Solutions Manual For A. Zee, Quantum Field Theory in A Nutshell, 2 Ed
Solutions Manual For A. Zee, Quantum Field Theory in A Nutshell, 2 Ed
Zee,
Quantum Field Theory in a Nutshell,
2nd Ed.
Yoni BenTov
The publisher would like to acknowledge the author of this volume for providing the camera-
ready copy from which this book was printed.
IX Part N 302
IX.1 N.2 Gluon Scattering in Pure Yang-Mills Theory . . . . . . . . . . . . . . . . 302
IX.2 N.3 Subterranean Connections in Gauge Theories . . . . . . . . . . . . . . . 311
These are the solutions to the problems in Quantum Field Theory in a Nutshell, 2nd Ed.,
that are not already solved in the back of the book. For most of the problems I provide a
detailed solution, while for others I sketch the solution and provide a reference to the litera-
ture for further details. Some problems are intentionally open-ended, serving as more of an
introduction to the literature rather than as a homework assignment. The goal of Zee’s text
is not only to teach quantum field theory but also to facilitate the transition from student to
researcher.
I thank my colleagues and teachers for helpful discussions in preparing these solutions. In
particular, I have benefitted from talking to Jen Cano, Gavin Hartnett, Kurt Hinterbichler,
Josh Ilany, Yonah Lemonik, Eugeniu Plamadeala, Yinbo Shi, Joe Swearngin, Benson Way
and Chiu-Tien Yu. I am grateful to Joshua Feinberg for providing some of the solutions. I
am also indebted to the faculty at the University of California at Santa Barbara for their pa-
tience in answering my questions. In particular, I thank David Berenstein, Andreas Ludwig,
Ben Monreal, Joe Polchinski, Mark Srednicki, and my thesis adviser, Tony Zee.
Finally, I thank my editor Ingrid Gnerlich at Princeton University Press for providing the
opportunity to undertake this project. I also thank my friends and family for their support
during the completion of this work.
Yoni BenTov
Apr. 9, 2012
v
I Motivation and Foundation
I.2 Path Integral Formulation
1. Verify (5) Z RT
−iHT dt[ 12 mq̇ 2 −V (q)]
hqF |e |qI i = Dq e i 0 (5)
Solution:
1 2
Start with the Hamiltonian Ĥ = 2m p̂ + V (q̂). Let T = N , with N → ∞, → 0, T
fixed. In this way, split up the time evolution operator into N pieces:
Use this decomposition in the transition amplitude hqF |e−iĤT |q0 i, and insert one copy of the
identity matrix between each pair of e−iĤ s:
The superscript is just a label to keep track of the fact that we have inserted N − 1 identity
matrices. It is convenient to resolve each identity matrix in a complete set of position
eigenstates: Z ∞
1(i) = dqi |qi ihqi |
−∞
1
So to O(2 ) → 0 we can perform the above manipulations.
1 2
Now insert an identity matrix to the right of e−i 2m p̂ , but this time resolve it in a com-
plete set of momentum eigenstates:
Z ∞
dp
1= |pihp|
−∞ 2π
1 2
hqi |e−iĤ |qj i = e−iV (qj ) hqi |e−i 2m p̂ 1 |qj i
Z ∞
−iV (qj ) 1 2
−i 2m p̂ dp
=e hqi |e |pihp| |qj i
−∞ 2π
Z ∞
−iV (qj ) dp 1 2
=e hqi |e−i 2m p̂ |pihp|qj i
2π
Z−∞ ∞
dp −i 1 p2
= e−iV (qj ) e 2m hqi |pi hp|qj i
−∞ 2π | {z } | {z }
eiqi p e−iqj p
Z ∞
dp −i 1 p2 +ip(qi −qj )
= e−iV (qj ) e 2m
−∞ 2π
Now do all of this for the original transition amplitude. This will require N − 1 resolutions
of the identity in position eigenstates, as already indicated, and it will require N resolutions
of the identity in momentum eigenstates.
2
∞
dpN −1 −i p2N −1 +ipN −1 (qF −qN −1 )
Z
−iĤ −iV (qN −1 )
• hqF |e |qN −1 i = e e 2m
−∞ 2π
Z ∞
−iĤ −iV (qN −2 ) dpN −2 −i p2N −2 +ipN −2 (qN −1 −qN −2 )
• hqN −1 |e |qN −2 i = e e 2m
−∞ 2π
...
Z ∞
−iĤ −iV (q1 ) dp1 −i p21 +ip1 (q2 −q1 )
• hq2 |e |q1 i = e e 2m
2π
Z−∞
∞
dp0 −i p20 +ip0 (q1 −q0 )
• hq1 |e−iĤ |q0 i = e−iV (q0 ) e 2m
−∞ 2π
Note that we need N resolutions into momentum eigenstates instead of just N − 1: after
inserting N − 1 resolutions of the identity into position eigenstates, we need one set of
momentum eigenstates for each position “ket,” |q0 i, ..., |qN −1 i, including the initial position
q0 over which we do not integrate. In any case, the amplitude is now:
hqF |e−iĤT |q0 i =
2 2
Z ∞ p p0
dp0 dpN −1 −i N2m−1 −pN −1 (qF −qN −1 ) −i 2m −p0 (q1 −q0 )
dq1 ...dqN −1 ... e ... e e−iV (qN −1 ) ...e−iV (q0 )
−∞ 2π 2π
Completing the square for each term in brackets will yield one-dimensional Gaussian integrals:
p2i
2 2m
− pi (qi+1 − qi ) = p − pi (qi+1 − qi )
2m 2m i
m 2 m2
2
= pi − (qi+1 − qi ) − 2 (qi+1 − qi )
2m
2
m 2 m qi+1 − qi
= pi − (qi+1 − qi ) −
2m 2
3
Remember that the limits N → ∞ and → 0 with T fixed are supposed to be implied
throughout. With this in mind, define the integral over paths and go to a continuum notation:
Z Z r N
N −1 m
Dq ≡ lim d q
q(0) = q0 2πi
N →∞
q(T ) = qF
→0
T fixed
−1
N
" 2 # " 2 #
T
qj+1 − qj
Z
X m m dq
− V (qj ) → dt − V (q)
j=0
2 0 2 dt
2. Derive (24) X
hxi xj ...xk x` i = (A−1 )ab ...(A−1 )cd (24)
Wick
Solution:
4
In general, for an n-point function, we have:
∂ ∂ ∂ ~ |~
hxi1 xi2 ... xin i = ... Z[J] J=0
∂Ji1 ∂Ji2 ∂Jin
Take the case n = 4 to see how this works. To clean up the notation, define G ≡ M −1 and
use the following shorthand:
N X
X N
J · G · J ≡ J~ T GJ~ = Jα Gαβ Jβ ≡ Jα Gαβ Jβ
α=1 β=1
∂ + 1 J·G·J 1 1 ∂
e 2 = e+ 2 J·G·J (Jα Gαβ Jβ )
∂Ji4 2 ∂Ji4
1 1
= e+ 2 J·G·J (δi4 α Gαβ Jβ + Jα Gαβ δβi4 )
2
+ 21 J·G·J
=e Gi4 α4 Jα4
The last line follows from the symmetry Gαβ = Gβα . Also, we relabeled the dummy index
to α4 just to associate it with i4 for later convenience. Before proceeding, note that setting
J = 0 here gives zero, which means that hxi4 i = 0. Now take another derivative, this time
with respect to Ji3 :
∂2 1 ∂ + 1 J·G·J
e+ 2 J·G·J = e 2 Gi4 α4 Jα4
∂Ji3 ∂Ji4 ∂Ji3
+ 21 J·G·J ∂ 1 1 ∂
=e Jα Gαβ Jβ Gi4 α4 Jα4 + e+ 2 J·G·J Gi4 α4 Jα
∂Ji3 2 ∂Ji3 4
1
= e+ 2 J·G·J (Gi3 α3 Jα3 Gi4 α4 Jα4 + Gi4 i3 )
Again, before proceeding try setting J = 0. This time, a non-zero piece is left over. We see
therefore that hxi3 xi4 i = Gi4 i3 = (M −1 )i4 i3 . Now differentiate with respect to Ji2 :
∂3 1 ∂ h + 1 J·G·J i
e+ 2 J·G·J = e 2 (Gi3 α3 Jα3 Gi4 α4 Jα4 + Gi4 i3 )
∂Ji2 ∂Ji3 ∂Ji4 ∂Ji2
1 1
= e+ 2 J·G·J (Gi2 α2 Jα2 ) (Gi3 α3 Jα3 Gi4 α4 Jα4 + Gi4 i3 ) + e+ 2 J·G·J (Gi3 i2 Gi4 α4 Jα4 + Gi3 α3 Jα3 Gi4 i2 )
5
As in the case for just one derivative, setting J = 0 gives 0, which means that hxi2 xi3 xi4 i = 0.
You can now see a general result for this generating function:
Having taken enough derivatives to see how this works, while taking the next and final
derivative, with respect to Ji1 , keep only the terms that have no powers of J left over, so
that you don’t have to bother keeping track of terms that will go to zero anyway. This gives:
∂4 1
e+ 2 J·G·J |J=0 = Gi2 i1 Gi4 i3 + Gi3 i2 Gi4 i1 + Gi3 i1 Gi4 i2
∂Ji1 ∂Ji2 ∂Ji3 ∂Ji4
On the right-hand side, all possible pairings of the indices {i1 , i2 , i3 , i4 } appear exactly once.
Solution:
d4 k e ik·x
Z
D(x) =
(2π)4 k 2 − m2 + iε
we see that the integral over k 0 can be obtained by the contour integral
0
e izx
I
dz
I≡
C 2π z 2 − (E − iε)2
with simple poles at z = ±(E − iε). For x0 > 0 we can use a semicircular contour in the
upper half plane to pick up the residue at z = −E + iε and obtain
Z ∞ 0 0 0
! 0
dk e ik x e izx e−iEx
I= 0 2 2
= i lim = −i , x0 > 0 .
−∞ 2π (k ) − (E − iε) z→ −E z−E 2E
6
For x0 < 0 we can use a semicircular contour in the lower half plane to pick up the residue
at z = +E − iε and, being careful to take into account the orientation of the contour, we
obtain
Z ∞ 0 0 0
! 0
dk e ik x e izx e+iEx
I= 0 )2 − (E − iε)2
= (−1)i lim = −i , x0 < 0 .
−∞ 2π (k z→+E z + E 2E
7
2. Work out the propagator D(x) for a free field theory in (1+1)-dimensional spacetime and
study the large x1 behavior for x0 = 0.
Solution:
d2 k e ik·x
Z
D(x) = .
(2π)2 k 2 − m2 + iε
where we have used the definition of the modified bessel function from problem I.3.1. For
mr 1, we obtain r
1 π −mr
D(x, 0) ≈ −i e ∼ e−mr
2π 2mr
as in (3+1) dimensions. In contrast to (3+1) dimensions, for m → 0 we have
K0 (mr) ≈ − ln(mr/2) − γ
where γ ≈ 0.577.
d4 k e ik(x−y)
Z
Dadv (x − y) =
(2π)4 k 2 − m2 − i sgn(k0 )ε
is nonzero only if x0 > y 0 . In other words, it only propagates into the future. [Hint: Both
poles of the integrand are now in the upper half of the k0 -plane.] Incidentally, some authors
prefer to write (k0 − iε)2 − ~k 2 − m2 instead of k 2 − m2 − isgn(k0 )ε in the integrand. Similarly,
show that the retarded propagator
d4 k e ik(x−y)
Z
Dret (x − y) =
(2π)4 k 2 − m2 + i sgn(k0 )ε
8
Solution:
Now evaluate this integral using the complex plane. Define the complex variable z ≡ k 0 + i β,
where β is real, and consider the following integral:
0
eizx
Z
0 dz
I(x ) ≡
c 2π [z − (ω + i sgn(Re{z}))] [z + (ω + i sgn(Re{z}))]
So far the contour c is unspecified.R The intention is to provide a contour such that integration
along the real axis will yield the dk 0 that we’re actually interested in evaluating, and that
the contributions along the imaginary parts of the complex plane will evaluate to zero.
So now we’ve distinguished the two cases: for x0 > 0, we must close the contour in the
upper half-plane, and for x0 < 0, we must close the contour in the lower half-plane. The
relevant contours for this problem, which will yield the desired real-valued integral, are shown
below:
9
Im z
C>
Re z
Im z
Re z
C<
c> is what we use for x0 > 0, and c< is what we use for x0 < 0. If you evaluate the integrand
for the piece of either c> or c< that is on the real axis, you will recover the integral you
are actually interested in evaluating (Re{z} = k 0 ). The ×s in the above diagram indicate
the locations of the poles of the integrand. A pole is where the integrand goes to infin-
ity, which in this case corresponds to a zero of the denominator. The pole for Re{z} > 0 is
z = ω +i sgn(Re{z})ε = ω +iε. The pole for Re{z} < 0 is z = −ω −i sgn(Re{z})ε = −ω +iε.
So both of the poles have positive imaginary parts and are therefore in the upper half-plane.
The residue theorem says that if you choose a contour that encircles the locations of the
poles of the integrand, then you will get some nonzero number for the integral. If, however,
your contour does not encircle any of the poles, then you will get zero. For the purposes of
this problem all you have to do is decide whether the integral is zero.
2. If x0 > 0, we close the contour in the upper half-plane. If x0 < 0, we close in the lower
half-plane.
Therefore, the integral is nonzero if x0 > 0, and the integral is zero if x0 < 0. Therefore, the
advanced propagator is only nonzero if x0 > 0.
10
For the retarded propagator, everything is the same except the poles of the integrand are in
the lower half-plane. So the retarded propagator is only nonzero if we close the contour in
the lower half-plane, which we can only do when x0 < 0.
Solution:
Taking the analog of equation (1) on page 24 of the book, we have in (d + 1) dimensions:
Z Z
iW [J] 1
Z[J] ∝ e , W [J] = − dd+1 x dd+1 y J(x)D(x − y)J(y)
2
dd+1 k −eik(x−y)
Z
D(x − y) =
(2π)d+1 −k 2 + m2 − i
As in the book, consider J(x0 , ~x) = δ (d) (~x − ~x1 ) + δ (d) (~x − ~x2 ), and look only at the cross
term ∼ δ (d) (~x − ~x1 )δ (d) (~x − ~x2 ) in W [J]:
W [J]cross
0 0 0 ~
d k dk 0 −eik (x −y ) e−ik·(~x−~y) (d)
Z Z Z Z Z d Z
0 d 0 d
= − dx d x dy d y δ (~x − ~x1 )δ (d) (~x − ~x2 )
(2π) d 0 2 ~
2π −(k ) + |k| + m − i
2 2
0 0 0 ~
dd k dk 0 −eik (x −y ) e−ik·(~x1 −~x2 )
Z Z Z Z
0 0
= − dx dy
(2π)d 2π −(k 0 )2 + |~k|2 + m2 − i
0 0 ~
dd k dk 0 −eik x e−ik·(~x1 −~x2 )
Z Z Z Z
0 0 −ik0 y 0
= − dx dy e
(2π)d 2π 0 2 ~2 2
| {z } −(k ) + |k| + m − i
2πδ(k 0 )
~
dd k −e−ik·(~x1 −~x2 )
Z Z
0
=− dx
(2π)d |~k|2 + m2 − i
~
dd k e−ik·(~x1 −~x2 )
Z
= +T
(2π)d |~k|2 + m2
11
Now set m to zero and use dimensional analysis:
Z ikr Z Z
d−1 e d−3 ikr du u d−3 iu 1
E ∼ dk k 2
= dk k e = e ∼ d−2
k r r r
The notation is intended to extract the dimensional information only; for instance, ~k · ~x =
kr cos θ, etc, but for the purposes of this problem the angular information is not important.
For d = 3, the above gives E ∼ 1/r, as expected.
However, for d = 2 this gives E ∼ r0 = 1, which may naively suggest that the force be-
tween the sources is zero, but that is incorrect. Define r ≡ |~x1 − ~x2 | and consider the force
between the sources:
~
d2 k e−i k·(~x1 −~x2 )
Z
dE d
F =− =+
dr dr (2π)2 |~k |2
Z ∞ Z 2π
1 d e−ikr cos θ
= 2 k dk dθ
4π dr 0 k2
Z ∞ Z 2π 0
1 dk
= 2 dθ e−ikr cos θ (−ik cos θ)
4π 0 k 0
Z ∞ Z 2π
−i
= 2 du dθ e−iu cos θ cos θ
4π r 0 0
1
∼
r
The force is not zero; it goes as 1/r. The potential energy is therefore E ∼ ln r. However
this is still not quite right, since the argument of the log must be dimensionless. So really
we get E ∼ ln(r/r0 ) for some cutoff r0 , which explains our naive estimate that the potential
energy is dimensionless. Let’s study the properties of this cutoff.
R 2π
The integral 2π1
0
dθ e−ikr cos θ = J0 (kr), where J0 (z) is the Bessel function of the first
kind of order 0 as a function of z. We may now use Mathematica’s Bessel function R input
BesselJ[n, z] ≡ Jn (z) to plot the integral. Plotting the dimensionless function du J0 (u)/u
gives:
12
Ù âu J0 HuLu
0.1
u
2 4 6 8 10
-0.1
-0.2
Identifying u = kr, we see that the integral over k is uneventful at high k but diverges as
k → 0. This suggests we regularize the integral by imposing a cutoff kmin ≡ 1/a on the lower
bound of integration. So really, the potential energy is:
Z ∞
dk 2π
Z
1
E(r) = − dθ e−ikr cos θ
4π 1/a k 0
Remember that in our units of ~ = c = 1, momentum has units of inverse length. If we take
the potential energy as written above and change variables to u = kr, then the potential
energy is: Z ∞
du 2π
Z
1
E(r) = − dθ e−iu cos θ
4π r/a u 0
Now we see that the expression for the potential energy is more sensible: the reason we naively
guessed that E does not depend on r is that we had improperly chosen the lower bound of
the k integral to be 0. We also now see what the r0 is in E ∼ ln(r/r0 ): r0 = a = 1/kmin is
the inverse of the minimum momentum, or equivalently the maximum separation between
the particles that we are willing to consider.
R∞ R 2π
Finally, the double integral in the force is 0
du 0
dθ e−iu cos θ cos θ = −2π i, so the force
between the particles is
1
F =− .
2πr
13
I.5 Coulomb and Newton: Repulsion and Attraction
P (a) (a)
1. Write down the most general form for a εµν (k)ελσ (k) using symmetry repeatedly. For
example, it must be invariant under the exchange {µν ↔ λσ}. You might end up with
something like
Solution:
The ingredients we have are kµ and Gµν (k) ≡ gµν − m12 kµ kν . G has the properties that
Gµν = Gνµ and k µ Gµν = k µ gµν − m12 kµ kν = kν − m12 (m2 )kν = 0.
(a)
Each polarization tensor εµν is symmetric under interchange of µ and ν, so we need fµνλσ =
fνµλσ and fµνλσ = fµνσλ . This leads us to write:
However, because the components of the polarization tensors are just ordinary numbers, we
know that εµν ελσ = ελσ εµν , which means that f should be invariant under the interchange
of pairs of indices, (µν) ↔ (λσ): fµνλσ = fλσµν . This sets C̃ = C.
Another thing we can do is stagger the indices across Gs, by which I mean something involv-
ing the term Gµλ Gνσ . If we take that and interchange µ and ν, we get Gνλ Gµσ . We therefore
have another term to add to f :
We can also use the same idea but instead of using two Gs, use two ks and one G, such as
kµ kλ Gνσ . If we take that and interchange µ and ν, we get kν kλ Gµσ . If we then interchange λ
and σ, we get kν kσ Gµλ . Interchanging µ and ν again gives us something different: kµ kσ Gνλ .
These four terms should be included in f :
Finally, there is the possibility of multiplying together four momenta to yield a term fµνλσ ⊃
14
E kµ kν kλ kσ . We therefore have:
Now we need to fix the constants A, B, C, D and E. To do that, we use the fact that
k µ εµν (k) = 0, which implies that k µ fµνλσ = 0. Recalling that k µ Gµν = 0 and k µ kµ = m2
gives the following:
To proceed we recall the other trace-free condition on the polarization tensor, namely that
g µν εµν = 0. In anticipation of multiplying fµνλσ by g µν , we first compute the following:
µν µν 1
g Gµλ =g gµλ − 2 kµ kλ
m
1
= δλν − 2 k ν kλ
m
1
=⇒ g µν Gµν = δνν − 2 k ν kν = 4 − 1 = 3
m
15
2
fµνλσ = B − Gµν Gλσ + (Gµλ Gνσ + Gλµ Gσν )
3
As suggested in the book, fix the constant B by imposing the normalization f1212 = 1 for all
k. If this is supposed to hold for all k, then try k = 0. Since Gij (k = 0) = 1 for i, j = 1, 2, 3,
we get:
Solution:
To clarify what this question is asking: The original point of this compact extra dimen-
sion setup is to propose that the scale of quantum gravity in the (n+3+1)-dimensional space
is actually similar to the weak scale, so that our guessing that the Planck mass is so high is
just an artifact of measuring in (3+1) dimensions.
So, really, we are interested in answering the question: “Can one extra dimension fix the
hierarchy problem?”
With that in mind, we set n = 1 and try MPl(5) ∼ 102 GeV. Using MPl ∼ 1019 GeV,
we predict a value for the characteristic length scale of the proposed 5th extra dimension:
2 2
102
1 MPl(5)
= MPl(5) 2
∼ 10 GeV = 10−32 GeV = 10−23 eV ∼ 10−42 J
R MPl 1019
R 1042 1042
10−26 J · m = 1016 m ∼ 1 parsec
R→ ∼ =⇒ R ∼
~c J J
We conclude that adding only one extra dimension is not a valid way to fix the hierarchy
problem.
16
I.7 Feynman Diagrams
1. Work out the amplitude corresponding to figure I.7.11 in (24).
Solution:
Since this is a 3-loop amplitude, we will just write down the answer using the Feynman
rules and then get the symmetry factor by looking at the diagram. Using the propagator
i∆(p) = p2 −mi 2 +iε and the vertex −iλ, we have
M=
d4 p d4 q d4 r
Z
1
4 4 4
[−iλ]4 [i∆(p)][i∆(k1 +k2 −p)][i∆(q)][i∆(p−q−r)][i∆(k1 +k2 − r)][i∆(r)]
S (2π) (2π) (2π)
where S is the symmetry factor. Up to the symmetry factor, this matches (24). Symmetry
factors arise in loop diagrams for the following reason. Consider the pictorial representation
of the vertices and of the propagators with sources:
1 A
4 2
B
3
We normalize the vertex as L = − 4!1 λϕ4 so that the permutations of the labels 1, 2, 3, 4 on
the “hands” contribute 4! terms and cancel out the 4!1 in the Lagrangian. We normalize
the kinetic term as L = − 12 ϕ(∂ 2 + m2 )ϕ so that swapping the labels A, B on the “blobs”
contribute 2 terms and cancel out the 12 in the Lagrangian.
Suppose we have a diagram in which hands 1 and 2 eat blobs A and B respectively:
17
1,A
4
2,B
From our previous argument, this diagram takes into account the 2 terms from swapping
hands 1 and 2, and also the 2 terms from swapping blobs A and B, leading to a total of 4
terms.
But swapping 1 and 2 and simultaneously swapping A and B gives exactly the same di-
agram with which we started, so this bookkeeping overcounts by a factor of S = 2.
Now instead of rearranging individual hands on a given vertex, and individual blobs on
a given propagator, consider rearranging the vertices themselves and the propagators them-
selves. Consider the path integral
Z R 4
Z(J) = Dϕ e i d x(L+Jϕ)
∞
" Z 4 #V X∞ Z P
X 1 4 1 1 δ 1 4 4 1
= −i d x λ i d x d y J(x)∆(x − y)J(y) .
V =0
V ! 4! i δJ(x) P =0
P ! 2
The quantity in the first square brackets raised to the power V is the vertex diagram drawn
above, and the quantity in the second square brackets raised to the power P is the barbell
diagram drawn above.
Thus we see (as explained in the main text) that a general diagram is given by pasting
together a bunch of different vertices with a bunch of different propagator lines. The blobs
δ
(the J(x)) on the propagator lines get eaten by the hands (the δJ(x) ) on the vertices.
Just as before, we might think that the V vertices can be arranged in V ! ways, and the
P propagators can be arranged in P ! ways, so that the V !P ! cancel out of the path integral.
But if exchanging two vertices and simultaneously exchanging two propagators gives you the
same diagram you started with, then counting those two swaps separately would overcount
the contribution of that term to the sum in Z(J).
The mismatch between the naive counting the correct counting of matching derivatives to
sources is the symmetry factor. The last thing to take note of is that symmetry factors
18
coming from external blobs are canceled once those blobs are given fixed external values. In
other words, tree diagrams do not have symmetry factors.
Now let us apply this reasoning to figure I.7.11, which we repeat below for convenience:
W X
3 4
1
C A E G
2
D B F H
6 5
Y Z
Here we omit the momentum arrows and labels, since they play no role in counting the sym-
metry factor.1 We label internal lines by numbers, and “hands” on the vertices by letters.
Suppose we swap the vertices labeled by (ABCD) and (EF GH). Can we swap the in-
ternal lines in such a way that we recover the original diagram? The answer is yes: swap
lines 3 ↔ 4 and 5 ↔ 6, and then also the hands W ↔ X and Y ↔ Z. This contributes a
factor of 2 to S.
This is also the same as keeping the vertices fixed, but swapping the hands A ↔ B and
E ↔ F , while also swapping the lines 1 ↔ 2. This contributes another factor of 2 to S.
19
3. Draw all the diagrams describing two mesons producing four mesons up to and including
order λ2 . Write down the corresponding Feynman amplitudes.
Solution:
= −iλ
The 2 → 4 scattering diagrams up to O(λ2 ) are:
1’
1
p 2’
3’
2 4’
1’
1
2’
p
3’
2
4’
There are also the diagrams with different permutations of the outgoing lines, which corre-
spond to different momenta flowing in the internal propagator. Reading the first diagram
from right to left, the corresponding amplitude is:
λ2
−i
iMfirst = [−iλ] [−iλ] = +i
−p2 + m2 − i −p2 + m2 − i
20
In the first diagram, the propagator’s momentum is p = p1 + p2 − p40 . The amplitude for the
second diagram has the same form as the first one, except p = p1 − (p10 + p20 ). So the total
amplitude is, suppressing the iεs:
4 0 4 X 4 0 0
X 2 1 X 1
+ O(λ4 ) .
M=λ
2 2
+ 2 2
n = 10
−(p 1 + p 2 − p n0 ) + m 0 j=10
−(p 1 − p i 0 − p j 0 ) + m
i=1
i0 <j 0
Solution:
dD k h
Z i
−iωk t+i~k·~
x † iωk t−i~k·~
x
ϕ(~x, t) = a~ e + a~k e
(2π)D 2ωk k
Solution:
a. This follows almost immediately from the following property of the delta function:
X 1
δ (F (x)) = δ(x − x∗ ) , where F (x∗ ) ≡ 0
x∗
|F 0 (x∗ )|
For this problem, define F (k 0 ) ≡ (k 0 )2 − ωk2 . The zeros of this function occur at k 0 = ±ωk ,
and the derivative of the function evaluated at these zeros is:
F 0 (k 0 ) = 2k 0 = ±2ωk
21
Therefore,
1
δ F (k 0 ) = δ(k 0 − ωk ) + δ(k 0 + ωk )
2ωk
We may now prove the desired result:
Z Z Z ∞
0 ~
4 µ 2 0
d k δ(k kµ − m )θ(k )f (k , k ) = d k 3
dk 0 δ((k 0 )2 − |~k |2 − m2 )θ(k 0 )f (k 0 , ~k )
−∞
Z Z ∞
= d3 k dk 0 δ(F (k 0 ))θ(k 0 )f (k 0 , ~k )
Z Z−∞∞
1
3 0 0 0
θ(k 0 ) f (k 0 , ~k )
= dk dk δ(k − ωk ) + δ(k + ωk )
−∞ 2ω k | {z }
picks out k0 >0 only
Z Z ∞
1
= d3 k dk 0 δ(k 0 − ωk ) f (k 0 , ~k )
−∞ 2ωk
Z
1
= d3 k f (ωk , ~k )
2ωk
b. Consider Lorentz transformations that do not change the signRof the time component
of a vector. The matrices Λ ≡ ∂x/∂x0 have determinant +1, so d4 k δ(k 2 − m2 )θ(k 0 ) is
unchanged under a Lorentz transformation. The function f (k) has no Lorentz indices and
therefore
R 4 does not transform under Lorentz transformations. Therefore the whole integral
d k δ(k − m2 )θ(k 0 )f (k) is Lorentz invariant.
2
If the integral d3 k 2ω1k f (ωk , ~k ) is to be Lorentz invariant, and if the function f (ωk , ~k ) is
R
Lorentz invariant, then the only conclusion we can make is that the measure d3 k/(2ωk ) is
also Lorentz invariant.
c. The point is to find the new commutation relations if you do not keep the square root in
the definition of the Fourier coefficients a~k and a~†k . Assume the relation [ak , a†p ] = δ(k − p)
p
then shift ak → ak / (2π)D 2ωk to get:
If you like, we can compute this explicitly as well. Using the square-root integration measure,
the Fourier expansion of ϕ(~x, t) is:
dD k
Z h i
~ ~
ϕ(~x, t) = p a~k e−i(ωk t−k·~x ) + a~†k e i(ωk t−k·~x )
(2π)D 2ωk
The canonical commutation relations are between ϕ and π ≡ ∂L/∂ ϕ̇ = ϕ̇, so we also need
the Fourier expansion of π:
22
r
dD k
Z
ωk h −i(ωk t−~k·~
x) † i(ωk t−~k·~
x)
i
π(~x, t) = −i p a~k e − a~k e
(2π)D 2
Now that we’ve taken the time derivative we may as well set t = 0 to simplify the calculation.
We need to change basis from (ϕ, π) to (a, a† ), and to do that we can invert the Fourier
transform:
dD k
Z Z Z h i
~ ~
D
d xe −i~
p·~
x
ϕ(~x, 0) = p dD x a~k ei(k−~p)·~x + a~†k e−i(k+~p)·~x
(2π)D 2ωk
s
(2π)D h (D) ~
Z i
= dD k a~k δ (k − p~) + a~†k δ (D) (~k + p~)
2ωk
s
(2π)D
= ap~ + a†−~p
2ωp
p
Note that ωp = |~p |2 + m2 , so the sign of p~ does not affect ωp . The algebra goes through
in exactly the same way for π except with the relative minus sign:
r
(2π)D ωp
Z
D −i~
p·~
x †
d xe π(~x, 0) = −i ap~ − a−~p
2
s Z
2ωp
ap~ + a†−~p
= dD x e−i~p·~x ϕ(~x, 0)
(2π)D
s Z
† 2
ap~ − a−~p = i dD x e−i~p·~x π(~x, 0)
(2π)D ωp
Organized in this way it’s pretty clear that what we want to do is to add the two equations
and divide by 2. Once we have ap~ we can just Hermitian conjugate it to get ap†~ . Therefore:
Z
1
ap~ = p dD x e−i~p·~x [ωp ϕ(~x, 0) + iπ(~x, 0)]
(2π)D 2ωp
The non-zero equal-time canonical commutation relation is [ϕ(~x, t), π(~y , t)] = iδ (D) (~x − ~y ).
The other commutators are zero. Therefore we can compute [ap~ , a~k ]:
23
Z Z
1 −i~
p·~ ~
[ap~ , a~k ] = D √
D
d xe x
dD y e−ik·~y [ ωp ϕ(~x) + iπ(~x) , ωk ϕ(~y ) + iπ(~y ) ]
(2π) 2 ωp ωk
Z Z
1 −i~
p·~ ~
= √
D
d xe x
dD y e−ik·~y 0 + iωp [ϕ(~x), π(~y )] +iωk [π(~x), ϕ(~y )] +0
D
(2π) 2 ωp ωk | {z } | {z }
(D) (D)
i δ (~x − ~y ) −i δ (~x − ~y )
Z
1 ~
= √ dD x e−i(~p+k)·~x (−ωp + ωk )
(2π)D 2 ωp ωk
1
= √ δ (D) (~k + p~) (−ωp + ωk ) = 0 since ω−~p = ω+~p
2 ω p ωk
To get a non-zero result, we’re going to have to change the sign on one term but not the
other; but that’s exactly what Hermitian conjugation will do for us. So, without spending
much extra effort we can change the sign in the eipx factor and the sign of ωp to get:
1
[ap~ , a~†k ] = √ δ (D) (~k − p~)(+ωp + ωk ) = δ (D) (~k − p~)
2 ωp ωk
3. For the complex scalar field discussed in the text calculate h0|T [ϕ(x)ϕ† (0)]|0i.
Solution:
We want to calculate h0|T [ϕ(~x, t)ϕ(~0, 0)]|0i, h0|T [ϕ(~x, t)ϕ† (~0, 0)]|0i and h0|T [ϕ† (~x, t)ϕ† (~0, 0)]|0i.
For convenience, let’s write the Fourier expansions for ϕ and ϕ† again:
d3 k
Z
−i(ω~k t−~k·~
x) † i(ω~k t−~k·~
x)
ϕ(~x, t) = a~ e + b~k e
(2π)3 2ω~k k
d3 k † +i(ω~ t−~k·~x)
Z
† −i(ω~k t−~k·~
x)
ϕ (~x, t) = a e k + b~k e
(2π)3 2ω~k ~k
The vacuum state is defined by a~k |0i = b~k |0i = 0. Since a commutes with b† , we see
immediately that the two-point function of ϕ with itself is zero. Similarly, the two-point
function of ϕ† with itself is zero. So we just have to calculate the two point function for ϕ
with its Hermitian conjugate. For t > 0, we have:
24
3
d3 p
Z Z
† d k
−i(ω~k t−~k·~ † i(ω~k t−~k·~
†
h0|ϕ(~x, t)ϕ (~0, 0)|0i = h0| a~k e x)
+ b~k e x)
ap~ + bp~ |0i
(2π)3 2ωk (2π)3 2ωp
d3 k d3 p
Z Z
~
= 3 3
e−i(ω~k t−k·~x) h0|a~k ap†~ |0i
(2π) 2ωk (2π) 2ωp | {z }
(2π) 2ωp δ (3) (~k − p~)
3
d3 k
Z
~
= 3
e−i(ωk t−k·~x)
(2π) 2ωk
d3 k d3 p
Z Z
† ~ † −i(ω~k t−~k·~x) † i(ω~k t−~k·~
x)
h0|ϕ (0, 0)ϕ(~x, t)|0i = h0| ap~ + b p
~ a~k e + b~k e |0i
(2π)3 2ωk (2π)3 2ωp
d3 k d3 p
Z Z
~
= 3 3
ei(ω~k t−k·~x) h0|bp~ b~†k |0i
(2π) 2ωk (2π) 2ωp | {z }
(2π) 2ωp δ (3) (~k − p~)
3
d3 k
Z
~
= 3
e i(ωk t−k·~x)
(2π) 2ωk
h0|T [ϕ(~x, t)ϕ† (~0, 0)]|0i ≡ θ(t)h0|ϕ(~x, t)ϕ† (~0, 0)|0i + θ(−t)h0|ϕ† (~0, 0)ϕ(~x, t)|0i
d3 k
Z
−i(ωk t−~k·~
x) i(ωk t−~k·~
x)
= θ(t)e + θ(−t)e
(2π)3 2ωk
µ
d4 k e ikµ x
Z
=i
(2π)4 k 2 − m2 + iε
25
I.9 Disturbing the Vacuum
1. Choose the damping function g(v) = 1/(1 + v)2 instead of the one in the text. Show that
this results in the same Casimir force. [Hint: To P
sum the resulting series,Rpass to an integral
∞ R∞ ∞
representation H(ξ) = − ∞ −(1+nξ)t
= 0 dt e−t /(1 − eξt ).
P
n=1 1/(1 + nξ) = − n=1 0 dt e
Note that the integral blows up logarithmically near the lower limit, as expected.]
Solution:
If g(v) = (1 + v)−2 , then h(v) = dv g(v) = −(1 + v)−1 up to some constant that we can
R
set equal to zero. Define ε ≡ πa/d. Using the hint, we can rewrite the function f (d) as an
integral:
∞ ∞
1 ∂ X 1 ∂ X 1
f (d) = h (nε) = −
2 ∂a n = 1 2 ∂a n = 1 1 + nε
∞
1 ∂ X ∞
Z
=− dt e−(1+nε)t
2 ∂a n = 1 0
Z ∞ ∞
1 ∂ −t
X n
=− dt e e−εt
2 ∂a 0 n=1
Z ∞ −t
1 ∂ e
=− dt
2 ∂a 0 1 − e−εt
Now we can Taylor expand the denominator, being careful to keep higher order terms:
1 1
e−εt = 1 − εt + (εt)2 − (εt)3 + O(ε4 )
2 6
We’d like to get that series out of the denominator, so use the expansion (1 + δ)−1 = 1 − δ +
26
δ 2 + O(δ 3 ) with δ = − 12 εt + 16 (εt)2 + O(ε3 ):
Z ∞ " 2 #
e−t
1 ∂ 1 1 1
f (d) = − dt 1 − − εt + (εt)2 + O(ε3 ) + − εt + O(ε2 ) + O(ε3 )
2 ∂a 0 εt 2 6 2
Z ∞
e−t
1 ∂ 1 1
=− dt 1 + εt + (εt)2 + O(ε3 )
2 ∂a 0 εt 2 12
Z ∞
1 ∂ −t 1 1 1 2
=− dt e + + εt + O(ε )
2 ∂a 0 εt 2 12
Z ∞
1 ∂ −t d 1 πat 2
=− dt e + + + O(a )
2 ∂a 0 πat 2 12 d
1 ∞
Z
−t d πt
=− dt e − 2 +0+ + O(a)
2 0 πa t 12 d
Notice that differentiating with respect to a kills the second term and turns the O(a) term
into a finite piece; that is why we
R∞needed to goR to third order in the original expansion. There
∞
is one integral we have to do: 0 dt e−t t = −t
R ∞0 dt−te = 1. The other integral is just some
number that will cancel out, so call it I ≡ 0 dt e /t and proceed:
Id π
f (d) = + 2
−
2πa 24 d
We have taken the limit a → 0 to kill the finite a-dependent terms. Now we’re in business.
Differentiate with respect to d to get:
I π
f 0 (d) = 2
+
2πa 24 d2
Just as in the book for the case of the exponential damping factor, the infinite term is
independent of the distance d, so subtracting off f 0 (L − d) removes this term and leaves only
the finite contribution. Remembering that the force is minus the finite piece of the above
derivative, we get the same Casimir force as before:
π
F (d) = −f 0 (d) + lim f 0 (L − d) = −
L→∞ 24 d2
2. Show that with the regularization used in the appendix, the 1/d expansion of the force
between two conducting plates contains only even powers.
Solution:
27
π
where bα = Λα
→ 0 regulate the series. The constants cα and bα are subject to the constraints
X cα X cα X
=0, =0, cα = 1 .
α
bα α
b2α α
The force is F = −[f 0 (d) − (d → L − d)] with L → ∞, where in the limit L → ∞ the second
term serves simply to cancel off the leading divergent piece. The derivative of f (d) is
Z ∞
0 π X e bα t/d
f (d) = − 2 cα dt e−t t 2 .
2d α 0 (e bα t/d − 1)
The function
x2 e x
g(x) ≡ x 2
= 12 x2 csch2 x
(e − 1)
is even in x. Therefore its Taylor series contains only even powers of x. The function we are
interested in is x−2 g(x) with x = bα t/d, with the first term to be canceled by subtracting
f 0 (L − d). Therefore the series is even in 1/d.
3. Show off your skill in doing integrals by calculating the Casimir force in (3+1)-dimensional
spacetime. For help, see M. Kardar and R. Golestanian, Rev. Mod. Phys. 71: 1233,1999; J.
Feinberg, A. Mann, and M. Revzen, Ann. Phys. 288:103, 2001.
Solution:
As given in the book, the energy per unit plate-area is given by adding up the energies
of all the individual modes:
X
ε(d) = ω(kx , ky , kz )
kx ,ky ,kz
For a massless scalar field quantized in the x-direction, the dispersion relation is:
r
nπ 2
q
ω(kx , ky , kz ) = |~k |2 = + ky2 + kz2
d
Since ky and kz are continuous, the discrete sum indicated above should actually be an
integral: ky → dk
P R y
2π
, and the same for kz . Therefore, the energy per unit area between the
plates at x = 0 and x = d is:
∞ Z ∞ r
dky ∞ dkz nπ 2
X Z
ε(d) = + ky2 + kz2
n=1 −∞
2π −∞ 2π d
As in the (1 + 1)-dimensional case, we’ve made a mistake in writing down this formula:
28
we haven’t accounted for the important physical fact that we can’t contain arbitrarily high
frequencies within the plates. Regularizing with the function F (ω, a) = e−a ω , the energy per
unit area between the plates at x = 0 and x = d is:
∞ Z ∞ r
dky ∞ dkz nπ 2
Z q
2
−a ( nπ
d )
X
2 2 +ky2 +kz2
ε(d) = lim + ky + kz e
a→0
n=1 −∞
2π −∞ 2π d
First let’s do the integrals. Since the integrand depends only on the combination k 2 ≡ ky2 +kz2 ,
change to polar coordinates:
∞ Z 2π r
dφ ∞ dk k nπ 2
Z q
2
−a ( nπ
d )
X
2 +k2
ε(d) = lim +k e
a→0 2π
n=1 | 0 {z } 0
2π d
=1
Let c ≡ nπ/d for convenience, and while doing the integral over k don’t forget that c is a
function of n. Change integration variables to x2 ≡ c2 + k 2 =⇒ k dk = x dx:
∞ Z ∞ ∞ Z ∞
X dx x −a x 1 X
ε(d) = lim xe = lim dx x2 e−a x
a→0
n=1 c(n) 2π 2π a→0 n=1 c(n)
The parameter a has dimensions of length, so to Taylor expand we should define a dimen-
sionless ratio, α ≡ a/D, where D ≡ d/π. The π was included in the α purely for notational
29
convenience. In terms of these variables, c = n/D and a = αD, so the energy is:
∞
1 1 X 2 2n 2
ε(d) = 3
lim n + + 2 e−αn
2πD α→0 α n=1 α α
∞
1 1 2 2 2 X −αn
= lim ∂α − ∂α + 2 e
2π(d/π)3 α→0 α α α n=1
∞
π2
1 1 2 1 1 X −αn
= 3 lim ∂ − ∂α + 2 e
d α→0 α 2 α α α n=1
π2
1 1 2 1 1 1
ε(d) = 3 lim ∂α − ∂α + 2 α
d α→0 α 2 α α e −1
−eα
1
∂α α = α
e −1 (e − 1)2
e2α + eα
1
∂α2 =
eα − 1 (eα − 1)3
π2 e2α + eα eα
1 1
ε(d) = 3 lim α 3
+ α 2
+ 2 α
d α→0 α 2(e − 1) α(e − 1) α (e − 1)
30
Recalling the definition α = a/d gives:
π2 3 d4 d3 π2
ε(d) = lim 3 − −
a→0 d a4 2 a3 720 d3
π2
3d 1
= lim π 2 − −
a→0 a4 2 a3 720 d3
The force per unit area between the pairs of plates at x = 0 and x = d and at x = d and
x = L as a function of the location x = d of the middle plate is:
The divergent piece is independent of the distance d; when we subtract the contribution to
the force from the plate at x = L, the infinity arising from lima→0 1/a4 will cancel out:
2
∂ ε(d) 3π π2
= lim +
∂d d→L−d a→0 a4 240 (L − d)4
2
π2
3π d
= lim + 1 + O
a→0 a4 240 L4 L
2
3π
→ lim for L → ∞
a→0 a4
The force per unit area between two parallel conducting plates a distance d apart is:
π2
f (d) = −
240 d4
The force is attractive, meaning that it takes effort to pull apart two conducting plates.
Restoring the factors of ~ and c to give an idea for how large this force is gives:
π 2 ~c −27
m 4
f (d) = − ≈ − 1.3 × 10 N · m−2
240 d4 d
31
So the force (per unit area) between plates of size ∼ 1 m2 separated a distance d ∼ 1 mm
= 10−3 m apart is ∼ 10−15 N. Since 1 N is about the gravitational force exerted by the Earth
on an apple, the Casimir force is extremely small, as expected from the fact that it is an
effect of quantum origin.
Note: Each harmonic oscillator mode contributions 12 ω to the vacuum energy, so for a real
scalar field, which has only one component, this answer should be divided by 2. We keep it as
is because it is the usual result quoted for the electromagnetic field, which has 2 polarizations.
Also, if we decide to use a complex scalar field instead of a real scalar field, then we have a
theory with two real scalar fields and for that case the answer would be correct as written.
