
xf (x) = y ⇒ x = y+x(1−f (x)) = y−f1 x2 h(x) , h(x) = xn .
n≥0
f1
(15.265)
As explained in appendix 15.15.12, the Lagrange expansion is basically just
the iteration of this equation, and we can (formally) write
X 1 d n−1 n
x = y+ −f1 y 2 h(y)
n≥1
n! dy
n−1 X
X 1 n d fm+1 m+2n
∼ (−f1 ) y
n≥1
(n − 1)! dy m≥0
f 1
XX 1
∼ − (−f1 )n−1 fm+1 mn−1 y m+n+1
n≥1 m≥0
(n − 1)!
k n−1
XX
k 1 f1
∼ −y fk y −
k≥0 n=1
(n − 1)! cf
∼ −yf (y) exp(−f1 /cf ) . (15.266)
It follows that H2m H2n if m > n. That in its turn implies G2n =
H2n /H0 ∼ H2n and C2n ∼ G2n , leading to the counter-intuitive result that
at very high loop order, almost all diagrams are connected ! Moreover, in low
order, we have C2 = 1 − u/2 + O (u2 ) and C4 = −u + 7u2 /2 + O (u3 ) so that
û = u(1 − 5u/2 + O (u2 )), and û(u) ∼ −H4 (u) up to an overall factor. This
gives us u = û(1+5û/2+O (û2 )), giving g1 = 5/2 in Eq.(15.264). In addition,
Eq.(15.267) tells us that cf = −2/3, where cf is defined in Eq.(15.258). Since
Hp (u) û(u) for p = 6, 8, 10, . . . we then find
Z∞
Γ(z) = dt tz−1 exp(−t) , (15.269)
0
but that works only if <(z) > 0. A slightly less well-known but very useful
representation, that works for almost all z, is a special limit46 :
n
Y
Γ(z) = lim n! nz (k + z)−1 . (15.270)
n→∞
k=0
(−)m
Γ(z) ≈ . (15.272)
m! (z + m)
46
Found by — who else ? — Euler.
512 August 31, 2019
Performing the integral over t and then integrating with respect to z yields
1 X Bm
ψ(z) = log(z) − − z −m , (15.279)
2z m≥2 m
where the constant of integration vanishes as we can check from the Stirling
approximation of section 15.15.2. Finally, from the fact that t/(et − 1) + t/2
is actually symmetric in t we can see that Bm vanishes for odd m ≥ 3.
for all test functions f (x). Viewed as some kind of function it therefore has
properties analogous to those in Eq.(15.281) :
Z∞
δ(x) = 0 for x 6= 0 , dx δ(x) = 1 . (15.283)
−∞
49
Colloquially, the Dirac delta function, but it is really a distribution in the sense of
Schwartz.
50
A test function has compact support and is inifinitely many times differentiable :
simplistically, it is a nice function.
514 August 31, 2019
The tiny region where the singularity resides is cut away. If we view distri-
butions as limiting cases of functions, we can write
1 1
lim ↓0 = P + iπ δ(x) . (15.287)
x − i x
If we have n − 1 objects distributed into k groups, we can let the nth object
form its own group, or add to one of the existing groups in k different ways.
This gives us the recursion
1 z k
φk (z) = e −1 , (15.292)
k!
so that
X zn X z −1)
B(n) = φk (z) = e(e . (15.293)
n≥0
n! k≥0
Z∞
exp(−y)
E1 (z) = dy . (15.294)
y
z
516 August 31, 2019
The modified Bessel function of the second kind, of order ν, can be defined
by its integral representation :
Z∞
1 ν−1 z 1
Kν (z) = dx x exp − x+ , (15.296)
2 2 x
0
We can see (by x → 1/x) that Kν (z) = K−ν (z). For small z and ν > 0, the
integral will be dominated by the large-x region, and we can approximate
Z∞ ν
1 ν−1
zx 1 2
Kν (z) ∼ dx x exp − = Γ(ν) . (15.297)
2 2 2 z
0
51
In the penultimate line of Eq.(15.295) we use the fact that, for n → ∞, the last
sum over k is dominated by its highest terms, and for large k the binomial coefficient is
approximately k ` /`!.