I.10 Symmetry
1. Some authors prefer the following more elaborate formulation of Noether’s theorem. Sup-
pose that the action does not change under an infinitesimal transformation δϕa (x) = θA VaA
[with θA some parameters labeled by A and VaA some function of the fields ϕb (x) and possibly
also of their first derivatives with respect to x.] It is important to emphasize that when we
say the action S does not change we are not allowed to use the equations of motion. After all,
the Euler-Lagrange equations of motion follow from demanding that δS = 0 for any variation
δϕa subject to certain boundary conditions. Our scalar field theory example nicely illustrates
this point, which is confused in some books: δS = 0 merely because S is constructed using
the scalar product of O(N ) vectors.
Now let us do something apparently a bit strange. Let us consider the infinitesimal change
written above but with the parameters θA dependent on x. In other words, we now con-
sider δϕa (x) = θA (x)VaA . Then of course there is no reason for δS to vanish; but, on the
other hand,
R 4 we know that since δS does vanish when θA is constant, δS must have the form
δS = d x J (x)∂µ θA (x). In practice, this gives us a quick way of reading off the current
µ
Consider the following real scalar field Lagrangian with global O(N ) symmetry:
1 1 1
L = ∂µ ϕα ∂ µ ϕα − m2 ϕα ϕα − λ(ϕα ϕα )
2 2 4
The field index α runs from 1 to N . The global O(N ) symmetry is:
32
N (N −1)/2
Let {T A }A=1 be the generators of the Lie algebra of the group O(N ) in the N -dimensional
(“defining”) representation. Given these generators, the N × N orthogonal matrix R can be
written as:
h A Ai β
Rαβ = eθ T = δα β + θA (T A )αβ + O(θ2 )
α
The goal is to find the conserved current corresponding to the O(N ) symmetry by promoting
the symmetry to a local one and finding the coefficient of ∂µ θA in the variation of the action.
Afterwards, we’ll compare the result of this method to the result of the formal procedure for
generating the Noether current and see that they are equivalent. The action is:
Z
4 1 µ α 1 2 α 1 α
S[ϕ] = d x ∂µ ϕα ∂ ϕ − m ϕα ϕ − λ(ϕα ϕ )
2 2 4
S[ϕ + δϕ] =
Z
4 1 µ α 1 2 α 1 α
d x ∂µ (ϕ + δϕ)α ∂ (ϕ + δϕ) − m (ϕ + δϕ)α (ϕ + δϕ) − λ((ϕ + δϕ)α (ϕ + δϕ) )
2 2 4
Z
= S[ϕ] + d4 x ∂µ ϕα ∂ µ (δϕα ) − m2 ϕα δϕα − λϕβ ϕβ ϕα δϕα
δϕα = θA (T A )αβ ϕβ
The symmetry is global, but we can find the conserved current by imagining that the sym-
metry were local and finding a term of the form:
Z
δS = d4 xj µ ∂µ θ
If we use the explicit expression for δϕ in the first-order change in the action, we get:
33
Z
δS ≡ S[ϕ + δϕ] − S[ϕ] = d4 x ∂µ ϕα ∂ µ (δϕα ) + ...
Z
= d4 x ∂µ ϕα ∂ µ θA (T A )αβ ϕβ + ...
Z
= d4 x ∂µ ϕα (T A )αβ ϕβ ∂ µ θA
We know the other terms must be zero because the action is invariant under global O(N )
transformations. Therefore, the conserved current corresponding to the global O(N ) sym-
metry is:
jµA = ∂µ ϕα (T A )αβ ϕβ
You might worry that the current doesn’t look hermitian, since (T A )T = −T A , but once you
transpose everything you will also switch the order of the ∂ϕ and the ϕ, which means you
have to integrate by parts and therefore generate a second minus sign to get back to the
original form. So jµA is hermitian.
To verify the result using Noether’s theorem, we compute the current directly:
∂L
θA jµA = δϕα = ∂µ ϕα δϕα
∂(∂ µ ϕα )
= ∂µ ϕα θA (T A )αβ ϕβ
jµA = ∂µ ϕα (T A )αβ ϕβ
4. Add a Lorentz scalar field η transforming as a vector under SO(3) to the Lagrangian in
exercise I.10.3, maintaining SO(3) invariance. Determine the Noether currents in this theory.
Using the equations of motion, check that the currents are conserved.
Solution:
34
The index α runs from 1 to 3, and the indices are contracted with usual identity ma-
trix.2 ϕ transforms under the 5-representation of SO(3), which can either be written as
a 5-dimensional column vector or as a 3 × 3 symmetric traceless matrix. For this problem,
we choose the latter description and write:
A C D
1 1
ϕ = ϕαβ = (ϕαβ + ϕβα ) − (ϕγδ δ γδ ) δαβ = C B E
2 3
D E −(A + B)
Note that the term Tr(ϕϕϕϕ) is not included, because in this case it is proportional to
[Tr(ϕϕ)]2 . For further details about ϕ, see the solution to problem 9.3 in the book.
To find the Noether current, we need an infinitesimal version of the above transformations.
If T A are the generators of the 3-dimensional representation of SO(3), where A runs from 1
to 3(3 − 1)/2 = 3, then the matrix R can be written:
h A Ai β β
β
Rα = eθ T = 1 + θA T A + O(θ2 ) α = δα β + θA (T A )αβ + O(θ2 )
α
h i
ϕαβ → δα γ + θA (T A )αγ + O(θ2 ) δβ δ + θB (T B )β δ + O(θ2 ) ϕγδ
h i
= δα γ δβ δ + θA (T A )αγ δβ δ + δα γ (T A )β δ + O(θ2 ) ϕγδ
h i
A δ
= ϕαβ + θ (T )α ϕγβ + (T )β ϕαδ +O(θ2 )
A A γ
| {z }
A
= δϕαβ
35
The Noether current is:
∂L ∂L
jµA = δϕAαβ + δη A
µ
∂(∂ ϕαβ ) ∂(∂ µ ηα ) α
= ∂µ ϕαβ δϕA α
αβ + ∂µ η δηα
A
h i
= ∂µ ϕαβ (T A )αγ ϕγβ + (T A )β δ ϕαδ + ∂µ η α (T A )αβ ηβ
Now we have to check that this current is in fact conserved, or in other words that ∂ µ jµA = 0.
The equations of motion are:
2
∂ + m2 + λ Tr(ϕϕ) ϕαβ + κ ηα ηβ = 0
2
∂ + µ2 + g (~η · η~ ) δαβ + 2κ ϕαβ ηβ = 0
The important point is that ∂ 2 ϕαβ = a(ϕ) ϕαβ − κ ηα ηβ and ∂ 2 ηα = b(η) ηα − 2κ ϕαβ ηβ , where
a and b are SO(3)-scalar functions of ϕ and η, respectively. Therefore, the 4-divergence of
the current is:
h i
µ A αβ A γ αβ A δ
∂ jµ = a(ϕ) ϕ (T )α ϕγβ + ϕ (T )β ϕαδ
+ b(η) η α (T A )αβ ηβ
− 2κ ηα [(T A )αβ ϕβγ + ϕαβ (T A )βγ ]ηγ
Each of the three terms is zero individually by symmetry. To see why, look at the first term:
Now the symmetry argument comes in: {T A }3A=1 are the generators of 3-dimensional rota-
tions
and are therefore antisymmetric matrices. Meanwhile, (ϕϕ) is a symmetric matrix, so
Tr (T A )(ϕϕ) = 0. Similarly, η α (T A )αβ ηβ = Tr (T A )(ηη) = 0.
36
I.11 Field Theory in Curved Spacetime
1. Integrate by parts to obtain for the scalar field action
√ √
Z
4 1 1 µν 2
S = − d x −g ϕ √ ∂µ −g g ∂ν + m ϕ
2 −g
and write the equation of motion for ϕ in curved spacetime. Discuss the propagator of the
scalar field D(x, y) (which is of course no longer translation invariant, i.e., it is no longer a
function of x − y).
Solution:
We will specialize to the maximally symmetric de Sitter spacetime, which has scalar curvature
R = n(n − 1)/a2 in n spacetime dimensions. Here a is the de Sitter length. We will follow
B. Allen and T. Jacobson, “Vector Two-Point Functions in Maximally Symmetric Spaces,”
Commun. Math. Phys. 103, 669-692 (1986).
Therefore the defining differential equation for the free scalar propagator in maximally sym-
metric curved space can also be written as an ordinary differential equation in one variable,
µ. Writing D(x, y) = D(µ(x, y)), the equation of motion for ϕ implies
for x 6= y. Here 0 = d
dµ
and A(µ) = 1
a
cot(µ/a). Change variables to
µ
2
z = cos
2a
to put the differential equation in the form
d2
d
z(1 − z) 2 + [c − (a + b + 1)z] − ab D(z) = 0
dz dz
37
where
h p i
1
a= 2
n − 1 + (n − 1)2 − (2ma)2
h p i
1 2 2
b = 2 n − 1 − (n − 1) − (2ma)
c = 12 n .
This is the defining equation of the hypergeometric function 2 F1 (a, b, c, z). Since a + b + 1 −
2c = 0, the differential equation is invariant under z → 1 − z. Two independent particular
solutions are therefore 2 F1 (a, b, c, z) and 2 F1 (a, b, c, 1 − z), and the general solution is a linear
combination of these:
38
II Dirac and the Spinor
II.1 The Dirac Equation
7. Show explicitly that (25) violates parity.
Solution:
ψL ≡ PL ψ → PL iγ 0 ψ = iγ 0 PR ψ = iγ 0 ψR .
Thus parity transforms left-handed spinors into right-handed spinors, and so the Lagrangian
in (25) violates parity.
See VII.7.7.
12. Work out the Dirac equation in (2+1)-dimensional spacetime. Show that the appar-
ently innocuous mass term violates parity and time reversal. [Hint: The three γ µ s are just
the three Pauli matrices with appropriate factors of i.]
Solution:
Multiply the whole equation by γ 2 and anticommute it through to the right to get:
Divide by −1 to get:
(iγ 0 ∂0 + iγ 1 ∂1 − iγ 2 ∂2 + m)γ 2 ψ(x) = 0
Defining the parity transformation P : x ≡ (x0 , x1 , x2 ) → x0 ≡ (x0 , x1 , −x2 ) puts the above
equation into the form:
39
Therefore the spinor ψ 0 (x0 ) ≡ γ 2 ψ(x) satisfies the Dirac equation with the wrong sign for the
mass term: (i6 ∂x0 + m)ψ 0 (x0 ) = 0.
At this point we should ask why the same argument doesn’t hold in (3+1) dimensions.
After all, if we parity transform using P : x ≡ (x0 , x1 , x2 , x3 ) → x0 ≡ (x0 , x1 , −x2 , x3 ), we get
the exact same result as above: (i6 ∂x0 + m)ψ 0 (x0 ) = 0. Where does the argument fail?
There are two answers. One answer is that in (odd+1) dimensions, the operation of flip-
ping one spatial coordinate is related by a rotation to the operation of flipping all of the
spatial coordinates, while in (even+1) dimensions that is not true. See the footnote on page
98 of the book for more details.
The second answer is that in (3+1) dimensions, we can multiply (i 6 ∂x0 + m)ψ 0 (x0 ) = 0
on the left by the γ 5 matrix, which anticommutes with the other 4 gamma matrices, to get
(−i6 ∂x0 + m)γ 5 ψ 0 (x0 ) = 0 =⇒ (i6 ∂x0 − m)γ 5 ψ 0 (x0 ) = 0, and the field redefinition ψ → γ 5 ψ
leaves the path integral unchanged. Therefore we conclude that the sign of the Dirac mass
term does not matter in (3+1) dimensions.
However, this operation is not possible in (2+1) dimensions. As noted in the partial solution
to this question in the back of the book, in (2+1) dimensions we can define the gamma
matrices to be the Pauli matrices: γ 0 = σ 3 , γ 1 = iσ 2 , γ 2 = −iσ 1 . Now try to define γ 5 :
In (2+1) dimensions, γ 5 is just the identity matrix. In other words, the Dirac spinor rep-
resentation of SO(2, 1) is irreducible. Therefore, in (2+1) dimensions, the sign of the Dirac
mass term does matter and therefore the mass term violates parity.
Time reversal is going to work the same way. Take the Dirac equation and multiply by
γ 0 , then anticommute γ 0 all the way to the right. This yields
Divide by −1 to get
(−iγ 0 ∂0 + iγ i ∂i + m)γ 0 ψ(x) = 0
Define the time reversed coordinates x0 ≡ (−x0 , ~x) to write the above as
Therefore the spinor ψ 0 (x0 ) ≡ γ 0 ψ(x) satisfies the Dirac equation with the wrong sign for the
mass term. Again, in (2+1) dimensions the Dirac spinor is irreducible.
40
II.2 Quantizing the Dirac Field
2. Quantize the Dirac field in a box of volume V and show thatR the vacuum energy E0 is
indeed proportional to V . [Hint: The integral over momentum d3 p is replaced by a sum
over discrete values of the momentum.]
Solution:
P
This problem was worded incorrectly in the book, since the vacuum energy is E0 = (#) p ωp
with no factor of volume (as required by dimensional analysis). The intention is to place the
Dirac field in a box and to calculate its contribution to the energy of the vacuum, and show
that the factors of volume work themselves out. The solution proceeds according to the
discussion on pp. 111 and 139, so we outline the salient steps:
• When quantizing in a box, momentum integrals are replaced with sums as:
d3 p
Z
1 X
→
(2π)3 V
p
~
• Each harmonic oscillator for a Lorentz-scalar field contributes + 12 ωp~ to the vacuum
energy.
• Each harmonic oscillator for a Lorentz-spinor field contributes − 21 ωp~ to the vacuum
energy.
• The Dirac field contains two Lorentz-spinor fields, for example the electron and the
positron, so the above gets multiplied by 2.
• Each Lorentz-spinor field contains two spin states, so the above gets multiplied by yet
another factor of 2.
P
Therefore, a Dirac field contributes a total of −2 p~ ωp~ to the energy of the vacuum.
41
II.3 Lorentz Group and Weyl Spinors
1. Show by explicit computation that ( 12 , 21 ) is indeed the Lorentz vector.
Solution:
Let χ transform under the ( 12 , 12 ) representation of the Lorentz group: χaċ , where a {1, 2}
denotes the two-dimensional representation of SU (2)L and ċ {1̇, 2̇} denotes the two-
dimensional representation of SU (2)R .
Since the Pauli matrices have the property {σ i , σ j } = 2δ ij , these transformations simplify:
iθ n̂·( ~σ2 ) θ θ
e = 12×2 cos + i n̂ · ~σ sin
2 2
θ
cos 2 0 x 0 1 y 0 −i z 1 0 θ
= θ +i n +n +n sin
0 cos 2 1 0 i 0 0 −1 2
θ θ − θ
z
cos 2 + i n sin 2 in sin z 2 , n± ≡ nx ± i ny
= θ
+
i n sin 2 cos 2 − i n sin 2θ
θ
Using cos(−i x) = cosh(x) and i sin(−i x) = sinh(x), the boost transformation matrix can
be obtained from the above with θ → −iβ:
cosh β2 + nz sinh β − β
β n̂·( ~σ2 ) 2 n sinh 2
e =
n+ sinh β2 cosh β2 − nz sinh β2
As indicated on page 114 of the book, the transformations for SU (2)R are the same as those
for SU (2)L except that the boosts get an extra minus sign from the replacement β → −β. We
can now compute explicitly how the four components of χaċ transform under rotations and
under boosts. First specialize to the case n̂ = (1, 0, 0), for which the transformations on χ are:
42
Rotation through angle θ about x̂-axis:
i bh
~ ċ
h i
~
χaċ → eiθ n̂·J eiθ n̂·J χb ė
a ė
b ċ
c is c is
= χb ė
is c a is c ė
b 1̇ 2̇
ė T ċ
c is χ1 χ1 c is
=
i s c a χ2 1̇ χ2 2̇ i s c ė
b
2 1̇
c χ1 − s2 χ2 2̇ + i cs(χ1 2̇ + χ2 1̇ ) c2 χ1 2̇ − s2 χ2 1̇ + i cs(χ1 1̇ + χ2 2̇ )
= 2 1̇
c χ2 − s2 χ1 2̇ + i cs(χ1 1̇ + χ2 2̇ ) c2 χ2 2̇ − s2 χ1 1̇ + i cs(χ1 2̇ + χ2 1̇ )
θ θ
c ≡ cos , s ≡ sin
2 2
Some trigonometric identities will be useful. Since cos(2x) = 2 cos2 x − 1 = 1 − 2 sin2 x and
sin(2x) = 2 sin x cos x, we have c2 = 12 (1 + cos θ), s2 = 12 (1 − cos θ) and cs = 12 sin θ. With
these, the above transformations simplify to:
1 1 i
χ1 1̇ → (1 + cos θ)χ1 1̇ − (1 − cos θ)χ2 2̇ + sin θ(χ1 2̇ + χ2 1̇ )
2 2 2
1 1̇ 1 i
= (χ1 − χ2 2̇ ) + cos θ(χ1 1̇ + χ2 2̇ ) + sin θ(χ1 2̇ + χ2 1̇ )
2 2 2
1 1 i
χ1 2̇ → (1 + cos θ)χ1 2̇ − (1 − cos θ)χ2 1̇ + sin θ(χ1 1̇ + χ2 2̇ )
2 2 2
1 2̇ 1 i
= (χ1 − χ2 1̇ ) + cos θ(χ1 2̇ + χ2 1̇ ) + sin θ(χ1 1̇ + χ2 2̇ )
2 2 2
1 1 i
χ2 1̇ → (1 + cos θ)χ2 1̇ − (1 − cos θ)χ1 2̇ + sin θ(χ1 1̇ + χ2 2̇ )
2 2 2
1 2̇ 1 i
= − (χ1 − χ2 1̇ ) + cos θ(χ1 2̇ + χ2 1̇ ) + sin θ(χ1 1̇ + χ2 2̇ )
2 2 2
1 1 i
χ2 2̇ → (1 + cos θ)χ2 2̇ − (1 − cos θ)χ1 1̇ + sin θ(χ1 2̇ + χ2 1̇ )
2 2 2
1 1̇ 1 i
= − (χ1 − χ2 2̇ ) + cos θ(χ1 1̇ + χ2 2̇ ) + sin θ(χ1 2̇ + χ2 1̇ )
2 2 2
These all appear in ± pairs, so add and subtract some of these transformations:
χ1 2̇ − χ2 1̇ → χ1 2̇ − χ2 1̇
χ1 1̇ − χ2 2̇ → χ1 1̇ − χ2 2̇
χ1 1̇ + χ2 2̇ → cos θ(χ1 1̇ + χ2 2̇ ) + i sin θ(χ1 2̇ + χ2 1̇ )
χ1 2̇ + χ2 1̇ → cos θ(χ1 2̇ + χ2 1̇ ) + i sin θ(χ1 1̇ + χ2 2̇ )
43
In terms of these new labels, then above transformation laws are:
t t
v 1 0 0 0 v
v x 0 1 0 0 v x
y →
v 0 0 cos θ sin θ v y
vz 0 0 − sin θ cos θ vz
That is exactly how a 4-vector v µ transforms under rotations about the x-axis.
iθ n̂·(~
σ /2) c s
For n̂ = (0, 1, 0), the transformation matrix is e = , so the components
−s c
of χ transform as:
χ1 1̇ → c2 χ1 1̇ + s2 χ2 2̇ + cs(χ1 2̇ + χ2 1̇ )
1 1 1
= (1 + cos θ)χ1 1̇ + (1 − cos θ)χ2 2̇ + sin θ(χ1 2̇ + χ2 1̇ )
2 2 2
1 1̇ 1 1
= (χ1 + χ2 2̇ ) + cos θ(χ1 1̇ − χ2 2̇ ) + sin θ(χ1 2̇ + χ2 1̇ )
2 2 2
χ1 2̇ → c2 χ1 2̇ − s2 χ2 1̇ − cs(χ1 1̇ − χ2 2̇ )
1 1 1
= (1 + cos θ)χ1 2̇ − (1 − cos θ)χ2 1̇ − sin θ(χ1 1̇ − χ2 2̇ )
2 2 2
1 2̇ 1 1
= (χ1 − χ2 1̇ ) + cos θ(χ1 2̇ + χ2 1̇ ) − sin θ(χ1 1̇ − χ2 2̇ )
2 2 2
χ2 1̇ → c2 χ2 1̇ − s2 χ1 2̇ − cs(χ1 1̇ − χ2 2̇ )
1 1 1
= (1 + cos θ)χ2 1̇ − (1 − cos θ)χ1 2̇ − sin θ(χ1 1̇ − χ2 2̇ )
2 2 2
1 2̇ 1 1
= − (χ1 − χ2 1̇ ) + cos θ(χ1 2̇ + χ2 1̇ ) − sin θ(χ1 1̇ − χ2 2̇ )
2 2 2
χ2 2̇ → c2 χ2 2̇ + s2 χ1 1̇ − cs(χ1 2̇ + χ2 1̇ )
1 1 1
= (1 + cos θ)χ2 2̇ + (1 − cos θ)χ1 1̇ − sin θ(χ1 2̇ + χ2 1̇ )
2 2 2
1 1̇ 1 1
= (χ1 + χ2 ) − cos θ(χ1 − χ2 ) − sin θ(χ1 2̇ + χ2 1̇ )
2̇ 1̇ 2̇
2 2 2
χ1 1̇ + χ2 2̇ → χ1 1̇ + χ2 2̇
χ1 1̇ − χ2 2̇ → cos θ(χ1 1̇ − χ2 2̇ ) + sin θ(χ1 2̇ + χ2 1̇ )
χ1 2̇ − χ2 1̇ → χ1 2̇ − χ2 1̇
χ1 2̇ + χ2 1̇ → cos θ(χ1 2̇ + χ2 1̇ ) − sin θ(χ1 1̇ − χ2 2̇ )
44
Using the same definitions from before, v t ≡ χ2 1̇ −χ1 2̇ , v x ≡ χ2 2̇ −χ1 1̇ , v y ≡ −i(χ1 1̇ +χ2 2̇ ), v z ≡
χ1 2̇ + χ2 1̇ , we get: t t
v 1 0 0 0 v
v x 0 cos θ 0 − sin θ v x
y →
v 0 0 1 0 v y
vz 0 sin θ 0 cos θ vz
That is exactly how a 4-vector v µ transforms under rotations about the y-axis.
iθ/2
iθ n̂·(~
σ /2) c + is 0 e 0
For n̂ = (0, 0, 1), the transformation matrix is e = = ,
0 c − is 0 e−iθ/2
so the transformation of χ is easy:
Rearrange these to make them look like the previous two cases:
χ1 2̇ − χ2 1̇ → χ1 2̇ − χ2 1̇
χ1 2̇ + χ2 1̇ → χ1 2̇ + χ2 1̇
χ1 1̇ + χ2 2̇ → cos θ(χ1 1̇ + χ2 2̇ ) + i sin θ(χ1 1̇ − χ2 2̇ )
χ1 1̇ − χ2 2̇ → cos θ(χ1 1̇ − χ2 2̇ ) + i sin θ(χ1 1̇ + χ2 2̇ )
That is exactly how a 4-vector v µ transforms under rotations about the z-axis. Therefore,
we know that the object χaċ , which transforms under the ( 12 , 12 )-representation of the Lorentz
group, can be repackaged into an object v µ , which transforms under the 4-dimensional (“vec-
tor”) representation of the Lorentz group.
Before proceeding to check the boosts, let us reflect on the repackaging of the components of
45
χ. The definitions we used are:
v t ≡ χ2 1̇ − χ1 2̇
v x ≡ χ2 2̇ − χ1 1̇
v y ≡ −i(χ1 1̇ + χ2 2̇ )
v z ≡ χ1 2̇ + χ2 1̇
Note the following very important point: since SU (2) indices are raised and lowered with
0 1 0 −1
the antisymmetric tensors ab = and ab = , the diagonal components
−1 0 +1 0
of χaċ correspond to the off-diagonal components of χaċ and vice versa, and all sorts of minus
signs appear. For example, χ2 1̇ = 2a χa1̇ = 21 χ11̇ = −12 χ11̇ = +12 χ11̇ = χ11̇ . Raising all of
the indices in that way results in the following for the definitions of (v t , ~v ):
v t ≡ χ2 1̇ − χ1 2̇ = +χ11̇ + χ22̇ = Iaȧ χaȧ
v x ≡ χ2 2̇ − χ1 1̇ = +χ12̇ + χ21̇ = σaxȧ χaȧ
v y ≡ −i(χ1 1̇ + χ2 2̇ ) = +iχ21̇ − iχ12̇ = σayȧ χaȧ
v z ≡ χ1 2̇ + χ2 1̇ = −χ22̇ + χ11̇ = σazȧ χaȧ
We therefore have discovered a way to turn the pair aȧ {11̇, 12̇, 21̇, 22̇} of SU (2)L ⊗ SU (2)R
indices into the single SO(3, 1) 4-vector index, µ {0, 1, 2, 3}: for any value of µ {x, y, z},
use the corresponding Pauli matrix to contract the aȧ indices; for µ = 0, use the identity
matrix. This is often written as σaµȧ , where σ 0 ≡ I.
Back to the problem at hand, we still have to check that the boosts work, but now we
have two options as to how to go about doing that. We could proceed as before, by writing
out the boost matrices explicitly, calculating the transformation of each component of χ,
repackaging them into a 4-vector, and showing that it’s the same 4-vector that we got before.
However, we now have a second option: if it is true that we can use the machine σaµȧ to change
the ( 21 , 12 ) indices to one 4-index, then it is already manifest that the object χµ ≡ σaµȧ χaȧ trans-
forms as a 4-vector under rotations and boosts, and therefore that the two representations
are equivalent. A valid way to approach this problem is simply to check that the symbol σaµȧ
is invariant under simultaneous Lorentz transformations for all of its indices.
It’s easiest to work with an object that has one lower SU (2)L index and one upper SU (2)R
index, so define the quantity sµa ċ ≡ −σaµȧ ȧċ whose components are, numerically:
sta ċ = (−iσ y )aċ
sxa ċ
= (σ z )aċ
sya ċ
= (−iσ t )aċ
sza ċ
= (−σ x )aċ
46
Let’s take a boost in the x-direction, meaning n̂ = (1, 0, 0), and apply it to sµa ċ :
h β x i b h β x iċ
sµa ċ → e 2 σ e− 2 σ Λµν sνb ė
a ė
µ
cosh β − sinh β 0 0
b ċ
cosh β2 sinh β2 cosh β2 − sinh β2
− sinh β cosh β 0 0
= β β β β
sνb ė
sinh 2 cosh 2 a − sinh 2 cosh 2 ė
0 0 1 0
0 0 0 1 ν
Λt ν sbν ė
= cosh β(iσ y )b ė − sinh β(σ z )b ė
ė ė
0 −1 1 0
= cosh β − sinh β
1 0 b 0 −1 b
ė
− sinh β − cosh β
=
cosh β sinh β b
Therefore,
ċ
cosh β2 sinh β2 cosh β2 − sinh β2
− sinh β − cosh β
sta ċ →
sinh β2 cosh β2 cosh β sinh β − sinh β2 cosh β2 a
ċ
0 −1
= = (−iσ y )a ċ = sta ċ X
1 0 a
For µ = x, we have:
Λxν sbν ė
= − sinh β(−iσ y )b ė + cosh β(σ z )b ė
ė ė
0 −1 1 0
= − sinh β + cosh β
1 0 b 0 −1 b
ė
cosh β sinh β
=
− sinh β − cosh β b
Therefore,
ċ
cosh β2 sinh β2 cosh β2 − sinh β2
cosh β sinh β
sxa ċ →
sinh β2 cosh β2 − sinh β − cosh β − sinh β2 cosh β2 a
ċ
1 0
= = (σ z )a ċ = sxa ċ X
0 −1 a
For µ = y, we have:
ė
y ν ė t ė −i 0
Λ ν sb = (−iσ )b =
0 −i b
47
Therefore,
ċ
cosh β2 sinh β2 cosh β2 − sinh β2
−i 0
sya ċ →
sinh β2 cosh β2 0 −i − sinh β2 cosh β2 a
ċ
−i 0
= = (−iσ t )a ċ = sya ċ X
0 −i a
For µ = z, we have:
ė
z ν ė x ė 0 −1
Λ ν sb = (−σ )b =
−1 0 b
Therefore,
ċ
cosh β2 sinh β2 cosh β2 − sinh β2
0 −1
sza ċ →
sinh β2 cosh β2 −1 0 − sinh β2 cosh β2 a
ċ
0 −1
= = (−σ x )a ċ = sza ċ X
−1 0 a
Since ab is invariant under SU (2)L (if it weren’t, we couldn’t use it to raise and lower SU (2)L
indices), the invariance of sµa ċ implies the invariance of σaµċ . We have checked only the case
of a Lorentz boost in the x-direction, but since we have already checked for rotational invari-
ance, we can just argue by symmetry that boosts in the y- and z- directions also leave the
symbol σaµċ invariant. Therefore, the ( 21 , 12 )-representation of SU (2)L ⊗ SU (2)R is equivalent
to the 4-vector representation of SO(3, 1).
2. Work out how the six objects contained in the (1, 0) and (0, 1) transform under the
Lorentz group. Recall from your course on electromagnetism how the electric and mag-
~ and B
netic fields E ~ transform. Conclude that the electromagnetic field in fact transforms as
(1, 0)⊕(0, 1). Show that it is parity that once again forces us to use a reducible representation.
Solution:
48
Now we can find out how the object X(ab) in the 3-dimensional representation of SU (2)
transforms, since each index gets its own transformation matrix. Explicitly, for a rotation
through an angle θ about an axis n̂, the transformation is:
h i ch i d
iθ n̂·( ~σ2 ) iθ n̂·( ~σ2 )
X(ab) → e e X(cd)
a b
h i c h iT d
iθ n̂·( ~σ2 ) iθ n̂·( ~σ2 )
= e X(cd) e
a b
in− s
c + inz s X11 X(12) c + inz s in+ s
=
in+ s c − inz s X(12) X22 in− s c − inz s
θ θ
where c ≡ cos , s ≡ sin
2 2
We’ve just been referencing SU (2), but remember that we’re talking about the Lorentz group,
SO(3, 1) ∼
= SU (2)L ⊗ SU (2)R . The discussion above applies for SU (2)L , so the transforma-
tions we found are those for an object in the (1, 0)-representation, meaning that said object
transforms under the 3-dimensional representation of SU (2)L and is unaffected by SU (2)R .
What about the (0, 1)-representation?
As for the case of (0, 21 ) versus ( 12 , 0) from the previous problem, the (0, 1) transforms in
exactly the same way as the (1, 0) except for the critical difference that K ~ (1,0) = +~σ /(2i)
whereas K~ (0,1) = −~σ /(2i). In practice that means we can copy the results from above except
with β → −β, which implies ch → ch and sh → −sh .
Let Y transform under the (0, 1) representation of the Lorentz group: Y = Y (ȧḃ) . The
Lorentz group transformations for Y are:
49
β β
ch ≡ cosh , sh ≡ sinh
2 2
Now consider a quantity Φ that transforms under the (1, 0) ⊕ (0, 1) representation of the
Lorentz group defined by the following:
Φ1 Φ 3 0 0
X(ab) 02×2 Φ3 Φ2 0 0
Φ≡X ⊕Y = ≡
02×2 Y (ȧḃ) 0 0 Φ4 Φ6
0 0 Φ6 Φ5
This quantity Φ is a 4 × 4 matrix with 6 independent components. Using the above work, it
transforms in the following way under boosts:
Φ→
ch +nz sh n− sh 0 0 Φ1 Φ3 0 0 ch +nz sh n+ sh 0 0
+ z − z
n sh ch −n sh −n
0 0 Φ3
Φ2 0 0
n s h c h s h 0 0
z − z +
ch −n sh −n sh ch −n sh −n sh
0 0 0
0 Φ4 Φ6
0 0
+ z − z
−n sh ch +n sh 0 −n sh ch +n sh
0 0 0 Φ6 Φ5 0 0
To understand what’s going on, specialize to a boost in the x-direction: n̂ = (1, 0, 0). For
that case, the field Φ transforms as:
ch s h 0 0 Φ1 Φ3 0 0 ch sh 0 0
Φ→ s h ch
0 0 Φ3
Φ2 0 0 s h
c h 0 0
0 0 ch −sh 0 0 Φ4 Φ6 0 0 ch −sh
0 0 −sh ch 0 0 Φ6 Φ5 0 0 −sh ch
50
Φ2 → c2h Φ2 + s2h Φ1 + 2sh ch Φ3
1 1
= (cosh β + 1)Φ2 + (cosh β − 1)Φ1 + sinh β Φ3
2 2
1 1
= cosh β (Φ1 + Φ2 ) − (Φ1 − Φ2 ) + sinh β Φ3
2 2
The fields appear in the combinations Φ± ≡ 12 (Φ1 ±Φ2 ) and Φ3 . From adding and subtracting
the transformations for Φ1 and Φ2 , we get:
Φ+ → cosh β Φ+ + sinh β Φ3
Φ3 → cosh β Φ3 + sinh β Φ+
Φ− → Φ−
To collect and rearrange the results, boosting in the x-direction with speed v = tanh β
51
transforms the components of Φ as follows:
Φ− → Φ−
Φ+ → cosh β Φ+ + sinh β Φ3
Φ6 → cosh β Φ6 − sinh β Φ̃+
Φ̃− → Φ̃−
Φ̃+ → cosh β Φ̃+ − sinh β Φ6
Φ3 → cosh β Φ3 + sinh β Φ+
Ex → Ex
E y → cosh β E y + sinh β B z
E z → cosh β E z − sinh β B y
Bx → Bx
B y → cosh β B y − sinh β E z
B z → cosh β B z + sinh β E y
These are precisely the transformation rules of an electric field E ~ and a magnetic field B ~
under a Lorentz boost in the x-direction with speed v = tanh β. We see that the electromag-
netic field transforms under the reducible representation (1, 0) ⊕ (0, 1) of the Lorentz group.
That fact is typically repackaged in a different way. A spin-1 representation corresponds to the
3-dimensional representation of the rotation group, which corresponds to the 4-dimensional
representation of the Lorentz group, denoted by the index µ {0, 1, 2, 3}. Therefore, an ob-
ject that transforms under two copies of spin-1 implies that the object gets two such indices,
µ and ν.
T
cosh β sinh β 0 0 0 F 01 F 02 F 03 cosh β sinh β 0 0
sinh β cosh β 0 0 −F 01 0 F 12 F 13 sinh β cosh β 0 0
F µν → 02 12 23
0 0 1 0 −F −F 0 F 0 0 1 0
03 13 23
0 0 0 1 −F −F −F 0 0 0 0 1
52
F 01 → F 01
F 02 → cosh β F 02 + sinh β F 12
F 03 → cosh β F 03 + sinh β F 13
F 23 → F 23
F 13 → cosh β F 13 + sinh β F 03
F 12 → cosh β F 12 + sinh β F 02
We know that parity interchanges SU (2)L with SU (2)R . How can we get an object that
transforms as spin-1 under the rotation group (that is, as a “3-vector” under spatial rota-
tions) that is also invariant under parity? As discussed, an object that transforms as (1, 0)
of SU (2)L ⊗ SU (2)R does in fact transform as spin-1 under the rotation group, but a parity
operation would change such an object to (0, 1). If we want the theory of the electromagnetic
field to preserve parity and to be Lorentz invariant, then it must transform as (1, 0) ⊕ (0, 1)
under Lorentz transformations.
3. Show that
e−~ϕ·~σ
−i~ ~ 0 +i~
ϕ·K ~
ϕ·K 0
e γ e =
e+~ϕ·~σ 0
and
e ϕ~ ·~σ = cosh ϕ + ~σ · ϕ̂ sinh ϕ
with the unit vector ϕ̂ ≡ ϕ
~ /ϕ. Identifying p~ = mϕ̂ sinh ϕ, derive the Dirac equation. Show
that
i 0 σi
γ = .
−σi 0
Solution:
If D is a diagonal matrix, then eD = diag(eD11 , eD22 , ...). Page 117 tells us that
~σ
~ 2
0 0 0 I
iK = and γ =
0 − ~σ2 I 0
53
so by direct computation we have
~σ /2 0 ~σ /2 0
−~
ϕ · ϕ·
+~
~ ~
e−~ϕ·(iK) γ 0 e+~ϕ·(iK) =e 0 −~σ /2 γ 0 e 0 −~σ /2
−~ϕ·(~σ/2) +~ϕ·(~σ/2)
e 0 0 I e 0
=
0 e+~ϕ·(~σ/2) I 0 0 e−~ϕ·(~σ/2)
−~ϕ·(~σ/2)
e−~ϕ·(~σ/2)
e 0 0
=
0 e+~ϕ·(~σ/2) e+~ϕ·(~σ/2) 0
−~
ϕ·~
σ
0 e
= +~ϕ·~σ
e 0
By the same argument, any odd power of (ϕ̂ · ~σ ) is equal to itself. This allows us to write
another simple expression: sin(θ ϕ̂ · ~σ ) = ϕ̂ · ~σ sin θ. Therefore, we have just proved that
Since sin(ix) = i sinh x, we can let θ = −iϕ to get the desired result
Next, we start with the form of the boosted Dirac equation written above equation (17) on
p. 118:
~ ~
e−i~ϕ·K γ 0 e+i~ϕ·K − I4 ψ(p) = 0
Using what we just derived in the first part of this problem gives
−I2 e−~ϕ·~σ
ψ(p) = 0
e+~ϕ·~σ −I2
54
p p
Identifying p~ = ϕ̂m sinh ϕ gives cosh ϕ = 1 + |~p |2 /m2 = m2 + |~p |2 /m = p0 /m, so after
multiplying through by m, the above equation becomes
−m I2 p0 − ~σ · p~
ψ(p) = 0
I2 p0 + ~σ · p~ −m
Define the 4-vectors σ µ ≡ (I2 , ~σ ) and σ̄ µ ≡ (I2 , −~σ ) (the bar is part of the name and does
not denote any sort of complex conjugation operation) to rewrite the above as
−m σ µ pµ
ψ(p) = 0
σ̄ µ pµ −m
µ 0 σµ
Defining γ = gives the Dirac equation:
σ̄ µ 0
(γ µ pµ − m)ψ(p) = 0
i
0 σ
This shows that γ i = , which completes the problem.
−σ i 0
Solution:
55
II.5 Vacuum Energy, Grassmann Integrals, and Feynman Dia-
grams for Fermions
1. Write down the Feynman amplitude for the diagram in figure II.5.1 for the scalar theory
(19). The answer is given in chapter III.3.
Solution:
= if
k −i
= 2 2
−k
p −i( p +m)
=
−p 2 + m 2
d4 k
−i −i (6 k+6 p + m)
Z
(Fig. II.5.1) = ū(p, s) (if ) (if ) u(p, s)
(2π)4 −k 2 + µ2 − i −(k + p)2 + m2 − i
d4 k
Z
2 1 (6 k+6 p + m)
= f ū(p, s) u(p, s)
(2π)4 k 2 − µ2 (k + p)2 − m2
2. Applying the Feynman rules for the vector theory (22) show that the amplitude for
the diagram in figure II.5.3 is given by
d4 k
6 p +6k + m µ
Z
2 2 1 kµ kν
(ie) i 4 2 2 2
− gµν ū(p)γ ν γ u(p) . (26)
(2π) k − µ µ (p + k)2 − m2
56
Solution:
−i
k
= ( −g + k 2k )
−k 2 2
The fermion propagator is the same as for the previous question. Again, reading the diagram
from left to right gives:
(Fig. II.5.3) =
Z 4
dk µ −i kµ kν −i (6 k+6 p + m)
ū(p, s) 4
(−ieγ ) 2 2
−gµν + 2 2 2
(−ieγ ν ) u(p, s)
(2π) −k + µ − i µ −(k + p) + m − i
4
6 k+6 p + m
Z
dk 1 kµ kν
= e2 4 2 2
−gµν + 2 ū(p, s)γ µ γ ν u(p, s)
(2π) −k + µ µ −(k + p)2 + m2
d4 k
6 k+6 p + m
Z
2 1 kµ kν
=e 4 2 2
−gµν + 2 ū(p, s)γ ν γ µ u(p, s)
(2π) k − µ µ (k + p)2 − m2
57
1
P
Summing over spins as s uū = 2m
(6 p + m) gives
2
21X 1 4πα
|M|2 = tr γ 0 (6 p + m)γ 0 (6 P + m)
M ≡
2 s,S 8m2 k2
tr γ 0 (6 p + m)γ 0 (6 P + m) = tr(γ 06 pγ 06 P + m2 I)
= tr(−6 p6 P + 2p0 γ 06 P + m2 I)
= 4(−pµ P µ + 2p0 P 0 + m2 )
= 4(p0 P 0 + p~ · P~ + m2 )
1 d3 P
dσ = 2π δ(p0 − P 0 ) M2
γ|~v | (2π)3 (P 0 /m)
p
where the relativistic factor γ ≡ 1/ 1 − |~v |2 is included in the denominator since we calcu-
late in the lab frame. The energy conserving delta function implies |~p | = |P~ |, as expected:
only the direction of the 3-momentum should change while the magnitude stays fixed. Choos-
ing not to integrate over the solid angle but integrating over |P~ |, we get
! p !
|~p |2 |~p |2 + m2 1 4πα 2
dσ m
(|~p |2 +m2 )+|~p |2 cos θ+m2
= 2
p 2 2
dΩ 2(2π) γ|~v | |~p |2 + m2 |~p | m k
2
|~p | 4πα 2 2
= 2 |~
p | (1 + cos θ) + 2m
8π γ|~v |m k 2
Using k 2 = −4|~p |2 sin2 (θ/2) from p. 133 along with cos θ = 1 − 2 sin2 (θ/2), we get
2
dσ |~p | πα
2 |~p |2 (1 − sin2 (θ/2)) + m2
= 2 2
dΩ 8π γ|~v |m |~p | sin (θ/2)
2
α2 1 2 2 2 2
= 4 |~
p | + m − |~
p | sin (θ/2)
4γ|~v ||~p |2 sin (θ/2) m|~p |
p
The relativistic 3-momentum is p~ = γm~v , with γ ≡ 1/ 1 − |~v |2 as before. From now on,
since there are no more 4-vectors, let v ≡ |~v | and p ≡ |~p |. The quantity in square brackets
58
simplifies as
p2 + m2 − p2 sin2 (θ/2) = γ 2 m2 v 2 + m2 − γ 2 m2 v 2 sin2 (θ/2)
2
v2
2 v 2
=m +1− sin (θ/2)
1 − v2 1 − v2
m2 2 2
= 1 − v sin (θ/2)
1 − v2
= γ 2 m2 [1 − v 2 sin2 (θ/2)]
The cross section is therefore
dσ α2 1
= 2 4 γ 2 m2 [1 − v 2 sin2 (θ/2)]
dΩ 4γvp sin (θ/2) m(γmv)
α2
= 2 2 4 [1 − v 2 sin2 (θ/2)] .
4v p sin (θ/2)
2. To order e2 the amplitude for positron scattering off a proton is just minus the amplitude
(3) for electron scattering off a proton. Thus, somewhat counterintuitively, the differential
cross sections for positron scattering off a proton and for electron scattering off a proton are
the same to this order. Show that to the next order this is no longer true.
Solution:
Consider the 1-loop corrections to the propagator of the exchanged photon. These will
all contribute with the same sign to the amplitudes for e+ p → e+ p and e− p → e− p. Since the
tree level amplitudes differ by a sign, the amplitudes including 1-loop effects are different.
3. Show that the trace of a product of an odd number of gamma matrices vanishes.
Solution:
σαµα̇
µ 0
The gamma matrices can be written in the Weyl basis as γ = . In this
σ̄ µα̇α 0
representation, it is immediately clear that the trace of an odd number of gamma matrices
is zero simply because such a quantity cannot be written down. For example,
µ ν α̇β
µ ν σαα̇ σ̄ 0 µ ν tr(σ µ σ̄ ν ) 0
γ γ = =⇒ tr(γ γ ) =
0 σ̄ µα̇α σαν β̇ 0 tr(σ̄ µ σ ν )
but !
0 σαµα̇ σ̄ ν α̇β σβρ γ̇
γ µγ ν γ ρ =
σ̄ µα̇α σαν β̇ σ̄ ρβ̇γ
cannot be traced over, since the spinor indices cannot be contracted.
59
6. For those who relish long calculations, determine the differential cross section for electron-
electron scattering without taking the relativistic limit.
Solution:
d3 p
P R
For this problem, we use s us (~p )ūs (~p ) =6 p + m for the spin sum and (2π)3 2ωp~
for the
measure. The relevant Feynman rules are:
6p + m
electron propagator: i
p2 − m2 + iε
−i ηµν
photon propagator: (Feynman gauge)
p2 + iε
vertex: = −ieγ µ
An incoming electron gets a us (~p ) and an outgoing electron gets a ūs0 (~p 0 ). We will use the
notation ui ≡ usi (~pi ) and primes for the outgoing particles. There are two diagrams that
contribute to M at tree level:
1 1’ 1 2’
q q’
2 2’ 2 1’
Remembering the relative minus sign between these diagrams gives the amplitude:
−iηµν −iηµν
M = (ū10 ieγ µ u1 ) 2 (ū20 ieγ ν u2 ) − (ū20 ieγ µ u1 ) 0 2 (ū10 ieγ ν u2 ) + O(e4 )
q + iε (q ) + iε
µ µ
2 (ū1 γ u1 )(ū2 γµ u2 )
0 0 (ū2 γ u1 )(ū1 γµ u2 )
0 0
= ie 2
− 2
+ O(e4 )
(p1 − p10 ) + iε (p1 − p20 ) + iε
Now we need |M|2 . Remember that M is a Lorentz-scalar (it is just a number), so the
conjugation can be carried out as M∗ = M† . However, we are used to working with u and
ū rather than u and u† , as required by Lorentz invariance. We have defined a barred spinor
as ψ̄ ≡ ψ † γ 0 , so for consistency we define a barred matrix as M ≡ γ 0 M † γ 0 . For instance,
(γ µ ) = γ 0 (γ µ )† γ 0 = γ 0 (γ 0 γ µ γ 0 )γ 0 = γ µ .
60
Just as (AB)† = B † A† , barring also reverses the order. So to compute the conjugate of the
amplitude, we write |M|2 = MM :
A B C +D
|M|2 /e4 = + −
[(p1 − p10 )2 ]2 [(p1 − p20 )2 ]2 (p1 − p10 )2 (p1 − p20 )2
Each set of parentheses denotes a Lorentz-scalar, meaning that all spinor indices are con-
tracted. To figure out how the Dirac traces work, we write out the indices explicitly and
rearrange everything in matrix multiplication order. Denoting a = 1, 2, 3, 4 as a Dirac 4-
spinor index, we rearrange as follows:
P
Now we would like to sum over all spins and use s ui ūi = (6 pi + m). To compute the
amplitude, we want to average over the initial spins, which gives a factor of (1/2) for each
incoming electron. We will multiply by this factor of (1/2)2 = 1/4 later. For now, we simply
define the spin-summed version of each of the above pieces:
X X X X
α≡ A , β≡ B , γ≡ C , δ≡ D
s1 ,s2 ,s10 ,s20 s1 ,s2 ,s10 ,s20 s1 ,s2 ,s10 ,s20 s1 ,s2 ,s10 ,s20
61
We now simplify the above with the following gamma matrix identities:
Tr[6 a6 b] = 4 (ab)
Tr[6 a6 b6 c6 d] = 4 [(ab)(cd) + (ad)(bc) − (ac)(bd)]
Tr[odd number of γs] = 0
γµ γ µ = 4 = η µν ηµν
γµ6 aγ µ = −26 a
γµ6 a6 bγ µ = 4(ab)
γµ6 a6 b6 cγ µ = −26 c6 b6 a
The notation (ab) = aµ bµ is used above. It is also useful to define a further condensed
notation: (ij) ≡ pi µ pµj . We will only use the latter notation when there is no risk of confusion.
Let us now simplify the first term:
The parentheses on the first term inside each [...] is there just to show that the term is
symmetric in µ and ν, and it is thus useful to organize the calculation using this grouping.
Continuing:
α = 16[(pµ10 pν1 + pν10 pµ1 )(p2µ p20 ν + p2ν p20 µ ) − 2(p2 p20 )(p10 p1 ) + 4m2 × 2(p1 p10 )
− 2(p1 p10 )(p2 p20 ) + (p1 p10 )(p2 p20 ) η µν ηµν −4m2 (p1 p10 )η µν ηµν + 4m2 × 2(p2 p20 )
| {z }
= +4
2 µν 4 µν
− 4m (p2 p20 )η ηµν + 16m η ηµν ]
= 16[(10 2)(120 ) + (10 20 )(12) + (10 20 )(12) + (10 2)(120 ) − 2(220 )(10 1) + 8m2 (10 1)
− 2(110 )(220 ) + 4(110 )(220 ) − 16m2 (110 ) + 8m2 (220 ) − 16m2 (220 ) + 64m4 ]
62
A quick check from dimensional analysis: Each (ij) has dimensions of m2 , so each term has
dimensions of m4 , so the result is at least dimensionally correct. The next term is:
= 16[(pµ1 pν20 + pν1 pµ20 )(p2µ p10 ν + p2ν p10 µ ) − 2(120 )(210 ) + 4m2 × 2(120 )
− 2(120 )(210 ) + 4(120 )(210 ) − 4m2 (120 ) × 4 + 8m2 (210 ) − 4m2 (210 ) × 4 + 16m4 × 4]
63
γ = Tr[γ µ (6 p1 + m)γ ν (6 p20 + m)γµ (6 p2 + m)γν (6 p10 + m)]
= Tr[(γ µ6 p1 + mγ µ )(γ ν6 p20 + mγ ν )(γµ6 p2 + mγµ )(γν6 p10 + mγν )]
= Tr[(γ µ6 p1 γ ν6 p20 + mγ µ6 p1 γ ν + mγ µ γ ν6 p20 + m2 γ µ γ ν )
× (γµ6 p2 γν6 p10 + mγµ6 p2 γν + mγµ γν6 p10 + m2 γµ γν )]
= −32(p1 p2 )(p10 p20 )+16m2 [(p20 p1 )+(p2 p1 )+(p1 p10 )+(p2 p20 )+(p10 p20 )+(p2 p10 )] − 32m4
64
One more:
δ = Tr[γ µ (6 p10 + m)γν (6 p2 + m)γµ (6 p20 + m)γ ν (6 p1 + m)]
= −32(p10 p20 )(p1 p2 ) + 16m2 [(p10 p2 ) + (p10 p20 ) + (p1 p10 ) + (p2 p20 ) + (p1 p2 ) + (p1 p20 )] − 32m4
We see that this equals the previous term: γ = δ. Now we have all the terms we need to
compute the amplitude-squared. Define theP amplitude-squared after averaging over initial
1
spins and summing over final spins: M ≡ 4 {all spins} |M|2 . We have:
2
e4
2 α β 2δ
M = + −
4 [(p1 − p10 )2 ]2 [(p1 − p20 )2 ]2 (p1 − p10 )2 (p1 − p20 )2
65
Until now, we have not used any information about the actual kinematics of the scattering
process. Remembering that all external particles are on-shell with equal masses, we know
p2i = m2 , where m is the
Pmass of the electron. We now define the Mandelstam variables s, t, u
such that s + t + u = i mi = 4m2 :
2
1
s ≡ (p1 + p2 )2 = (p21 + p22 + 2(12)) = 2m2 + 2(12) =⇒ (12) = (s − 2m2 )
2
= (p10 + p20 )2 =⇒ (10 20 ) = (12)
1
t ≡ (p1 − p10 )2 = (p21 + p210 − 2(110 )) = 2m2 − 2(110 ) =⇒ (110 ) = − (t − 2m2 )
2
= (p2 − p20 )2 =⇒ (220 ) = (110 )
1
u ≡ (p1 − p20 )2 = (p21 + p220 − 2(120 )) = 2m2 − 2(120 ) =⇒ (120 ) = − (u − 2m2 )
2
= (p2 − p10 )2 =⇒ (210 ) = (120 )
Now that we have gotten this far, we will drop the mass of the electron to check with the
results in the text. Setting m = 0 simplifies the dot products: (12) = +s/2 , (110 ) =
−t/2 , (120 ) = −u/2, where now s + t + u = 0. The pieces of the amplitude-squared are now:
1 1
α = 32[ u2 + s2 ] = 8(s2 + u2 )
4 4
1 2 1 2
β = 32[ s + t ] = 8(s2 + t2 )
4 4
1 2
δ = −16[ s ] = −8s2
2
The amplitude-squared is therefore:
e4 α
2 β 2δ
M = + −
4 t2 u2 tu
e4 8(s2 + u2 ) 8(s2 + t2 ) 2(−8s2 )
= + −
4 t2 u2 tu
2 2 2 2 2
s +u s +t 2s
= 2e4 2
+ 2
+
t u tu
66
2
From the definition of s, we see that s = ECM , where ECM is the total energy in the center-
of-mass frame. We now calculate the differential cross section, whose formula is given in
Appendix C. The differential cross section in the center-of-mass frame is:
1 d3 p10 d3 p20
dσCM = √ M2 (2π)4 δ 4 (p1 + p2 − p10 − p20 )
4|~p1 |CM s (2π)3 2p010 (2π)3 2p020
p
where p0i = |~pi |2 + m2 = |~pi | for m = 0. For 4 identical particles with zero mass, the
prefactor can be simplified:
√ 1
q
1
|~p1 |CM s = s2 − 2(m21 + m22 )s + (m21 − m22 )2 = s
2 2
Now, the 4-dimensional delta function is a product of a one dimensional energy-conserving
delta function and a 3-dimensional momentum-conserving delta function. As discussed, √ the
0 0
total energy of the incoming electrons in the center-of-mass frame is p1 + p2 = s . Fur-
thermore, the center-of-mass frame is defined such that p~1 + p~2 = ~0, so the differential cross
section is:
1 2 d3 p10 d3 p20
dσCM = M (2π)4 δ 4 (p1 + p2 − p10 − p20 )
2s (2π)3 2p010 (2π)3 2p020
M2 d3 p10 d3 p20 √
= 3 2
δ( s − |~p10 | − |~p20 |) δ 3 (~p10 + p~20 )
2 (2π) s |~p10 | |~p20 |
We see that the 3-dimensional delta function simply sets p~20 = −~p10 :
Z
d3 p20 f (~p10 , p~20 ) δ 3 (~p10 + p~20 ) = f (~p10 , −~p10 )
Therefore, integrating over p~20 along with the fact that δ(2x) = 12 δ(x) gives:
M2 d3 p1 0 √
dσCM = δ( s − |~p10 |)
24 (2π)2 s |~p10 |2
Now, using spherical coordinates for p~10 gives d3 p10 = dΩ dp p2 , where Ω is the solid angle
over which you are not asked to integrate, and p ≡ |~p10 |. The differential cross section per
solid angle is now a 1-dimensional integral:
Z ∞ √
dσ 1 2 s
= dp M δ −p
dΩ CM 24 (2π)2 s 0 2
√
The delta function just sets p = s /2, so all we have to do now is figure out how these
integrals over delta functions affected M2 . We can do this by looking at the Mandelstam
67
variables:
Solution:
We will work out the case n = 2, which will make the generalization to arbitrary n clear.
There are now (2 + 1)! = 6 diagrams, corresponding to the different ways in which the photon
lines can cross. We will compute them in groups of three.
68
p’ p’ p’
2
k2 k2
2
p’−k 2 k k2
p’−k 2
p’−k
k1 k1
1
k 1
k1
p+k p+k1 p+k2
1 2
p k
p p
where we choose the convention in which the internal photon momenta always flow towards
the left. The three ks are related by p0 − p = k + k1 + k2 . The amplitude given by adding
(bose statistics) these three diagrams is:
0 λ2 1 λ1 1 λ2 1 1 λ1 1 λ1 1 λ2
M1 = ū γ γ 6k + γ 6k γ + 6k 0 γ γ u.
6 p 0 − 6 k2 6p +6k 6 p 0 − 6 k2 6 p + 6 k1 6p −6k 6 p + 6 k2
Here we have dropped the fermion masses since they play no role in the proof. We have also
suppressed all irrelevant factors in the amplitude, retaining only the kµ term in the photon
propagator, as explained in the text.
In the numerator of the first term, write 6 k = (6 p +6 k)−6 p, and in the third term write
6 k = −(6 p 0 − 6 k)+ 6 p 0 . The Dirac equation is now 6 pu = 0 and 6 p 0 u0 = 0, so that the 6 p gives
zero acting to the right, and 6 p 0 gives zero acting to the left. We have
0 λ2 1 λ1 λ2 1 1 λ1 λ1 1 λ2
M1 = ū γ γ +γ 6k γ −γ γ u
6 p 0 − 6 k2 6 p 0 − 6 k2 6 p + 6 k1 6 p + 6 k2
0 λ2 1 λ1 λ2 1 1 λ1 λ1 1 λ2
= ū γ γ +γ 6k γ −γ γ u
6 p + 6 k+ 6 k1 6 p + 6 k+ 6 k1 6 p + 6 k1 6 p + 6 k2
where in the second line we have written p 0 − k2 = p + k + k1 . In the numerator of the middle
term, write k = (p + k1 + k) − (p + k1 ) to get:
0 λ2 1 λ1 λ2 1 λ1 λ2 1 λ1 λ1 1 λ2
M1 = ū γ γ +γ γ −γ γ −γ γ u
6 p + 6 k+ 6 k1 6 p + 6 k1 6 p + 6 k+ 6 k1 6 p + 6 k2
0 λ2 1 λ1 λ1 1 λ2
= ū γ γ −γ γ u
6 p + 6 k1 6 p + 6 k2
where we have canceled the first and third terms from the first line. We see that the remain-
ing two terms do not cancel; indeed, this is one way to discover additional diagrams in case
you had forgotten them. That is, it is obvious that there are at least 3 diagrams but it may
have escaped your attention that there are actually a total of 3! = 6 diagrams.
69
p’ p’ p’
1 k2 1
k1 k1 k k2
p’−k 1 p’−k 1 p’−k
2 k 2
As before, two of the resulting terms will cancel each other and the remaining terms will
cancel the leftover terms in M1 . The amplitude from these three new diagrams is:
λ1 1 λ2 1 λ1 1 1 λ2 1 λ2 1 λ1
M2 = ū γ γ 6k + γ 6k γ + 6k 0 γ γ u.
6 p 0 − 6 k1 6p +6k 6 p 0 − 6 k1 6 p + 6 k2 6p −6k 6 p + 6 k1
In the numerator of the first term, write k = (p + k) − p and in the third term write
k = −(p 0 − k) + p0 . Again 6 p acting to the right gives zero, and 6 p 0 acting to the left gives
zero. The amplitude is
λ1 1 λ2 λ1 1 1 λ2 λ2 1 λ1
M2 = ū γ γ +γ 6k γ −γ γ u
6 p 0 − 6 k1 6 p 0 − 6 k1 6 p + 6 k2 6 p + 6 k1
λ1 1 λ2 λ1 1 1 λ2 λ2 1 λ1
= ū γ γ +γ 6k γ −γ γ u
6 p + 6 k+ 6 k2 6 p + 6 k+ 6 k2 6 p + 6 k2 6 p + 6 k1
where we have used p0 − k1 = p + k + k2 . In the numerator of the middle term, write
k = (p + k + k2 ) − (p + k2 ) to get:
λ1 1 λ2 λ1 1 λ2 λ1 1 λ2 λ2 1 λ1
M2 = ū γ γ +γ γ −γ γ −γ γ u
6 p + 6 k+ 6 k2 6 p + 6 k2 6 p + 6 k+ 6 k2 6 p + 6 k1
λ1 1 λ2 λ2 1 λ1
= ū γ γ −γ γ u = −M1
6 p + 6 k2 6 p + 6 k1
Thus the contribution from the offending term kµ kν /µ2 in the photon propagator contributes
a total of M = M1 + M2 = 0 to the amplitude. In general, the photon with momentum k
can connect to the left-most line in (n + 1) ways, which results in (n + 1)! different diagrams,
whose constituent terms will cancel in pairs.
2. You might have worried whether the shift of integration variable is allowed. Rationalizing
the denominators in the first integral
d4 p
Z
ν 1 σ 1 λ 1
tr γ γ γ
(2π)4 6 p2 − m 6 p1 − m 6 p − m
70
in (13) and imagining doing the trace, you can convince yourself that this integral is only
logarithmically divergent and hence that the shift is allowed. This issue will come up again
in chapter IV.7 and we are anticipating a bit here.
Solution:
Solution:
Let us first derive a general Lorentz-covariant definition of the sum over photon polarizations.
For a circularly polarized3 photon moving in the +ẑ direction, the polarization vectors can
be taken as
0 0
1 1 1 1
εµ1 = √ µ
, ε 2 = √ .
2 +i 2 −i
0 0
Explicitly computing the sum over polarizations gives
0 0 0 0
0 1 0 0
X
εµa εν∗
a =
.
0 0 1 0
a = 1,2
0 0 0 0
3
We are using circular polarization to show an example with complex polarization vectors.
71
We would like to write this in terms of Lorentz-covariant notation. The 4-momentum of the
photon is k µ = |~k |(1, 0, 0, 1)T , so that
1 0 0 1
kµkν 0 0 0 0
= .
~
|k |2 0 0 0 0
1 0 0 1
Define the spacelike unit vector in the direction of propagation of the photon: nµ = (0, 0, 0, 1)T .
Then we have
0 0 0 1
k µ nν + k ν nµ 0 0 0 0
= .
~
|k | 0 0 0 0
1 0 0 2
Furthermore, we can write |~k | = −k · n. Therefore, using the Lorentz-invariant metric
η µν = diag(+1, −1, −1, −1) we can write
X k µ nν + k ν nµ kµkν
εµa (k)ενa (k)∗ = −η µν + + .
a = 1,2
k·n (k · n)2
We may now proceed with the problem. We can use the above result to write:
" #" #
X X
S≡ εµa (k)ενa (k)∗ εa0 µ (k 0 )εa0 ν (k 0 )∗ =
a = 1,2 a0 = 1,2
(k 0 · n) (k · n0 ) 1 (k · k 0 )2
2 0 0 0 0
(k · n )(k · n)+(k · k ) (n · n )+ 0 0 + + .
(k · n)(k 0 · n0 ) (k · n ) (k · n) 2 (k · n)(k 0 · n0 )
We have:
k · k 0 = ωω 0 (1 − cos θ) , n · n0 = − cos θ
k · n = −ω , k 0 · n0 = −ω 0
k · n0 = −~k · n̂0 = −ω cos θ , k 0 · n = −~k 0 · n̂ = −ω 0 cos θ .
72
Therefore (let c ≡ cos θ):
= 2{c2 + c − c2 + 21 (1 − 2c + c2 )}
= 2{c + 21 − c + 12 c2 }
= 1 + c2
= 2 − sin2 θ .
The problem wants us to take equation (9) on p. 154 and replace it with
X X XX
M2 ≡ 12 1
2
|M| 2
a = 1,2 a0 = 1,2
0
e4
ω ω 1
= 2 + − 2 + 4 × 2S
(2m)2 ω ω0
0
e4
ω ω 2
= 2 + − 4 + 2(2 − sin θ)
(2m)2 ω ω0
0
e4
ω ω 2
= 2 + − 2 sin θ
(2m)2 ω ω0
0
e4
ω ω 2
= 2 + − sin θ .
(2m)2 ω ω0
Solution:
The explicit forms of the photon polarization vectors never entered into the derivation of
the Klein-Nishina formula on p. 155; the only assumption was that the polarization vectors
were real. For circularly polarized photons, everything goes through as before except that the
εµa are complex. The result is equation (12) on p. 155 with the replacement (εε0 )2 → |ε∗ ε0 |2 .
73
III Renormalization and Gauge Invariance
III.1 Cutting Off Our Ignorance
1. Work through the manipulations leading to (9) without referring to the text.
Solution:
Λ2 Λ2 Λ2
M = −iλbare + iCλ2bare log + log + log + O(λ3bare )
s t u
Here we write the physical coupling as λ and the unphysical bare coupling as λbare .
We now pass to the notation of equations (5) and (6), where the sum of logs in the square
brackets is denoted simply by L:
74
III.3 Counterterms and Physical Perturbation Theory
1. Show that in (1 + 1)-dimensional spacetime the Dirac field ψ has mass dimension 12 , and
hence the Fermi coupling is dimensionless.
Solution:
R
The action S = dd x L is dimensionless, so L has dimensions of (mass)d , denoted by [L] = d.
The derivative ∂µ has mass dimension +1, so the kinetic term ψ̄γ µ ∂µ ψ gives:
d−1
L = iψ̄6 ∂ψ =⇒ [L] = [∂] + 2[ψ] =⇒ d = 1 + 2[ψ] =⇒ [ψ] =
2
Therefore, for d = 1 + 1 = 2, we have [ψ] = 1/2.
The Fermi interaction has the form L = G (ψ̄ψ)(ψ̄ψ). As above, this implies:
d−1
[L] = [G] + 4[ψ] =⇒ d = [G] + 4 =⇒ [G] = 2 − d
2
Solution:
As in the book, we are really doing physical (“dressed”) perturbation theory, in which case
we should have renormalizing factors for each term or, equivalently, we should add a series of
counterterms. For this problem, all that matters is the structure of the Lagrangian, so this
issue is not important.
Introduce the sources J for ϕ, η for ψ̄ and η̄ for ψ and expand the path integral with
sources in a Taylor series, as in equation (4) on page 44 and as described below equation (20)
75
on page 128:
Z
d4 x(Jϕ+η̄ψ+ψ̄η)
R
Z(η, η̄, J) = DψDψ̄Dϕ e iS[ψ,ψ̄,ϕ]+i
Z
d4 x Lint (ϕ→−iδJ ,ψ→−iδη ,ψ̄→+iδη̄ )
R
i
=e DψDψ̄Dϕ e iS[ψ,ψ̄,ϕ]
∞
X X ∞
X ∞
X
∼ [δJ4 ]Vλ [δJ δη δη̄ ]Vf [J Gϕ J]BI [η̄ Gψ η]FI
Vλ = 0 Vf = 0 BI = 0 FI = 0
The tilde means we are being schematic and ignoring all numerical factors, which are irrel-
evant for this problem. Gϕ is the propagator for ϕ, and Gψ is the propagator for ψ, both
of which appear from performing the Gaussian integrals as usual. Vλ , Vf , BI and BF are, at
this level, simply dummy summation variables.
These names are chosen for a reason: Vλ counts the number of λϕ4 vertices, Vf counts
the number of f ϕψ̄ψ vertices, BI counts the number of scalar propagators (“internal boson
lines”), and FI counts the number of fermion propagators (“internal fermion lines”).
Similarly, each vertex kills Vf ηs, as can be read off from the term [...δη ...]Vf , and each
fermion propagator goes with one η, as can be read from [...η]FI , so the number of ηs killed
is Vf − FI . The number of η̄s killed is the same, so FE = 2(Vf − FI ).
76
Plugging in the values for Vf and Vλ gives:
1 1
D = 2BI + 3FI − BE + 2BI − FI − FE − 4 FI + FE + 4
2 2
1
= (2 − 2)BI + (4 − 4)FI − BE + FE − 4FI − 2FE + 4
2
3
= −BE − FE + 4
2
Now consider the Fermi theory in (1+1) dimensions:
5. Show that the result P = L − 1 holds for all the theories we have studied.
Solution:
77
III.5 Field Theory without Relativity
1. Obtain the Klein-Gordon equation for a particle in an electrostatic potential (such as
that of the nucleus) by the gauge principle of replacing (∂/∂t) in (2) by ∂/∂t − ieA0 . Show
that in the nonrelativistic limit this reduces to the Schrödinger’s equation for a particle in
an external potential.
Solution:
To add a photon to the relativistic free field theory L = (∂Φ† )(∂Φ) − m2 Φ† Φ, replace the
partial derivatives with covariant derivatives as ∂µ Φ → Dµ Φ ≡ (∂µ − ieAµ )Φ. (We also need
the photon kinetic term − 41 Fµν F µν , but it will play no role in what follows, so we drop it.)
The Lagrangian is
L = (Dµ Φ)† Dµ Φ + m2 Φ† Φ = ∂µ Φ† ∂ µ Φ + m2 Φ† Φ + ieAµ (Φ† ∂µ Φ − Φ∂µ Φ† ) + e2 Φ† ΦAµ Aµ
with a purely electric potential independent of time: Aµ (~x, t) = V (~x)δ µ0 . Now take the
non-relativistic limit as in the text:
1 1
Φ(~x, t) = √ e−imt ϕ(~x, t) =⇒ ∂t Φ = √ (−im ϕ + ∂t ϕ)e−imt
2m 2m
Therefore
1 +imt † 1 1
Φ† ∂t Φ = √ e ϕ √ (−im ϕ + ∂t ϕ)e−imt = (−im ϕ† ϕ + ϕ† ∂t ϕ)
2m 2m 2m
and so we have
1 1
Φ† ∂t Φ − Φ∂t Φ† = (−im ϕ† ϕ + ϕ† ∂t ϕ) − (+im ϕ† ϕ + ϕ∂t ϕ† )
2m 2m
1
= −i ϕ† ϕ + (ϕ† ∂t ϕ − ϕ∂t ϕ† ) .
2m
The term linear in the potential is therefore
µ † † † 1 † †
ieA (Φ ∂µ Φ − Φ∂µ Φ ) = ieV (~x) −i ϕ ϕ + (ϕ ∂t ϕ − ϕ∂t ϕ )
2m
ie
= eV (~x)ϕ† ϕ + V (~x)(ϕ† ∂t ϕ − ϕ∂t ϕ† )
2m
ie
= eV (~x)ϕ† ϕ + V (~x)ϕ† ∂t ϕ + (total time derivative)
m
To this we add the term quadratic in the potential,
e2 †
e2 Φ† ΦAµ Aµ = ϕ ϕV (~x)2
2m
Adapting the previously obtained Lagrangian from the book (equation (6) on p. 191), we
obtain (up to total derivatives)
e2
† 1 2 ie 2
L = ϕ i∂t + ∇ + V (~x)∂t + eV (~x) + V (~x) ϕ
2m m 2m
78
δS
Set δϕ†
= 0 to get the equation of motion:
h
e i 1 2 h e i
1 + V (~x) i∂t + ∇ + 1+ V (~x) eV (~x) ϕ(~x, t) = 0
m 2m 2m
In the non-relativistic limit m ∂t ϕ ϕ, and neglecting the term quadratic in the potential,
we obtain
1 2
i∂t + ∇ + eV (~x) ϕ(~x, t) = 0
2m
∇ 2
which is precisely the Schrödinger equation (i∂t + H)ϕ = 0 with H = 2m + eV the correct
non-relativistic Hamiltonian for a particle in the presence of an external potential eV .
3. Given a field theory we can compute the scattering amplitude of two particles in the non-
relativistic limit. We then postulate an interaction potential U (~x) between the two particles
and use nonrelativistic quantum mechanics to calculate the scattering amplitude, for example
in Born approximation. Comparing the two scattering amplitudes we can determine U (~x).
Derive the Yukawa and the Coulomb potentials this way. The application of this method to
the λ(Φ† Φ)2 interaction is slightly problematic since the delta function interaction is a bit
singular, but it should be all right for determining whether the force is repulsive or attractive.
Solution:
79
III.6 The Magnetic Moment of the Electron
3. By Lorentz invariance the right hand side of (7) has to be a vector. The only possibilities
are ūγ µ u, (p + p0 )µ ūu, and (p − p0 )µ ūu. The last term is ruled out because it would not be
consistent with current conservation. Show that the form given in (7) is in fact the most
general allowed.
iσ µν qν
0 0 µ 0 0 µ 2
hp , s |J (0)|p, si = ū(p , s ) γ F1 (q ) + F2 (q ) u(p, s) , q ≡ p0 − p
2
(7)
2m
Solution:
First, for notational convenience, write u(p, s) = u and u(p0 , s0 ) = u0 . Lorentz invariance
dictates that if we have one free µ index on the left, we need one free µ index on the right.
The only such vectors we have are γ µ , q µ and (p + p0 )µ . However, since Lorentz invariance
also dictates that we have no free spinor indices on the right, any vector we write down is
going to be sandwiched between ū0 and u. Therefore, we may use the Gordon decomposition
rearranged as:
ū0 (p + p0 )µ u = ū0 (2m γ µ − iσ µν qν ) u
So between the spinors, (p + p0 ) is a linear combination of γ and q. Therefore anytime we
write a function of (p + p0 ) we may as well write it as a function of q. Equation (7) follows
immediately.
4. In chapter II.6, when discussing electron-proton scattering, we ignored the strong in-
teraction that the proton participates in. Argue that the effects of the strong interaction
could be included phenomenologically by replacing the vertex ū(P, S)γ µ u(p, s) in (II.6.1) by
iσ µν qν
µ µ 2 2
hP, S|J (0)|p, si = ū(P, S) γ F1 (q ) + F2 (q ) u(p, s) (17)
2m
Careful measurements of electron-proton scattering, thus determining the two proton form
factors F1 (q 2 ) and F2 (q 2 ), earned R. Hofstadter the 1961 Nobel Prize. While we could ac-
count for the general behavior of these two form factors, we are still unable to calculate them
from first principles (in contrast to the corresponding form factors for the electron.) See
chapters IV.2 and VII.3.
Solution:
This form follows from parity and gauge invariance. Consult the literature, for example
L. N. Hand, D. G. Miller and R. Wilson, “Electric and magnetic form factors of the nu-
cleon,” Rev. Mod. Phys, Vol. 35, No. 2, Apr. 1963 and J. D. Bjorken and E. A. Paschos,
“Inelastic Electron-Proton and γ-Proton Scattering and the Structure of the Nucleon,” Phys.
Rev. Vol. 185, No. 3, 25 Sep. 1969. For a modern development regarding the form factors
for the proton, see V. Barger, C. W. Chiang, W. Y. Keung and D. Marfatia, “Proton Size
Anomaly,” Phys. Rev. Lett. 106:153001, 2011 (arXiv:1011.3519v2 [hep-ph]).
80
III.7 Polarizing the Vacuum and Renormalizing the Charge
1. Calculate Πµν (q) using dimensional regularization.
Solution:
For this problem, we will use the metric signature g = (−, +, +, +). The scheme of di-
mensional regularization is to calculate everything in d dimensions, then write d = 4 − ε
and take ε → 0+ . The divergent piece will show up as a term proportional to 1/ε. The
polarization tensor is:
The simplification followed from remembering that the trace of an odd number of gamma
matrices is zero. In d dimensions, the gamma matrices have dimension 2d/2 , so the trace
identities become:
Tr[γ µ γ ν ] = −2d/2 g µν
Tr[(6 k+6 p)γ ν6 pγ ν ] = 2d/2 [(k + p)µ pν + (k + p)ν pµ − g µν (k + p) · p ]
dd p (k + p)µ pν + (k + p)ν pµ − g µν (k + p) · p − m2 g µν
Z
µν d/2
iΠ (k) = −2
(2π)d [(k + p)2 + m2 ] [p2 + m2 ]
81
1
dd p (k + p)µ pν + (k + p)ν pµ − g µν (k + p) · p − m2 g µν
Z Z
µν d/2
iΠ (k) = −2 dx
(2π)d 0 {x [(k + p)2 + m2 ] + (1 − x)(p2 + m2 )}2
1
dd p (k + p)µ pν + (k + p)ν pµ − g µν (k + p) · p − m2 g µν
Z Z
d/2
= −2 dx
(2π)d 0 [x(k 2 + p2 + 2k · p + m2 ) + p2 + m2 − xp2 − xm2 ]2
1
dd p (k + p)µ pν + (k + p)ν pµ − g µν (k + p) · p − m2 g µν
Z Z
d/2
= −2 dx
(2π)d 0 [p2 + 2x k · p + xk 2 + m2 ]2
1
dd p (k + p)µ pν + (k + p)ν pµ − g µν (k + p) · p − m2 g µν
Z Z
d/2
= −2 dx
(2π)d 0 [(p + xk)2 + x(1 − x)k 2 + m2 ]2
R R
Now interchange the order of integration so that the dd p comes before the dx, and shift
the integration variable to q ≡ p + xk =⇒ p = q − xk. For notational convenience, also
define D ≡ x(1 − x)k 2 + m2 . We get:
iΠµν (k) =
Z 1 Z d
d/2 d q [q+(1−x)k]µ (q−xk)ν +(µ ↔ ν)−g µν [q+(1−x)k]·(q−xk)−m2 g µν
−2 dx
0 (2π)d [q 2 + D]2
Z 1 Z
dd q 2q µ q ν − 2x(1 − x)k µ k ν − g µν [q 2 − x(1 − x)k 2 ] − m2 g µν
= −2d/2 dx
0 (2π)d [q 2 + D]2
We have dropped terms in the numerator that are R linear in q, because now the denominator
is a function of the Lorentz-invariant q , so that d q q /(q 2 + D)2 = 0. On a related note:
2 d µ
Z Z
d µ ν 2 µν
d q q q f (q ) = C g dd q q 2 f (q 2 ) by Lorentz invariance
Z Z
d µ ν 2 µν d 2 2
gµν d q q q f (q ) = C g d q q f (q )
Z Z
dd q q 2 f (q 2 ) = C Tr(Id ) dd q q 2 f (q 2 )
1=Cd
Z Z
1
=⇒ d q q q f (q ) = g µν
d µ ν 2
dd q q 2 f (q 2 )
d
82
iΠµν (k) =
dd q d2 q 2 g µν − 2x(1 − x)k µ k ν − g µν [q 2 − x(1 − x)k 2 ] − m2 g µν
Z 1 Z
d/2
= −2 dx
0 (2π)d [q 2 + D]2
dd q 2( d1 − 1) q 2 g µν − 2x(1 − x)k µ k ν + g µν x(1 − x)k 2 − m2 g µν
Z 1 Z
= −2d/2 dx
0 (2π)d [q 2 + D]2
Z 1
d/2 2 µν
µ ν µν 2 2 µν
= −2 dx − 1 g I1 + −2x(1 − x)k k + g x(1 − x)k − m g I0
0 d
As you can see, we have started to plug in d = 4 − ε. Before continuing, note a subtlety.
In d = 4 spacetime dimensions, the gauge field Aµ has mass dimension +1 and the Dirac
spinor ψ has mass dimension +3/2, so the gauge coupling e is dimensionless. But we are no
longer working in d = 4 but rather in d = 4 − ε, so we need to find out how e changes with
dimension. Taking typical terms in a Lagrangian, we get:
d−2
L = −(∂A)2 =⇒ [A] =
2
d−1
L = iψ̄ 6 ∂ψ =⇒ [ψ] =
2
d ε
L = e Aµ ψ̄γ µ ψ =⇒ [e] = 2 − =
2 2
To make this mass dimension explicit, we rewrite the gauge coupling as e → eµ̃ε/2 , where
the parameter µ̃ has dimensions of mass, so that e remains dimensionless for all d. The
important point here is that since Π appears with an e2 , this consideration sends Π → µ̃ε Π.
So in the expression for the integral In , the (4π/D)ε/2 term is changed to (4π µ̃2 /D)ε/2 .
83
ε
2d/2 = 22−ε/2 = 4 × 2−ε/2 = 4 × 1 − ln 2 + O(ε2 )
2
2 2 1 ε 1 ε
−1= −1= 1 + + O(ε2 ) − 1 = − + + O(ε2 )
d 4−ε 2 4 2 8
ε ε
Γ(3 − 2 ) ε Γ(2 − 2 )
ε
ε = 2− ε =2−
Γ(2 − 2 ) 2 Γ(2 − 2 ) 2
ε 2
Γ −1 + = − + γ − 1 + O(ε)
2 ε
ε 2
Γ = + − γ + O(ε)
2 ε
2 ε/2
4π µ̃2
4π µ̃ ε
= 1 + ln + O(ε2 )
D 2 D
ε/2
4π µ̃2
i ε ε
I1 = 2− Γ −1 + D
16π 2 2 2 D
4π µ̃2
i ε −2 ε 2
= 1− + γ − 1 + O(ε) D 1 + ln + O(ε )
8π 2 4 ε 2 D
4π µ̃2
i −2 1 ε 2
= + γ − 1 + + O(ε) D 1 + ln + O(ε )
8π 2 ε 2 2 D
4π µ̃2
i −2
= D − ln + γ − 1 + O(ε)
8π 2 ε D
4π µ̃2
i −2 γ 1
= D − ln + ln e − + O(ε)
8π 2 ε D 2
2
−i 2 4π µ̃ 1
= 2
D + ln γ
+ + O(ε)
8π ε e D 2
i ε 4π µ̃2 ε/2
I0 = Γ
16π 2 2 D
4π µ̃2
i 2 ε 2
= − γ + O(ε) 1 + ln + O(ε )
16π 2 ε 2 D
4π µ̃2
i 2
= + ln − γ + O(ε)
16π 2 ε D
4π µ̃2
i 2
= + ln + O(ε)
16π 2 ε eγ D
84
Define µ2 ≡ 4π µ̃2 /eγ for convenience. After all, µ̃ was defined arbitrarily anyway, just as long
as it has dimensions of mass. Notice that now everything has an overall factor of i, so divide
by that factor and plug in all of the results so far into the expression for the polarization
tensor:
Πµν (k) =
Z 1
d/2 2 µν
µ ν µν 2 2 µν
−2 dx − 1 g I1 /i + −2x(1 − x)k k + g x(1 − x)k − m g I0 /i
0 d
Z 1 2
ε 2 −1 ε µν −1 2 µ 1
= −[4 − ln 2 + O(ε )] dx + g D + ln + + O(ε)
2 0 2 8 8π ε D 2
ε
− [4 − ln 2 + O(ε2 )]×
Z 12 2
µ ν µν 2 2 µν
1 2 µ
dx −2x(1 − x)k k + g x(1 − x)k − m g + ln + O(ε)
0 16π 2 ε D
1 h ε iZ 1
1 1
2
µ 1 1
2 µν
= + 2 1 − ln 2 + O(ε ) dx g D − − ln − + + O(ε)
2π 8 0 ε 2 D 4 4
Z 1
1 h ε 2
i µ ν µν 2 2 µν 2
− 1 − ln 2 + O(ε ) dx −2x(1 − x)k k + g x(1 − x)k − m g
4π 2 8 ε
Z0 1 2
1 h ε 2
i µ ν µν 2 2 µν
µ
− 2
1 − ln 2 + O(ε ) dx −2x(1 − x)k k + g x(1 − x)k − m g ln
4π 8 0 D
+ O(ε)
Πµν
ε−1 (k) =
Z 1
1 µν 1
2
dx −g D + 2x(1 − x)k µ k ν − g µν x(1 − x)k 2 + m2 g µν
2π 0 ε
Z 1
1 1
dx −g µν [x(1 − x)k 2 + m2 ] + 2x(1 − x)k µ k ν − g µν x(1 − x)k 2 + m2 g µν
= 2
2π 0 ε
Z 1
1 1
dx x(1 − x) k µ k ν − g µν k 2
= 2
π 0 ε
1 1
= 2 k µ k ν − g µν k 2
6π ε
The divergent piece is indeed transverse. What about the finite piece?
85
Πµν
ε0 (k) =
Z 1
ln 2 µν 1
2
−1
Z
1 µν µ
dx g D ln + g dx D
2π 2 0 2 D 16π 2 0
Z 1
ln 2
dx −2x(1 − x)k µ k ν + g µν x(1 − x)k 2 − m2 g µν
+ 2
16π 0
Z 1 2
1 µ ν µν 2 2 µν
µ
− 2 dx −2x(1 − x)k k + g x(1 − x)k − m g ln
4π 0 D
Z 1 2
1 1 µν 2 2
µ ν 1 µν 2 1 2 µν µ
= 2 dx − 2 g x(1−x)k + m + x(1−x)k k − 2 g x(1−x)k + m g ln
2π 0 2 D
ln 2 1 2 1 1
+ 2
g µν k + m2 − k µ k ν + g µν k 2 − m2 g µν
16π 6 3 6
Z 1 2
1 µ ν 2 µν
µ
= 2 dx x(1 − x) k k − k g ln
2π 0 D
ln 2 2 µν
k g − kµkν
+ 2
48π
1
µ2
Z
1 ln 2
= 2 k µ k ν − k 2 g µν
dx x(1 − x) ln +
2π 0 D 24
The finite piece is also transverse. We have shown that dimensional regularization preserves
the transverse structure of the polarization tensor and therefore preserves gauge invariance.
with D ≡ x(1 − x)k 2 + m2 , and with A being the counterterm. That is exactly the form of
equation (13) on page 187 from using Pauli-Villars regularization.
86
d3 q e i~q·~x /{~q 2 [1+
R
2. Study the modified Coulomb’s law as determined by the Fourier integral
e2 Π(~q 2 )]}.
Solution:
e2 e i~q·~r e2
Z
Vcoul (~r) = 2 d3 q = (as usual)
2π |~q |2 r
2 Z 1
e2 e i~q·~r m2
Z
3
δV (~r) = − dq dx x(1 − x) ln
2π 2 |~q |2 0 m2 + x(1 − x)|~q |2 − iε
We have used the on-shell renormalized result for Π(k 2 ) from the previous problem, setting
k 0 = 0 and ~k = ~q.
First notice that the log has a branch cut at |~q |2 /4 + m2 < 0, which causes the log to
have a nonzero imaginary piece, which means that with enough energy a virtual electron-
positron pair can become real. This can happen if the energy is greater than 2m, which is
precisely the branch point of the log. This is enough to show that there is a characteristic
length scale of 1/(2m) in the potential.
Now let us compute the integrals explicitly. First evaluate the angular part of the Fourier
integral:
Z q ·~
i~ r Z ∞ Z 1
3 e 2 2 1 2
dq f (|~q | ) = 2π dq q 2 f (q ) dη e iqrη
|~q |2 0 q −1
Z ∞
e iqr
− e−iqr
= 2π dq f (q 2 )
0 iqr
2 R
e2 1
Therefore we have δV = − 2π 2 0
dx x(1 − x) I, where
∞
e i~q·~r m2 e+iu − e−iu
Z Z
3 2πi 1
I≡ dq ln =− du ln
|~q |2 m2 + x(1 − x)|~q |2 − iε r 0 u 1 + βu2 − iε
with β ≡ x(1 − x)/(mr)2 . Now we will evaluate the integral over u by the method of contour
integration. Consider the integral
e iz
I
J≡ dz ln(1 + βz 2 )
C z
over the complex variable z, with the contour C = Γ2 + C3 + γ1 + C2 + γ2 + C4 + Γ1 + C1 as
shown in the picture:
87
Im z
C4 2
1 C3
2m
C2
C1
Re z
−R +R
1 2
As stated
√ previously and as shown in the diagram, the log has a branch cut starting at
z = i/ β extending along the imaginary axis to z → +i∞. Taking the radius R of the
contour to infinity and the radius of the small semicircle C1 to zero gives
Z ∞
e iu
Z
= du ln(1 + βu2 ) .
Γ1 +Γ2 −∞ u
R
The residue of the pole at C1 is ln 1 = 0, so C1 = 0.
Along γ2 , the log function can be written ln z = ln |z| + iθ, where θ = π/2 is the loca-
tion of the cut. The line γ1 is on the other side of the cut, so the angle picks up an extra
factor of 2π and hence on γ1 , we write ln z = ln |z| + i(θ + 2π).
88
and ∞ ∞
e+iu − e−iu e+iu − e−iu
Z Z
1
du f (u2 ) = du f (u2 )
0 u 2 −∞ u
so therefore: ∞ ∞
e iu − e−iu e−y
Z Z
du ln(1 + βu2 ) = 2πi √
dy .
−∞ u 1/ β y
Thus we find ∞ ∞
e−y 4π 2 e−y
Z Z
+2πi
I= 2πi √
dy =− √
dy .
r 1/ β y r 1/ β y
The 1-loop correction to the Coulomb potential is now written as two nested integrals over
the real parameters x and y:
Z 1 Z ∞
e4 e−y
δV (r) = + 2 dx x(1 − x) dy
π r 0 a(x) y
√ p
where a(x) ≡ 1/ β = mr/ x(1 − x) . In the short-distance limit, we can Taylor expand
around mr = 0 to get
Z ∞
e−y 1 mr
dy = − ln(mr) + ln[x(1 − x)] − γ + p + O(mr)2
a(x) y 2 x(1 − x)
R∞
where γ ≡ − 0
dx e−x ln x ≈ 0.577. Using the integrals
Z 1 Z 1
5 p π
dx x(1 − x) ln[x(1 − x)] = − , dx x(1 − x) =
0 18 0 8
R1
and 0
dx x(1 − x) = 16 , we have:
1 ∞
e−y
Z Z
1 5 3π 2
dx x(1 − x) dy =− ln(mr) + + γ − mr + O(mr)
0 a(x) y 6 6 4
e4
5 3π
δV (r) ≈ − 2 ln(mr) + + γ − mr .
6π r 6 4
89
III.8 Becoming Imaginary and Conserving Probability
1. Evaluate the imaginary part of the vacuum polarization function, and by explicit calcu-
lation verify that it is related to the decay rate of a vector particle into an electron and a
positron.
Solution:
We first evaluate the imaginary part of the vacuum polarization function by the method
leading to equation (8) on p. 211. The vacuum polarization function is
Z 1
e2 2
h z i
Π(z) = − 2 M dx x(1 − x) ln 1 − x(1 − x) 2
2π 0 m
where we have included a factor of e2 from using the Lagrangian L = 14 Fµν F µν instead of
L = 4e12 Fµν F µν , and we have included a factor of M 2 since we are shifting the pole from
k 2 = M 2 rather than from k 2 = 0.