August 31, 2019 517
Im z
Γ
Re z
x
since also f (z)/(z − x) is analytic on, and inside, Γ. Splitting the integral
into its various contributions, we therefore have
Zx− Z+∞ I
f (z) f (z) 1 f (z)
0= dz + dz − dz , (15.306)
z−x z−x 2 z−x
−∞ x+ z∼x
where we have assumed that the big half-circle does not contribute since
f (z)/(z − x) vanishes fast enough. The number is infinitesimal, and the
sum of the first two terms is called the principal value integral :
x−
Z+∞ Z Z+∞
f (z) f (z) f (z)
P dz ≡ lim dz + dz . (15.307)
z−x →0 z−x z−x
−∞ −∞ x+
and by inspecting the real and imaginary parts separately we arrive at the
Kramers-Kronig relations
Z+∞ Z+∞
1 =f (z) −1 <f (z)
<f (x) = P dz , =f (x) = P dz . (15.309)
π z−x π z−x
−∞ −∞
where the integration contour should not cross the cut in the logarithm (this
is usually chosen to be the real axis at z values larger than 1). Figure 15.6
shows the function Li2 (z + iη) for real values of z. Along to the cut we have
which is handy for evaluating the dilogarithm for small arguments. We also
have Li2 (1) = π 2 /6 (cf section 15.15.11). There are a number of useful
identities for the dilogarithm53 , of which we give several below. You can
prove them by differentiating the left-hand and right-hand sides with respect
to z, and additionally checking them at z = 0 or z = 1 or so.
1
Li2 (z) + Li2 (−z) = Li2 (z 2 ) ,
2
π2
Li2 (z) + Li2 (1 − z) = − log(z) log(1 − z) ,
6
π2 1 2
Li2 (z) + Li2 (1/z) = − − log(−z) ,
6 2
53
A word of caution ! Sometimes the definition of the dilogarithm differs from the one
given here. In particular, in the computer-algebra system MAPLE the function dilog(z)
corresponds to our Li2 (1 − z).
August 31, 2019 521
1 π2
Li2 (−z) − Li2 (1 − z) + Li2 1 − z 2 = − − log(z) log(1 + z) ,
2 12
1 2
Li2 (1 − z) + Li2 (1 − 1/z) = − log(z) . (15.313)
2
With all these relations, small wonder that expressions involving dilogarithms
can be made to look bewilderingly different ! Some other special values can
also be derived using the above identitites :
π2 1 2
Li2 (0) = 0 , Li2 (1/2) = − log(2) ,
12 2
π2 π2
Li2 (−1) = − , Li2 (2) = − iπ log(2) . (15.314)
12 4
As for the numerical evaluation of the dilogarithm function : by applying
Eqs.(15.313) we can always arrange for the argument z to have |z| ≤ 1 and
0 ≤ <(z) ≤ (1/2) ; and then we can judiciously change the integration
variable from x to y = − log(1 − x), to see that
− log(1−z)
Z
y X Bn n+1
Li2 (z) = dy = − log(1 − z) , (15.315)
ey − 1 n≥0 (n + 1)!
0
and the last term of this expression cancels against the last term in Eq.(15.322).
We are left with the following :
X 1 ∂ n−1
ξ = x+ f (x)n . (15.323)
n≥1
n! ∂x
ξ = x ,
ξ = x + f (x) ,
ξ = x + f (x + f (x)) ≈ x + f (x) + f 0 (x)f (x) ,
ξ = x + f (x + f (x) + f (x)f 0 (x))
1
≈ x + f (x) + f (x)f 0 (x) + f (x)f 0 (x)2 + f (x)2 f 00 (x) (15.324)
2
and so on, assuming f (x) and its derivatives to be small enough to warrant
Taylor expansion. This reproduces the Lagrange expansion.
1
ξ = x + ξ2 . (15.325)
2
The solution that vanishes with x is
√
ξ =1− 1 − 2x . (15.326)
524 August 31, 2019
where X
Rp = λj1 λj2 · · · λjp , (15.332)
1≤j1 <j2 <···<jp ≤n
526
August 31, 2019 527
vacuum bubbles
absence, 46
definition, 42
vacuum polarization
hadronic, 326
leptonic, 324
vertices, 37
zero-dimensional space, 29