The log goes imaginary when its argument is negative. Since ln(−1) = iπ, we find
Z 1
e2 2 σ
ImΠ(σ + iε) = − 2 M dx x(1 − x)(−π) θ[x(1 − x) 2 − 1]
2π m
2 Z 0x+
e
= + M2 dx x(1 − x)
2π x−
e2 2 2 2 3 x+ 1 2 2
M x+ − x2− − 23 x3+ − x3−
= M (x − 3 x ) x− =
4π 4π
p
where x± ≡ 12 1 ± 1 − 4m2 /σ are the two roots of the quadratic equation x2 −x+m2 /σ =
2 2
p
0 found by setting
p the argument of the step function to zero. We have x + −x − = 1 − 4m2 /σ
and x3+ − x3− = 1 − 4m2 /σ (1 − m2 /σ), and so
r
2m2
2 2 2 3 3 1 4m2
x+ − x− − 3 (x+ − x− ) = 1− 1+ .
3 σ σ
The decay rate is given on p. 212 as Γ = ImΠ(M 2 )/M , so we find
r
e2 2m2
4m2
Γ= M 1− 1+ .
12π M2 M2
Now we will calculate the decay rate of a vector particle into a fermion-antifermion pair using
Feynman diagrams. Consider the Lagrangian for quantum electrodynamics with a massive
photon:
6 ψ − mψ̄ψ + 12 M 2 Aµ Aµ − 14 Fµν F µν
L = ψ̄ i D
The Dirac field ψ represents the electron with mass m and electric charge −e, as spec-
ified by the covariant derivative Dµ ψ = ∂µ ψ − ie(−1)Aµ ψ = ∂µ ψ + ieAµ ψ. This im-
plies a vertex −ieγ µ and an amplitude iM(γ → e+ e− ) = ∗µ (p0 )ū2 [−ieγ µ ]v1 . (Notation:
90
0 = photon, 1 = outgoing positron, 2 = outgoing electron). The magnitude squared is
|M|2 = e2 ∗µ (p0 )ν (p0 )(ū2 γ µ v1 )(v̄1 γ ν u2 ) = e2 ∗µ (p0 )ν (p0 )tr[γ µ (v1 v̄1 )γ ν (u2 ū2 )].
P
Next
P we sum over the spins of the outgoing fermions using s uū = (6 p + m)/(2m) and
s vv̄ = (6 p − m)/(2m), and average over the three spin states of the incoming massive
photon using X pµ pν
∗(a)µ (p)(a)ν (p) = −ηµν + .
a
M2
The spin-summed amplitude squared M2 ≡ 13 a,s1 ,s2 |M|2 is
P
2 h
2 21
1 p0µ p0ν i
M = −e ηµν − tr [γ µ (6 p1 − m)γ ν (6 p2 + m)]
3 2m M2
e2
tr [γ µ (6 p1 − m)γµ (6 p2 + m)] − tr γ 0 (6 p1 − m)γ 0 (6 p2 + m)
=− 2
12m
where we have used pµ0 = (M, ~0) in the rest frame of the decaying photon.
we obtain
e2
M2 = 3p01 p02 − p~1 · p~2 + 3m2
3m 2
91
p
where p0i = |~pi |2 + m2 .
1 d3 p1 d3 p2
dΓ = (2π)4 δ 4 (p0 − p1 − p2 )M2
2M (2π)3 p01 /m (2π)3 p02 /m
where δ 4 (p0 − p1 − p2 ) = δ(M − p01 − p02 )δ 3 (~p1 + p~2 ) in the rest frame of the parent photon.
Integrating over the 3-dimensional delta function sets p~2 = −~p1 , for which the squared
amplitude becomes
e2
M2 = [3(p01 )2 + p~12 + 3m2 ]
3m2
e2
= [3(~p12 + m2 ) + p~12 + 3m2 ]
3m2
e2
= [4~p 2 + 6m2 ] .
3m2 1
The decay rate is
1 m2 d3 p1
Z q
Γ= δ(M − 2 p~12 + m2 ) M2 |p~2 =−~p1
2M (2π)2 p~12 + m2
1 1 e2 d3 p1
Z q
= 2 2 2
δ(M − 2 p~12 + m2 ) (4~p12 + 6m2 )
2M 4π 3 p~1 + m
2 Z
e p2 p
= dp 2 2
δ(M − 2 p2 + m2 ) (2p2 + 3m2 )
3πM p +m
e2 p2∗
M
Γ= 2 2
(2p2∗ + 3m2 )
3πM p∗ + m 4p∗
e2
p∗ 2 3 2
= p∗ + m
6π p2∗ + m2 2
2
2
e 4p∗ M 2 3 2
= −m + m
6π M 2 4 2
2 2
e 2m
= p∗ 1 +
6π M2
r
e2 2m2
4m2
= M 1− 1+ .
12π M2 M2
92
This agrees with the result obtained previously.
2. Suppose we add a gϕ3 term to our scalar ϕ4 theory. Show that to order g 4 there is
a “box diagram” contributing to meson scattering p1 + p2 → p3 + p4 with the amplitude
d4 k
Z
4 1
I=g
(2π) (k − m )[(k + p2 ) − m ][(k − p1 )2 − m2 ][(k + p2 − p3 )2 − m2 ]
4 2 2 2 2
(where m2 above is understood as m2 −iε. Note that this is a typo in the problem in the text.)
Calculate the integral explicitly as a funtion of s = (p1 +p2 )2 and t = (p3 −p2 )2 . Study the an-
alyticity property of I as a function of s for fixed t. Evaluate the discontinuity of I across the
cut and verify Cutkosky’s cutting rule. Check that the optical theorem works. What about
the analyticity property of I as a function of t for fixed s? And as a function of u = (p3 −p1 )2 ?
Solution:
We add the term L = − 3!1 gϕ3 to the Lagrangian, which generates the cubic vertex −ig.
The diagram in question is displayed below:
p1 p4
k−p 1
k k + p 2 −p 3
k+p2
p2 p3
Each of the external momenta is on-shell, meaning p2i = m2 for i = 1, 2, 3, 4. Using the scalar
propagator
i
i∆(p) = 2
p − (m2 − iε)
we obtain the amplitude
d4 k
Z
4
I = (−ig) [i∆(k)][i∆(k + p2 )][i∆(k − p1 )][i∆(k + p2 − p3 )]
(2π)4
d4 k
Z
4 1 1 1 1
=g
(2π) [k − m ] [(k + p2 ) − m ] [(k − p1 ) − m ] [(k + p2 − p3 )2 − m2 ]
4 2 2 2 2 2 2
93
Note that the integral has four powers ofR k in the numeratorR and 4 × 2 = 8 powers of k in
the denominator, so the integral goes as dk k 3 /[(k 2 )4 ] ∼ dk/k 5 , which converges at high
energy and thus does not need to be regularized.
Now that we have the integral, we should be clear as to what exactly this problem is asking.
The steps are as follows:
First compute the integral using Feynman parameters, and evaluate the discontinuity of
I across the cut at s = 4m2 . Second, “verify Cutkosky’s cutting rule” by returning to the
original integral, replacing the propagators with on-shell delta functions with positive fre-
quency, evaluating the resulting integral and showing that we get the same result for the
imaginary part as we got before.
Third, to “check the optical theorem,” we are to compute the tree-level scattering ampli-
tude obtained by cutting the box diagram down the middle, and plug it into the right-hand
side of equation (19) on p. 215, which we repeat below for convenience:
!
X Y 1
ImF(i → f ) = 21 (2π)4 δ (4) (Pni ) 2
F(i → n) [F(f → n)]∗
n n
ρn
where in our case F(i → f ) = −iI. By evaluating the right-hand side explicitly, we are
to observe that it equals the left-hand side, namely the imaginary part of the original box
diagram that we computed previously in two different ways.
Finally, all results as a function of s for fixed t should be the same as those obtained as
a function of t for fixed s.
To evaluate the integral directly, one introduces Feynman parameters and integrates over
the loop momentum to put the integral in the form
We will now verify Cutkosky’s cutting rule and the optical theorem.
We compute the imaginary part of the integral by replacing the cut internal propagators
with momentum-conserving delta functions and thereby verify Cutkosky’s cutting rule, as
described on p. 215. For convenience, we collect the results from the example on pp. 216-217.
94
Given the amplitude
d4 k
Z
2
Π = +ig [i∆(µ) (k)][i∆(m) (q − k)]
(2π)4
where the subscript on the scalar propagator denotes the location of the pole in momentum-
squared (that is, the mass), Cutkosky tells us that twice the imaginary part of the amplitude
is given by replacing the propagators i∆ with on-shell delta functions with positive frequency:
d4 k 0
Z
1 2
2
θ(k )2πδ(k 2 − µ2 ) θ (q − k)0 2πδ (q − k)2 − m2
Im Π = 2 g 4
.
(2π)
As explained on p. 215, in our case we identify I with iΠ. We first shift the loop momentum
as k → k + p1 to obtain
d4 k
Z
4 1
I=g
(2π) [k − m ][(k + p1 ) − m ][(k + p4 )2 − m2 ][(k + p1 + p2 )2 − m2 ]
4 2 2 2 2
where we have used p1 + p2 − p3 = p4 . From this point on, we follow5 P. van Nieuwenhuizen,
“Muon-Electron Scattering Cross Section to Order α3 ,” Nucl. Phys. B28 (1971) 429-454.
We will cut the diagram through the propagators labeled by momenta k and k + p1 + p2 .
Cutkosky tells us that this entails replacing i∆(k) → θ(k 0 )2πδ(k 2 −m2 ) and i∆(k+p1 +p2 ) →
θ[−(k 0 + p01 + p02 )]2πδ[(k + p1 + p2 )2 − m2 ]. Note the minus sign in the second theta function;
we are supposed to ensure that the momentum flows in one direction through the loop when
we cut it. The imaginary part of the diagram I ≡ Im(−iI) is
1 g4 0 2 2 0 0 0 2
4 θ(k )δ(k − m ) θ[−(k + p1 + p2 )]δ[(k + p1 + p2 ) − m ]
2
Z
I= d k .
2 (2π)2 [(k + p1 )2 − m2 ][(k + p4 )2 − m2 ]
√ √
g4 d3 k
Z
θ[−(ωk + s )]δ[s + 2ωk s ]
= √ √
2(2π)2 2ωk [ m2 + ωk s − 2~k · p~1 ][ m2 + ωk s − 2~k · p~4 ]
where we have defined s ≡ (p1 + p2 )2 and have chosen to work in the center-of-mass frame:
µ 1√ µ 1√
p1 = ( 2 s , +~p ), p2 = ( 2 s , −~p ) =⇒ p~1 + p~2 = 0.
5
The reference uses the notation qm = p1 , qe = p2 , qe0 = p3 , qm = p4 , ∆ = k and P = p1 + p2 in Eq.
(3.9). We treat the case for which all internal masses are equal, so M = λ = m. Also their metric is mostly
positive, whereas ours is mostly negative.
95
Now let us define coordinates carefully in order to perform the integral. First fix the di-
rection of the incoming momentum p1 to point along the ẑ-axis:
√
p~1 = |~p1 |ẑ , |~p1 | = 12 s − 4m2 .
Then fix the scattering plane as the (x, z)-plane, and define the scattering angle θ0 in the
center-of-mass frame as follows:
√
p~4 = |~p4 | (x̂ sin θ0 + ẑ cos θ0 ) , |~p4 | = |~p1 | = 21 s − 4m2 .
We will perform the integral in spherical coordinates, so that the 3-vector to be integrated
over is
~k = |~k | (x̂ sin θ cos ϕ + ŷ sin θ sin ϕ + ẑ cos θ) .
In these coordinates, we have
~k · p~1 = |~k ||~p1 | cos θ
~k · p~4 = |~k ||~p4 | (sin θ0 sin θ cos ϕ + cos θ0 cos θ) .
√ √
The delta function sets ω = − 12 s , so that k = 21 s − 4m2 = p, and therefore kp =
1
4
(s − 4m2 ). The integral is
r
g4 4m2
I= 1 −
8(2π)2 s
Z
1
× dΩ .
[s − 2m2 + (s − 4m2 ) cos θ][s − 2m2 + (s − 4m2 )(sin θ0 sin θ cos ϕ + cos θ0 cos θ)]
The integral over angles is in the standard form6
√ !
aA − bB + X
Z
1 2π
dΩ = √ ln √
(a + b cos θ)(A + B cos θ + C sin θ cos ϕ) X aA − bB − X
6
See Appendix A of W. Beenakker and A. Denner, “Infrared Divergent Scalar Box Integrals with Appli-
cations in the Electroweak Standard Model,” Nucl. Phys. B338 (1990) 349-370.
96
where
a = s − 2m2
b = s − 4m2
A = s − 2m2 = a
B = (s − 4m2 ) cos θ0 = b cos θ0
C = (s − 4m2 ) sin θ0 = b sin θ0
g4
1 1 − Q(s, t, m)
I= ln
4π P (s, t, m) 1 + Q(s, t, m)
where
s
4m2 2
3m2
P (s, t, m) ≡ t 1− s − 4m 1 −
t t
s
4m2 s − 3m2
Q(s, t, m) ≡ 1− .
t s − 4m2
Finally, let us verify the optical theorem. If we cut the box diagram vertically, we arrive at
the amplitudes ML and MR for the left and right halves respectively:
p1 k1 k1 p4
ML = k = p1+ k1 MR = k’ = k1 + p 4
p2 k2 k2 p3
We have
g2 g2
ML = −i , M R = −i .
(k1 + p1 )2 − m2 (k1 + p4 )2 − m2
We can also cut the diagram horizontally. After suitably shifting the internal momenta, this
second cut will contribute an identical term to the imaginary part of the amplitude.
97
P
Define F ≡ −iM as on p. 215. The sum n over intermediate states becomes
Z 3 Z 3
X d k1 d k2
→ 3
n
(2π) (2π)3
where the factor of accounts for the other possible cut, as discussed. In the center-of-mass
frame, we have p~1 + p~2 = 0 and therefore
√
δ 4 (k1 + k2 − p1 − p2 ) = δ(ω1 + ω2 − s ) δ 3 (~k1 + ~k2 ) .
πg 4 √
Z 3
dk 1 1
I= 3 2
δ(2ωk − s )
4(2π) ωk [(k + p1 ) − m ] [(k + p4 )2 − m2 ]
2 2
q √
where ωk ≡ |~k |2 + m2 . As before, we can write d3 k = dΩ dω ω ω 2 − m2 and integrate
√ √
over the remaining delta function using δ(2ω − s ) = 12 δ(ω − 12 s ). We have:
√
g4 ω 2 − m2
Z
1 1
I= dΩ
8(2π)2 ω [(k + p1 ) − m ] [(k + p4 )2 − m2 ]
2 2
r
g4 4m2
Z
1 1
= 2
1− dΩ .
8(2π) s [(k + p1 ) − m ] [(k + p4 )2 − m2 ]
2 2
This is exactly the expression we got when applying the Cutkosky cutting rule.
3. Prove (28). R R
4
[Hint: Do unto dP x e iqx h0|[O(x), O(0)]|0i what we did to d4 x e iqx h0|T (O(x)O(0))|0i,
namely insert 1 = n |nihn| (with |ni a complete set of states) between O(x) and O(0)
in the commutator. Now we don’t have to bother with representing the step function.]
98
Solution:
Z Z
d x e h0|[O(x), O(0)]|0i = d4 x e iq·x (h0|O(x)O(0)|0i − h0|O(0)O(x)|0i)
4 iq·x
Z
= d4 x e iq·x h0|e+iP ·x O(0)e−iP ·x O(0)|0i − h0|O(0)e+iP ·x O(0)e−iP ·x |0i
Z
= d4 x e iq·x h0|O(0)e−iP ·x O(0)|0i − h0|O(0)e+iP ·x O(0)|0i
Z ! ! !
X X
= d4 x e iq·x h0|O(0) |nihn| e−iP ·x O(0)|0i − h0|O(0) |nihn| e+iP ·x O(0)|0i
n n
Z !
X X
= d4 x e iq·x h0|O(0)|nihn|O(0)|0i e−iPn ·x − h0|O(0)|nihn|O(0)|0i e+iPn ·x
n n
X Z Z
2 4 i(q−Pn )·x 4 i(q+Pn )·x
= |h0|O(0)|ni| d xe − d xe
n
X
= (2π)4 |O0n |2 δ (4) (q − Pn ) − δ (4) (q + Pn )
(†)
n
Thus for q satisfying the delta function δ (4) (q+Pn ), equation (28) also follows. This concludes
the problem.
7
Explicitly, use O(x) = e iP ·x O(0)e−iP ·x and Pµ |0i = 0 to get h0|[O(x), O(0)]|0i = h0|O(x)O(0)|0i −
h0|O(0)O(x)|0i = h0|O(0)e−iP ·x O(0)|0i − h0|O(0)e+iP ·x O(0)|0i = h0|O(0)O(−x)|0i − h0|O(−x)O(0)|0i =
h0|[O(0), O(−x)]|0i.
99
IV Symmetry and Symmetry Breaking
IV.1 Symmetry Breaking
2. Construct the analog of (2) with N complex scalar fields and invariant under SU (N ).
Count the number of Nambu-Goldstone bosons when one of the scalar fields acquires a
vacuum expectation value.
1 1
L= ~ )2 − m2 ϕ
(∂ ϕ ~ 2 − λ(~
ϕ 2 )2 (2)
2 4
Solution:
Let ϕ be a complex scalar field that transforms under the N -dimensional representation
of SU (N ). Its Lagrangian is
λ
L = ∂µ ϕ†a ∂ µ ϕa + µ2 ϕ†a ϕa − (ϕ†a ϕa )2
2
The index a runs from 1 to N . Note that this Lagrangian has U (N ) symmetry, rather than
just SU (N ) symmetry. Actually, this Lagrangian secretly has an even larger symmetry group
that is at this level opaque because of our choice of field coordinates.
Decompose each field into its real and imaginary parts: ϕa = √1 (χa + iη a ). The U (N )-
2
invariant scalar product is
1 1
ϕ†a ϕa = (χa − iηa )(χa + iη a ) = (χa χa + ηa η a )
2 2
Repackage the N χs and the N ηs into a 2N -dimensional column vector:
χ
~
φA ≡
~η
Since the U (N )-invariant Lagrangian above depends only on this scalar product, it is actually
invariant under the symmetry group O(2N ) (for more on this point, see p. 407):
2N 2N
!2
X 1 λ X
L= ∂ µ φA ∂ µ φA + µ 2 φA φA − φA φA
A=1
2 8 A=1
The group O(2N ) has 2N (2N − 1)/2 = N (2N − 1) generators, as compared with the N 2
generators of U (N ). Now the problem is identical to IV.1.1, in which you are asked to show
that O(N ) breaks to O(N − 1) and yields N − 1 Nambu-Goldstone bosons. Before computing
the Lagrangian explicitly, we first count the number of Nambu-Goldstone bosons from group
theory, as described in the section “Counting Nambu-Goldstone bosons” on p. 199.
We start with the group O(2N ), which has 2N (2N −1)/2 = N (2N −1) generators. We break
100
it to O(2N − 1), which has (2N − 1)(2N − 2)/2 = (2N − 1)(N − 1) generators. Therefore,
the number of Nambu-Goldstone bosons is N (2N − 1) − (N − 1)(2N − 1) = 2N − 1.
Now let us compute the Lagrangian explicitly. It is more straightforward to work with
the explicitly O(2N )-invariant theory, repeated below
1 λ
L = (∂µ φA ∂ µ φA + µ2 φA φA ) − (φA φA )2
2 8
We have written indices up and down just for convenience; since O(N ) has the invariant
tensor δAB , up and down make no difference. (Contrast with SU (N ), which does not have
the invariant tensor δAB but only the invariant tensors εA1 ...AN , εA1 ...AN and δ AB .)
Use the O(2N ) freedom to rotate the vacuum expectation value (VEV) of φ into the 2N th
component: hφA i = vδA,2N . For clarity, let us define a new index i = 1, ..., 2N − 1 to single
out the component with nonzero VEV. Let us also denote the shifted value of the last field
by h(x), so that we write:
φA = 2N = v + h(x) , φA6=2N = φi
1 µ2 λ 2
L = (∂µ h∂ µ h + ∂µ φi ∂ µ φi ) + (v + h)2 + φi φi − (v + h)2 + φi φi
2 2 8
We have:
(v + h)2 = v 2 + 2vh + h2
We display this explicitly because the numerical factors are critical to get the right answer.
The Lagrangian is:
101
v2 λv 2
2 1
L= µ − + (∂µ h∂ µ h + ∂µ φi ∂ µ φi )
2 4 2
2 2 2
µ λ 3 µ λ 2 2 µ λ 2
+ 2v − (4v ) h + − (6v ) h + − (2v ) φi φi
2 8 2 8 2 8
λ
4vh3 + h4 + 4vhφi φi + 2h2 φi φi + (φi φi )2
−
8
v2 λv 2
2 1
= µ − + (∂µ h∂ µ h + ∂µ φi ∂ µ φi )
2 4 2
2
1 2 3λv 2 2 1 2 λv 2
2 λv
+v µ − h+ µ − h + µ − φi φi
2 2 2 2 2
λv 3 λv λ λ λ
− h − hφi φi − h2 φi φi − h4 − (φi φi )2
2 2 4 8 8
You see now what is going to happen: minimizing with respect to h at the point h = 0 and
φi = 0, along with v 6= 0, yields the condition µ2 = λv 2 /2. (Alternatively, you can minimize
the Lagrangian with respect to v, set h = φi = 0, then solve for v.) This also sets the
coefficient of the term quadratic in φi to zero, resulting in 2N − 1 massless Nambu-Goldstone
bosons. As you can see, the resulting theory has O(2N − 1) symmetry, along with cubic
interaction terms, indicating that an h-particle can decay into two Nambu-Goldstone bosons.
Solution:
102
Thus the regularized 1-loop effective potential is
2
1 2 2 1 4 1 1 2 Λ 2
Veff (ϕ) = (m + B)ϕ + λϕ + λϕ ln −m S
2 4! 8π 2 m2
where we have defined the sum
∞ 2 n
X (−1)n λϕ λϕ2
S≡ = (1 + x) ln(1 + x) − x , x ≡ .
n=2
n(n − 1) 2m2 2m2
We now require renormalization conditions. We would like to follow p. 241 in the text and
00
impose “on-shell” renormalization conditions, for which Veff (0) = m2 . This fixes
2
λ Λ
B=− ln .
8π m2
Putting this B back into the effective potential, we find
2
1 2 2 1 4 1 2 2 2 λϕ
Veff (ϕ) = m ϕ + λϕ + λϕ − (λϕ + 2m ) ln +1 .
2 4! 16π 2m2
Taking m2 arbitrarily small but nonzero, we have
λϕ2
1 1
Veff (ϕ) = λϕ4 − λϕ2 ln .
4! 16π 2 2m2
2
1 2 2m
For ϕ → 0, the potential behaves as Veff (ϕ) → + 16π 2 λϕ ln λϕ2
> 0. The Z2 : ϕ → −ϕ
symmetry of the classical potential is not spontaneously broken.8
Now consider the case for which we start with a massless theory, for which the 1-loop effective
potential is
∞ n
d2 k X 1 λϕ2
Z
1 4 1 2 i
Veff (ϕ) = λϕ + Bϕ +
4! 2 2 (2π)2 n = 1 n 2k 2
Z 2
λϕ2
1 4 1 2 1 d kE
= λϕ + Bϕ + ln 1 + 2
4! 2 2 (2π)2 2kE
where in the second line we have summed the series and rotated to Euclidean momentum.
Regularizing the integral with an arbitrarily large UV cutoff Λ, we have
Z 2
λϕ2
2
d kE 1 2 2Λ
2
ln 1 + 2 = λϕ ln
(2π) 2kE 8π λϕ2
where we have dropped terms that go to zero for Λ → ∞. The regularized effective potential
is 2
1 4 1 2 1 2 2Λ
Veff (ϕ) = λϕ + Bϕ + λϕ ln .
4! 2 16π λϕ2
8
Actually the ϕ → −ϕ symmetry is not broken in (3 + 1) dimensions either. See the addendum at the
end of this section.
103
The second derivative of this has a log divergence at ϕ = 0. Instead, define the curvature at
00
an arbitrary field point ϕ = µ as Veff (µ) ≡ m2 (µ). This fixes
2
2 λ 2Λ 2
B = m (µ) − ln + 4πµ − 3 .
8π λϕ2
The idea is to approach the origin ϕ = 0 carefully by keeping ξ small but nonzero, taking
ϕ → 0 and then taking ξ → 0.
9
Note also that, as in d = (3 + 1) dimensions, all logs of λ have disappeared after renormalization. In
Appendix A.2 of the Coleman-Weinberg paper, it is shown that by rescaling the loop momenta k → λ1/2 k,
d
and therefore dd k → λd/2 dd k, the contribution to the effective potential is of the form ϕ4 f (ϕ/M )λV + 2 L−I ,
where V is the number of vertices, L is the number of loops, and I is the number of internal lines. For d = 4,
the contribution goes as λV +2L−I = λL+1 , and thus the 1-loop correction goes as λ2 , as shown in the text.
For the present case of d = 2, the contribution goes as λV +L−I = λ, which is independent of the number of
loops L. This explains our result that the 1-loop correction to massless ϕ4 theory in d = (1 + 1) dimensions
goes as λ.
104
00
We are interested in “massless” ϕ4 theory, so we want m2 (µ → 0) → 0. (i.e., Veff (0) = 0.)
From the RG equation, we have
1
m2 (ϕ) = m2 (ξ 1/2 ϕ) + λ[ 12 (1 − ξ)ϕ2 + ln ξ] .
8π
λ
Consider ξ > 0, but ϕ → 0. Then m2 (ϕ) ≈ m2 (ξ 1/2 ϕ) + 8π
ln ξ, which implies m2eff (ξ) ≈
3 3
m2 (ϕ) + 8π λ. If m2 (ϕ → 0) → 0, then m2eff (ξ) → 8π λ > 0.
1
So near the origin ϕ → 0, taking 0 < ξ 12 in massless ϕ4 theory implies that Veff (ϕ) ≈
1
4!
λ ϕ4 . The Z2 symmetry is not spontaneously broken.
Solution:
In the Landau gauge (see p. 267 with ξ = 0), computing the photon contribution to the
effective potential constitutes summing up the one-loop diagrams depicted below:
+ + + + ...
The external lines are scalar lines with no external momenta. The diagrams (from left to
right) contain n = 1, 2, 3, 4, ... powers of the photon propagator. The reason for choosing
Landau gauge is that otherwise there would be diagrams with both photons and scalars in
the internal lines.
105
where we have recognized the scalar propagator
i
i∆(k) =
k2 − M (ϕ)2
and we have defined a projection operator
kµ kν
Pµν (k) ≡ ηµν − .
k2
This is a projection operator in the sense that Pµν (k)P νρ (k) = Pµ ρ (k), so that tr(P n ) =
trP = 3 for any positive integer n.
Because of this, and since every diagram in the above one-loop expansion contains an equal
number of e2 ϕ† ϕAµ Aµ → 2ie2 η µν vertices and photon propagators i∆µν (k) = −i∆(k)Pµν (k),
we can simply replace the series by an equivalent series with internal scalar lines instead of
photon lines, as long as we include a minus sign in the vertex and multiply the sum by an
overall factor of 3. In terms of the scalar coupling λ, the replacement is λ → 2e2 .
In other words, we can immediately jump to equation (14) in the text with the replace-
ment V 00 (ϕ) = M (ϕ) and an overall factor of 3 multiplying the integral.
The photon contribution is equal to (+3) real scalar contributions, which makes sense given
the 3 polarization states of a massive vector boson (the photon obtains an effective mass in
the ϕ background).
Now compare this to the contribution of a fermion (equation (26) on p. 243, with slight
notational adjustments):
d4 k
2
k − M (ϕ)2
Z
1
Veff (ϕ) = V (ϕ) + 4 × 2 i ln .
(2π)4 k2
The Dirac fermion contribution is equal to (−4) real scalar contributions to the effective
potential.
On p. 242 of the text, it is shown that the one-loop correction +ϕ4 ln ϕ2 overwhelms the
classical +ϕ4 potential near ϕ = 0. However, the conclusion to draw from this is not that
quantum fluctuations break the discrete symmetry ϕ → −ϕ but rather that perturbation
106
theory is not valid near ϕ = 0. The issue is not whether ~ is small (indeed we have set it to
1), but rather, as explained on p. 242, that the expansion parameter is λ ln ϕ rather than just
λ. For small ϕ, the second term in the expansion becomes larger than the first term, and the
third term will be larger than second term, and so forth, which means that the perturbation
expansion breaks down near ϕ = 0 and cannot tell us whether ϕ = 0 is a maximum or a
minimum of the potential.
We now use the renormalization group to obtain a perturbative expansion that is valid near
ϕ = 0 and show that the ϕ → −ϕ symmetry is not spontaneously broken by one-loop effects.10
The one-loop beta function for the quartic coupling λ is (equation (19) on p. 242)
dλ(M ) 3
M = 2
[λ(M )]2 + O[λ(M )]3 .
dM 16π
The solution to this equation, integrated from some arbitrary scale M0 , is
λ(M0 )
λ(M ) = .
3 M2
1− 32π 2
λ(M0 ) ln M02
The question now is whether we can trust perturbation theory near ϕc = 0. Rearranging the
solution of the flow equation for λ, we find
2
3 M
λ(M ) = λ(ϕc ) 1 + 2
λ(ϕc ) ln .
32π ϕ2c
10
This argument follows section V of S. Coleman and E. Weinberg, “Radiative Corrections as the Origin
of Spontaneous Symmetry Breaking,” Phys. Rev. D, Vol. 7 No. 6, 15 Mar 1973 and Section 18.2 in S.
Weinberg, Volume II.
107
We are interested in taking ϕc → 0, so that we may consider M close to but larger than ϕc
and thereby take ln(M 2 /ϕ2c ) > 0. For perturbation theory to be valid, we need
2
3 M
2
λ(ϕc ) ln 1
32π ϕ2c
to keep the second term in the expansion smaller than the first term.
Assuming that this is true by taking M sufficiently close to ϕc (so that the log is not large)
and by taking λ(ϕc ) sufficiently small, then the solution to the flow equation implies that
λ(M ) and λ(ϕc ) have the same sign. For vacuum stability, we take λ(M ) positive. Therefore,
the effective potential is
1
Veff (ϕc ) = λ̃(ϕc ) ϕ4c
4!
with
25
λ̃(ϕc ) ≡ 1 − λ(ϕc ) λ(ϕc ) > 0 .
64π 2
The minimum of the effective potential is at ϕc = 0, and the symmetry ϕc → −ϕc of the
classical potential is not broken.
For a scalar theory in which a classical symmetry truly is broken by one-loop effects, see
problems IV.3.5 and IV.6.9.
108
Addendum 2: Effective Potential using Dimensional Regularization
It is pedagogically instructive to repeat the calculation of the effective potential using di-
mensional regularization, in contrast to the cutoff regularization used in the text and in the
original paper by Coleman and Weinberg.
Lagrangian:
1 1 1
L = Z(∂ϕ)2 − Zm m2 ϕ2 − Zλ λµ̃ε ϕ4
2 2 4!
1-loop effective potential:
∞ 1 ε 2 n
dd k X 1 λµ̃ ϕc
Z
1 1
V (ϕc ) = Zm m2 ϕ2c + Zλ λµ̃ε ϕ4c + i 2
2 4! (2π)d n = 1 2n k 2 − m2 + iε
d
Specialize to d = 4 − ε =⇒ 2
= 2 − 2ε , 1 − d
2
= −1 + 2ε , 2 − d
2
= + 2ε .
Only the n = 1 and n = 2 integrals diverge, and their divergences will be absorbed in
Zm and Zλ .
109
• n = 1 (note k 2 = −kE2 so an overall (−1) gets factored out of the integral):
dd k
Z Z d
1 ε 2 1 1 ε 2 d kE 1
λµ̃ ϕc d 2 2
= λµ̃ ϕc (−i) 2
2 (2π) k − m + iε 2 (2π) kE + m2
d
1 ε 2 Γ(1 − d2 ) 2 −(1− d )
= λµ̃ ϕc (−i) (m ) 2
2 (4π)d/2
Γ(−1 + 2ε ) 4π ε/2 2
1 ε 2
= λµ̃ ϕc (−i) m
2 (4π)2 m2
ε/2
m2 ε 4π µ̃2
1 2
= λϕc (−i) Γ(−1 + )
2 (4π)2 2 m2
m2 4π µ̃2
1 2 2 ε
= λϕc (−i) (−1)[ − γ + 1 + O(ε)][1 + ln + O(ε2 )]
2 (4π)2 ε 2 m2
m2 2 4πe−γ µ̃2
1 2
= λϕc (+i) [ + ln + 1 + O(ε)]
2 (4π)2 ε m2
So in total we have
1 ε 2
dd k 1 m2 4πe−γ µ̃2
λµ̃ ϕc
Z
2 2 2
i =− λϕ + ln +1 .
(2π)d 2 k 2 − m2 + iε 4(4π)2 c ε m2
• n = 2 (since the denominator is squared this time, the (−1) doesn’t matter):
dd k
Z Z d
1 ε 2 2 1 1 ε 2 2 d kE 1
( λµ̃ ϕc ) d 2 2 2
= ( λµ̃ ϕc ) (+i) 2
2 (2π) (k − m + iε) 2 (2π) (kE + m2 )2
d
1 Γ(2 − d2 ) 2 −(2− d )
= ( λµ̃ε ϕ2c )2 (+i) (m ) 2
2 (4π)d/2
Γ( 2ε ) 4π ε/2
1 ε 2 2
= ( λµ̃ ϕc ) (+i)
2 (4π)2 m2
ε/2
ε 4π µ̃2
ε 1 2 2 1
= µ̃ ( λϕc ) (+i) Γ( )
2 (4π)2 2 m2
4π µ̃2
ε 1 2 2 1 2 ε
= µ̃ ( λϕc ) (+i) [ − γ + O(ε)][1 + ln + O(ε2 )]
2 (4π)2 ε 2 m2
4π e−γ µ̃2
ε 1 2 2 1 2
= µ̃ ( λϕc ) (+i) [ + ln + O(ε)]
2 (4π)2 ε m2
So in total we have
1 ε 2 2
dd k 1 µ̃ε 4π e−γ µ̃2
λµ̃ ϕc
Z
2 2 4 2
i =− λ ϕc + ln .
(2π)d 2 · 2 k 2 − m2 + iε 16(4π)2 ε m2
110
The 1-loop effective potential is (define µ2 ≡ 4π e−γ µ̃2 ):
V (ϕc ) =
1 ε 2 1 ε 2 2
dd k 1 λµ̃ ϕc dd k 1 λµ̃ ϕc
Z Z
1 2 2 1 ε 4 2 2
Zm m ϕc + Zλ λµ̃ ϕc + i d 2 2
+i
2 4! (2π) 2 k − m + iε (2π) 2 · 2 k − m2 + iε
d 2
∞ 1 ε 2 n
dd k X 1 λµ̃ ϕc
Z
2
+i
(2π) n = 3 2n k − m2 + iε
d 2
2 2
1 λ 2 µ 2 2 1 λ 2 µ
= Zm − 2
+ ln 2
+ 1 m ϕc + Zλ − 2
+ ln 2
λµ̃ε ϕ4c
2 64π ε m 4! 256π ε m
d ∞ 1 ε 2 n
λµ̃ ϕc
Z
d k X 1 2
+i
(2π)d n = 3 2n k 2 − m2 + iε
At this point we are free to take µ̃ε → 1 everywhere. The regularized 1-loop effective potential
is:
Vreg (ϕc ) =
2 2
1 λ 2 µ 2 2 1 λ 2 µ
Zm − 2
+ ln 2
+ 1 m ϕc + Zλ − 2
+ ln 2
λϕ4c
2 64π ε m 4! 256π ε m
Z 4 ∞ n
1 2 n
d kE X (−1) 2
λϕc
− 4 2
(2π) n = 3 2n kE + m2
We have Wick rotated the integral to Euclidean momentum and set d = 4 in all of the
convergent integrals. We can now perform the integrals and then resum the series. For each
n ≥ 3, the integral is:
Z 4
d kE 1 1 Γ(n − 2) 2 −(n−2) m4 1 1
4 2 2 n
= 2
(m ) =
(2π) (kE + m ) (4π) Γ(n) 16π (n − 1)(n − 2) (m2 )n
2
111
Therefore, the dimensionally regularized and resummed 1-loop effective potential is (now
dropping the subscript on the field):
Vreg (ϕ) =
2 2
1 λ 2 µ 2 2 1 λ 2 µ
Zm − 2
+ ln 2
+1 m ϕ + Zλ − 2
+ ln 2
λϕ4
2 64π ε m 4! 256π ε m
" #
2
m4 λϕ2 3λϕ2 λϕ2 λϕ2
− + 2 − 2 1 + ln 1 +
128π 2 2m2 2m2 2m2 2m2
Now we are ready to pick renormalization conditions. We choose the “on-shell” renormaliza-
tion scheme defined by:
d2 Vreg (ϕ) d4 Vreg (ϕ)
≡ m2 , ≡ λ(µ)
dϕ2 ϕ=0 dϕ4 ϕ=µ
This is the same as the scheme chosen in Coleman and Weinberg’s paper and in the text.
Imposing these renormalization conditions fixes:
µ2
λ 2
Zm = 1 + + 1 + ln 2
64π 2 ε m
2
2
µ2
λ 2 µ λµ 2
Zλ = 1 + + ln 2 − 24 ln + 1 − 48 2 λ + O(λ ) .
256π 2 ε m 2m2 m
Here we are dropping O(λ3 ) and higher terms, but not Taylor expanding the logs. The
reason is that in the m → 0 limit, the λ 1 in the numerator is offset by m → 0 in the
denominator, so that the log is not well approximated by any finite order in its Taylor series.
Putting these back into Vreg (ϕ) gives the renormalized effective potential:
2
1 3 2 λϕ
Vren (ϕ) = { m ln + 1 + 12ϕ m22
24 8π 2 2m2
2
3 λϕ 1 2 2 4
+ ln +1 − m ϕ +ϕ λ
8π 2 2m2 2
2
λϕ + 2m2
3 4 3 2
+ 2
ϕ ln 2 2
− λ + O(λ3 )} .
32π λµ + 2m 2
In the limit m → 0, this becomes:
λ(µ)2 4 ϕ2 25
1 4
Vren (ϕ) = λ(µ)ϕ + ϕ ln 2 −
24 256π 2 µ 6
which matches equation (20) on p. 242. This is a useful check on the arithmetic: although
we have regularized the integrals differently, if we choose the same renormalization scheme
then we must get the same answer.
112
IV.4 Magnetic Monopole
2. Show by writing out the components explicitly that dF = 0 expresses something that you
are familiar with but disguised in a compact notation.
Solution:
Choose µ = i, ρ = j, ν = 0 to get
it is possible to repeat this exercise with d(∗ F ) = 0 as follows. The electromagnetic action
on a D-dimensional spacetime M is:
Z
1 ∗ ∗
S= − F F +A j
M 4
A is the 1-form potential, F = dA is its 2-form field strength, and j is a 1-form current that
couples to the potential A. If j is a 1-form, then ∗ j is a (D − 1)-form, so that the term A ∗ j
is a D-form and thus can be integrated over the space M . Varying the action with respect to
A as δS ≡ S[A + δA] − S[A] = 0 implies d(∗ F ) = ∗ j. So d(∗ F ) = 0 should be the source-free
Maxwell equations ∇~ ·E ~ = 0 and ∇ ~ ×B ~ − ∂t E
~ = 0.
∗ 1
F = εµνρσ F ρσ dxµ dxν
2
113
Now take the derivative:
1
d(∗ F ) = εµνρσ ∂λ F ρσ dxλ dxµ dxν
2
∗
The operation is defined such that operating twice gives you back the original thing (hence
the name “dual”), so the equation d(∗ F ) = ∗ j can be written ∗ d(∗ F ) = j, which is more
convenient. The left-hand side is:
∗ ∗ 1 1 µνρσ λ
d( F ) = εαµνλ ε ∂ Fρσ dxα
2 2
Using
εµναλ εµνρσ = 2(δαρ δλσ − δλρ δασ )
we get:
∗ 1
d(∗ F ) = (∂ σ Fασ − ∂ ρ Fρα )dxα = ∂ µ Fαµ dxα
2
∗ ∗
So the equation d( F ) = j in components becomes just
∂µ F αµ = j α
Writing the 4-current as j α = (ρ, J~ ), setting α = 0 and defining again E i ≡ F 0i in the above
gives ∂i F 0i = j 0 =⇒ ∇~ ·E ~ = ρ. (In verifying this remember F 00 = 0 by antisymmetry.)
Now instead of α = 0, set α = i to get:
∂µ F iµ = ∂0 F i0 + ∂j F ij = j i =⇒ −∂t E i + ∂j F ij = J i
Solution:
114
The coordinate transformation from spherical to cartesian coordinates is
so
z z y
cos θ = =p , tan ϕ =
r x2 + y 2 + z 2 x
dz z 1 z
d cos θ = − 2 dr = (dz − dr)
r r r r
p 1
dr = d x2 + y 2 + z 2 = p (x dx + y dy + z dz)
x2 + y 2 + z 2
z2
1 z x dx + y dy + z dz 1 z
d cos θ = dz − = 1 − 2 dz − 2 (x dx + y dy)
r r r r r r
1
dϕ = (−y dx + x dy)
x2 + y 2
z2
1 z 2 2
d cos θ dϕ = 2 1− dz(−y dx + x dy) − (x dx dy − y dy dx)
(x + y 2 )r r2 r2
z2
1 z 2 2
= 2 1− dz(−y dx + x dy) − (x + y )dx dy
(x + y 2 )r r2 r2
z2 1 1
1− 2
= 2 (r2 − z 2 ) = 2 (x2 + y 2 )
r r r
1
d cos θ dϕ = 3 [dz(−y dx + x dy) − z dx dy]
r
dz dx = ε dy = +dy , dz dy = ε321 dx = −dx , dx dy = ε123 dz = +dz
312
1 1
d cos θ dϕ = 3 (−y dy − x dx − z dz) = − 3 (x dx + y dy + z dz)
r r
1
The unit vector in the r direction is precisely dr = r (x dx + y dy + z dz), so
g g
F = d cos θ dϕ = − dr
4π 4πr2
That is a radial magnetic field from a point charge −g.
115
potential. The generalization to a relativistic particle is straightforward and only affects
kinetic terms, and that is the first quantized version of field theory. If we are still working in
“natural” units, with ~ = c = 1, the Lagrangian is
1
L = mẋ2 − eφ(x) + e~x˙ · A(~
~ x)
2
~ + ~v × B)
with e the charge of the particle. This would give the familiar Lorentz force e(E ~ →
~ ~
e(E + (~v /c) × B), where we have restored c in Heaviside-Lorentz units (in SI units, we would
instead change the dimensions of the magnetic field). Therefore, we need to take the vector
potential coupling to
~ x) → e ~x˙ · A(~
e~x˙ · A(~ ~ x) .
c
In the case of the monopole, the needed gauge transformation is
A ~ − 1 ∇Λ,
~→A ~ Λ=
eg
φ,
e 2π
so our Lagrangian for the particle goes to
1˙ ~
L→L− ~x · ∇Λ .
c
The text states that we must have e iΛ (2π) = e iΛ (0) for the gauge transformation to make
sense. Physically, this comes from the fact that the wavefunction of our particle undergoes
ψ → e−iΛ ψ
in natural units, and the wavefunction must remain single-valued. We should now check
how gauge transformations alter the wavefunction in Heaviside-Lorentz units. Restoring ~,
the amplitude to go from initial wavefunction ψ1 (~xi ) to a final wavefunction ψ2 (~xf ) in path
integral notation is
Z Z Z ~xf
i
R
∗
hψ2 |ψ1 i ≡ dxf ψ2 (~xf ) dxi ψ1 (~xi ) Dx e ~ dt L
~
xi
where L is the one with c restored as discussed, and Dx is the path integral measure. Doing
the gauge transformation shifts the phase in the path integral by
Z
1 ~ = − 1 [Λ(~xf ) − Λ(~xi )]
− dt ~x˙ · ∇Λ
~c ~c
because ~x˙ · ∇Λ
~ = ∂t Λ(~x(t)). The amplitude is therefore left invariant if the wavefunctions
transform as
ψ → e−iΛ/(~c) ψ .
The rest follows as in the text, and we find the condition
hcn
g= .
e
116
where h = 2π~ as usual. Note also that this is consistent with comparing equations (2) and
(3) on p. 307.
Solution:
6. Let g(x) be the element of a group G. The 1-form v = gdg † is known as the Cartan-Maurer
form. Then trv N is trivially
R closed on an N -dimensional manifold since it is already an N -
form. Consider Q = S N trv N with S N the N -dimensional sphere. Discuss the topological
meaning of Q. These considerations will become important later when we discuss topology
in field theory in chapter V.7. [Hint: Study the case N = 3 and G = SU (2).]
Solution:
Let’s consider the case for which G is some group whose manifold has no pathologies, for
instance S N , so that infinitesimal deviations from any point in G are sufficient to determine
the global structure of object Q. So consider letting g → g + δg. Then v changes as
So v transforms as
v → v + δg dg −1 − gd(g −1 δg g −1 )
= δg dg −1 − d(δg g −1 ) − dg g −1 δg g −1
= δg dg −1 − d(δg)g −1 + δg dg −1 − dg g −1 δg g −1
= −d(δg)g −1 + dg g −1 δg g −1
= −d(gg −1 δg)g −1 + dg g −1 δg g −1
= −g d(g −1 δg)g −1
117
So under g → g + δg, we have δv ≡ v[g + δg] − v[g] = −g d(g −1 δg)g −1 , and therefore:
Z Z Z Z
δQ = tr δv v...v + tr vδv...v + ... + tr vv...δv = N tr vv...v δv
Z
= −N tr vv...v gd(g −1 δg)g −1
Z
= −N tr (gdg −1 )(gdg −1 )...(gdg −1 )g d(g −1 δg)g −1
Z
= −N tr dg −1 gdg −1 ... gdg −1 gd(g −1 δg) (by cyclicity of trace)
Z
= +N tr dg −1 g dg −1 g ...gdg −1 (gg −1 )dg d(g −1 δg) (by dg −1 g = d(g −1 g)−g −1 dg = 0−g −1 dg)
Z
= +N tr dg −1 g dg −1 g ...gdg −1 dg d(g −1 δg) (since g −1 g = I)
Z
N −1
= (−1) N tr dg dg ... dg d(g −1 δg) (by repeating the above another (N − 1) times)
Z
N −1
N d tr dg dg ... dg g −1 δg (since d2 = 0)
= (−1)
=0
So the quantity Q calculated at a point g on the group space G and the quantity Q calcu-
lated at a point g + δg on G are the same. Therefore, barring any unforeseen pathologies on
the space G, we can compound the infinitesimal transformations and conclude that we can
calculate Q using any point on the group manifold and get the same answer. Thus Q is a
topological quantity, which depends only on the particular group G we pick. Now that we
know Q is topological, let’s try to figure out what it means.
First consider the case G = U (1) and N = 1. Then g(x) = einθ(x) , so v = gdg −1 =
neiθ d(e−iθ ) = −nidθ, and the Cartan-Maurer form is
Z Z
Q= v = −in dθ = −i2πn , n Z
S1 S1
In conclusion, for this case the quantity Q/(−2πi) counts the number of times the spatial
circle wraps around the group circle:
Q
Π1 (S 1 )
−2πi
For any N , the object Q properly normalized counts the number of times the spatial N -sphere
wraps around the group manifold G:
Q ΠN (G)
where Q is suitably normalized. In particular, for G = S N , ΠN (S N ) = Z. This mathematical
fact, that Q is proportional to an integer determined purely by topology, tells us that the
chiral anomaly does not get renormalized (see Chapter IV.7).
118
IV.5 Nonabelian Gauge Theory
3. In 4 dimensions εµνλρ trFµν Fλρ can be written as trF 2 . Show that dtrF 2 = 0 in any di-
mensions.
Solution:
where in the last step we have used the Bianchi identity DF ≡ dF + [A, F ] = 0.
5. For a challenge show that trF n , which appears in higher dimensional theories such as
string theory, are all total divergences. In other words, there exists a (2n − 1)-form ω2n−1 (A)
n
such
R 1 that trF = dω2n−1 (A). [Hint: A compact representation of the form ω2n−1 (A) =
0
dt f2n−1 (t, A) exists.] Work out ω5 (A) explicitly and try to generalize knowing ω3 and
ω5 . Determine the (2n − 1)-form f2n−1 (t, A). For help, see B. Zumino et al., Nucl. Phys.
B239:477, 1984.
Solution:
d (tr F n ) = n tr dF F n−1
= −n tr [A, F ]F n−1 (Bianchi identity)
n n−1
= −n tr AF − F AF
= −n tr (AF n − AF n ) (cyclicity of trace)
=0
Now let us find this (2n − 1)-form ω2n−1 (A). We follow the reference B. Zumino, Wu Yong-
Shi, A. Zee, “Chiral Anomalies, Higher Dimensions, and Differential Geometry,” Nucl. Phys.
B239:477 (1984).
119
Given a Yang-Mills potential A, define a 1-parameter class of potentials As ≡ sA and their
associated field strength tensors Fs = dAs + A2s = s dA + s2 A2 . Now differentiate:
d n dFs n−1
(tr Fs ) = n tr F
ds ds s
d 2 2 n−1
= n tr (s dA + s A )Fs
ds
= n tr (dA + 2sA2 )Fsn−1
But As = sA, so I can take the s from one A and put it in the other A to give
[As , A] = 2As A
So we have
d
(tr Fsn ) = n tr (dA + [As , A])Fsn−1
ds
= n tr Ds A Fsn−1
Ds = d + [As , · ] is the covariant derivative associated with the potential As . Using the
Bianchi identity Ds Fs = 0, the covariant derivative can act on the product AFsn−1 :
d
(tr Fsn ) = n tr Ds A Fsn−1
ds
= n d tr AFsn−1 + n tr As , AFsn−1
It is tempting to use the cyclicity of the trace to have those two terms add, but we have to
be careful. Let Ω be a (2n − 1)-form. Consider the following trace:
Cyclicity of the trace indeed lets us move the matrix Aµ to the left of the matrix Ωµ1 ...µ2n−1 .
But to repackage the expression into forms notation, we have to move the dxµ to the left of
all of the other dxµi , with each exchange picking up a minus sign. Since there are (2n − 1)
120
exchanges, we pick up an overall minus sign. So tr (ΩA) = −tr (AΩ), for Ω being an arbitrary
(2n − 1)-form.
π 1/2 Γ(n + 1)
ω2n−1 (A) = (−1)n−1 tr (g −1 dg)2n−1
22n−1 Γ(n + 1/2)
Thus for g G, and for 2n-dimensional Euclidean space (E2n ) whose boundary can be taken
as spherical, we have the result:
Z Z Z Z
n
tr F = dω2n−1 = ω2n−1 ∝ tr (g −1 dg)2n−1 Π2n−1 (G)
E2n E2n S 2n−1 S 2n−1
The quantity on the left is the nonabelian generalization of the chiral anomaly, and the quan-
tity on the right is an integer determined purely by topology.
121
IV.6 Anderson-Higgs Mechanism
1. Consider an SU (5) gauge theory with a Higgs field ϕ transforming as the 5-dimensional
representation: ϕi , i = 1, ..., 5. Show that a vacuum expectation value of ϕ breaks SU (5) to
SU (4). Now add another Higgs field ϕ0 , also transforming as the 5-dimensional representa-
tion. Show that the symmetry can either remain at SU (4) or be broken to SU (3).
Solution:
Now consider an SU (N ) gauge theory with two Higgs fields in the N -dimensional repre-
sentation. SU (N ) is the collection of transformations that leave the norm
N
X
(ϕ, ϕ0 )N ≡ ϕ† i ϕ0i
i=1
122
invariant. Suppose again that the SU (N ) symmetry is broken by ϕ obtaining a vacuum ex-
pectation value (VEV). As before, we can choose a coordinate system in field space such that
this VEV points in the N th direction: hϕi = (0, ..., 0, v). The question is, in which direction
does the VEV of the other Higgs field, ϕ0 , point? This direction is in principle given to us
by the potential.
If this VEV also happens to be in the N th direction, then hϕ0 i = (0, ..., 0, v 0 ). If this
VEV happens to be perpendicular to the N th direction, then we can still choose coordi-
nates such that hϕ0 i is aligned along one of the other directions, say the (N − 1)th direction:
hϕ0 i = (0, ..., 0, v 0 , 0), where now there are (N − 2) zeros to the left of v 0 rather than (N − 1)
zeros as in the previous case. It may be the case that this VEV is somewhere in between, nei-
ther parallel nor perpendicular to the first VEV. We therefore parameterize the most general
case as:
ϕ01
ϕ1
ϕ2
ϕ02
..
.. 0 .
ϕ= . , ϕ =
0
ϕN −1
ϕ N −2
v 0 cos β + ϕ0N −1
v + ϕN
v 0 sin β + ϕ0N
When β = 0, the second VEV is perpendicular to the first, and when β = π/2, the second
VEV is parallel to the first. Putting this parameterization into the SU (N )-invariant norm
gives
N
X −2
0
(ϕ, ϕ )N = ϕ† i ϕ0i + ϕ†N −1 (v 0 cos β + ϕ0N −1 ) + (v + ϕN )† (v 0 sin β + ϕ0N )
i=1
= (ϕ, ϕ0 )N −2 + ϕ†N −1 ϕ0N −1 + ϕ†N ϕ0N + v 0 cos β ϕ†N −1 + v † ϕ0N + v 0 sin βϕ†N + v † v 0 sin β
For generic β 6= π/2, there are terms linear in ϕN −1 and in ϕN , which are not invariant under
SU (N ) or SU (N − 1) transformations; this expression is only invariant under SU (N − 2),
as expressed by the SU (N − 2)-invariant norm (ϕ, ϕ0 )N −2 . For the special case β = π/2, the
term linear in ϕN −1 drops out, so we can repackage the ϕN −1 into an SU (N − 1)-invariant
norm.
So for the particular case N = 5, having two Higgs fields transforming as a 5 generically
breaks SU (5) to SU (3). In the special case for which the two VEVs happen to be aligned,
SU (5) instead breaks to SU (4).
123
2. In general, there may be several Higgs fields belonging to various representations labeled
by α. Show
P that the mass squared matrix for the gauge bosons generalizes immediately to
(µ ) = α g (Tα vα · Tαb vα ), where vα is the vacuum expectation value of ϕα and Tαa is the
2 ab 2 a
ath generator represented on ϕα . Combine the situations described in exercises IV.6.1 and
IV.6.2 and work out the mass spectrum of the gauge bosons.
Solution:
1
(µ2 )ab = g 2 v T {TRa , TRb }v
2
Now suppose there are α = 1, ..., N Higgs fields, each of which transforms under a represen-
tation Rα of the group G. Each has its own covariant derivative Dµ ϕα = ∂µ ϕα − igAaµ TRaα ϕα ,
so the kinetic term L = (Dϕ)† Dϕ yields
N
X N dim
X XG
L= (Dϕα )† Dϕα = g 2 Vα† TRaα TRb α Vα Aaµ Abµ + ...
α=1 α = 1 a,b = 1
Now let’s specialize to the situation from problem IV.6.1. We have N = 2 Higgs fields, both
of which transform under the 5-representation of G = SU (5). [The group SU (n) has n2 − 1
generators, so dim G = 52 − 1 = 24.] Taking the vacuum expectation value of the first field
ϕ to be hϕi = V (0, 0, 0, 0, 1)T , we can without loss of generality take the vacuum expectation
value of the second field ϕ0 to lie in the (ϕ4 , ϕ5 )-plane:
0
0
0 0
hϕ i = V 0
cos β
sin β
This gives
(µ2 )ab = g 2 c2 V 02 (tab )44 + 2csV 02 (tab )45 + (V 2 + s2 V 02 )(tab )55
124
where we have defined c ≡ cos β, s ≡ sin β and a set of symmetric matrices tab with compo-
nents:
(tab )ij ≡ {T5a , T5b }ij
where T5a is the ath generator of the 5-dimensional representation of SU (5).
As explained in IV.6.1, the SU (5) symmetry can be broken down to either SU (4) or SU (3),
depending on whether hϕi and hϕ0 i point along the same axis. The above mass matrix is
another manifestation of that statement; we see that if cos β = 0, then
so that whichever gauge bosons were massless with just the first Higgs field remain massless
with the addition of the second field as long as the two vacuum expectation values point in
the same direction. Alternatively, if sin β = 0 then
which shows that if the vacuum expectation values are perpendicular, each Higgs breaks a
separate direction in SU (5) and thereby gives mass to the corresponding gauge bosons.
4. In chapter IV.5 you worked out an SU (2) gauge theory with a scalar field ϕ in the
I = 2 representation. Write down the most general quartic potential V (ϕ) and study the
possible symmetry breaking patterns.
Solution:
Suppose that m2 < 0 and λ > 0 so that the potential exhibits spontaneous symmetry
breaking. By SO(5) invariance, we can choose field coordinates for which the vacuum expec-
~ points purely in the fifth direction: hφi
tation value of φ ~ = v(0, 0, 0, 0, 1)T . To study small
oscillations about this vacuum, write
χ1
χ2
~
φ= χ3
χ4
v+H
and study the SO(5) invariant norm
~·φ
φ ~=χ ~ + (v + H)2
~ ·χ
where the vector arrow over χ runs over only 4 indices, χ ~ = (χ1 , ..., χ4 )T . The analysis is just
~·φ
as in IV.6.1: if v 6= 0, then the norm φ ~ is invariant only under SO(4) transformations on
the χ~ fields. With only one scalar field in the theory, any negative mass squared instability
will break SO(5) to SO(4).
The situation is more complicated with additional scalar fields. If we stick with SO(5),
meaning introduce φ ~ and φ
~ 0 both transforming as 5-vectors under SO(5), then the analysis
proceeds analogously to that in problem IV.6.1: the SO(5) symmetry can break either to
SO(4) or to SO(3). If instead we insist on starting with traceless symmetric tensors ϕij and
ϕ0ij of SO(3), then we have to worry about cross terms of the form
V (ϕ, ϕ0 ) = λ1 [tr(ϕ2 )][tr(ϕ02 )] + λ2 tr(ϕ2 ϕ02 ) + λ3 tr(ϕϕ0 ϕϕ0 ) + λ4 tr(ϕ3 ϕ) + λ5 tr(ϕϕ03 )
in which case the analysis is more complicated.
5. Complete the derivation of the Feynman rules for the theory in (3) and compute the
amplitude for the physical process χ + χ → B + B.
Solution:
The Lagrangian is
1 1 1 1
L = − Bµν B µν + M 2 Bµ B µ + (∂χ)2 − m2 χ2 + Lint
4 2 2 2
√
where Bµν = ∂µ Bν − ∂ν Bµ , M = ev, m = 2λ v and
1 1
Lint = eM χBµ B µ + e2 χ2 Bµ B µ − λvχ3 − λχ4
2 4
126
where we have dropped an additive constant. The propagator for χ (solid line) is
i
i∆(q) =
q2 − m2 + iε
and the propagator for the massive photon Bµ (wavy line) is
i qµ qν
i∆µν (q) = 2 −ηµν + .
q − M 2 + iε M2
1
The cubic χBB vertex comes from L = eM χB 2 = 2
(2eM η µν )χBµ Bν and is represented
diagrammatically as:
Similarly, the quartic χχBB vertex comes from L = 21 e2 χ2 B 2 = 14 (2e2 η µν )χ2 Bµ Bν and is
represented diagrammatically as:
= i2e 2
The cubic and quartic self-couplings for χ come from L = −λvχ3 − 41 λχ4 = 1
3!
(−3!λv)χ3 +
1
4!
(−3!λ)χ4 , which are represented diagrammatically as:
= v
M1 =
p2 p 2’
127
This gives
M1 = ε10 µ ε20 ν (i2e2 )η µν .
Next we have a contribution from s-channel exchange of a χ:
p1 p 1’
k
M2 =
p2 p 2’
This gives
1
= iε10 µ ε20 ν (12λM 2 ) η µν
s − M2
where we have defined k = p1 + p2 = p10 + p20 and s ≡ k 2 , and we have used M = ev.
M3 = q
p2 p 2’
This gives
where q = p1 − p10 = p20 − p2 and t ≡ q 2 = (p1 − p10 )2 = (p20 − p2 )2 . We can use gauge
invariance in the form of ε10 ·p10 = 0 and ε20 ·p20 = 0 to write ε10 ·q = ε10 ·p1 and ε20 ·q = −ε20 ·p2 ,
so that this part of the amplitude becomes
pµ1 pν2
2 2 1 µν
M3 = iε10 µ ε20 ν (4e M ) η + .
t − M2 M2
The fourth is the crossed version of M3 , which results in
pµ2 pν1
2 2 1 µν
M4 = iε10 µ ε20 ν (4e M ) η +
u − M2 M2
128
where u ≡ (p20 − p1 )2 = (p2 − p10 )2 .
6. Derive (14). [Hint: The procedure is exactly the same as that used to obtain (III.4.9).]
Write L = 12 Aµ Qµν Aν with Qµν = (∂ 2 + M 2 )g µν − [1 − (1/ξ)]∂ µ ∂ ν or in momentum space
Qµν = −(k 2 − M 2 )g µν + [1 − (1/ξ)]k µ k ν . The propagator is the inverse of Qµν .
−i kµ kν
(14) gµν − (1 − ξ) 2
k 2 − M 2 + iε k − ξM 2 + iε
Solution:
The hint gave us the Fourier transform, so all we have to do is to find the matrix inverse
of Qµν in the Lorentz-index space. That is, solve Qµν (Q−1 )νρ = δρµ for (Q−1 )νρ . Lorentz
invariance tells us that (Q−1 )νρ = a gνρ + b kν kρ , so plug this into the matrix inverse equation:
The term multiplying δρµ must equal 1, and the term multiplying k µ kρ must equal zero. This
tells us that a = 1/(−k 2 + M 2 ) and
Therefore, we have
−1 −1 1 −1
(Q )νρ = gνρ + (ξ − 1) 2 kν kρ
k2 − M 2 k − ξM 2 k 2 − M 2
−1 kν kρ
= 2 gνρ + (ξ − 1) 2 .
k − M2 k − ξM 2
129
7. Work out the (...) in (13) and the Feynman rules for the various interaction vertices.
µ4 1 1 1 1
(13) L= − Fµν F µν + M 2 Aµ Aµ − M Aµ ∂ µ ϕ2 + [(∂ϕ01 )2 − 2µ2 (ϕ01 )2 ] + (∂ϕ2 )2 + ...
4λ 4 2 2 2
Solution:
We start with the U (1)-invariant scalar QED Lagrangian in Cartesian notation ϕ = √1 (ϕ1 +
2
iϕ2 ) as given in equation (12) of the text, on page 240:
1 1 µ2 λ
L = − F 2 + [(∂ϕ1 + eAϕ2 )2 + (∂ϕ2 − eAϕ1 )2 ] + (ϕ21 + ϕ22 ) − (ϕ21 + ϕ22 )2
4 2 2 4
Now it’s a matter of algebra:
Now break the symmetry by writing ϕ1 = v + h (h ≡ ϕ01 ) and performing some more algebra:
1 1 1 µ2 + e2 A2 2
L = − F 2 + (∂h)2 + (∂ϕ2 )2 + (v + 2vh + h2 + ϕ22 )
4 2 2 2
λ 4
v + 4v 3 h + 6v 2 h2 + 4vh3 + h4 + 2v 2 ϕ22 + 4vhϕ22 + 2h2 ϕ22 + ϕ42
−
4
+ eAµ (ϕ2 ∂ µ h − h∂ µ ϕ2 ) − evAµ ∂ µ ϕ2
Minimizing the potential with respect to v at the point for which all the fields are zero gives
∂L
= µ2 v − λv 3 = 0 =⇒ µ2 = λv 2
∂v
130
2 2
Therefore, µ2 2vh − λ4 4v 3 h = 0, which kills the terms linear in h, and µ2 h2 − λ4 6v 2 h2 =
1 2 2 1
√ 2 2
√
2
(1 − 3)λv h = − 2
( 2λ v) h , so the mass of the physical h particle is mh = 2λ v. The
Lagrangian is:
1 1 1 1 λ
L = − F 2 + (ev)2 A2 + (∂h)2 − m2h h2 − (4vh3 + h4 + 4vhϕ22 + 2h2 ϕ22 + ϕ42 )
4 2 2 2 4
1 1
+ (∂ϕ2 )2 − evA∂ϕ2 + eA(ϕ2 ∂h − h∂ϕ2 ) + e2 A2 (2vh + h2 + ϕ22 )
2 2
As explained on p. 240, we add the gauge fixing term Lgf = − 2ξ1 (∂A+ξM ϕ2 )2 , where M = ev
is the mass of the photon after spontaneous symmetry breaking. Since
1
Lgf = − [(∂A)2 + 2ξM ∂Aϕ2 + ξ 2 M 2 ϕ22 ]
2ξ
1 1
= − (∂A)2 − M ∂Aϕ2 − ξM 2 ϕ22
2ξ 2
1 1
= − (∂A)2 + M A∂ϕ2 − ξM 2 ϕ22 + (total derivative)
2ξ 2
the A∂ϕ2 term cancels from the Lagrangian (by design), and we get:
1 1 1
L = − F 2 − (∂A)2 + M 2 A2 ← (yields massive vector boson propagator in Rξ gauge)
4 2ξ 2
1 1
+ (∂h)2 − m2h h2 ← (yields scalar propagator for h)
2 2
1 1
+ (∂ϕ2 )2 − ξM 2 ϕ22 ← (yields scalar propagator for Goldstone boson ϕ2 )
2 2
1 1
− λv h3 − λ h4 − λϕ42 ← (cubic and quartic self-interactions for h, quartic for ϕ2 )
4 4
1
+ eM hAµ Aµ + e2 (h2 + ϕ22 )Aµ Aµ ← (photon-scalar interactions)
2
1
− λv hϕ22 − λ h2 ϕ22 ← (h-ϕ2 interactions)
2
+ e(ϕ2 ∂µ h − h∂µ ϕ2 )Aµ ← (photon coupling to scalar electromagnetic current)
Using solid lines for h, dashed lines for ϕ2 and wavy lines for Aµ , the interaction vertices in
momentum space are:
131
v
= +i(2e 2
= +i(2e 2
=
v
k2
= + e ( k 1 −k 2 )
k1
where in the last diagram, the arrows indicate that both momenta flow into the vertex.
8. Using the Feynman rules derived in exercise IV.6.7 calculate the amplitude for the physical
process ϕ01 + ϕ01 → A + A and show that the dependence on ξ cancels out. Compare with the
result in exercise IV.6.5. [Hint: There are two diagrams, one with A exchange and the other
with ϕ2 exchange.]
Solution:
132
and the ϕ2 propagator is
i
i∆µν (q) ≡ dashed line = .
q 2 − ξM 2 + iε
M5 = ϕ2 -exchange + Aµ -exchange
= ε10 µ ε20 ν [e(p1 + q)µ i∆(q) e(p2 − q)ν + (i2eM η µρ )i∆ρσ (q)(i2eM η σν )]
(2p1 − p10 )µ (2p2 − p20 )ν qµqν
2 2 1 µν
= ie ε10 µ ε20 ν + (2M ) 2 η + (ξ − 1) 2
q 2 − ξM 2 q − M2 q − ξM 2
M 2 µν M2 pµ1 pµ2
= 4ie2 ε10 µ ε20 ν η + 1 − (ξ − 1)
t − M2 t − M 2 t − ξM 2
where t ≡ q 2 = (p1 − p10 )2 = (p20 − p2 )2 , and again we have used gauge invariance in the form
of ε10 · p10 = 0 and ε20 · p20 = 0 to write ε10 · q = ε10 · p1 and ε20 · q = −ε20 · p2 . Now simplify
the quantity in parentheses:
M2 t − M 2 − (ξ − 1)M 2 t − ξM 2
1 − (ξ − 1) = = .
t − M2 t − M2 t − M2
The ξ-dependent numerator cancels the t − ξM 2 in the denominator of pµ1 pν2 , so we find
M2 pµ1 pµ2
2 µν
M5 = i4e ε10 µ ε20 ν η + .
t − M2 M2
M2 pµ2 pµ1
0 2 µν
M5 = i4e ε10 µ ε20 ν η +
u − M2 M2
133
9. Consider the theory defined in (12) with µ = 0. Using the result of exercise IV.3.5 show
that
ϕ2
1 4 1 2 4 4 25
Veff (ϕ) = λϕ + (10λ + 3e )ϕ ln 2 − + ...
4 64π 2 M 6
where ϕ2 = ϕ21 + ϕ22 . This potential has a minimum away from ϕ = 0 and thus the gauge
symmetry is spontaneously broken by quantum fluctuations. In chapter IV.3 we did not have
the e4 term and argued that the minimum we got there was not to be trusted. But here we
can balance the λϕ4 against e4 ϕ4 ln(ϕ2 /M 2 ) for λ of the same order of magnitude as e4 . The
minimum can be trusted. Show that the spectrum of this theory consists of a massive scalar
boson and a massive vector boson, with
m2scalar 3 e2
= .
m2vector 2π 4π
For help, see S. Coleman and E. Weinberg, Phys. Rev. D7: 1888, 1973.
Solution:
The Lagrangian is
where Lct includes the counterterms. Since the result will depend only on ϕ2c = ϕ21c + ϕ22c ,
we need only to consider diagrams whose external lines are ϕ1 . The diagrams we need to
consider are of the form
+ + + + ...
The text already calculated the scalar contributions to the effective potential. In problem
IV.3.5, we showed that the photon contribution is the same as the scalar contribution, except
with the replacement λ → 2e2 and an overall factor of 3. Thus the photon contribution to
Veff is
ϕ2c
(A) 1 2 2 4 25
Veff (ϕc ) = 3 × [2e ] ϕc ln 2 −
256π 2 M 6
4 2
3e 4 ϕ 25
= 2
ϕc ln c2 − .
64π M 6
134
where the couplings are renormalized at the scale M .
Now treat λ as a coupling of order e4 , so that the term of order λ2 = O(e8 ) should be
dropped. Then we have
3e4 4 ϕ2c
1 4 25
Veff (ϕc ) = λϕc + ϕ ln 2 −
4! 64π 2 c M 6
0
Let the value ϕc = v be the location of the minimum of Veff (ϕc ), or in other words Veff (v) = 0.
0
Furthermore, let us choose M = v. Then the equation Veff (v) = 0 implies
33e4
λ= .
8π 2
Putting this back into the effective potential results in
3e4 4 ϕ2c
1
Veff (ϕc ) = ϕ ln 2 − 2 .
64π 2 c v
00 3e4 2
m2S ≡ Veff (v) = v .
8π 2
The mass of the vector after spontaneous symmetry breaking can be read off from the La-
grangian as
mV = ev .
Dividing these, we arrive at the relation
m2S 3α
2
=
mV 2π
where α ≡ e2 /(4π).
135
IV.7 Chiral Anomaly
1. Derive (11) from (9). The momentum factors k1λ and k2σ in (9) become the two derivatives
in Fµν Fλσ in (11).
i µνλσ
qλ ∆λµν (a, k1 , k2 ) = ε k1λ k2σ (9)
2π 2
e2 µνλσ
∂µ J5µ = ε Fµν Fλσ (11)
(4π)2
Solution:
We will derive this relation by comparing the matrix elements of both operators in the
canonical formalism, and finding that they are the same up to numerical factors. We are
interested in processes for which an operator causes two photons to pop out of the vacuum,
or in other words we want matrix elements of the form
hk, a; k 0 , b|O(0)|0i
where a labels the polarization of the outgoing photon with momentum k, and b labels
the polarization of the outgoing photon with momentum k 0 . By translation invariance,
0
hk, a; k 0 , b|O(x)|0i = e−i(k+k )x hk, a; k 0 , b|O(0)|0i, so it is sufficient to consider the opera-
tor at the origin x = 0. The first matrix element we will consider is for the operator
O(x) = −i∂µ J5µ (x), for which all of the work has already been done in the chapter:
e2 µνλσ
hk, a; k 0 , b|(−i)∂µ J5µ (0)|0i = εaµ (k)∗ εbν (k 0 )∗ e2 (−i)qλ ∆λµν (a, k, k 0 ) = εaµ (k)∗ εbν (k 0 )∗ 2
ε kλ kσ0 .
2π
The factor of −i came from the Fourier replacement ∂µ → iqµ . Now consider the matrix
element for O(x) = εµνρσ ∂µ Aν (x)∂ρ Aσ (x). We will need the plane-wave expansion for the
photon field:
XZ d3 p a −ipx † a ∗ +ipx
Aν (x) = a p ε ν (p) e + a ε
p ν (p) e .
a
(2π)3 2ωp
The spacetime derivative at the origin is
XZ d3 p a † a ∗
∂µ Aν (0) = (−ip µ ) a p ε ν (p) − a p ε ν (p) .
a
(2π)3 2ωp
We will need the following properties of the creation and annihilation operators acting on
the vacuum state:
ak |0i = 0
ak a†k0 |0i = (2π)3 (2ωk )δ 3 (~k − ~k 0 )
hk, k 0 |0i = h0|ak0 ak |0i = 0
hk, a; k 0 , b|p, c; p0 , di =
h i
3 3 ac bd 3 ~ 3 ~0 0 ad bc 3 ~ 0 3 ~0
[(2π) 2ωk ][(2π) 2ωk0 ] δ δ δ (k−~p )δ (k −~p )+δ δ δ (k−~p )δ (k −~p )
136
3
In what follows, define (dp) ≡ (2π)d 3p2ωp and suppress the polarization labels on the outgoing
states to save room. Now compute the matrix element:
= −2 ε µνρσ
kµ kρ0 εaν (k)∗ εbσ (k 0 )∗
= −2 ερµσν kρ kσ0 εaµ (k)∗ εbν (k 0 )∗
= −2 (−1)3 εµνρσ kρ kσ0 εaµ (k)∗ εbν (k 0 )∗
= +2 εµνρσ εaµ (k)∗ εbν (k 0 )∗ kρ kσ0
so in total we have
hk, a; k 0 , b|εµνρσ Fµν (0)Fρσ (0)|0i = +8 εµνρσ εaµ (k)∗ εbν (k 0 )∗ kρ kσ0 .
Therefore
e2 e2
1
hk, a; k 0
, b|∂µ J5µ (0)|0i = 2 εµνλσ εaµ (k)∗ εbν (k 0 )∗ kλ kσ0 = 2 0
hk, a; k , b|εµνρσ
Fµν (0)Fρσ (0)|0i .
2π 2π 8
We therefore have the operator equation
e2 µνρσ
∂µ J5µ = ε Fµν Fρσ .
16π 2
2. Following the reasoning in chapter IV.2 and using the erroneous (10) show that the
decay amplitude for the decay π 0 → γ + γ would vanish in the ideal world in which the π 0
is massless. Since the π 0 does decay and since our world is close to the ideal world, this
provided the first indication historically that (10) cannot possibly be valid.
Solution:
137
where k µ is the momentum of the decaying neutral pion. This amplitude has one initial
hadron and no final hadrons. By translation invariance, we have
As in chapter IV.2, we have hγ1 γ2 |J5µ (0)|π 0 (k)i = f0 k µ and so kµ hγ1 γ2 |J5µ (0)|π 0 (k)i = f0 k 2 =
f0 m2π0 .
Since hγ1 γ2 |∂µ J5µ (x)|π 0 (k)i = −ikµ hγ1 γ2 |J5µ (0)|π 0 (k)ie−ik·x = −if0 m2π0 e−ik·x , we deduce (in-
correctly) that ∂µ J5µ (x) = 0 if and only if f0 m2π0 = 0.
In other words, mπ0 → 0 would seem to imply kµ hγ1 γ2 |J5µ (0)|π 0 (k)i → 0. Since A(π 0 → γγ)
is proportional to that, we deduce incorrectly that the decay π 0 → γγ cannot occur.
3. Repeat all the calculations in the text for the theory L = ψ̄(iγ µ ∂µ − m)ψ.
d4 p
Z
λµν λ 5 1 ν 1 µ 1
∆ (k1 , k2 ) = i tr γ γ γ γ + (µ, k1 ↔ ν, k2 ) ,
(2π)4 6 p− 6 q − m 6 p− 6 k1 − m 6 p − m
which is linearly divergent by power counting. Therefore, we have to specify how we will
regulate and evaluate it. Following the text, we will set up all divergent integrals as integrals
of total derivatives, using the formula
Z
d4 p[f (p + a) − f (p)] = lim i aµ P µ (2π 2 P 2 )f (P )
P →∞
evaluated symmetrically over the 3-sphere at infinity (so that P µ P ν /P 2 → 41 g µν , etc.). Fur-
thermore, we will exchange (µ, k1 ↔ ν, k2 ) at the last step in any calculation, and we define
the momentum space amplitude as the amplitude given above with loop momentum p shifted
by a chosen so that the vector currents are conserved: k1µ ∆λµν = 0 and k2ν ∆λµν = 0. Tech-
nically, also, when P → ∞, we should be taking that limit in Euclidean space, so we don’t
get complications from null vectors, and we can always take P 2 m2 .
So let us find the difference of the shifted and unshifted amplitudes, as in the text:
i 2N
∆λµν (a) − ∆λµν (0) = (i2π 2 ρ
)a lim P ρ P
(2π)4 P →∞ D
where the numerator is
N = tr γ λ γ 5 (6 P − 6 q + m)γ ν (6 P − 6 k1 + m)γ µ (6 P + m)
138
and the denominator is
D = [(P − q)2 − m2 ][(P − k1 )2 − m2 ][P 2 − m2 ] .
To this we also add (µ, k1 ↔ ν, k2 ) as in the text. Only the terms in the numerator with six
powers of P survive, so we need tr[γ λ γ 56 P γ ν6 P γ µ6 P ] = 2P µ tr[γ λ γ 56 P γ ν6 P ] − P 2 tr[γ λ γ 56 P γ ν γ µ ]
using the gamma-matrix anticommutator. Then tr[γ 5 γ σ γ ν γ µ γ λ ] = 4iεσνµλ , so the first term
vanishes. Then
4i
∆λµν (a) − ∆λµν (0) = 2 aρ εσνµλ lim Pρ Pσ + (µ, k1 ↔ ν, k2 )
8π P →∞
i
= 2 εσνµλ aσ + (µ, k1 ↔ ν, k2 ) .
8π
Now we need to find the shift variable a so that the vector currents are conserved. The only
independent momenta available in the problem are k1 and k2 , and Bose symmetry implies
that we take a = β(k1 − k2 ). Thus
iβ λµνσ
∆λµν (a) − ∆λµν (0) = ε (k1 − k2 )σ
4π 2
as in the text. Now we can enforce vector current conservation. Using 6 k1 = (6 p − m)−
(6 p−6 k1 − m) = (6 p− 6 k2 − m) − (6 p−6 q − m), we can write
d4 p
Z
λµν λ 5 1 ν 1 λ 5 1 ν 1
k1µ ∆ (0) = i tr γ γ γ −γ γ γ
(2π)4 6 p− 6 q − m 6 p− 6 k1 − m 6 p− 6 q − m 6 p − m
d4 p
Z
λ 5 1 ν 1 λ 5 1 ν 1
+i tr γ γ γ −γ γ γ
(2π)4 6 p− 6 q − m 6 p − m 6 p− 6 k2 − m 6 p − m
d4 p ∂
Z
1 1
= −ik1ρ λ 5
tr γ γ γ ν
(2π)4 ∂pρ 6 p− 6 k2 − m 6 p − m
d4 p ∂ (p − k2 )σ pτ
Z
= 4k1ρ εσντ λ
(2π)4 ∂pρ [(p − k2 )2 − m2 ][p2 − m2 ]
4i ρ Pρ Pτ
=− 2
k1 k2σ ελνστ lim
8π P →∞ P 2
i λντ σ
=+ ε k1τ k2σ .
8π 2
In the third equality, we have used the fact that traces of γ 5 with fewer than four other
gamma matrices vanish. Putting this all together, we obtain:
iβ λµνσ
k1µ ∆λµν (a) = k1µ ∆λµν (0) + ε k1µ k2σ
4π 2
i
= 2
(1 + 2β)ελµνσ k1µ k2σ
8π
139
Demanding vector current conservation means setting this quantity to be zero, so as in the
text we need to set β = −1/2.
6 q = (6 p − m) − (6 p − 6 q + m) + 2m
where the difference in sign of m in the second term is due to anticommuting the slashed
momenta with γ 5 . After some algebra,
d4 p ∂
Z
λµν ρ 5 1 ν 1
qλ ∆ = −ik1 4 ρ
tr γ γ γ + (µ, k1 ↔ ν, k2 ) + 2m∆µν
µ
(2π) ∂p 6 p− 6 k2 − m 6 p − m
where
d4 p
Z
µν 5 1 ν 1 µ 1
∆ (k1 , k2 , m) = i tr γ γ γ .
(2π)4 6 p− 6 q − m 6 p− 6 k1 − m 6 p − m
The first two terms evaluate just as in the book to give
4i ρ Pσ Pτ i
2
k1 (−k2σ )εσντ µ lim 2
+ (µ, k1 ↔ ν, k2 ) = 2 εµνστ k1σ k2τ .
8π P →∞ P 4π
Thus, our properly regulated amplitude is
i µνστ
qλ ∆λµν (a) = ε k1σ k2τ + 2m∆µν (k1 , k2 , m) .
2π 2
Now we just need to calculate ∆µν (k1 , k2 , m). We’ll work on the first integral and then
perform the Bose swap. We have
From the trace, the only terms that don’t vanish are linear in m, or
m tr γ 5 γ ν (6 p− 6 k1 )γ µ 6 p + γ 5 (6 p− 6 q)γ ν γ µ 6 p + γ 5 (6 p− 6 q)γ ν (6 p− 6 k1 )γ µ ,
Thus
∆µν (k1 , k2 , m) = 4m εµνστ k1σ k2τ [I(q, k1 , m) + I(q, k2 , m)]
where
d4 p
Z
1
I(q, k, m) =
(2π) [(p − q) − m ][(p − k)2 − m2 ][p2 − m2 ]
4 2 2
140
is superficially convergent, so we can just evaluate it and shift variables as desired. Using
Feynman parameters and shifting variables, we get
Z 1
d4 p δ(x + y + z − 1)
Z
I(q, k, m) = 2 dx dy dz
0 (2π) [p − m − 2p · qx − 2p · ky + q 2 x + k 2 y]3
4 2 2
Z 1
d4 ` δ(x + y + z − 1)
Z
=2 dx dy dz
0 (2π)4 [`2 − m2 + q 2 x(1 − x) + k 2 y(1 − y) + 2q · kxy]3
for ` ≡ p − qx − ky. Integrating gives
Z 1
i δ(x + y + z − 1)
I(q, k, m) = − 2
dx dy dz 2 .
16π 0 m − q x(1 − x) − k 2 y(1 − y) − 2q · kxy
2
It is relatively easy to put the expression back together again. The extra term is just what
we would expect from hψ̄γ 5 ψJ µ J ν i.
given in lowest orders by triangle diagrams with axial currents at each vertex. [Hint: Call
the momentum space amplitude ∆λµν 5 2
5 (k1 , k2 ).] Show by using (γ ) = 1 and Bose symmetry
that
1 λµν
∆λµν
5 (k1 , k2 ) = [∆ (a, k1 , k2 ) + ∆µνλ (a, k2 , −q) + ∆νλµ (a, −q, k1 )].
3
Now use (9) to evaluate qλ ∆λµν
5 (k1 , k2 ).
The momentum space amplitude, by analogy with the one studied in the text, is
d4 p
Z
λµν λ 5 1 ν 5 1 µ 51
∆5 = i tr γ γ γ γ γ γ + (µ, k1 ↔ ν, k2 ) .
(2π)4 6 p−6 q 6 p−6 k1 6p
We have only the two terms because there are only two circular permutations of the 3 vertices.
This is of course a divergent integral, so we must specify how we will regulate it. First off,
by Bose symmetry, since the 3 currents are identical, we must choose a regulator which is
invariant under (λ, −q ↔ µ, k1 ) and (λ, −q ↔ ν, k2 ) as well as (µ, k1 ↔ ν, k2 ). (The sign on
q is because q is incoming as opposed to outgoing.) A simple way to implement this type
of regularization is to define ∆λµν
5 (a) as the expression ∆5
λµν
above but with the integration
variable p shifted to p + a and then define our regulated amplitude as
1
∆λµν
5 = [∆λµν (0) + ∆µνλ νλµ
5 (k1 ) + ∆5 (q)]
3 5
141
with the permutation of indices indicating that we have also used the cyclic property of the
trace. Using the second or third terms would be equivalent to evaluating the Feynman dia-
gram starting at a different vertex.
From looking at the integral expression for ∆λµν 5 above, we can anticommute the second
γ matrix through 1/(6 p−6 k1 ) and 1/(6 p−6 k2 ) and then γ µ and γ ν , and use (γ 5 )2 = 1 to see
5
that
∆λµν
5 (0) = ∆
λµν
(k1 , k2 ) , ∆µνλ
5 (k1 ) = ∆
µνλ
(k2 , −q) , ∆νλµ
5 (q) = ∆
νλµ
(−q, k1 ) .
We should now use the appropriately shifted versions of the ∆ amplitudes that conserve
vector currents, so
1 −1 −1
∆λµν
5 = [∆λµν (a = (k1 − k2 ), k1 , k2 ) + ∆µνλ (a = (k2 + q), k2 , −q)
3 2 2
1
+ ∆νλµ (a = (q + k1 ), −q, k1 )] .
2
We see that
1 i
qλ ∆λµν
5 = qλ ∆λµν (a, k1 , k2 ) = 2 εµνλσ k1λ k2σ .
3 6π
The anomaly is spread evenly over the three vertices.
7. Define the fermionic measure Dψ in (16) carefully by going to Euclidean space. Cal-
culate the Jacobian upon a chiral transformation and derive the anomaly. [Hint: For help,
see K. Fujikawa, Phys. Rev. Lett. 42: 1195, 1979.]
Solution:
We will proceed somewhat differently from Fujikawa’s original paper. First, we will use
the two-component spinor notation instead of the 4-component Dirac notation. Second, we
will stay in Minkowski signature for most of the discussion and rotate to Euclidean space
only to compute an integral. For this problem we use the metric convention η = (−, +, +, +).
Consider the Lagrangian for one massless 2-component spinor ψ charged under a U (1) gauge
symmetry:
1
L = iψȧ† σ̄ µȧa (∂µ − igAµ )ψa − Fµν F µν
4
The covariant derivative Dµ ψ ≡ (∂µ − igAµ )ψ is present to preserve gauge invariance. For
later convenience, we define the following notation:
vaȧ ≡ σaµȧ vµ , v̄ ȧa ≡ σ̄ µȧa vµ for any Lorentz vector v µ
Consider the U (1) gauge transformation on ψ:
Z
0 iθ(x)
ψa (x) = e ψa (x) = d4 y e iθ(x) δ 4 (x, y)δa b ψb (y)
142
We thus read off the Jacobian that defines this transformation:
δψa0 (x)
Ja b (x, y) ≡ = e iθ(x) δ 4 (x, y)δa b
δψb (y)
x0 = Jx implies dx0 = dx det J.
R R
For commuting numbers, the coordinate transformation
For anticommuting numbers, x0 = Jx implies dx0 = dx (det J)−1 . So the above change of
R R
variables gives Z Z Z
Dψ = Dψ (det J) = Dψ e−tr ln J
0 −1
Imagine putting the field theory on a lattice, so that you are comfortable with δ 4 (x, y) being
simply the identity matrix in spacetime. The log is then:
ln J = ln e iθ I = ln [(1 + iθ + ...)I] = iθ I + O(θ2 )
R
Putting this into the trace, we run into the immediate problem that tr I = d4 x δ 4 (x, x) = ∞,
but this is really no surprise at all. When we evaluate the path integral for a free scalar field
theory, for example, we get a determinant:
Z Z
∗ ∗
−1
4 2 2
DφDφ exp i d x φ ∂ − m φ ∝ det ∂ 2 − m2
The determinant is a product of eigenvalues, and the eigenvalues can be arbitrarily large,
which causes the determinant to diverge. When we can treat the determinant as an overall
factor for the path integral, we don’t have to worry about this. When we need to extract
physical content from the determinant, we need to regulate it. Phrased in this way, the
natural way to regulate the determinant is to impose an upper cutoff on the possible size of
the eigenvalues.
We are now tempted to apply the above reasoning for the spinor field: when we evaluate the
path integral over the fermion ψ, we get a determinant:
Z Z
† 4 †
DψDψ exp i d x iψ D̄ψ ∝ det D̄
However, continuing this line of thought fails, because the operator D̄ does not have eigen-
values. Recall that an eigenvalue equation for the operator O is O~v = λ~v , where λ is just a
number. That is, there exists a vector ~v for which applying the matrix O returns something
proportional to that same vector ~v .
Look at the operator D̄, defined as D̄ȧa ≡ σ̄ µȧa Dµ . The fact that one index is dotted
while the other isn’t tells you everything you need to know: the operator D̄ acts on a vector
in SU (2)L but returns a vector in SU (2)R – it can’t possibly have an eigenvalue equation.
We remedy this problem by constructing operators that can have eigenvalue equations.
Namely, consider the operators
(DD̄)ac ≡ Daȧ D̄ȧc and (D̄D)ȧċ ≡ D̄ȧa Daċ
143
As you can see from the indices, the operator DD̄ acts on a vector in SU (2)L and returns
a vector in SU (2)L . Similarly, the operator D̄D acts on a vector in SU (2)R and returns a
vector in SU (2)R . We are thus free to regulate these operators by imposing upper cutoffs on
their eigenvalue spectra.
R
Back to the matter at hand: we wish to compute tr ln J = tr iθ I = d4 x iθ(x) tr δ 4 (x, x)I2 ,
where Ja b (x, y) = eiθ(x) δ 4 (x, y)δa b is the Jacobian for the transformation ψa0 (x) = eiθ(x) ψa (x).
We wish to regulate this by cutting off the eigenvalues of the operator DD̄, which is an
operator that acts purely within SU (2)L . Let |λi be an eigenstate of DD̄ with eigenvalue λ,
meaning (DD̄)|λi = λ|λi. Let us now regulate:
!
X X
δ 4 (x, x) = lim δ 4 (x, y) ≡ lim hx|yi = lim hx| |λihλ| |yi = lim hx|λihλ|yi
y→x y→x y→x y→x
λ λ
−λ/Λ2 −DD̄/Λ2
X X
regulate
→ lim hx|λihλ|yi e = lim hx| e |λihλ|yi
y→x y→x
λ λ
!
2 2
X
= lim hx|e−DD̄/Λ |λihλ| |yi = lim hx|e−DD̄/Λ |yi
y→x y→x
λ
d4 k d4 k
Z
2 2
= lim hx|e−DD̄/Λ 4
|kihk| |yi = 4
lim hx|e−DD̄/Λ |kihk|yi
y→x (2π) (2π) y→x
The momentum states |ki in the position basis are hy|ki = e+iky . For any function f , we
know how to write D acting on it in the position basis:
∂
hx|Dµ |f i = − igAµ (x) f (x)
∂xµ
We therefore know how to write any function of D acting on any function in the position
basis, so in particular:
2 2
hx|e−DD̄/Λ |ki = e−(DD̄/Λ )(x) e ikx
The argument (x) for the operator DD̄ is there to remind you that these are derivatives
with respect only to x, not y, so we can move the hk|yi = hy|ki∗ = e−iky to the left of the
2
hx|e−DD̄/Λ |ki. Taking the limit y → x, we arrive at the regulated delta function:
Now let’s put the operator DD̄ into a nicer form. Introduce the usual notation M(µν) ≡
1
2
(Mµν + Mνµ ) and M[µν] ≡ 12 (Mµν − Mνµ ), so that any matrix can be written as the sum
of its symmetric and antisymmetric parts: Mµν = M(µν) + M[µν] . For any function f (x), we
have:
(µ [µ
(DD̄)ab f (x) = σaµȧ σ̄ ν ȧb Dµ Dν f (x) = (σaȧ σ̄ ν)ȧb + σaȧ σ̄ ν]ȧb )Dµ Dν f (x)
(µ
By Lorentz invariance, we know σaȧ σ̄ ν)ȧb = Cδa b η µν , where C is some constant. For µ = ν =
0, the left-hand side is just the identity matrix δa b , while the right-hand side equals Cδa b η 00 .
144
So the constant C simply cancels the 00 piece of the metric: η 00 = −1 =⇒ C = −1. To
summarize:
(µ
σaȧ σ̄ ν)ȧb = −δa b η µν
The antisymmetric piece has no neat simplification as-is, since it is just proportional to the
generator of rotations. Recalling that the commutator of covariant derivatives defines the
field strength, [Dµ , Dν ] = −igFµν , we proceed:
[µ
(DD̄)ab f (x) = −δa b η µν Dµ Dν f (x) + σaȧ σ̄ ν]ȧb Dµ Dν f (x)
1 [µ
= −δa b Dµ Dµ f (x) + σaȧ σ̄ ν]ȧb [Dµ , Dν ]f (x)
2
b 2 1 [µ ν]ȧb
= −δa D − ig σaȧ σ̄ Fµν f (x)
2
Since this expression appears exponentiated, you may expect that we’ll need to raise it to
various powers. Actually we will only need the square of the term with the Fµν . For later
convenience, let’s simplify that term now. By Lorentz invariance, we know:
Before worrying about the terms that don’t go to zero (and in fact look divergent), let’s
remind ourselves that the original Lagrangian was L = iψ † D̄ψ, so in the path integral we
need to integrate over ψ † in addition to ψ. All of the previous work was to regulate the
Jacobian J from the transformation
ψa0 (x) = e iθ(x) ψa (x) =⇒ Ja c (x, y) = e iθ(x) δ 4 (x, y)δa c
But we also need to include the Jacobian J˜ from the conjugate transformation
0
ψ † ȧ (x) = e−iθ(x) ψ †ȧ (x) =⇒ J˜ȧċ (x, y) = e−iθ(x) δ 4 (x, y)δ ȧċ
This leads to a total change in the path integral measure:
Z Z Z
˜ ˜
Dψ D(ψ ) = DψDψ (det J) (det J) = DψDψ † e−tr ln J−tr ln J
0 0 † † −1 −1
There are two things to notice here. First, most of the manipulations carry through for J˜
exactly as for J, so the opposite sign for the phase kills most of the terms upon the addition
˜ The second thing to notice is that the opposite sign does not kill all of the
tr ln J+tr ln J.
terms when adding the two Jacobians.
How can this be? To regularize J, we needed the operator (DD̄), which acts purely within
˜ we need the operator (D̄D), which acts purely within SU (2)R .
SU (2)L . To regularize J,
Everything carries through in exactly the same way, except that squaring DD̄ led to the
expression
(σ [µ σ̄ ν] σ [ρ σ̄ σ] )ac = δa c (−2η ρ[µ η ν]σ + i εµνρσ )
whereas squaring D̄D leads to the expression
(σ̄ [µ σ ν] σ̄ [ρ σ σ] )ȧċ = δ ȧċ (−2η ρ[µ η ν]σ − i εµνρσ )
146
The minus sign is critical: tracing and subtracting the two leaves a nonzero piece +4i εµνρσ .
Therefore, subtracting the two regularized delta functions (recall there is a trace over each
set of spinor indices sitting out front) gives:
The first trace is over SU (2)L indices, and the second trace is over SU (2)R indices, which of
course is the way it must be for the expression to make any sense.
Now you see what is happening here. The lowest order term (just the identity matrix)
drops out. The first order term also drops out, except for the subtraction σ [µ σ̄ ν] − σ̄ [µ σ ν] .
Since σ [µ σ̄ ν] = 12 (σ µ σ̄ ν − σ ν σ̄ µ ), the cyclic property of the trace over SU (2)L indices causes
that term to be zero. Similarly, the trace over SU (2)R indices yields tr(σ̄ [µ σ ν] ) = 0.
So expanding the exponentials yields terms that will cancel each other out, trace to zero, or
will go to 0 for Λ → ∞. As discussed, the only exception is:
So tracing over the delta functions and subtracting them, and then taking the cutoff to
infinity, gives
2
d4 k −k2 1 ig
Z
4 4
tr δ (x, x)reg,J − tr δ (x, x)reg,J˜ = e 4i εµνρσ Fµν Fρσ
(2π)4 2 2
d4 k −k2
Z
1 2 µνρσ
= − ig ε Fµν Fρσ e
2 (2π)4
∞ 4 4
d4 k −k2 d4 kE −kE2
Z
−i
Z Z
dk −k2 1
e = −i e = −i e = −i √ =
(2π)4 (2π)4 −∞ 2π 2 π 16π 2
4 4 g 2 µνρσ
tr δ (x, x)reg,J − tr δ (x, x)reg,J˜ = − ε Fµν Fρσ
32π 2
R
Now recall tr ln J = d4 x iθ(x) tr δ 4 (x, x)reg , so
g 2 µνρσ
Z
˜ 4
tr ln J + tr ln J = i d x θ(x) − ε Fµν Fρσ
32π 2
147
We have finally managed to compute the change in the path integral measure. Under the
change of integration variables
0
ψa0 (x) = e iθ(x) ψa (x) , ψ † ȧ (x) = e−iθ(x) ψ †ȧ (x)
We see that the measure is not invariant under this change of variables (in other words, the
Jacobian is not 1).
Finally, we are ready to state the result. The infinitesimal transformation ψ → ψ + δψ, ψ † →
ψ † + δψ † with δψa (x) = iθ(x) ψa (x), δψ †ȧ = −iθ(x) ψ †ȧ changes the path integral as:
Z Z
† µȧa 1
→ DψDψ † e i d x L+i d x θ(x)[∂µ (ψȧ σ̄ ψa )+ 32π2 g ε Fµν Fρσ ]
R 4 R 4 2 µνρσ
† i d4 x L
R
DψDψ e
R∞ †
Just as the integral −∞ dx f (x) is not a function of x, the path integral DψDψ † e iS(ψ,ψ ) is
R
not a function of ψ or ψ † . Therefore the path integral should remain completely unchanged
under the infinitesimal transformation described above. Moreover, since this transformation
is for an arbitrary infinitesimal function of spacetime θ(x), we conclude that
1 2 µνρσ
∂µ ψȧ† σ̄ µȧa ψa = − g ε Fµν Fρσ
32π 2
This entire analysis was for one complex Weyl spinor. Suppose we introduce a second complex
Weyl spinor ξ. The Lagrangian density is
1
L = iψȧ† σ̄ µȧa (∂µ − igAµ )ψa + iξȧ† σ̄ µȧa (∂µ − igAµ )ξa − Fµν F µν
4
This Lagrangian exhibits a classical U (1) gauge symmetry
iθ(x)
ψa e ψa
→
ξa e iθ(x) ξa
148
with the associated Noether current
Classically Noether’s theorem says ∂µ j µ = 0, but we have just derived that ψ and ξ each
contributes a nonzero contribution to this classical conservation law, so that at the quantum
level we have
1 2 µνρσ
∂µ j µ = − g ε Fµν Fρσ
16π 2
Thus this theory does not actually have the U (1) gauge symmetry, even though it appears
as though it did at the classical level. That is the statement of the chiral anomaly.
Note the relative minus sign between the two terms: the contributions to ∂µ j µ cancel exactly.
Thus the theory truly does have this U (1) gauge symmetry, even quantum mechanically.
To compare this to the usual treatment in terms of Dirac spinors, let us package the 2-
component fields ψa and ξa into a 4-component Dirac spinor:
ψ
Ψ ≡ †aȧ
ξ
is the transformation Ψ → e iθ(x) Ψ. The current associated with this transformation is the
usual electromagnetic current, which is indeed conserved.
149
8. Compute the pentagon anomaly by Feynman diagrams in order to check remark 6 in
the text. In other words, determine the coefficient c in ∂µ J5µ = ... + cεµνλσ trAµ Aν Aλ Aσ .
Solution:
The pentagon diagram has 5 fermion propagators, so the integral goes as Λ d4 k 6k15 ∼ Λ1 ,
R
which is convergent as Λ → ∞ and thereby naively appears not to contribute to the anomaly.
However, gauge invariance relates the pentagon diagram to the triangle and box diagrams,
so that only the set of all three is physically meaningful. Demanding that the box diagram
conserves vector currents then generates a pentagon anomaly in the form
1 µνλσ
∂µ J5aµ = ... + ε tr{(T5a )[ 14 vµν vλσ + 12
1
aµν aλσ ]
4π 2
+ i 23 (aµ aν vλσ + vµν aλ aσ + aµ vνλ aσ ) − 83 aµ aν aλ aσ ]} .
Here we have split up the gauge field into vector and axial vector components, 6 A =6 v + γ 56 a,
and defined the corresponding field strengths vµν = ∂µ vν − ∂ν vµ − i[vµ , vν ] − i[aµ , aν ] and
aµν = ∂µ aν − ∂ν aµ − i[vµ , aν ] − i[aµ , vν ]. The matrix T5a is the generator associated with the
axial coupling to fermions: aµ = aaµ T5a .
See W. A. Bardeen, “Anomalous Ward Identities in Spinor Field Theories,” Phys. Rev.
Vol. 184 No. 5, 25 Aug 1969 for further details.
to (1). Derive the corresponding equation of motion for ϕ. The equation, known as the
Gross-Pitaevski equation, has been much studied in recent years in connection with the
Bose-Einstein condensate.
1
L = iϕ† ∂0 ϕ − ∂i ϕ† ∂i ϕ − g 2 (ϕ† ϕ − ρ̄)2 (1)
2m
Solution:
Adding the term Lpot = −W (~x)ϕ† ϕ implies the addition of a term −W (~x)ϕ to the equation
of motion found from varying L with respect to ϕ† .
150
V.2 Euclid, Boltzmann, Hawking, and Field Theory at Finite Tem-
perature
1. Study the free field theory L = 12 (∂ϕ)2 − 21 m2 ϕ2 at finite temperature and derive the
Bose-Einstein distribution.
Solution:
One way to derive the Bose-Einstein distribution is to show that average quantities are
computed with the correct probability distribution.11
We first compute the thermodynamic partition function for the relativistic free Bose gas,
given by:
Z Rβ R 3 1 1
Z= Dφ e− 0 dτ d x LE , LE = ∂µ φ ∂µ φ + M 2 φ2 .
φ(0,~
x)=φ(β,~
x) 2 2
Here we have dropped an overall constant, which corresponds to an additive constant in the
argument of the exponent. The trace is the sum of eigenvalues of the operator −∂E2 + M 2 ,
which is diagonalized in Matsubara frequency / momentum space. So we have12
∞ Z 3
1 X d kV 2πn
ln Z = − 3
ln(ωn2 + ~k 2 + M 2 ) , ωn = .
2 n = −∞ (2π) β
While this sum diverges, its derivative is finite and can be computed:
∞
0
X 2x 1 2π
f (x) = = −π − + .
n=1
n2 +x 2 x 1 − e−2πx
dx f 0 (x), we find:
R
Using f (x) =
∞
X
ln(n2 + x2 ) = 2πx + 2 ln 1 − e−2πx .
n = −∞
11
We follow p. 394 of “Lattice Gauge Theories - An Introduction,” 3rd Ed, by Heinz J. Rothe.
12
We have implicitly put the system into a box to quantize the spatial momenta and then taken the
P R d3 k
continuum limit k → (2π) 3 V , where V is the volume of the box. This treatment misses the formation of
151
Therefore we can compute the sum over Matsubara frequencies to obtain, up to an irrelevant
additive constant,
d3 k
Z q
−βε(~k)
ln Z = −V ln 1 − e , ε(~k) ≡ ~k 2 + M 2 .
(2π)3
∂
The average energy is given by hEi = − ∂β ln Z, so that the average energy per unit volume
is given by:
d3 k
Z
E ∂
−βε(~k)
= ln 1 − e
V ∂β (2π)3
d3 k ~
Z
1
= 3
ε(k) ~
.
(2π) e βε( k) −1
Thus the average energy is computed with the probability distribution
1
n(ε) =
e βε − 1
which is the Bose-Einstein distribution.
2. It probably does not surprise you that for fermionic fields the periodic boundary con-
dition (6) is replaced by an antiperiodic boundary condition ψ(~x, 0) = −ψ(~x, β) in order
to reproduce the results of chapter II.5. Prove this by looking at the simplest fermionic
functional integral. [Hint: The clearest exposition of this satisfying fact may be found in
appendix A of R. Dashen, B. Hasslacher, and A. Neveu, Phys. Rev. D12: 2443, 1975.]
Solution:
Consider the case of free field theory in (0 + 1)-spacetime dimensions (that is, the quan-
tum mechanics of a fermionic harmonic oscillator).
The Lagrangian is L = ψ̄(t)(i∂t − m)ψ(t). The system has two possible single-particle
energy states, ε0 and ε1 , corresponding to whether or not a fermion is present. The energy
difference ε1 − ε0 = m is fixed, while the energies themselves are determined only up to an
overall additive constant: ε0 = E0 − 12 m and ε1 = E0 + 21 m.
Any path integral representation of the partition function must reproduce this basic result.
152
The standard tricks give us the path integral representation
Z RT
Z= DψDψ̄ e i 0 dt ψ̄(t)(i∂t −m)ψ(t)
±BC
where the ±BC indicates periodic or antiperiodic boundary conditions, respectively: ψ(t +
T ) = ±ψ(t).
Z = C det(i∂t − m)
where C is a normalization factor that is determined by the overall additive constant in the
energy:
Z RT
−iH(m=0)T
tr(e )= DψDψ̄ e i 0 dt ψ̄(t)i∂t ψ(t) = C det(i∂t ) ≡ 2e−iE0 T
±BC
2e−iE0 T
=⇒ C = .
det(i∂t )
det(i∂t − m)
Z = 2e−iE0 T .
det(i∂t )
The determinant of an operator is the product of its eigenvalues. The eigenvalue equation
(i∂t − m)ϕn (t) = −ωn (m)ϕn (t) subject to ± boundary conditions implies the eigenvalues
(n = 0, ±1, ±2, ...):
2nπ
T
+m (periodic)
ωn (m) =
(2n+1)π
T
+m (antiperiodic)
Since ∞
det(i∂t − m) Y ωn (m)
=
det(i∂t ) n = −∞
ωn (0)
we have a choice between two infinite products. Since
∞
Y x πx
1+ = cos
n = −∞
2n + 1 2
we deduce that only antiperiodic boundary conditions will give us the correct answer:
∞ (2n+1)π ∞
det(i∂t − m) Y
T
+ m Y mT mT
= (2n+1)π
= 1+ = cos .
det(i∂t ) −BC n = −∞ n = −∞
(2n + 1)π 2
T
153
By the usual analytic continuation T = −iβ, we conclude that fermions at finite temperature
obey antiperiodic boundary conditions.
3. It is interesting to consider quantum field theory at finite density, as may occur in dense
astrophysical objects or in heavy ion collisions. (In the previous chapter we studied a system
of bosons at finite density and zero temperature.) In statistical mechanics we learned to
go from the partition function to the grand partition function Z = tr e−β(H−µN ) , where a
chemical potential µ is introduced for every conserved particle number N . For example, for
noninteracting relativistic fermions, the Lagrangian is modified to L = ψ̄(i 6 ∂ − m)ψ + µψ̄γ 0 ψ.
Note that finite density, as well as finite temperature, breaks Lorentz invariance. Develop
the subject of quantum field theory at finite density as far as you can.
Solution:
First let us clarify that finite density breaks Lorentz invariance because the term ψ̄γ µ ψ
is a Lorentz 4-vector, and we add to the Lagrangian only the µ = 0 component of that term,
ψ̄γ 0 ψ = ψ † ψ.
To work with a theory at nonzero temperature T ≡ β −1 , take the Minkowski theory and com-
pactify the imaginary time direction: t → iτ , 0 ≤ τ ≤ β. The gamma matrices of SO(3, 1)
satisfy the relation {γ µ , γ ν } = 2η µν , so that (γ 0 )2 = +1 while (γ 1 )2 = (γ 2 )2 = (γ 3 )2 = −1.
The gamma matrices of SO(4) satisfy {γEµ , γEν } = 2δ µν , so that (γE0 )2 = (γE1 )2 = (γE2 )2 =
(γE3 )2 = +1, where the subscript E stands for “Euclidean.” We can therefore write
γE0 = γ 0 , γEi = iγ i
for the Euclidean gamma matrices13 . The Dirac operator in Euclidean space is
The thermal two-point function S(x) ≡ hψ(x)ψ̄(0)i is defined as the inverse of the operator
6 ∂E + m + µγE0 :
6 ∂E + m + µγE0 S(x) = Iδ 4 (x)
where I is the R4 × 4 identity matrix in Dirac spinor space. Multiply this equation by e−ipµ xµ
and integrate β d4 x to get
(+i 6 pE + m + µγE0 )S(p) = I
13
We continue to label the spacetime indices by 0,1,2,3 in Euclidean spacetime, instead of the often-used
convention 1,2,3,4.
154
where we have defined the Fourier transform
Z
S(p) ≡ d4 x e−ipµ xµ S(x) .
β
where we have defined p0 = ωn ≡ βπ (2n + 1) and summed instead of integrated because the
τ ≡ x0 direction is compact. Everything proceeds as in the case for which µ = 0, except with
the replacement p0 → p0 + iµ.
Solution:
~ ·M
a) Adding −H ~ to (1) gives the free energy
~2−H
G = (aM ~ ·M
~ )V + O[(M
~ 2 )2 ]
Stripping off the δij and integrating over space gives the susceptibility
√
∞
e− a |~x |
Z Z √ 1 1
χ= 3
dx = dr r e− a r
= ∼
4π|~x | 0 a T − Tc
Again we find γ = 1.
155
V.4 Superconductivity
1. Vary (1) to obtain the equation for A and determine the London penetration length more
carefully.
1 b
(1) F = Fij2 + |Di ϕ|2 + a|ϕ|2 + |ϕ|4 + ...
4 2
Solution:
The equation of motion found from setting δF = F[A + δA] − F[A] = 0 is therefore
m2γ †
−δij ∂~ + ∂i ∂j + † ϕ ϕ δij Aj = −2ie(ϕ† ∂i ϕ − ϕ∂i ϕ† )
2
hϕ ϕi
where we have defined the mass of the photon inside the superconductor via m2γ ≡ 2(2e)2 hϕ† ϕi.
The extra factor of two is there because the mass term for a real vector boson is F =
1 2
m AA.
2 A i i
Choose the gauge ∂j Aj = 0. When ϕ assumes its vacuum expectation value, we have
(−∂~ 2 + m2γ )Ai = 0 . The physically relevant solution is the decaying exponential. Sub-
~
stitute the form Ai (~x) = C e−k·~x to get
We find an exponential decay Ai (r) ∼ e−mγ r . The London penetration depth is defined as
the length scale over which the field penetrates into the material, meaning
1 1 1
`L = =p = √
mγ 2(2e)2 v 2 2 2 ev
p
where v ≡ hϕ† ϕi .
156
2. Determine the coherence length more carefully.
Solution:
Use the free energy from problem V.4.1, but this time vary ϕ† to get an equation of mo-
tion for ϕ. The relevant part of the free energy is
m2γ † b
F = ∂i ϕ† ∂i ϕ + 2ieAi (ϕ† ∂i ϕ − ϕ∂i ϕ† ) + 2
ϕ ϕAi Ai + aϕ† ϕ + (ϕ† ϕ)2 .
2v 2
If we choose the gauge ∂i Ai = 0, then we can integrate by parts to move the derivative off of
ϕ† and onto ϕ. The variation δF = F[ϕ† + δϕ† ] − F [ϕ† ] immediately gives the equation of
motion for ϕ:
m2γ 2
~ 2
−∂ + 4ie A ~
~·∂+ ~ + a + bϕ ϕ ϕ = 0 .
A †
2v 2
Deep inside the superconductor (or more accurately, at depths much greater than `L derived
in the previous problem), we can set the vector potential to zero. Assuming that the self-
interaction term governed by the parameter b is negligible, we obtain
(−∂~ 2 + a)ϕ = 0
Recall that for this entire discussion to work, we must have a < 0 (the analog of the “negative
mass squared” instability from chapter IV.) We therefore determine an oscillating solution
ϕ(r) ∼ e±ir/`ϕ , where
1
`ϕ = √
−a
is the coherence length, defined as the characteristic length scale over which ϕ varies.
157
V.6 Solitons
3. Compute the mass of the kink by the brute force method and check the result from the
Bogomol’nyi inequality.
Solution:
Solution:
158
2. In the vortex, study the length scales characterizing the variation of the fields ϕ and
A. Estimate the mass of the vortex.
Solution:
159
Therefore, the mass of the soliton is
( 0 2 )
Z 2
v 1 g (r)
M = d2 x 2 [g(r) − 1]2 f (r)2 + r2 [f 0 (r)]2 + λv 2 r2 [f (r)2 − 1]2 +
r 2 ev
Since [v 2 ] = +1, each term in the square brackets must have mass dimension +2. The reader
may verify using the formulas collected previously that this is true. Note that the combina-
tion ξ ≡ evr is dimensionless.
Let us now switch to dimensionless variables. Define the functions of dimensionless vari-
ables F (ξ) ≡ f (r) and G(ξ) ≡ g(r), so that f 0 (r) = ev F 0 (ξ) and g 0 (r) = ev G0 (ξ). Also
define the dimensionless coupling β ≡ λ/e2 . In terms of these, the mass of the vortex is given
by
Z evR
2 1 2 2 2 0 2 2 2 1 0 2
M = 2πv dξ [G(ξ) − 1] F (ξ) + ξ [F (ξ)] + β ξ [F (ξ) − 1] + [G (ξ)]
eva ξ 2
where a is the size of the vortex and R is the size of the spatial region. (We are using the
notation of the discussion about the Kosterlitz-Thouless phase transition on p. 310.)
We now vary with respect to the functions F (ξ) and G(ξ) to find the equations of motion.
Minimizing with respect to F (ξ) gives
d d 1
ξ F (ξ) = βξ[F (ξ) − 1] + [G(ξ) − 1]2 F (ξ)
dξ dξ ξ
and minimizing with respect to G(ξ) gives
d 1 d
ξ G(ξ) = 2[G(ξ) − 1]F (ξ)2 .
dξ ξ dξ
Recall that the boundary conditions on F (ξ) and G(ξ) are that they go to 1 for ξ → ∞.
With this in mind, consider the ansatze
F (ξ) = 1 − e−αF ξ and G(ξ) = 1 − e−αG ξ .
Since ξ = evr, the quantities `ϕ ≡ 1/(evαF ) and `A ≡ 1/(evαG ) characterize the length
scales over which the fields ϕ and A, respectively, vary. With the above forms, the equations
of motion become
1 −2αG ξ +αF ξ ξ→∞
αF (1 − ξαF ) = −βξ + e [e − 1] αF2 = β (provided that αF < 2αG )
ξ →
1 ξ→∞
(αG − )αG = 2[1 − e−αF ξ ]2 2
αG =2
ξ →
Thus, recalling that β = λ/e2 , we obtain the lengths
1 1
`ϕ = √ and `A = √ .
λv 2 ev
160
Note that the gauge coupling e has dropped out of the length scale of ϕ, and that the length
scale of A does not depend on the quartic coupling λ. We could have arrived at this conclu-
sion simply on the basis of dimensional analysis: both length scales are determined by ∼ 1/v,
but v has mass dimension +1/2 and so another power of mass dimension +1/2 is needed √ for
each field. Since λ has mass dimension +1 and is associated with ϕ, we guess `ϕ ∼ 1/( λ v).
Since e has mass dimension +1/2 and is associated with A, we guess `A ∼ 1/(ev).
Note also that the condition αF < 2αG implies λ < 8e2 or λ < 32πα, where α ≡ e2 /(4π) as
usual. This implies that the configuration makes sense only if the quartic coupling for the
scalar field is less than ∼ 100α.
√
As for√the mass of the vortex, we take the above forms for F (ξ) and G(ξ) with αF = β and
αG = 2 , and perform the integral. As discussed on p. 310, we need to cut off the integral
at the low end by the size a of the vortex. On the other hand, the boundary conditions
discussed previously allow us to take R → ∞. Asking Mathematica to perform the integral,
we obtain:
" √ √ 2! #
1 ( β + 2 2 ) 1
M = 2πv 2 ln + ln √ √ − γ + + O(eva)
eva 16( β + 2 ) 2
where γ ≈ 0.577. Perhaps a more useful quantity to compute is the mass difference between
a vortex of size a0 and one of size a:
a
0 2
M (a ) − M (a) = 2πv ln 0
a
which shows the amount of energy required to increase the size of the vortex from a size a
to a size a0 > a. In any case, the mass of the vortex is estimated to be of order M ∼ 2πv 2 .
Recall the remark on p. 309 that the mass of the topological monopole in (3 + 1) di-
mensions comes out to be of order M ∼ MW /α, where α ≡ e2 /(4π). In our case, we have
MW ≡ `−1 2 2 2
A ∼ ev =⇒ v ∼ MW /e and thus in (2 + 1) dimensions we find M ∼ MW /e ,
2 2
1
which is dimensionally correct since e now has mass dimension + 2 .
3. Consider the vortex configuration in which ϕ(r → ∞) → ve iνθ with ν an integer. Calcu-
late the magnetic flux. Show that the magnetic flux coming out of an antivortex (for which
ν = −1) is opposite to the magnetic flux coming out of a vortex.
Solution:
ϕ∞ = v e inθ
161
where n is an integer. The gauge field at infinity is
†
−i ϕ∞ ∂i ϕ∞ −i v(v in∂i θ) n
A∞i = 2
= 2
= ∂i θ
e |ϕ∞ | e v e
Everything
R 2 proceedsH iprecisely as in the text, except multiplied by the integer n. The flux
Φ ≡ d x F12 = dx Ai is
2π
Φ=n = nΦ0
e
where Φ0 ≡ 2π e
is the fundamental unit of flux in equation (2) on p. 307. As discussed, a
vortex corresponds to n = +1, while an antivortex corresponds to n = −1. Immediately we
have
Φvortex = Φ0 = −Φantivortex .
6. Display explicitly the map S 2 → S 2 , which wraps one sphere around the other twice.
Verify that this map corresponds to a magnetic monopole with magnetic charge 2.
Solution:
~
An arbitrary element g of SO(3) can be parametrized as g(~x) = e ~x·T , where the matrices
0 0 0 0 0 1 0 −1 0
1
T ≡ 0 0 −1 , T 2 ≡ 0 0 0 , T 3 ≡ 1 0 0
0 1 0 −1 0 0 0 0 0
satisfy [T a , T b ] = εabc T c (with ε123 ≡ +1) and thereby generate SO(3), and the coordinates
satisfy ~x ·~x = 1 and thereby manifestly parameterize the surface of a two-sphere. The number
n is an integer whose significance we will derive shortly. The function g(~x) thereby consti-
tutes a map S 2 → S 2 .
We now compute the winding number of this configuration, namely the quantity
Z
1
νg ≡ − tr[(g −1 dg)3 ]
8π R3
162
~
where we are using forms notation. Using the notation g = e~x·T , we have g −1 dg = d~x · T~ . We
therefore compute the integral:
Z Z
−1 3
tr[(g dg) ] = dxa dxb dxc tr(T a T b T c )
3 3
R
ZR
= dxa dxb dxc 21 tr([T a , T b ]T c )
3
ZR
= dxa dxb dxc 12 εabd tr(T d T c )
3
ZR
= dxa dxb dxc 21 εabd (−2δ dc ) ← [check explicitly from the above T a ]
RZ3
Therefore we have
Z 2π Z π
1
νg = dφ dθ εabc εAB xa ∂A xb ∂B xc .
8π 0 0
Using our expressions for xa given above, we compute the partial derivatives:
Next we expand out the epsilon tensor εabc corresponding to the SO(3) algebra directions:
εabc xa ∂A xb ∂B xc = x1 ∂A x2 ∂B x3 + x3 ∂A x1 ∂B x2 + x2 ∂A x3 ∂B x1
− x3 ∂A x2 ∂B x1 − x1 ∂A x3 ∂B x2 − x2 ∂A x1 ∂B x3
163
Now we compute:
εabc xa ∂θ xb ∂φ xc
= 0 + [cos θ][cos θ cos(nφ)][n sin θ cos(nφ)] + [sin θ sin(nφ)][− sin θ][−n sin θ sin(nφ)]
− [cos θ][cos θ sin(nφ)][−n sin θ sin(nφ)] − [sin θ cos(nφ)][− sin θ][n sin θ cos(nφ)] − 0
= n sin θ .
Thus the integer n is itself the winding number for the gauge field configuration. If we desire
the map that wraps one sphere around the other twice, then we choose n = 2 in the definition
of the SO(3) algebra coordinates xa .
Next we have to compute the magnetic charge of the monopole described by this field con-
figuration. Equation (7) on p. 308 provides us with the definition of the gauge-invariant
electromagnetic field:
1
Fij ≡ F~ij · ~x − εabc xa (Di x)b (Dj x)c
e
where (Di x) = ∂i x + e ε Ai x and Ai = ∂i xb as before, and we have used xa xa = 1. (We
a a abc b c b
are labeling the field variables by xa – as we have been throughout this problem – to empha-
size that the parameters on the group manifold should be thought of as coordinates like any
other. We use i, j, k to denote spatial R3 , and A, B to denote the S 2 of spatial infinity.)
The gauge-covariant field strength is Fija = ∂i Aaj − ∂j Aa + e εabc Abi Acj = e εabc ∂i xb ∂j xc , so
F~ij · ~x = e εabc xa ∂i xb ∂j xc .
Next we have to contract two covariant derivatives with an epsilon. This will require using
εabc εade = δ bd δ ce − δ be δ cd , which follows from SO(3) invariance and checking with b = d = 2
and c = e = 3. We compute:
164
xa εabc (Di x)b (Dj x)c = xa [εabc ∂i xb ∂j xc + e∂i xb (εabc εcpq )∂j xp xq
+ e(εabc εbef )∂i xe xf ∂j xc + e2 (εabc εbef εcpq )∂i xe xf ∂j xp xq ]
= (1 + e2 )εabc xa ∂i xb ∂j xc
1 2π
Z Z π
g=− dφ dθ εabc εAB xa ∂A xb ∂B xc
2e 0 0
1
= − (8πn)
2e
4π
=− n.
e
As explained in the book’s solution for V.7.5, the fundamental unit of charge is actually 21 e,
so that we have Dirac’s quantization condition
2π
g=− n
(e/2)
as expected. As in V.7.5, the sign just corresponds to which we call the monopole and which
one we call the antimonopole.
165
9. Discuss the dyon solution. Work it out in the BPS limit. [B. Julia and A. Zee, Phys. Rev.
D11: 2227, 1975.]
Solution:
We follow M. K. Prasad and C. M. Sommerfield, “Exact classical solution for the ’t Hooft
monopole and the Julia-Zee dyon,” Phys. Rev. Lett. Vol. 35, No. 12, 22 Sep 1975.
r2 K 00 = K(K 2 − 1) + K(H 2 − J 2 )
r2 J 00 = 2JK 2
λ
r2 H 00 = 2HK 2 + 2 (H 2 − C 2 r2 )H
e
where C ≡ µe/λ1/2 . The boundary conditions are:
166
The equations of motion admit the solution
Cr
K=
sinh(Cr)
J = sinh γ [Cr coth(Cr) − 1]
H = cosh γ [Cr coth(Cr) − 1] .
Given the electromagnetic field (see equation (7) on p. 308 of the book)
1 a a 1 abc a
Fµν = φ Fµν − ε φ (Dµ φ)b (Dν φ)c
|φ| e|φ|3
we find the electric field Ei ≡ −F0i to be
1 − K2
xi d J(r) xi
Ei = = sinh γ .
r dr er r er2
This can be used to find the electric charge:
8π ∞ JK 2
Z Z
3 4π
Q = d x ∂i Ei = dr = sinh γ .
e 0 r e
The case γ = 0 corresponds to the monopole solution discussed in the text. See both refer-
ences for further discussion.
10. Verify explicitly that the magnetic monopole is rotation invariant in spite of appear-
~ are covariant
ances. By this is meant that all physical gauge invariant quantities such as B
b
under rotation. Gauge variant quantities such as Ai can and do vary under rotation. Write
down the generators of rotation.
Solution:
First a small but important typographical error: All physical gauge invariant quantities
are invariant under rotation, not covariant. Covariant means transforms as a vector, while
invariant means does not transform. Gauge transformations (rotations within the gauge
group) are conceptually nothing but changing coordinates (in field space), and physics must
be invariant under a change of coordinates.
Since the symbols δ ab and εabc are invariant under SO(3), the electromagnetic field Fµν
is manifestly gauge invariant. We will now verify this explicitly using infinitesimal SO(3)
transformations.
and satisfy the commutation relations [T a , T b ] = εabc T c , where ε123 ≡ +1. (By the way, this
immediately shows that εabc is invariant under SO(3), since the structure constants of a Lie
algebra are invariant symbols of the corresponding group.)
To show explicitly that Fµν is invariant under SO(3), we need to show explicitly that δ ab and
εabc are invariant under simultaneous SO(3) transformations on all of their indices, since the
first term of Fµν is constructed from δ ab and the second term is constructed from εabc (and
the magnitude |ϕ| is also constructed from δ ab ).
since (T a )bc = −(T a )cb , as can be seen above by the explicit construction of the matrices
{T a }3a = 1 .
168
Specialize to the case (abc) = (123), since all elements of εabc are defined in terms of ε123 .
We have
ε123 → ε123 + iθ~n · T~ 1d εd23 + T~ 2d ε1d3 + T~ 3d ε12d
= ε123 + iθ~n · T~ 11 ε123 + T~ 22 ε123 + T~ 33 ε123
= ε123 X
since all diagonal entries of the T a are equal to zero. This completes the calculation.
Solution:
The Lagrangian is
L = c1 εij ai ∂0 aj + c2 εij a0 ∂i aj .
2. By thinking about mass dimension, convince yourself that the Chern-Simons term domi-
nates the Maxwell term at long distances. This is one reason that relativistic field theorists
find anyon fluids so appealing. As long as they are interested only in long distance physics
they can ignore the Maxwell term and play with a relativistic theory (see exercise VI.1.1).
Note that this picks out (2+1)-dimensional spacetime as special. In (3+1)-dimensional space-
time the generalization of the Chern-Simons term εµνλσ fµν fλσ has the same mass dimension
as the Maxwell term f 2 . In (4+1)-dimensional space the term ερµνλσ aρ fµν fλσ is less impor-
tant at long distances than the Maxwell term f 2 .
Solution:
169
There are two standard ways to assign mass dimensions to gauge fields, which of course
yield equivalent results. If we normalize the gauge field by L = − 4g12 f 2 with the covariant
derivative Dµ ψ = (∂µ − iaµ )ψ (for some matter field ψ), then [a] = [∂] = +1 for any choice of
the spacetime dimension d. The Chern-Simons term εµνρ aµ ∂ν aρ therefore has mass dimen-
sion [a∂a] = 2[a] + [∂] = 3. Meanwhile, the field strength fµν = ∂µ aν − ∂ν aµ has dimension
[f ] = [∂a] = 2, so the Maxwell term f 2 has dimension 4. Therefore [a∂a] < [f 2 ].
Instead suppose we normalize the gauge field by L = − 41 f 2 with the covariant derivative
Dµ ψ = (∂µ − igaµ )ψ. Since the Lagrangian
R d always has dimension [L] = d in d spacetime
dimensions to make the action S = d x L dimensionless, we have [f 2 ] = 2[∂a] = 2(1+[a]) =
d =⇒ [a] = (d − 2)/2. The Chern-Simons term has dimension [a∂a] = 2[a] + 1 = d − 2 + 1 =
d − 1. Since the Maxwell term f 2 has the same mass dimension as that of the Lagrangian L,
namely d, we see again that [a∂a] < [f 2 ].
Solution:
First a typo: The problem meant to contrast this with the previous exercise, VI.1.2.
4. Consider L = γaε∂a − (1/4g 2 )f 2 . Calculate the propagator and show that the gauge
boson is massive. Some physicists puzzled by fractional statistics have reasoned that since
in the presence of the Maxwell term the gauge boson is massive and hence short ranged, it
can’t possibly generate fractional statistics, which is manifestly an infinite ranged interaction.
(No matter how far apart the two particles we are interchanging are, the wave function still
acquires a phase.) The resolution is that the information is in fact propagated over an infinite
range by a q = 0 pole associated with a gauge degree of freedom. This apparent paradox
is intimately connected with the puzzlement many physicists felt when they first heard of
the Aharonov-Bohm effect. How can a particle in a region with no magnetic field whatso-
ever and arbitrarily far form the magnetic flux know about the existence of the magnetic flux?
170
Solution:
The Lagrangian is
1
L=− 2
fµν f µν + γεµνρ aµ ∂ν aρ
4g
Since fµν f µν µ 2 ν
= 2a (−ηµν ∂ + ∂µ ∂ν )a , we have
1 µ 1
L= a (ηµν ∂ − ∂µ ∂ν ) + 2γεµλν ∂ aν
2 λ
2 g2
The propagator G is the inverse of the thing in brackets, so
1
2
(ηµν ∂ − ∂µ ∂ν ) + 2γεµλν ∂ Gνρ (x) = δµρ δ 3 (x)
2 λ
g
Multiplying by e−ikx , integrating d3 x and moving derivatives around via integration by
R
parts gives
1 2 λ
− 2 (ηµν k − kµ kν ) + 2iγεµλν k G̃νρ (k) = δµρ
g
If we try to solve for G̃νρ now we will run into trouble. To see why, use Lorentz invariance
to write
G̃νρ (k) = A(k 2 )η νρ + B(k 2 )k ν k ρ + C(k 2 )ενρσ kσ .
Putting this into the equation for G̃νρ gives
A C 2
2 ρ ρ
− 2iγC k δµ − k kµ + 2iγA + 2 k εµρλ kλ = −δµρ
g 2 g
which has no solution. (Note that B dropped out.) This is just the usual gauge fixing
problem, which we can resolve by adding a term 2ξ1 ηµν to the equation for G̃νρ :
1 1
− 2 (ηµν k − kµ kν ) + 2iγεµλν k + ηµν G̃νρ (k) = δµρ
2 λ
g 2ξ
Now plugging in the form G̃ = Aη + Bkk + Cεk results in the conditions
g2 A g2
2 2 A B C 2
k − − 2iγCk = −1 , 2 + − 2iγC = 0 , 2iγA + 2 k − =0.
2ξ g 2 g 2ξ g 2ξ
Solving these gives
g2
−(k 2 − 2ξ
)g 2 ξ→∞ −g 2
A=
(k 2 − g2 2
2ξ
) − (2γg 2 )2 k 2 → k2 − (2γg 2 )2
−g 2
2ξ 2ξ
ξ→∞
1 − 2 k 2 − (2γg 2 )2
B=
(k 2 − g2
2ξ
)2 − (2γg 2 )2 k 2 g → k2
2iγg 4 ξ→∞ 2iγg 4
C=
(k 2 − g2
2ξ
)2 − (2γg 2 )2 k 2 → k2 (k2 − (2γg 2 )2 ) .
171
The propagator for the gauge boson has a simple pole at k 2 = (2γg 2 )2 which persists in the
limit ξ → ∞. Therefore, the gauge boson has a mass m = 2γg 2 . The propagator also has a
simple pole at k 2 = 0; notice that the terms involving the pole at k 2 = m2 are independent
of the gauge parameter ξ when taking ξ → ∞, while the pole at k 2 = 0 comes with a factor
of ξ in the numerator in the coefficient B. This is what is meant by the k 2 = 0 pole being
associated with a gauge degree of freedom. Even though the physical gauge boson has mass
m = 2γg 2 6= 0, there is still a long-range interaction from the nontrivial topology of the gauge
theory.
5. Show that θ = 1/4γ. There is a somewhat tricky factor of 2. So if you are off by a
factor of 2, don’t despair. Try again. [X. G. Wen and A. Zee, J. de Physique, 50: 1623,
1989.]
Solution:
As instructed at the top of p. 318, one approach is to take the Lagrangian La = γεµνλ aµ ∂ν aλ +
aµ j µ and integrate out aµ to get the nonlocal Hopf Lagrangian. In anticipation of fixing trans-
verse gauge ∂µ aµ = 0, add a term − 2ξ1 (∂a)2 to the Lagrangian to get, after integration by
parts and ignoring a total derivative, the gauge-fixed Lagrangian Lξa = 21 aµ (G−1 )µλ aλ , where
(G−1 )µλ = 2γεµνλ ∂ν + 1ξ ∂ µ ∂ λ . The effective action SHopf is computed by performing the
gauge-fixed path integral over aµ :
Z Z
i d3 x Lξa
R
SHopf = −i ln Da e = d3 x d3 y j λ (x)Gλρ (x − y)j ρ (y)
where G̃(k) ≡ d3 x e−ikx G(x). We now need to solve for G̃ in Lorentz-index space. By
R
Plugging this form into the above equation for G̃(k) and using εµνλ ελρσ = δρµ δσν − δσµ δρν results
in the three conditions:
1
2iγBk 2 = 1 , 2iγAεµνρ kν = 0 , 2iγB + Ck 2 = 0 .
ξ
172
Therefore A = 0 and
1 ξ
B= 2
, C=− 2 2 .
2iγk (k )
The momentum-space propagator is therefore
ξ→0
1 ξ 1
G̃λρ (k) = 2
ελρσ k σ − 2 2 kλ kρ → 2
ελρσ k σ .
2iγk (k ) 2iγk
R
Now the goal is to evaluate SHopf = d3 x d3 y j λ (x)Gλρ (x − y)j ρ (y) with
d3 k e+ik(x−y)
Z
Gλρ (x − y) = ελρσ k σ
(2π)3 2iγk 2
and the current j λ (x) = j1λ (x) + j2λ (x) describing the exchange of two particles. Let particle
1 with charge q1 be sitting at rest at the origin: j1λ (x) = q1 δ0λ δ 2 (~x ). We want particle 2 with
charge q2 to move in a half-circle of radius R around particle 1: j2λ (y) = q2 uλ (y)δ 2 (~y − ~r(t)),
where
π
t ≡ y 0 [0, ], ~r(t) = R(x̂ cos(ωt) + ŷ sin(ωt)), uλ (y) = (1, ~y˙ ).
ω
The speed ~y˙ can be taken arbitrarily small to ignore the relativistic factor (1 − |~y˙ |2 )−1/2 .
d3 k e+ik(x−y)
Z Z Z
cross 3 3
q1 δ0λ δ 2 (~x ) ελρσ k σ q2 uλ (y)δ 2 (~y − ~r(t))
SHopf = dx dy
(2π)3 2iγk 2
π/ω 0 0 ~
d3 k e+ik (x −t)+ik·~r(t)
Z Z Z
0
= q1 q2 dx dt ε0ij k j ui (t, ~r(t))
0
3 0 2 ~
(2π) 2iγ[(k ) − |k | ] 2
R∞ 0 +ik0 x0
Since
R 0 −∞ dx e = 2πδ(k 0 ), we can perform the integral over x0 and then the integral
dk , leaving behind the above expression with k 0 set to zero:
π/ω ~
d2 k e+ik·~r(t)
Z Z
cross
SHopf = q1 q2 dt ε0ij k j ṙi (t) .
0 (2π)2 −2iγ|~k |2
The arbitrary 2-momentum k j can be written in polar coordinates as ~k = k(x̂ cos ϕ + ŷ sin ϕ),
so that ~k · ~r(t) = kR[cos(ωt) cos ϕ + sin(ωt) sin ϕ] = kR cos(ωt − ϕ). Also, since ~r˙ (t) =
ωR(−x̂ sin(ωt) + ŷ cos(ωt)) we have
ε0ij ṙi (t)k j = ωkR[− sin(ωt) sin ϕ + (−1) cos(ωt) sin ϕ] = −ωkR cos(ωt − ϕ)
173
R R 2π R∞
where the (−1) comes from ε021 = −ε012 = −1. In polar coordinates, d2 k = 0 dϕ 0 dk k,
so all of the factors of k = |~k | cancel out in the integrand. We have
dϕ ∞ dk ikR cos(ωt−ϕ)
Z π/ω Z 2π Z
cross q1 q2
SHopf = (−ωR) dt e cos(ωt − ϕ)
−2iγ 0 0 2π 0 2π
Z ∞
q1 q2 ωR π/ω
Z Z 2π
dϕ 1 dk ∂ ikR cos(ωt−ϕ)
= dt e
iγ 0 0 2π iR 0 2π ∂k
Z π/ω Z 2π
q1 q2 ω dϕ 1 ikR cos(ωt−ϕ) ∞
= dt e k=0
(−1)2γ 0 0 2π 2π
Z π/ω Z 2π
q1 q 2 ω dϕ ikR cos(ωt−ϕ)
=− dt lim e − 1
4πγ 0 0 2π k→∞
| {z }
=0
q1 q 2 ω π q1 q2
=− − =+
4πγ ω 4γ
Therefore the statistics parameter θ = SHopf for two particles of charge q1 = q2 = +1 is
θ = 1/(4γ).
6. Find the nonabelian version of the Chern-Simons term ada. [Hint: As in chapter IV.6 it
might be easier to use differential forms.]
Solution:
The new feature in a non-abelian theory is that a3 6= 0. The Lagrangian will be of the
form L = γ tr(ada + βa3 ) where the coefficient β should be fixed by gauge invariance (up to
a total derivative). For a gauge transformation a → a + δa, we get δL = tr[(2 da + 3β a2 )δa],
where we have dropped a total derivative. A non-abelian gauge transformation δa = [θ, a]−dθ
with matrix-valued infinitesimal parameter θ(x) implies δL = γ tr[(2 − 3β)a2 dθ], where again
we have dropped total derivatives. Demanding that δL = 0 implies β = 23 , so the non-abelian
Chern-Simons Lagrangian is
L = γ tr ada + 32 a3 .
7. Using the canonical formalism of chapter I.8 show that the Chern-Simons Lagrangian
leads to the Hamiltonian H = 0.
Solution:
174
where in the second equality we have integrated by parts, dropped a total derivative and
defined B ≡ 21 ε0ij Fij . Notice that a0 has no time derivatives and is therefore constant; it acts
as a Lagrange multiplier to enforce the constraint B = 0:
∂L
= 0 =⇒ B = 0 .
∂a0
The Lagrangian is now
L = γε0ij ȧi aj
which is linear in time derivatives. The conjugate momenta to ai are
∂L
πi ≡ = γε0ij aj
∂ ȧi
and therefore
H ≡ π i ȧi − L = 0 .
8. Evaluate (6).
d3 p
Z
1 1
tr γν γµ (6)
(2π)3 6 p +6q − m 6 p − m
Solution:
First we need gamma matrix identities in 3 dimensions. As pointed out in problem II.1.12, in
(2+1) dimensions we can take the gamma matrices to be γ 0 = σ 3 , γ 1 = iσ 2 and γ 2 = −iσ 1 ,
where σ i are the usual 2-by-2 Pauli matrices. Let aµ , bµ , cµ and dµ be arbitrary 4-vectors.
From the defining relation of the Clifford algebra {γ µ , γ ν } = 2η µν I, where here I is the 2 × 2
identity matrix, and using Lorentz invariance we get the following gamma matrix relations:
tr(6 a6 b) = 2(ab) where (ab) ≡ a · b ≡ η µν aµ bν
tr(6 a6 b6 c) = −2i εµνρ aµ bν cρ
tr(6 a6 b6 c6 d) = 2 ((ab)(cd) + (ad)(bc) − (ac)(bd))
Note that unlike in four dimensions, the trace of three gamma matrices is not zero.
Next notice that this integral is linearly divergent, so we will have define it properly such
that gauge invariance is preserved (recall problem IV.7.3). We will adapt the method on p.
273 for d = 2 + 1 spacetime dimensions. For d = 3 + 1 dimensions, we had
Z
4 µ Pµ
d p[f (p + a) − f (p)] = lim ia f (P )2π 3 P 3
P →∞ P
where the expression on the right-hand side is to be averaged over the sphere at infinity, for
example replacing P µ P ν → 14 η µν P 2 . Adapting this to d = 2 + 1 dimensions we have
Z
3 µ Pµ
d p[f (p + a) − f (p)] = lim ia f (P )4πP 2 .
P →∞ P
175
Replacing f (p) with the integrand
1 1
f (p) ≡ tr γ ν
µν
γµ
6 p+ 6 q − m 6 p − m
d3 p µν
R
we find that the integral I µν (a) ≡ f (p
+ a) satisfies the relation
(2π)3
µν µν i µ Pµ
I (a) = I (0) + lim 2
a f µν (P )P 2 .
P →∞ 2π P
In this expression, f µν (P ) simplifies to
tr(γ ν 6 P γ µ 6 P ) 2P µ P ν − η µν P 2
µν ν 1 µ 1
f (P ) = tr γ γ = =
6P 6P (P 2 )2 (P 2 )2
2( 1 η µν P 2 ) − η µν P 2 1 µν
= 3 2 2
=− η .
(P ) 3P 2
Therefore, our integral satisfies
µν µν i λ Pλ
I (a) = I (0) − lim a η µν .
P →∞ 6π 2 P
The strategy now is to evaluate I µν (0) with a regulator in place, then to choose the vector
aµ such that I µν (a) is gauge invariant, meaning qµ I µν (a) ≡ 0 and qν I µν (a) ≡ 0.
Move the gamma matrices into the numerator using 1/(6 p − m) = (6 p + m)/(p2 − m2 ) to
get
d3 p N µν
Z
µν
I (0) = , N µν = tr [γ µ (6 p + m)γ ν (6 p + 6 q + m)] .
(2π)3 [p2 − m2 ][(p + q)2 − m2 ]
Using the gamma matrix identities and tracing, we find
N µν = 2[pµ (p + q)ν + pν (p + q)µ − η µν p · (p + q) − im eµνλ qλ + m2 η µν ] .
The denominator can be simplified using the Feynman trick
Z 1
1 1
= dx
AB 0 [xA + (1 − x)B]2
with A = p2 − m2 and B = (p + q)2 − m2 . After some arithmetic, this leads to
Z 1
1 1
2 2 2 2
= dx 2
[p − m ][(p + q) − m ] 0 (` + D)2
where ` ≡ p + (1 − x)q and D ≡ x(1 − x)q 2 − m2 . The denominator depends on ` only as `2 ,
so we can shift the integration variable from p to `, ignore terms linear in ` in the numerator,
and make the replacement `µ `ν → 31 η µν `2 . This leads to
µν 1 µν 2 µν 2 µ ν µνλ 2 µν
N = 2 − η ` + x(1 − x)(η q − 2q q ) − im ε qλ + m η .
3
176
R R
Now rotate to Euclidean momentum: d3 ` = i d3 `E and `2 = −`2E . Also define DE ≡
−D = x(1 − x)qE2 + m2 but leave q as Minkowski in the numerator. In total, we have
1
+ 13 η µν `2E + K µν
Z Z
µν i
I (0) = 3 dx d3 `E
4π 0 (`2E + DE )2
where
K µν ≡ x(1 − x)(η µν q 2 − 2q µ q ν ) − im εµνλ qλ + m2 η µν
does not depend on `E . The relevant integrals over `E are
π2
Z
1
d3 `E 2 = 1/2
(`E + DE )2 DE
and
`2E
Z
3 1/2
d `E 2 2
= 2πΛ − 3π 2 DE
(`E + DE )
p
where we have regulated the second integral by imposing a cutoff Λ ≡ ( qE2 )max . We now
have I µν (0) =
2Λ µν µν 1
Z Z 1
i +1/2 µν 2 µ ν −1/2
{ η −η dx DE + (η q −2q q ) dx x(1−x)DE
4π 3π 2 0 0
Z 1
−1/2
+ (−im εµνλ qλ + m2 η µν ) dx DE } .
0
p
Now we need the integrals over x. In terms of β ≡ 2m/ qE2 , the relevant integrals are:
Z 1
m
+1/2
dx DE [(1 + β 2 ) cot−1 β + β]
=
2β
Z0 1
−1/2 β
dx DE = cot−1 β
m
Z0 1
−1/2 β
dx x(1 − x) DE = [(1 − β 2 ) cot−1 β + β]
0 8m
We therefore have
µν i 2Λ µν µν β 2 −1 µ ν µνλ −1
I (0) = η +η R− [(1 − β ) cot β + β]q q − iε qλ β cot β
4π 3π 2 4m
where
m β
R≡− [(1 + β 2 ) cot−1 β + β] + q 2 [(1 − β 2 ) cot−1 β + β] + mβ cot−1 β .
2β 8m
The term R better simplify in such a way that q µ q ν and q 2 η µν appear only in the combination
q µ q ν − q 2 η µν if the integral is to be gauge invariant. Indeed, this will be the case. Since
177
qE2 = 4m2 /β 2 = −q 2 , we have q 2 = −4m2 /β 2 so
2
m −1 4m β
2
R = − [(1 + β ) cot β + β] − [(1 − β 2 ) cot−1 β + β] + mβ cot−1 β
2β β2 8m
m m
= − [(1 + β 2 ) cot−1 β + β] − [(1 − β 2 ) cot−1 β + β] + mβ cot−1 β
2β 2β
m
= − [(1 + β ) cot β + β + (1 − β 2 ) cot−1 β + β − 2β 2 cot−1 β]
2 −1
2β
m
= − [2 cot−1 β + 2β − 2β 2 cot−1 β]
2β
m
= − [(1 − β 2 ) cot−1 β + β]
β
Therefore we have
2
µν i 2Λ µν β 2 −1 µ ν µν 4m µνλ −1
I (0) = η − [(1 − β ) cot β + β][q q + η ] − iε qλ β cot β .
4π 3π 2 4m β2
Since β 2 = −4m2 /q 2 , we have
µν i 2Λ µν β 2 −1 µ ν µν 2 µνλ −1
I (0) = η − [(1 − β ) cot β + β][q q − η q ] − iε qλ β cot β .
4π 3π 2 4m
The finite part of I µν (0) therefore satisfies qµ I µν (0) = 0 and qν I µν (0) = 0. As discussed
previously, we want
µν µν i λ Pλ
I (a) = I (0) − lim a η µν
P →∞ 6π 2 P
to be gauge invariant. We therefore choose aµ such that
µ Pµ Λ
lim a =
P →∞ P π
which makes
µν i β 2 −1 µ ν µν 2 µνλ −1
I (a) = − [(1 − β ) cot β + β][q q − η q ] + iε qλ β cot β
4π 4m
p
satisfy qµ I µν (a) = 0 and qν I µν (a) = 0. Finally, let β = iα with α ≡ 2m/ q 2 to write the
integral in terms of the Minkowskian momentum q µ . Since
−1 1 α+1
cot (iα) = −i ln
2 α−1
we have
2 −1 1 2 α+1
β[(1 − β ) cot β + β] = α (1 + α ) ln −α
2 α−1
and
−1 1 α+1
β cot β = α ln .
2 α−1
p
Finally, in terms of the variable α ≡ 2m/ q 2 , the gauge invariant integral is
µν α 1 2 α+1 µ ν µν 2
µνλ α+1
I (a) = −i (1 + α ) ln − 2α q q − η q + iε qλ ln .
8π 4m α−1 α−1
178
VI.2 Quantum Hall Fluids
1. To define filling factor precisely, we have to discuss the quantum Hall system on a
sphere rather than on a plane. Put a magnetic monopole of strength G (which accord-
ing to Dirac can be only a half-integer or an integer) at the center of a unit sphere. The flux
through the sphere is equal to Nφ = 2G. Show that the single electron energy is given by
E` = ( 12 ~ωc )[`(`+1)−G2 ]/G with the Landau levels corresponding to ` = G, G+1, G+2, ...,
and that the degeneracy of the `th level is 2` + 1. With L Landau levels filled with non-
interacting electrons (ν = L) show that Nφ = ν −1 Ne − S, where the topological quantity S
is known as the shift.
Solution:
We follow X. G. Wen and A. Zee, “Shift and Spin Vector: New Topological Quantum Num-
bers for the Hall Fluids,” Phys. Rev. Lett., Vol. 69, No. 6, 10 Aug 1992.
The point of this problem is to recognize that since a curved space has a connection 1-
form ω (see p. 443), it is possible to write down a Chern-Simons-type interaction between ω
and the gauge potential aµ of the Hall fluid, which we remind the reader is defined as the po-
1 µνλ
tential for the conserved electromagnetic current: J µ = 2π ε ∂ν aλ . Thus to the Lagrangian
of equation (3) on p. 325, we add the interaction term
s
Ls = s ωµ J µ = ωµ εµνλ ∂ν aλ
2π
where s is a real number. For example, on the sphere the connection 1-form is ω ab =
−εab cos θ dϕ, where ε12 = +1. (See p. 444 in the text.)
Augmenting the discussion on pp. 326-327 with this new term, we find the electromagnetic
current
µ ∂Leff 1 µνλ
JEM =− =+ ε ∂ν (Aλ − sωλ )
∂Aµ 2πk
and the spin current
∂Leff s µνλ
Jsµ = = ε ∂ν (−Aλ + sωλ ) .
∂ωµ 2πk
The zero component of each R 2current is a number density of each type. Along
R 2 with the
0 0
number of electrons Ne = d xJEMR and the number of spin quanta Ns = d xJs , define
1 1
R
the number of flux quanta Nφ ≡ 2π F and the number of curvature quanta NR ≡ 2π R.
ij 2 ij 2
Upon recognizing ε ∂i Aj d x = dA = F and ε ∂i ωj d x = dω = R, integrating the µ = 0
components of the currents implies
Ne 1 1 −s Nφ
= 2 .
Ns k −s s NR
Inverting this matrix equation shows that
Nφ = kNe + sNR .
179
In the text it was shown that the filling factor is ν = 1/k (p. 327), so we have obtained
Nφ = ν −1 Ne − S, with a shift
Z
R
S = −sNR = −s .
2π
1
R
The Gauss-Bonnet theorem says that 2π R = 2(1 − g), where g is the genus of the manifold
over which we integrate. In particular, the sphere has genus zero so the shift is S = −2s.
2. For a challenge, derive the effective field theory for Hall fluids with filling factor ν = m/k
with k an odd integer, such as ν = 25 . [Hint:
Pm You have to introduce m gauge potentials aIλ
µ µνλ
and generalize (2) to J = (1/2π)ε ∂ν I=1 aIλ . The effective theory turns out to be
m m
1 X X
L= aI KIJ ε∂aJ + aIµ j̃ Iµ + ...
4π I,J = 1 I =1
Solution:
For this problem and the next, we follow J. Fröhlich and A. Zee, “Large Scale Physics
of the Quantum Hall Fluid,” Nucl. Phys. B364 (1991) 517-540.
For notational convenience we sometimes use differential forms and work with the Lagrangian
volume-form L ≡ 4πL d3 x. The gauge potential aµ for the conserved electromagnetic current
1 ∗
J µ is defined by the equation J = 2π da, where ∗ denotes the Hodge-star operator that takes
1 µνλ
p-forms to (3 − p)-forms (see IV.4.2). The Chern-Simons Lagrangian L = 4π ε aµ ∂ν aλ is
written L = a da. (We set the coefficient of the Chern-Simons theory that results from the
long-distance physics of the non-interacting Hall fluid to 1.)
Let I = 1, 2, ..., m label the m levels. For each conserved electromagnetic current
PmJI , we in-
1 ∗
troduce a gauge potential aI defined by JI = 2π daI . The new Lagrangian is L = I = 1 aI daI .
Now turn on interactions between the electrons, which couple the different Landau lev-
els. As discussed in the text, the resulting long-distance theory will be described by another
180
Chern-Simons Lagrangian. The interactions should only involve the total electromagnetic
current !
m m
X 1 ∗ X
J= JI = daI
I =1
2π I =1
Denote the current of quasiparticles (the “vortex current”) in the I th Landau level by jIµ .
This couples to the Chern-Simons gauge potential of each level as
m
X
L1 = aIµ jIµ
I =1
by definition. (“Here we define the quasiparticles as the entities that couple to the gauge
potential...” above equation (5) on p. 326.)
Now subject the system to additional external electromagnetic fields described by the vector
potential Aµ . This couples to the Chern-Simons gauge potentials through the electromagnetic
current interaction for each level:
m m m
X µ
X 1 µνλ X 1 µνλ
L2 = JI Aµ = ε ∂ν aIλ Aµ = − ε aIλ ∂ν Aµ
I =1 I =1
2π I =1
2π
181
The effective current j̃Iµ = jIµ − 2π1 µνλ
ε ∂ν Aλ is called the “reduced vorticity current.” To
understand the long-distance physics described by L, we integrate out the gauge fields aI to
obtain a matrix-valued version of the Hopf Lagrangian of equation (6) on p. 326:
m
X
−1 µνλ 1
Leff = π j̃Iµ (K )IJ ε ∂ν 2
j̃Jλ .
I,J = 1
∂
with 1 − p/(1 + mp) on the diagonals and −p/(1 + mp) everywhere else. This is the field
theory whose consequences we will study in the next problem.
3. For the Lagrangian in (13), derive the analogs of (8), (9), and (11).
µ 1 µνλ
(8) Jem = ε ∂ ν Aλ
4πk
1
(9) L= Aµ j µ
k
θ 1
(11) =
π k
Solution:
182
Varying this with respect to A gives the expectation value of the total electromagnetic cur-
rent: !
m
µ 1 X
hJEM i = (K −1 )IJ εµνλ ∂ν Aλ .
2π I,J = 1
As explained on p. 327 of the text, the µ = 0 component of this equation implies that the
filling factor is
m
X m 1 2 3
ν= (K −1 )IJ = = , , , ...
I,J = 1
1 + mp 1 + p 1 + 2p 1 + 3p
This is what replaces k −1 in equation (8). Thus for p = 2 we have derived the theory of
quantum Hall fluids with filling factors
1 2 3
ν=
, , , ...
3 5 7
which realizes the claim of problem VI.2.2.
µ
From the spatial components of hJEM i ∝ εµνλ ∂ν Aλ , we also learn that an electric field
in thePy-direction produces a current in the x-direction with a proportionality constant
σH = m I,J = 1 (K
−1
)IJ = ν, known as the Hall conductance.
Next we want to compute the electric charge of each quasiparticle. To do this, consider
the coupling of the quasiparticle currents jI to the applied gauge field A:
m
X
LAj
eff = Aµ (K −1 )IJ jJµ ≡ Aµ JEM
µ
I,J = 1
where we have identified the total electromagnetic current induced by a quasiparticle cur-
rents jIµ . (All of these currents are actually current densities, but as is common we are being
sloppy with the language.)
Recall from earlier that the total electromagnetic current is comprised of a sum of m in-
dividually conserved electromagnetic currents for each Landau level:
m m
µ
X X 1 µνλ
JEM = JIµ = ε ∂ν aIλ .
I =1 I =1
2π
183
Since we have already computed the matrix inverse of K, this can be inverted immediately
to obtain m
X
JIµ = (K −1 )IJ jJµ .
J =1
The µ = 0 component of this equation determines the charge density, which we can integrate
to obtain the total electric charge induced in the I th Landau level:
Z m
X
qI ≡ d2 xJI0 = (K −1 )IJ ΦJ
J =1
R
where we have defined the “vorticities” ΦI = d2 x jI0 . In this language, qI is the electric
charge bound to the vorticity ΦI in the I th Landau band. Using the explicit form of K −1 ,
the charge is
m
p p X
qI = 1− ΦI − ΦJ
1 + mp 1 + mp J = 1 , J6=I
m
p X
= ΦI − ΦJ .
1 + mp J = 1
As in the case of just one filled Landau level, this interaction describes a generalized version
of the Aharonov-Bohm effect. An excitation of the system with vorticity vector (Φ1 , ..., Φm )
results in a statistics phase
m
θ X m
= ΦI (K −1 )IJ ΦJ =
π I,J = 1 1 + mp
where in the last line we have used ΦI = 1 for I = 1, ..., m. This is the generalization of
equation (11).
184
VI.4 The σ Models as Effective Field Theories
1. Show that the vacuum expectation value of (σ, ~π ) can indeed point in any direction with-
out changing the physics. At first sight, this statement seems strange since, by virtue of
its γ5 coupling to the nucleon, the pion is a pseudoscalar field and cannot have a vacuum
expectation value without breaking parity. But (σ, ~π ) are just Greek letters. Show that by a
suitable transformation of the nucleon field parity is conserved, as it should be in the strong
interaction.
Solution:
Before proceeding to the solution, it is useful to review the sigma model for nucleons and
pions using slightly different notation from the text.
This problem is about transformation properties under the global SU (2)L ⊗SU (2)R symmetry
whose diagonal piece SU (2)I ⊂ SU (2)L ⊗ SU (2)R is Heisenberg’s isospin. To remove possible
sources of confusion about the Dirac spinor indices and to unclutter the notation as little
as possible, let us write the Dirac field as two-component spinors: ψ = (χ, χ̄† )T . (Review
Appendix E now if you are unfamiliar with this notation.) The above Lagrangian can be
rewritten as
L = −g χa εab Πbċ χ̄ċ + h.c.
where we have defined the matrix
4
X 1 µ 1 σ + iπ3 i(π1 − iπ2 )
Πbċ ≡ πµ (τ )bċ = .
2 2 i(π1 + iπ2 ) σ − iπ3 bċ
µ=1
Also, τ 4 ≡ I is the 2 × 2 identity matrix, and π4 ≡ σ is the fourth meson field. (Review
Appendix B for the local isomorphism SU (2) ⊗ SU (2) ' SO(4), which necessitates the in-
clusion of the fourth meson field.)
The factors of 21 give the correctly normalized kinetic term L = tr(ΠΠ† ) = 12 [(∂σ)2 +(∂π3 )2 ]+
∂π + ∂π − , where π ± ≡ √12 (π1 ∓ iπ2 ).
Under SU (2)L , we have χa → (Lχ)a = Lab χb = χb Lab = χb (LT )b a and Πbċ → Lb c Πcċ ,
where L is a 2 × 2 unitary matrix with determinant 1. Since ~σ T ε = −ε~σ , we have LT ε =
−εL† . So SU (2)L acts as χεΠ → χεL† LΠ = χεΠ, or in other words the Lagrangian is
185
invariant under SU (2)L . Under SU (2)R , we have χ̄ċ → Rċ ė χ̄ė and Πbċ → Πbė (R−1 )ė ċ , so
Πχ̄ → ΠR−1 Rχ̄ = Πχ̄, so the Lagrangian is invariant under SU (2)R . This is what is meant
by stating that the Lagrangian is invariant under SU (2)L ⊗ SU (2)R .
Now it is clear that the vacuum alignment may be chosen to point in an arbitrary direction in
SO(4) ' SU (2)L ⊗ SU (2)R , since we can always perform SU (2)L ⊗ SU (2)R transformations
on the nucleon fields to leave the physics invariant. In other words, the vacuum expectation
values hπµ i just need to satisfy the constraint
3
X 3
X
hπµ i2 = hσi2 + hπi i2 ≡ v 2
µ=0 i=1
with v some constant with dimensions of mass. The text chooses hσi = v and hπi i = 0, so that
upon spontaneous symmetry breaking the Lagrangian contains the term L = gv χχ̄ + h.c. =
gv ψ̄L ψR + h.c. = gv ψ̄ψ, which implies a mass M = gv for the nucleons. But any other
choice of vacuum alignment can be transformed into this form using the SU (2)L ⊗ SU (2)R
transformation hΠi → LhΠiR−1 , as long as we also redefine the nucleon fields by χ → Lχ
and χ̄ → Rχ̄.
So upon the field redefinitions χ → −τx χ and χ̄ → τz χ̄, we recover the mass term L =
gv χχ̄ + h.c. = gv ψ̄L ψR + h.c. = gv ψ̄ψ, as we must by SO(4) ' SU (2)L ⊗ SU (2)R invariance.
186
VI.5 Ferromagnets and Antiferromagnets
1. Work out the two branches of the spin wave spectrum in the ferromagnetic case, paying
particular attention to the polarization.
Solution:
For this problem and the next we follow X. G. Wen and A. Zee, “Spin Waves and Topological
Terms in the Mean-Field Theory of Two-Dimensional Ferromagnets and Antiferromagnets,”
Phys. Rev. Lett., Vol. 61 No. 8, 22 Aug 1988.
where h(~k) ≈ 2|J|a2 k 2 for small k ≡ |~k | (and J < 0). Setting the determinant of the matrix
to zero gives the quadratic equation
r
2 g2 2 ~ ∓ g2 g2 16
ω ± ω − g h(k) = 0 =⇒ ω (k) = ∓ + 1 + 2 h(k) .
2 4 4 g
We have chosen the + root in the quadratic equation to keep only positive-frequency solutions.
For small k, we have:
r
+ g2 g2 16 g2
ω =+ + 1 + 2 h(k) = + + O(k 2 )
4 4 g 2
2 2
r
g g 16
ω− = − + 1 + 2 h(k) ≈ +2h(k) = +4|J|a2 k 2
4 4 g
so that there is a high frequency branch ω + = g 2 /2 + O(k 2 ) and a low-frequency branch
ω − ∝ k 2 , which is the typical dispersion relation for a nonrelativistic particle.
Now let us work out the polarizations for each branch. Plugging in ω + ≈ 21 g 2 , dropping
all terms of order k 2 and dividing through by −g 2 /4, the equations of motion become
x
1 +i ξ
=0.
−i 1 ξy
This is satisfied by (ξ x , ξ y ) ∝ (1, +i). Plugging in ω − ≈ 4|J|a2 k 2 and h(k) ≈ 2|J|a2 k 2 ,
dropping terms of higher order in k, and dividing through by 2|J|a2 k 2 , the equations of
motion become x
1 −i ξ
=0.
+i 1 ξy
This is satisfied by (ξ x , ξ y ) ∝ (1, −i).
14
The text uses the notation δni ≡ ξi . We choose to change notation in order to distinguish linearizing
about the ground state from the arbitrary variation ~ni → ~ni + δ~ni used to obtain the equations of motion.
But of course you are free to use whichever notation you like.
187
2. Verify that in the antiferromagnetic case the Berry’s phase term merely changes the spin
wave velocity and does not affect the spectrum qualitatively as in the ferromagnetic case.
As given on p. 346, linearizing around the ground state ~ni (t) = (−1)i êz + ξ~i (t) implies15
2
! !
− ωg2 + f (~k) − 21 iω x ~
ξ (k, ω)
1
+ 2 iω ω2 ~ ~
− g2 + f (k + Q) y ~
ξ (k + Q,~ ω) = 0
where f (~k) = 4J[2 + 2µ = 1 cos(kµ a)] and J > 0. Note that since Qµ = πa , and cos(x + π) =
P
where k ≡ |~k|.
Returning to the equation of motion, setting the determinant of the matrix equal to zero
results in a quadratic equation for ω 2 :
~ − 1 g4ω2 = 0 .
[ω 2 − g 2 f (~k)][ω 2 − g 2 f (~k + Q)]
4
This has the solutions:
v
4f (~k)f (~k + Q)
~
u
g2
~ + 1 g2
f (~k) + f (~k + Q)
2
u
ω± = 1±u 1 − i2 .
2 4
h
~ ~ ~
t
1 2
f (k) + f (k + Q) + 4 g
2 1
ω+ ≈ g 2 (16J + g 2 ) + O(k 2 )
4
2 2 2
2 32J a g
ω− ≈ k2 .
16J + 14 g 2
188
The question now is to track the effect of the Berry’s phase term. Without this term (namely
the ±iω terms in the equations of motion), setting the determinant of the matrix to zero in
the equations of motion implies
[ω 2 − g 2 f (~k)][ω 2 − g 2 f (~k + Q)]
~ =0
which is still a quadratic equation for ω 2 . The solutions are ω 2 = g 2 f (~k) ≈ g 2 (16J − 2Ja2 k 2 )
and ω 2 = g 2 f (~k + Q)
~ ≈ g 2 (2Ja2 k 2 ). Again, there is a high-frequency branch
2
ωhigh = 16g 2 J + O(k 2 )
and a low-frequency branch
2
ωlow = 2Ja2 g 2 k 2
which is a linear dispersion relation with spin-wave velocity
c = 2Ja2 g 2 .
We see that the spectrum is qualitatively the same, and the Berry’s phase only changes the
particular values of the parameters in the dispersion relations.
Here we fill in theR steps absent from the text in order to derive the equations of motion.
The action is S = dt L with Lagrangian
N
X † 1 X
L= i zi ∂t zi + 2 ∂t~ni · ∂t~ni − J ~ni · ~nj
i=1
2g
hiji
where ~ni = zi†~σ zi is the Pauli-Hopf map introduced in the text, and the sum over nearest
neighbors can be written explicitly as
N N d
X 1 XX X
= δi, j+êµ + δi, j−êµ .
2 i=1 j =1 µ=1
hiji
The number of lattice sites is N , and the number of lattice spatial dimensions is d. (We
specialize to d = 2.) The vector êµ is a unit vector pointing in the µth direction, and the
kronecker deltas are to be interpreted according to the example
N
X
δi, j+êµ f (~xi ) = f (~xj+êµ ) ≡ f (~xj + a êµ )
i=1
189
given in the text, we have
N
Z " N X d
#
X 1 1 2 X
δS = dt δ~ni · − ~ni × ∂t~ni − 2 ∂t ~ni − J δi, j+êµ + δi, j−êµ ~nj .
i=1
2 g j =1 µ=1
When deriving local equations of motion by setting δS = 0, we must assume that the
variations δ~ni (t) are arbitrary functions of i and t – that is, that the sum over i and the
integral over t are unconstrained. But here the sum is constrained by ~ni · ~ni = 1. We need
to incorporate this constraint into the path integral:
Z N
Y R
Z= Dn δ(~ni · ~ni − 1) e i dt L
i=1
Z YN Z R R
ni ·~
ni −1)ν
= Dn Dν e i dt (~
ei dt L
i=1
Z PN
dt [ ]
R
ni ·~
ni −1)ν+L
= DnDν ei i = 1 (~
Now the sum over i is totally unconstrained, so the equation of motion is the term in {...}
set equal to zero:
N
1 1 X
~ni × ∂t~ni + 2 ∂t2~ni + Jij ~nj = 0
2 g j =1
with
d
X
Jij ≡ J δi, j+êµ + δi, j−êµ − 2νδij .
µ=1
First consider the ferromagnet (J < 0). The ground state of the ferromagnet is the state for
which all spins point in the positive z-direction. Linearize about this state as ~ni (t) = êz + ξ~i (t)
to obtain
~ni × ∂t~ni = (êz + ξ~i ) × ∂t (êz + ξ~i ) = êz × ∂t ξ~i + O(ξ 2 ) .
As mentioned in the text, the constraint ~n2i = (êz + ξ~i )2 = 1 + 2 êz · ξ~i + O(ξ 2 ) = 1 implies
êz · ξ~i = 0, so that the linearized deviation from the ground state occurs only in the (xy)-plane.
190
Since ξ~ lies purely in the (x, y)-plane, the equation of motion in the z-direction is:
N
X
Jij = 0 =⇒ ν = Jd .
j =1
and as usual d = 2.
Now that the equation of motion is linear in ξ~i (t), we may Fourier transform using the
convention
N Z ∞
~
X
~
f (k, ω) ≡ dt e−i(ωt+k·~xi ) fi (t) .
i=1 −∞
We have
N Z ∞
~
X
dt e−i(ωt+k·~xi ) ∂t2 ξ~i (t) = −ω 2 ξ(
~ ~k, ω)
i=1 −∞
and
N Z N N Z
" d
#
∞ ∞
−i(ωt+~k·~ −i(ωt+~k·~
X X X X
xi )
Jij ξ~j = xi )
δi, j+êµ + δi, j−êµ − 2d δij ξ~j (t)
dt e dt e J
i=1 −∞ j =1 i=1 −∞ µ=1
N Z ∞
" d
#
−i~k·(~ −i~k·(~ −i~k·~
X X
= dt e −iωt
J e xj +aêµ )
+e xj −aêµ )
− 2d e xj
ξ~j (t)
j =1 −∞ µ=1
" d
# N Z
∞
−i~k·êµ a +i~k·êµ a ~
X X
=J e +e − 2d dt e−i(ωt+k·~xj ) ξ~j (t)
µ=1 j =1 −∞
" d
#
X
= 2J cos(kµ a) − d ~ ~k, ω)
ξ( , kµ ≡ ~k · êµ .
µ=1
~ ~k, ω) .
= +iω êz × ξ(
191
In the first step, the first minus sign is from moving the time derivative to act on the
exponential using integration by parts. The Fourier transformed equation of motion now
takes the simple form
( " d #)
ω2 X
~ ~k, ω) = − 1 iω êz × ξ(
~ ~k, ω) .
− 2 + 2J cos(kµ a) − d ξ(
g µ=1
2
Using êz × êx = +êy and êz × êy = −êx and moving everything to the left-hand side yields
2
! !
− ωg2 + h(~k) − 21 iω ξ x (~k, ω)
2 =0
+ 12 iω − ωg2 + h(~k) ξ y (~k, ω)
where
" d
#
X
h(~k ) ≡ 2|J| d − cos(kµ a)
µ=1
" d #
X 1 2 2
≈ 2|J| d − 1 − kµ a
µ=1
2
1~ 2 2
= 2|J| d − d − |k | a
2
2 2
= +|J|a k
where
P2 k ≡ |~k |. On p. 346 of the book, the function h(k) is defined as h(k) ≡ 4J[2 −
2 2
µ = 1 cos(kµ a)] ≈ 2|J|a k , which is a factor of two larger than the one we have derived.
This can be traced back to when we included a factor of 21 in the explicit form of the sum
over nearest neighbors. Had we not included this 21 , we would be counting each nearest
neighbor twice rather than once. In the reference, Wen and Zee have implicitly normalized
the coupling J such that the nearest neighbor sum does not include the 21 . Thus we can take
the results in the text and rescale J → 12 J in order to match our treatment.
For the antiferromagnet (J > 0), the ground state is where the spins alternate in orien-
tation, so we expand around the ground state as ~ni (t) = (−1)i êz + ξ~i (t). Let Q
~ ≡ ( π , π ) be a
a a
i ~ xi
iQ·~ µ
2-dimensional vector. We can write (−1) = e , where ~xi = xi êµ is the coordinate of the
192
ith lattice site. Thus, for the antiferromagnet we have:
N N
" d #
~
X X X
Jij (−1)j = δi,j+êµ + δi,j−êµ − 2ν δij e iQ·~xj
J
j =1 j =1 µ=1
d
~ xi −aêµ ) ~ xi +aêµ ) ~
X
iQ·(~ iQ·(~
=J e +e − 2ν e iQ·~xi
µ=1
" d
#
~ xi
X
iQ·~
= 2e J cos(Qµ a) − ν
µ=1
" d
#
~ xi
X
iQ·~
= 2e J cos π − ν
µ=1
~
= 2 e iQ·~xi (−1) (Jd + ν) .
Therefore the z-component of the equation of motion for the antiferromagnet fixes the La-
grange multiplier to be
ν = −Jd
rather than +Jd. The matrix Jij for the antiferromagnet is therefore:
" d #
X
Jij = J δi,j+êµ + δi,j−êµ + 2d δij .
µ=1
To put this into a matrix notation, we need to shift the second equation by ~k → ~k + Q
~
~ ~k + 2Q,
and use ξ( ~ ~k, ω). We obtain
~ ω) = ξ(
2
! !
− ωg2 + f (~k) − 21 iω ξ x (~k, ω)
+ 12 iω
2
− ωg2 + f (~k + Q)
~ ξ y (~k + Q,~ ω) = 0 .
Note the argument ~k + Q~ in the lower-right component of the above matrix. This is acciden-
tally absent in the book, but present in equation (7) of the reference.
193
VI.6 Surface Growth and Field Theory
3. Field theory can often be cast into apparently rather different forms by a change of
1
variable. Show that by writing U = e 2 gh we can change the action (7) to
Z 2
2 D −1 ∂ −1 2
S = 2 d x dt U U −U ∇ U
g ∂t
a kind of nonlinear σ model.
Z 2
1 D ∂ 2 g 2
(7) S= d x dt − ∇ h − (∇h)
2 ∂t 2
Solution:
Let h = 2
g
ln U . Then ∇j ln U = U −1 ∇j U =⇒ ∇i ∇j ln U = −U −2 ∇i U ∇j U + U −1 ∇i ∇j U
implies
2
g 2 g 2
∇ h + (∇h)2 = − −U −2 ∇i U ∇i U + U −1 ∇2 U +
2
(U −1 ∇i U )(U −1 ∇i U )
2 g 2 g
2
= U −1 ∇2 U
g
Since ∂t ln U = U −1 ∂t U , we have
Z
2 2
dD x dt U −1 ∂t − ∇2 U .
S=
g
Solution:
194
Similarly, we can also write
Z
1 R D 0 † 0 †
(y, x) = i lim Dϕ†− Dϕ− e i d x {∂ ϕ~ − ∂ ϕ~ − +[V (x )−w]~ϕ− ϕ~ − } ϕ−1 (y)ϕ†−1 (x)
w−H m→0
where we have separated out the terms linear in V (x) to prepare for averaging over disorder.
− 1
R D
R d x V (x)2
We now carry out the average hO(V )i = N DV e 2g2 O(V ) using the formula
Z
1 1 −1
N DV e− 2 V ·M ·V +j·V = e 2 j·M ·j
ϕ†+ ϕ
where in our case j = i(~ ~ †− ϕ
~+ + ϕ ~ − ) and M = 1/g 2 . We arrive at the expression
1 1
(x, y) (y, x) =
z−H w−H
Z R D 0 0
− lim Dϕ†+ Dϕ+ Dϕ†− Dϕ− e i d x L(x ) ϕ+1 (x)ϕ†+1 (y)ϕ−1 (y)ϕ†−1 (x)
n,m→0
195
2. As another example from the literature on disorder, consider the following problem. Place
N points randomly in a D-dimensional Euclidean space of volume V . Denote the locations
of the points by ~xi (i = 1, ..., N ). Let
~
dD k e ik·~x
Z
f (~x ) = −
(2π)D ~k 2 + m2
Consider the N by N matrix Hij = f (~xi − ~xj ). Calculate ρ(E), the density of eigenvalues
of H as we average over the ensemble of matrices, in the limit N → ∞, V → ∞, with the
density of points ρ0 ≡ N/V held fixed. Hint: Use the replica method and arrive at the field
theory action
Z " n #
X Pn 2
S = dD x |∇ϕa |2 + m2 |ϕa |2 − ρ0 e−(1/z) a = 1 |ϕa |
a=1
This problem is not entirely trivial; if you need help consult M. Mèzard et al., Nucl. Phy.
B559: 689, 2000, cond-mat/9906135.
Solution:
Then to “average over disorder,” we average over the locations of the N points {~xi }N
i = 1:
Z "YN
#
dD xi
h(...)i = (...) .
i=1
V
Just as we find the propagator by taking derivatives of the partition function, to find the
operator tr[1/(z − H)] we will instead compute the function
Z "Y N
#Z
dD xi Pn PN †
ξN ≡ dnN ϕ† dnN ϕ e i a = 1 i,j = 1 ϕai (Hij −zδij )ϕaj
i=1
V
N
X
φa (~x ) ≡ δ D (~x − ~xi )ϕai
i=1
196
and their hermitian conjugates, then
Z
dD x dD y φ†a (~x )f (~x − ~y )φa (~y ) =
Z N
! N
!
X X
dD x dD y δ D (~x − ~xi )ϕ†ai f (~x − ~y ) δ D (~y − ~xj )ϕaj
i=1 j =1
N
X N
X
= ϕ†ai f (~xi − ~xj )ϕaj = ϕ†ai Hij ϕaj ,
i,j = 1 i,j = 1
so we can replace ij ϕ†ai Hij ϕaj with dD x dD y φ†a (~x)f (~x − ~y )φa (~y ). To make this a useful
P R
substitution, we need to be able to treat the variables R φa (~x ) and φ†a (~x ) as unconstrained
field variables. In other words, we want to integrate Dφ† Dφ as if φa (~x ) and φ†a (~x ) were
usual bosonic quantum fields that transform as a vector underPan O(n) symmetry, and we
can do so as long as we include the constraint that φa (~x) − N D
i = 1 δ (~x − ~xi )ϕai = 0 and
†
P N D †
φa (~x) − i = 1 δ (~x − ~xi )ϕai = 0. To implement this, insert the number 1 into the path
integral as follows:
Z " N
# Z Z
~ PN
~ δ (~x − ~x ) = Dφ Dµ† e i d x µ~ (~x)·[φ(~x)− i = 1 ϕ~ i δ (~x−~xi )]
†
X R D (D)
(n) ~ (D)
1 = Dφ δ φ(~x ) − ϕ i i
i=1
We chose the minus sign in this version of 1 for later convenience. If you are getting bogged
down in functional integral notation, then orient yourself with the familiar example from
ordinary 1-dimensional calculus:
Z ∞ Z ∞ Z ∞
dµ ±iµ(x−c)
1= dx δ(x − c) = dx e
−∞ −∞ −∞ 2π
where c is some real number. In the functional integrals, the factors of 2π are swept into the
definitions of the measures. Inserting both of these factors of 1 into the integral for ξN , we
arrive at
ξN =
Z "Y N
#Z "Z n
#
dD xi
Z X
Dφ† DφDµ† Dµ dnNϕ† dnNϕ exp i dD x dD y φ†a (~x )f (~x − ~y )φa (~y ) ×
i=1
V a=1
" n N Z n
( N
)#
XX † X X
exp −i z ϕai ϕai + dD x i µ†a (~x )[φa (~x) − ϕai δ D (~x − ~xi )] + h.c. .
a=1 i=1 a=1 i=1
197
RWeDcan now D perform the Gaussian integral over the ϕai and ϕ†ai variables. First note that
d x µa (~x )δ (~x − ~xi ) = µa (~xi ), and then define the “sources” Ja (~x ) ≡ −iµa (~x ). Then the
integral we have to perform is
Z ( n N )
XXh i
I ≡ dnN ϕ† dnN ϕ exp −iz ϕ†ai ϕai + i µ†a (~xi )ϕai − i ϕ†ai µa (~xi )
a=1 i=1
Z ( n X N h
)
X i
= nN
d † nN
ϕd ϕ exp −iz ϕ†ai ϕai + Ja† (~xi )ϕai + ϕ†ai Ja (~xi )
a=1 i=1
n Y
Y N Z n o
= dϕ†ai dϕai exp −(iz)ϕ†ai ϕai + Ja† (~xi )ϕai + ϕ†ai Ja (~xi ) .
a=1 i=1
Using Z
† ϕ+J † ϕ+ϕ† J 2πi + 1 J † J
dϕ† dϕ e−α ϕ = e α
α
with α = iz, we have
n Y N
Y 2π 1 †
I= exp Ja (~xi )Ja (~xi )
a=1 i=1
z iz
nN ( n N
)
2π i XX †
= exp − J (~xi )Ja (~xi )
z z a=1 i=1 a
nN ( n X N
)
2π i X
= exp − µ†a (~xi )µa (~xi ) .
z z a=1 i=1
ξN =
nN Z " Y N
#Z "Z n
#
2π dD xi † † D D
X
†
Dφ DφDµ Dµ exp i d x d y φa (~x )f (~x − ~y )φa (~y ) ×
z i=1
V a=1
( n N Z n
)
i XX † X
µ (~xi )µa (~xi ) + i dD x µ†a (~x )φa (~x ) − φ†a (~x )µa (~x )
exp − .
z a=1 i=1 a a=1
Now we can perform the Gaussian integral over the fields φa (~x ) and φ†a (~x ):
Z R D D Pn † R D Pn † †
Dφ† Dφ e i d x d y a = 1 φa (~x)f (~x−~y)φa (~y)+ d x a = 1 [Ja (~x)φa (~x)+φa (~x)Ja (~x)]
Ja† (~
Pn
d D x dD y x)f −1 (~
R
= (det f )−1 e+i a=1 x−~
y )Ja (~
y)
,
dD k 1
R
where again Ja (~x ) = −iµa (~x ). Since f (x) = − (2π)D k2 +m2
e ikx , the inverse of f (x) is
dD k 2
Z
−1
f (x) = − (k + m2 ) e ikx .
(2π)D
198
To convince yourself this is right, compute
Z
dD z f −1 (x − z)f (z − y)
dD k dD p 2
Z Z Z
2 2 1 +ikx −ipy
= (−1) (k + m ) 2 e e dD z e+i(p−k)z
(2π)D (2π)D p + m2
dD k dD p 2
Z Z
1
= D D
(k + m2 ) 2 2
e+ikx e−ipy (2π)D δ D (p − k)
(2π) (2π) p +m
Z D
d k +ik(x−y)
= e = δ D (x − y) . X
(2π)D
To summarize our progress so far, we have
nN Z " Y N D
#Z " n X N
#
2π d xi i X
ξN = Dµ† Dµ exp − µ†a (~xi )µa (~xi ) ×
z i=1
V z a=1 i=1
" Z n
#
X
exp − tr ln f + i dD x dD y µ†a (~x )f −1 (~x − ~y )µa (~y ) .
a=1
Given the explicit form of f −1 (~x ), the last term simplifies (suppress the index a for clarity):
dD k ~ 2
Z Z Z
D D † −1 D D † 2 i~k·(~
x−~
y)
d x d y µ (~x )f (~x − ~y )µ(~y ) = d x d y µ (~x ) − (|k | + m ) e µ(~y )
(2π)D
dD k †
Z Z
~ ~
= − d xd yD D
D
µ (~x ) e+ik·~x (|~k |2 + m2 ) e−ik·~y µ(~y )
(2π)
dD k †
Z Z h i
+i~k·~ 2 −i~k·~
= − dD x dD y µ (~
x ) e x
(−∇ 2
y + m )e y
µ(~y )
(2π)D
dD k +i~k·(~x−~y )
Z Z
= −(−1) 2 D
d xd yD
D
e µ† (~x )(−∇2y + m2 )µ(~y )
(2π)
Z
= − dD x dD y δ D (~x − ~y )µ† (~x )(−∇2y + m2 )µ(~y )
Z
= − dD x µ† (~x )(−∇2 + m2 )µ(~x )
Z h i
= − dD x +∇µ ~ † (~x ) · ∇µ(~
~ x ) + m2 µ† (~x )µ(~x ) .
Now this is starting to approach the desired result. At this point we have the field theory
Z
†
ξN = C Dµ† Dµ AN e−S0 [ µ , µ ]
199
the function n Z
dD x −(i/z) Pna = 1 µ†a (~x )µa (~x )
2π
A= e ,
z V
and the irrelevant overall constant
C = e−tr ln f .
We will now drop C since it will drop out of all correlation functions as do all overall constants
in the path integral. We will also take n → 0 in the prefactor (2π/z)n in A. Passing to a
grand canonical formulation of the disorder16 , we compute
∞ ∞
Z ! Z
X 1 †
X 1 −S0 [µ† , µ] †
ξ≡ N
ξN α = Dµ Dµ (Aα) N
e = Dµ† Dµ e Aα−S0 [µ , µ] .
N =0
N! N =0
N!
We therefore arrive at the result ξ = Dµ† Dµ e−S , where the full action S = S0 − αA is
R
Z " n n
!#
X α i X
S[µ† , µ] = dD x i ~ †a (~x )· ∇µ
∇µ ~ a (~x )+m2 µ†a (~x )µa (~x ) − exp − µ†a (~x )µa (~x ) .
a=1
V z a=1
The last thing to determine is the meaning of the parameter α. The average number of points
is17 N = αhAi, so that in the limit n → 0 we have N = α and hence α/V = ρ0 , the density
of points. We have arrived at the action
Z " n n
!#
X
~ † (~x )· ∇µ
~ a (~x )+m2 µ† (~x )µa (~x ) −ρ0 exp − i X
S[µ† , µ] = dD x i ∇µ a a µ†a (~x )µa (~x ) .
a=1
z a=1
200
VI.8 Renormalization Group Flow as a Natural Concept in High
Energy and Condensed Matter Physics
1. Show that the solution of dg/dt = −bg 3 + ... is given by
1 1
= + 8πbt + ...
α(t) α(0)
where we defined α(t) = g(t)2 /4π.
Solution:
g t
dg 0
Z Z
0 1 1 1 1 1
= −b dt =⇒ − − 2 = −b t =⇒ = 2 + 2bt
g0 g 03 0 2 g 2 g0 g 2 g0
Define g 2 ≡ 4πα and multiply by 4π to get
1 1
= + 8πb t .
α α0
2. In our discussion of the renormalization group, in λϕ4 or in QED, for the sake of simplicity
we assumed that the mass m of the particle is much smaller than µ and thus set m equal
to zero. But nothing in the renormalization group idea tells us that we can’t flow to a mass
scale below m. Indeed, in particle physics many orders of magnitude separate the top quark
mass mt from the up quark mass mu . We might want to study how the strong interaction
coupling flows from some mass scale far above mt down to some mass scale µ below mt but
still large compared to mu . As a crude approximation, people often set any mass m below
µ equal to zero and any m above µ to infinity (i.e., not contributing to the renormalization
group flow). In reality, as µ approaches m from above the particle starts to contribute less
and drops out as µ becomes much less than m. Taking either the λϕ4 theory or QED study
this so-called threshold effect.
Solution:
First consider QED. Equation (6) on p. 358 gives the renormalization group flow equation
for the QED coupling:
d 1 d
µ e(µ) = − e(µ)3 µ Π(µ2 ) + O(e5 )
dµ 2 dµ
The lowest order solution to this is
1 1 1 µ
2
= 2 − 2 ln .
e (µ) e (µ0 ) 6π µ0
Suppose we have n + 1 electrons in this theory, one having mass m and the others massless,
and take the initial condition to be at some superheavy mass scale M .
201
For m < µ < M , all n + 1 particles contribute to the vacuum polarization function, and
we have
1 1 n+1 M
2
= 2 + ln .
e (µ) e (M ) 6π 2 µ
For 0 < µ < m, the lowest order approximation is to use
1 1 n M
= + ln .
e2 (µ) e2 (M ) 6π 2 µ
Including the threshold correction at the scale m amounts to replacing the lowest order
expression with the following:
1 1 n m
= + ln
e2 (µ) e2 (m) 6π 2 µ
1 n+1 M n m
= + ln + ln
e2 (M ) 6π 2 m 6π 2 µ
1 n M 1 M
= 2 + 2 ln + 2 ln .
e (M ) 6π µ 6π m
We see that there is an extra term that depends explicitly on m. This result is usually written
in terms of the µ-dependent coupling for µ > m, which we now denote by18 eG (µ). With
−2
e−2 (M ) = eG (µ) − 6π2 ln µ , we can rewrite the threshold-corrected coupling for µ < m
n+1 M
as:
1 1 n+1 M n M 1 M
2
= 2 − 2
ln + 2 ln + 2 ln
e (µ) eG (µ) 6π µ 6π µ 6π m
1 1 m
= 2 − 2 ln .
eG (µ) 6π µ
The analysis for ϕ4 scalar field theory proceeds in the same way. Equation (5) on p. 357
along with appendix 1 in chapter III.1 (equation (14) on p. 168) gives the renormalization
group flow equation for λϕ4 theory:
d 1 d
µ λ(µ) = − λ(µ)2 µ Π(µ2 ) + O(λ3 )
dµ 16 dµ
where we have defined the function
1
Λ2
Z
1
Π(s) ≡ 2 dx ln
2π 0 m2 − x(1 − x)s
18
We use this notation to evoke the connection to threshold corrections in grand unified theories. See
chapter VII.6 in the main text, as well as S. Weinberg, “Effective Gauge Theories,” Phys. Lett. 91B, No. 1
(1980) and L. Hall, “Grand Unification of Effective Gauge Theories,” Nucl. Phys. B178 (1981) 75-124.
202
using a notation intentionally evocative of the vacuum polarization function in QED. [Here Λ
is an arbitrary upper cutoff on the momentum, and m is the mass of the quanta (“mesons”)
of the scalar field ϕ.]
The analysis proceeds in exactly the same fashion as for QED: take n + 1 mesons, one
with mass m and the others massless, and assume an initial condition at M m.
As before, we relabel the high-energy coupling as λG (µ) and rewrite the low-energy coupling
using λ−1 (M ) = λ−1 n+1 M
G (µ) − 16π 2 ln µ :
1 1 n+1 M n M 1 M
= − 2
ln + 2
ln + 2
ln
λ(µ) λG (µ) 16π µ 16π µ 16π m
1 1 m
= − 2
ln .
λG (µ) 16π µ
203
4. In S̃(h) only derivatives of the field h can appear and not the field itself. (Since the trans-
formation h(~x, t) → h(~x, t) + c with c a constant corresponds to a trivial shift of where we
measure the surface height from, the physics must be invariant under this transformation.)
Terms involving only one power of h cannot appear since they are all total divergences. Thus,
S̃(h) must start with terms quadratic in h. Verify that the S̃(h) given in (17) is indeed the
most general. A term proportional to (∇h)2 is also allowed by symmetries and is in fact
generated. However, such a term can be eliminated by transforming to a moving coordinate
frame h → h + ct.
Solution:
The action is supposed to be the most general compatible with invariance under the Galilean
transformation
h(~x, t) → h0 (~x, t) = h(~x + g ~u t, t) + ~u · ~x + 21 gu2 t
with ~u a constant velocity. In the solution for VI.8.3 it is shown that the combination
g ~ 2
∂t h − (∇h)
2
is invariant under the Galilean transformation, as is ∇2 h. Thus in general the action should
be a linear combination of these terms:
g ~ 2
X ≡ α ∂t h − (∇h) − β∇2 h .
2
As discussed in the problem, the action should also not involve terms linear in h since all
such
R D terms2 are total divergences. Thus to lowest order in h the action is proportional to
d x dt X , which is the form given in (17).
6. Calculate the h propagator to one loop order. Extract the coefficients of the ω 2 and
k 4 terms in a low frequency and wave number expansion of the inverse propagator and de-
termine α and β.
Solution:
We need the free-field propagator and the cubic vertex (all momenta point into the vertex):
1
=
2 4
2 2 2
= 1 1
) k2 3 2 2
) k3 1 3 3
) k1 2
3 2
204
Fortunately the quartic vertex does not contribute to this order. The only diagram we need
to compute is
Here µ and Λ are arbitrary lower and upper cutoffs, respectively, on the integral. The idea is
to integrate over an infinitesimal shell, so that the lower cutoff can be taken as µ = (1−δL)Λ.
The integral is now
Z Λ d
d K 1 Λd−2
= δL ≡ f (Λ, d) δL .
d
µ (2π) K
2
2d−1 π d/2 Γ( d2 )
Adding the 1-loop self-energy to the zeroth-order propagator, we obtain the desired coeffi-
cients
g2
α=1− f (Λ, d) δL
8
g 2 4d2 − d − 6
β =1− f (Λ, d) δL .
8 d(d + 2)
See M. Karder and A. Zee, “Matrix generalizations of some dynamic field theories,” Nucl.
Phys. B464 (1996) 449-462 for further discussion.
205
VII Grand Unification
VII.1 Quantizing Yang-Mills Theory and Lattice Gauge Theory
3. Consider a lattice gauge theory in (D + 1)-dimensional space with the lattice spacing a
in D-dimensional space and b in the extra dimension. Obtain the continuum D-dimensional
field theory in the limit a → 0 with b kept fixed.
Solution:
l k
Uk l
Ul i b
P Uj k
Ui j
i j
a
We have S(P ) =Re trUij Ujk Uk` U`i , where
1 1
Uij = e+iaAµ (x) , µ̂ = (xj − xi ) , x = (xi + xj )
a 2
1 1 a b
Ujk = e+ibAν (y) , ν̂ = (xk − xj ) , y = (xj + xk ) = x + µ̂ + ν̂
b 2 2 2
−iaAµ (x0 ) 0
Uk` = e , x = x + bν̂
0 a b
U`i = e−ibAν (y ) , y 0 = y − aµ̂ = x − µ̂ + ν̂
2 2
Since Ujk and U`i are inverses when evaluated at the point x + 2b ν̂, let Ujk (x + 2b ν̂) ≡ U and
U`i (x + 2b ν̂) = Ui`† (x + 2b ν̂) ≡ U † . Taylor expanding in powers of a, we have
1
• Uij = e+iaAµ (x) = I + iaAµ (x) − a2 Aµ (x)Aµ (x) + O(a3 )
2
a b a 1 a 2
• Ujk (x + µ̂ + ν̂) = U + ∂µ U + ∂µ ∂µ U + O(a3 )
2 2 2 2 2
−iaAµ (x0 ) 1 2
• Uk` = e = I − iaAµ (x ) − a Aµ (x0 )Aµ (x0 ) + O(a3 )
0
2
a b † a † 1 a 2
• U`i (x − µ̂ + ν̂) = U − ∂µ U + − ∂µ ∂µ U † + O(a3 )
2 2 2 2 2
206
For brevity suppress the direction µ and write Aµ (x) ≡ A and Aµ (x0 ) ≡ A0 . The trace of the
product of these four matrices (ignoring terms of O(a3 ) and higher) is
a2 2 a † a2 2 †
1 a 0 1 2 02 †
= tr I + iaA − a2 A2 U + ∂U + ∂ U I − iaA − a A U − ∂U + ∂ U
2 2 8 2 2 8
1 1 2 1 2 2
= tr U + a ∂U + iAU + a ∂ U + iA∂U − A U ×
2 2 4
† 1 † 0 † 1 2 1 2 † 0 † 02 †
U −a ∂U + iA U + a ∂ U + iA ∂U − A U
2 2 4
† 1 † † 1 † 0 †
= tr[ U U + a ∂U U + iAU U − U ∂U − iU A U ]
2 2
2 1 1 2 † 0 † 02 † 1 1 † 0 †
+ a tr[ U ∂ U + iA ∂U −A U − ∂U +iAU ∂U +iA U
2 4 2 2
1 1 2
+ ∂ U +iA∂U −A U U † ]
2
2 4
Since U U † = I, the first term is a constant and can be dropped. Also, trAU U † = trA = 0.
Using the cyclicity of the trace, we also have trU A0 U † = trA0 U † U = trA0 = 0.
207
Therefore the O(a) terms in tr Uij Ujk Uk` U`i are zero. Dropping the O(a0 ) additive constant,
we have:
tr Uij Ujk Uk` U`i
1
= a2 tr U ∂ 2 U † + ∂ 2 U U † − 2∂U ∂U †
8
1
+ i a2 tr U A0 ∂U † + A∂U U † − AU ∂U † − ∂U A0 U †
2
2 1 02 † 0 † 1 2 †
+ a tr − U A U + AU A U − A U U
2 2
1
= − a2 tr ∂U ∂U †
2
1 2
+ i a tr (∂U U † − U ∂U † )A + (∂U † U − U † ∂U )A0
2
1 2 2
− a tr A + A02 − 2 AU A0 U †
2
Consider a derivative operator defined as DU = ∂U + iAU − iU A0 . This is the covariant
derivative appropriate for a field U transforming as ∼ N ⊗ N̄ under SU (N ) gauge transfor-
mations. To clarify this remark, return to ordinary continuum field theory for a moment and
consider a matter field ψ ∼ N ⊗ N̄ of SU (N ). This has one lower index a = 1, ..., N and one
upper index ā = 1, ..., N , meaning that the components of ψ are ψa ā . The gauge-covariant
derivative of ψ is
N 2 −1
X h i
ā ā
(Dµ ψ)a = ∂µ ψa + i AIµ (TNI )ab ψb ā + AIµ (TN̄I )āb̄ ψa b̄
I =1
where TNI are the generators of the N -representation (“fundamental”) of SU (N ), and TN̄I are
the generators of the N̄ -representation (“anti-fundamental”) of SU (N ). (As usual, the index
I = 1, ..., N 2 − 1 counts the number of generators.) It is true in general that TN̄I = −(TNI )∗ .
Since the generators are hermitian, we have (TNI )∗ = (TNI )T , or in components [(TNI )ab ]∗ =
(TNI )b a . Therefore:
Suppressing the matrix indices in the usual way and defining the matrix-valued gauge field
Aµ ≡ I AIµ TNI , we have
P
Dµ ψ = ∂µ ψ + iAµ ψ − iψAµ
208
which matches our lattice covariant derivative DU = ∂U + iAU − iU A0 .
tr(DU )(DU )†
= tr ∂U ∂U † − i∂U (U † A − A0 U † ) + i(AU − U A0 )∂U † + (AU − U A0 )(U † A − A0 U † )
= tr ∂U ∂U † − i tr ∂U U † − U ∂U † A + ∂U † U − U † ∂U A0
+ tr A2 − 2 AU A0 U † + A02
Therefore tr Uij Ujk Uk` U`i = − 12 a2 tr[(DU )(DU )† ] for ijk` bounding a rectangular plaquette.
The full lattice gauge theory action
X a2 X
S= Re tr Uij Ujk Uk` U`i + 2 Re tr Uij Ujk Uk` U`i
P square
b P rectangle
where bN is the length of the lattice in the extra dimension. This is the action of N copies
of a D-dimensional continuous SU (N ) Yang-Mills theory with a Lorentz-scalar U transform-
ing under the (N ⊗ N̄ )-dimensional representation. For further discussion in the context of
Kaluza-Klein theory, see http://arxiv.org/pdf/hep-th/0104005v1.
4. Study the alternative limit b → 0 with a kept fixed so that you obtain a theory on a
spatial lattice but with continuous time.
5. Show that for lattice gauge theory the Wilson area law holds in the limit of strong
coupling. [Hint: Expand (20) in powers of f −2 .]
Z hY i
1 − 1
P
S(P )
hW (C)i = dU e 2f 2 P W (C) (20)
Z
209
Solution:
where β ≡ 2N/g 2 , and the function KP is defined as KP ≡ Uij Ujk Uk` U`i for a plaquette P
whose four corners are the lattice sites xi , xj , xk , x` as in the diagram below:
l k
Uk l
Ul i P Uj k
Ui j
i j
Now consider a large closed rectangular curve C on the lattice whose sides have lengths R
and T . Define the Wilson loop WC for the curve C as
210
To find the leading contribution in the limit of small β, we have to understand how to carry
out integrals over the link variables Uij . The rules are as follows20 :
Z
DU UAB = 0
Z
DU UAB UCD = 0
Z
1 D B
DU UAB (U † )CD = δ δ
N A C
Z
1
DU UAB 1
UAB 2
... UAB N
= εA A ...A εB1 B2 ...BN
1 2 N
N! 1 2 N
for any N -by-N link variable U . (The indices A, B, ... run from 1 to N .) To get the leading
nonzero contribution to hWC i, we need to pair up each Uij in WC with exactly one Uij† from
Q P 1 n
(P in Σ) n n! (βSP ) , where Σ is the surface of minimal area whose boundary is the curve
C. Thus the leading contribution from the product over plaquettes in Σ is when n = 1 in
the sum above, and the integral can be approximated to leading order as21
Z n∗
β
hWC i ≈ DU WC (trK1† )...(trKn† ∗ )
2N
where n∗ ≡ RT /a2 is the number of plaquettes that fit inside the region Σ, and a2 is the area
of a square plaquette with sides of length a:
P P P P
P P P P T
P P P P
To understand how to compute the integral, first consider the overly simplified but illustrative
case for which the curve C encloses only 1 plaquette, bounded by the lattice sites (i, j, k, `)
as in the diagram above. In this case, the Wilson loop is simply
WC = tr Uij Ujk Uk` U`i = (Uij )AB (Ujk )BC (Uk` )CD (U`i )DA
20
See p. 90 and p. 222 of the text “Lattice Gauge Theories - An Introduction, 2nd Ed.” by Heinz. J. Rothe
for more details.
21
The denominator Z contributes only a factor of 1 to this order. To the next order in β, we have to worry
about nontrivial contributions from Z.
211
and the only K that contributes is K1 = tr Uij Ujk Uk` U`i , so that
K1† = U`i† Uk`
† †
Ujk Uij† = (U`i† )EF (Uk`
† †
)FG (Ujk )GH (Uij† )HE .
The expectation value of the Wilson loop to leading order22 is
Z
β
hWC i ≈ DU (Uij )AB (Ujk )BC (Uk` )CD (U`i )DA (U`i† )EF (Uk` † †
)FG (Ujk )GH (Uij† )HE
2N
Z Z
β B † E C † H
= DUij (Uij )A (Uij )H DUjk (Ujk )B (Ujk )G ×
2N
Z Z
D † G A † F
DUk` (Uk` )C (Uk` )F × DU`i (U`i )D (U`i )E
β 1 E B 1 H C 1 G D 1 F A
= δ δ δ δ δ δ δ δ
2N N A H N B G N C F N D E
β 1 A1 B1 C1 D
δ δ δ δ
2N N A N B N C N D
β
= .
2N
If we now consider the second simplest class of example, for which the curve C contains
β 2
, where again all of the factors of N1 will
exactly two plaquettes, we will find hWC i ≈ 2N
cancel. As soon as the curve C is large enough such that there can actually be an interior
to the surface Σ – that is, for which some plaquettes will not be along the perimeter C –
then the cancellation of N1 factors will be incomplete. After the integrations, there will be
n∗ − 1 = RT /a2 − 1 factors of (1/N ) left over, so the leading contribution to the expectation
value of the Wilson loop is
RT /a2 RT /a2 −1 RT /a2
β 1 β
hWC i ≈ =N .
2N N 2N 2
Recalling the definition β = 2N/g 2 , we obtain
RT /a2
1 RT 2
hWC i ≈ N 2
= N e− a2 ln(g N )
g N
The behavior hWC i ∼ e−(#)RT is called the area law. As explained on p. 377, the energy
between a quark and an antiquark is
1
E(R) = − lnhWC i = σ R + O(1/T )
T
where we have defined the string tension
ln(g 2 N )
σ= .
a2
We can neglect the constant of order 1/T in the limit of taking the curve C arbitrarily large.
22
R
Remember R that the
R Qintegral DU is written in a condensed notation to denote integration over all link
variables. So DU = [ hiji DUij ], where Uij connects the sites i and j on the lattice and hiji denotes that
sites i and j are nearest neighbors.
212
VII.2 Electroweak Unification
1. Unfortunately, the mass of the elusive Higgs particle H depends on the parameters in
the double well potential V = −µ2 ϕ† ϕ + λ(ϕ† ϕ)2 responsible for the spontaneous symmetry
breaking. Assuming that H is massive enough to decay into W + W − and ZZ, determine the
rates for H to decay into various modes.
Solution:
where H is a real scalar field whose quantum is the spin-0 Higgs boson. As described on p.
382, after using the above parameterization for ϕ in the Lagrangian L = (Dµ ϕ)† (Dµ ϕ) and
identifying the photon A and the Z-boson through the rotation
3
Z cos θ − sin θ W
=
A sin θ cos θ B
(HW W ) = −2iv −1 MW
2
gµν , (HZZ) = −2iv −1 MZ2 gµν , (H``) = −im` v −1
Note the factor of 2 in the HZZ vertex due to the interchange symmetry of the Zs.
which implies
2
Mz2
2
εµλ1 (k1 )ενλ1 (k1 )∗ [ελ2 µ (k2 )ελ2 ν (k2 )∗ ]
|M| = 4
v
213
Summing over outgoing polarizations using the rule
X µ kµkν
ελ (k)ενλ (k)∗ = g µν −
λ
MZ2
gives the polarization-averaged amplitude-squared M2 ≡ λ1 ,λ2 |M|2 :
P
2 2
k1µ k1ν
2 Mz µν k2µ k2ν
M =4 g − gµν −
v MZ2 MZ2
2 2 " µ 2 #
Mz 1 k1 k2µ
=4 4 − 2 (k12 + k22 ) +
v MZ MZ2
!2
2 2 0 0 ~ ~
=4
Mz 2 + k1 k2 − k1 · k2
v MZ2
where in the last line we have used the fact that the outgoing Z bosons are on-shell, meaning
k12 = k22 = MZ2 . Now we use equation (38) on p. 141 for the differential decay rate in the
center of mass frame:
1 d3 k1 d3 k2
dΓ = 3 0 3 0
(2π)4 δ (4) (K − k1 − k2 ) M2
2mH (2π) 2k1 (2π) 2k2
√
where mH ≡ λ v is the mass of the Higgs boson and K is its 4-momentum. When integrating
over the outgoing momenta, the fact that there are two identical outgoing Z bosons forces
1
R
us to divide by 2, meaning that the total decay rate is Γ = 2 dΓ. Now we compute:
Z 3 Z 3
1 1 1 d k1 d k2
Γ= 0
δ(mH − k10 − k20 ) δ (3) (~k1 + ~k2 ) M2
2 2mH 2
2 (2π) 2 k1 k20
4π dk k 2
Z
1
q
= 6 2 δ(m H − 2 k 2 + MZ2 ) M2
2 π mH k 2 + MZ2
p 1
Define f (k) ≡ mH − 2 k 2 + MZ2 and use the formula δ(k) = |f 0 (k ∗ )|
δ(k − k∗ ), where k∗ is
m2
defined by f (k∗ ) ≡ 0. f (k) = 0 =⇒ k 2 = 4H − MZ2 . The derivative of f (k) at this value of
k is
−2k 4k
f 0 (k) = p =−
2
k + MZ2 mH
where we wait until later to plug in the value for k. Using this, we have
1 k2 mH 2
Γ= 4 2 2
M
2 πmH k + MZ 4k
1 k
= 4 M2
2 π 4(k + Mz2 )
2
p
1 m2H /4 − MZ2
= 4 M2
2π m2H
s 2
1 2MZ
= 5 1− M2
2 πmH mH
214
If the Z bosons were spin-0 particles, p
then there would be no momentum dependence in M
and we would get the behavior Γ ∝ 1 − (2MZ /mH )2 we expect from the discussion of
1 → 2 meson decay on p. 142. The decay rate is only real if mH > 2MZ , which reflects the
perfectly sensible fact that the Higgs can only decay into two Z bosons if its mass is larger
than twice the mass of the Z boson.
Now let’s see what the spin-1 polarizations do. Using the integrations over the delta functions,
we have:
2 2 2
mH m2H m4H 2MZ2
0 0 ~ ~ 2 2 2 2 2
(k1 k2 − k1 · k2 ) = [(k + MZ ) + k ] = + 2
− MZ = 1− 2
4 4 4 mH
Therefore:
s 2 " 2 #
MZ4 m4H 2MZ2
1 2MZ
Γ= 5 1− 4 2 2+ 1− 2
2 πmH mH v 4MZ4 mH
2 2 s 2 " 2 #
m4H 2MZ2
mH MZ 2MZ
= 1− 1+ 1− 2
4π mH v mH 8MZ4 mH
It is often convenient to display these decay rates as polynomials in the variable x ≡ 4MZ2 /m2H .
Substituting MZ2 = xm2H /4 gives
2 2
m4H 2M 2 m4H 2(xm2H /4)
1+ 1 − 2Z =1+ 1−
32MZ4 mH 8(xm2H /4)2 m2H
2 x2
2 x 2 x 2
=1+ 2 1− = 2 + 1−
x 2 x 2 2
2 2
2 x x 1
= 2 +1−x+ = 2 (3x2 − 4x + 4)
x 2 4 2x
m3H √ 2 4MZ2
Γ(H → ZZ) = 1 − x (3x − 4x + 4) , x ≡
27 πv 2 m2H
√
When comparing to other sources23 substitute v for the Fermi constant GF ≡ √ 1/(v 2 2 ),
although be aware that different definitions of GF exist, for example without the 2 . To get
an idea for the size of the prefactor, take mH = v ≈ 246 GeV, for which m3H /(27 πv 2 ) ∼ 0.6
GeV and x ∼ 0.5 so that Γ ∼ 1.8 GeV.
To calculate the decay rate for H → W + W − , note that almost everything would go through
in exactly the same way as for H → ZZ with MZ replaced by MW . The only difference is
23
For example, B. W. Lee, C. Quigg and H. B. Thacker, “Weak Interactions at Very High Energies: the
Role of the Higgs Boson Mass,” FERMILAB-Pub-77/30-THY March 1977.
215
that since W + and W − are not identical particles, there is no factor of 1/2 when integrating
dΓ to get Γ. Therefore we can immediately write down
m3H p 4MW 2
Γ(H → W + W − ) = 1 − y (3y 2
− 4y + 4) , y ≡
26 πv 2 m2H
Note that Γ(H → W + W − )/Γ(H → ZZ) is not quite 2, since MW /MZ < 1.
Finally, we compute the decay width of the Higgs boson into a lepton pair `+ `− , which
becomes the dominant decay channel if mH < 2MW ∼ 190 GeV. The vertex is
¯ = imv −1
(H ``)
so the amplitude is simplyPM = imv −1 ū(p1 , s1 )v(p2 , s2 ), which implies that the spin-summed
amplitude squared M2 ≡ s1 ,s2 |M|2 is
m2
6 p1 + m 6 p2 − m 1
2
M = 2 tr = 2 ( pµ1 p2µ − m2 )
v 2m 2m v
The integrals over the momentum-conserving delta function will work out in exactly the
same way as for the H → ZZ mode, which implies pµ1 p2µ − m2 = 2p2 , where p ≡ |~p1 | = |~p2 |.
Therefore, as far as the decay rate is concerned, we have
4p2
M2 =
v2
The decay rate formula is
d3 p1 d3 p2
Z Z
1
Γ= (2π)4 δ(mH − p01 − p02 )δ (3) (~p1 + p~2 ) M2
2mH (2π)3 (p01 /m) (2π)3 (p02 /m)
We write the formula explicitly to remind you of the factor p0 /m for the fermions, although
everything else will proceed in the same way as before, leaving the result
3/2
m2` 4m2`
+ −
Γ(H → ` ` ) = mH 1 − 2
16πv 2 mH
√
Note that m` = f v/ 2 , where f is defined by the Lagrangian L = f ψ̄L ϕ `R + h.c. (using
the notation of pp. 380-381), so the limit m` → 0 is perfectly fine and just leaves behind
f2
Γ(H → `+ `− ) → mH .
32π
2. Show that it is possible to stay with the SU (2) gauge group and to identify W 3 as
the photon A, but at the cost of inventing some experimentally unobserved lepton fields.
This theory does not describe our world: For one thing, it is essentially impossible to incor-
porate the quarks. Show this! [Hint: We have to put the leptons into a triplet of SU (2)
216
instead of a doublet.]
Solution:
We take the gauge group to be G = SU (2), and we will use 2-component spinor notation.
Consider a lepton triplet written as a 2-by-2 symmetric matrix:
!
`+ √1 `0
2
` ∼ 2 ⊗S 2 =⇒ `ij = √1 0 − .
2
` `
The superscript labels indicate that we are anticipating Pa bit and denoting the electric charge
of each component. With the gauge bosons (Wµ )i j ≡ 3a = 1 Wµa ( 12 σ a )i j , the covariant deriva-
tive acting on `ij with gauge coupling g is
!
Wµ3 `+ + Wµ+ `0 √1 (W + `− + W − `+ )
2 µ µ
i(Dµ `)ij = i∂µ `ij + g √1 + − − + 3 − − 0 .
2
(W µ ` + W µ ` ) −W µ ` + W µ `
The currents defined by L = i`† σ̄ µ Dµ ` = i`† σ̄ µ ∂µ ` + Wµ3 J3µ + Wµ+ J −µ + Wµ− J +µ are therefore:
If Wµ3 remains massless, then we see that J3µ is the correct definition of the electromagnetic
current, with the electric coupling defined as e ≡ g.
Consider a Higgs field φ ∼ 2 ⊗S 2. Its covariant derivative is the same as that for `, and its
interactions with the gauge bosons comes from the kinetic term L = (Dµ φ)† Dµ φ. When φ0
obtains a vacuum expectation value, we find
+
0 Wµ 0
(Dµ φ)ij = −ighφ i + ...
0 Wµ−
and therefore
(Dµ φ)† Dµ φ = 2g 2 |hφ0 i|2 Wµ+ W −µ + ...
√
The charged bosons W ± get a mass mW = 2 g|hφ0 i| and the neutral boson W 3 remains
massless. Given this and the lepton currents derived previously, we can indeed identify W 3
as the photon.
217
However, there is a problem in assigning masses to the charged leptons while keeping the
neutral one massless (or at least approximately massless). From `ij ∼ 2 ⊗S 2, we can form
the singlet
1
` εik εj` `k` = − 12 `0 `0 + `− `+ .
2 ij
If we couple an SU (2)-singlet Higgs ϕ with a nonvanishing vacuum expectation value hϕi to
this term, then we find a Majorana mass for the neutral lepton with the same magnitude as
the Dirac mass for the charged lepton. This is clearly incompatible with the near massless-
ness of the neutrino that participates in nuclear beta decay.
For further discussion as well as for other models, we refer the reader to the literature.
Here we derive the electroweak SU (2) ⊗ U (1) covariant derivatives for a field transform-
ing as φ ∼ (2 ⊗S 2 = 3, y), for arbitrary values of the hypercharge y. (Here we will use the
convention for which the hypercharge generator is denoted by Y rather than 12 Y .)
218
g s
i[(D − ∂)φ]12 = [W a (σ a )1 i δ2 j + W a δ1 i (σ a )2 j + 2y Bδ1i δ2j ]φij
2 c
g a a i a a j s
= [W (σ )1 φi2 + W (σ )2 φ1j + 2y Bφ12 ]
2 c
g √ √ s
= [(W − W )φ12 + 2 W φ22 + 2 W − φ11 + 2ys(− Z + A)φ12 ]
3 3 +
2 c
s 1 + −
= g[ys(− Z + A)φ12 + √ (W φ22 + W φ11 )]
c 2
g s
i[(D − ∂)φ]22 = [W a (σ a )2 1 φ12 + W a (σ a )2 2 φ22 + W a (σ a )2 1 φ21 + W a (σ a )2 2 φ22 + 2y Bφ22 ]
2n c
√ s o
=g 2 W − φ12 + [−c(1 + y( )2 )Z + s(−1 + y)A]
c
To summarize:
[i(D − ∂)φ]ij =
2 √ !
[c(1 − y sc2 )Z + s(1 + y)A]φ11 + 2 W + φ12 ys(− sc Z + A)φ12 + √12 (W + φ22 + W − φ11 )
g 2 √
× [−c(1 + y sc2 )Z + s(−1 + y)A]φ22 + 2 W − φ12
Let us check that the couplings to the photon come out as planned. The generator of electric
0 0
3 i0 0 3 j0 0
0 0
charge is Q = T 3 + Y , or in components Qiji j = σ2 δj j + δi i σ2 + yδii δjj . We find:
i j
This Higgs triplet can couple to the left-handed lepton triplet `(i `j) to generate neutrino
masses.
219
VII.3 Quantum Chromodynamics
1. Calculate C in (9). [Hint: If you need help, consult T. Appelquist and H. Georgi,
Phys. Rev. D8: 4000, 1973; and A. Zee, Phys. Rev. D8: 4038, 1973.]
!
σ(e+ e− → hadrons)
X
2 2
(9) R(E) ≡ = 3 Qa 1+C 2 + ...
σ(e+ e− → µ+ µ− ) a
(11 − 3
nf ) ln(E/µ)
Solution:
The details of this two-loop calculation can be found on p. 415 (chapter 8-4-4) of the
quantum field theory text by Itzykson and Zuber. If you compute the integrals, you can
check your answer with the Particle Data Group24 :
!
X
2 αS (E) 2
R(E) = 3 Qa 1+ + O(αS )
a
π
2. Calculate (2).
dg 11 g3
(2) = − T2 (G)
dt 3 16π 2
Solution:
We will use the background field gauge. The goal is to compute the 1-loop effective po-
tential for pure Yang-Mills theory in the presence of a constant background gauge field.25
Including gauge-fixing and ghost terms (see p. 372), the full Yang-Mills Lagrangian with
counterterms is
L = LYM + Lgf + Lghost + Lct
where
1 a aµν a
LYM = − Fµν F , Fµν = ∂µ Aaν − ∂ν Aaµ + gf abc Abµ Acν
4
1
Lgf = − Ga Ga
2ξ
Lghost = (Dµ c̄)a (Dµ c)a , (Dµ c)a = ∂µ ca + gf abc Abµ cc
1 a
Lct = − (Zg − 1)Fµν F aµν + (Zc − 1)(Dµ c̄)a (Dµ c)a
4
24
http://pdg.ihep.su/2009/reviews/rpp2009-rev-qcd.pdf
25
As suggested on p. 388 of the text, here we follow the calculation of S. Weinberg.
220
We will choose the gauge fixing function Ga shortly.
gB = Zg1/2 g µ̃ε/2
between the bare coupling gB and the renormalized coupling g. Here we have anticipated
using dimensional regularization d = 4 − ε to regulate the forthcoming divergent integral,
and therefore replaced g → g µ̃ε/2 to keep g dimensionless in d 6= 0. The goal is now to
compute the O(g 2 ) contribution to Zg and thereby obtain the 1-loop beta function for the
gauge coupling g.
We now split up the gauge field Aaµ into a constant background field Āaµ and a fluctua-
tion A0a 0
µ : A = Ā + A . (See p. 504 in chapter N.3 of the text.) The full (infinitesimal) gauge
transformation δAaµ = ∂µ εa − f abc εb Acµ is split up into
so that Ā transforms as a gauge field while A0 transforms as a matter field in the adjoint
representation.
we arrive at
1 a 2
LYM = − 2
F̄µν + (D̄µ A0ν )a − (D̄ν A0µ )a + gf abc A0bµ A0cν
4g
1
Lgf = − [(D̄µ A0µ )a ]2
2ξ
Lghost = (D̄µ c̄)a [(D̄µ c)a − gf abc cb A0cµ ] .
a
Here F̄µν = gf abc Ābµ Ācν is the field strength of the constant background field (that is,
∂µ Āaν = 0), and D̄µ is the covariant derivative with respect to the background field, as
defined by the gauge fixing function Ga above.
Since Āaµ is to be taken as a fixed classical background field, we do not integrate over it
in the path integral. The theory is invariant under the formal transformation
221
irrespective of whether the transformation is performed before or after integrating over the
fluctuation fields A0a a a
µ , c , c̄ . Thus by this formal background gauge invariance, we can com-
pute the 1-loop correction to the gauge coupling by finding the coefficient of − 41 F̄µν
a
F̄ aµν ∼
a 4 0
(µ ) in the 1-loop effective action, obtained by keeping only quadratic terms in A , c, c̄ and
performing the resulting gaussian integrals.
where in the second line we have chosen the gauge ξ = 1 and defined the matrices
ab ca ∂ cda dα cb ∂
Mµν (x, y) = ηµν δ − gf Ā δ − gf Āα δ 4 (x − y)
ceb e
∂xα ∂y α
ca ∂ cda d cb ∂
− δ − gf Āν δ ceb e
− gf µ − (µ ↔ ν) δ 4 (x − y)
∂xν ∂y µ
+ gf cab F̄µν
c 4
δ (x − y)
and
ab ∂ ca cb ∂
N (x, y) = δ µ
− f cda Ādµ δ ceb eµ
− f Ā δ 4 (x − y) .
∂x ∂yµ
The one-loop effective action can now be obtained by integrating over the fluctuations A0
and c, c̄:
d4 p 1
Z Z
4
iΓ1-loop (Ā) = d x 4
− 2 tr ln M (p) + tr ln N (p)
(2π)
ab
where Mµν (p) and N ab (p) are defined through the Fourier transforms of Mab µν (x, y) and
ab
N (x, y) respectively:
ab
(p) = −ηµν ipα δ ca − gf cda Ādα ipα δ cb + gf ceb Āeα
Mµν
+ ipν δ ca − gf cda Ādν ipµ δ cb + gf ceb Āeµ − (µ ↔ ν)
+ gf cab F̄µν
c
a
We are interested in extracting the coefficient of the term proportional to F̄µν F̄ aµν in the
one-loop effective action, so we need all the terms quartic in Āaµ . Let On denote the part of
an operator O that is of order (Āaµ )n . Then the terms we are looking for are of the form:
for O = M, N . Each of these terms has a diagrammatic interpretation. The first term is
222
A A
O 0 −1
O2 O2
A O 0 −1 A
In other words, reading from right to left, the term tr(O0−1 O2 O0−1 O2 ) starts with an operator
with two external Ā lines, turns into an internal propagator O0−1 , meets another operator
with two external Ā lines, then closes with another internal propagator O0−1 . The symmetry
factors and minus signs are all automatically taken into account by the algebra at the level
of the effective action.
A
A
O1
O 0 −1
O2 O 0 −1
O 0 −1
A O1
A A
O 0 −1
O1 O1
O0 −1
O 0 −1
O1 O1
O 0 −1
A A
Applying this decomposition to the operators O = M and N and simplifying the momentum
integrals using pµ pν → 41 p2 and pµ pν pα pβ → 4!1 (η µν η αβ +η µα η νβ +η µβ η να )(p2 )2 in the integrals,
223
we obtain
d4 p
Z
5
4
[tr ln M (p)]4 = − I g 2 f abc f abd F̄µν
c
F̄ dµν
(2π) 3
Z 4
dp 1
4
[tr ln N (p)]4 = + I g 2 f abc f abd F̄µνc
F̄ dµν .
(2π) 12
Here we have defined a divergent integral
d4 p
Z
1
I≡ .
(2π) (p + i)2
4 2
For the color group SU (n), we have f abc f abd = n δ cd and therefore:
d4 p
Z
5 1
1
[− tr ln M (p) + tr ln N (p)] = + + I g 2 n F̄µν
a
F̄ aµν
(2π)4 2 6 12
11 g 2 1 1
2
µ 1 a aµν
=i − n 2 + ln − F̄µν F̄
12 2π ε 2 m2 4
We have therefore deduced the coefficient of − 14 F̄µν
a
F̄ aµν in the 1-loop effective action. Mul-
tiplying by −i, we find the renormalization factor Zg to O(g 2 ):
11 g 2 1
µ
Zg = 1 − n 2 + ln .
12 2π ε m
1/2
The bare coupling gB is related to the running coupling g by gB = Zg g µ̃ε/2 , so that
ln gB = O(g, ε) + ln g + 12 ε ln µ̃ .
where O(g, ε) ≡ 21 ln Zg . The bare coupling does not know about the parameter µ, so
differentiating the above relation gives
∂O(g, ε) dg
0= + 12 ε .
∂g d ln µ
In general, the function O(g, ε) will contain finite terms as well as poles in powers of 1/ε:
∞
X On (g)
O(g, ε) = O0 (g) + n
.
n=1
ε
224
dg
Plugging in this expansion, writing limε→0 d ln µ
= β(g) and matching powers of 1/ε we obtain
h i
1 2 0 0 0 2
β(g) = 2
g O1 (g) 1 − gO0 (g) + O (gO0 (g)) .
11 g 2 1 11 g 2 1
1 1 µ µ
O(g, ε) = ln Zg = ln 1 − n 2
2 2
+ ln = n 2 + ln
12 2π ε m 12 4π ε m
where we have dropped terms of O(g 4 ) and terms of higher powers in 1/ε, which are assumed
to cancel order by order. Thus matching to the general expansion of O(g, ε), we have
11 g 2 µ
O0 (g) = − n 2 ln
12 4π m
11 g 2
O1 (g) = − n 2 .
12 4π
Thus gO00 (g) is O(g 2 ) and can be dropped from the beta function at this order, since it
multiplies g 2 O10 (g) = O(g 3 ). This is consistent with the general property of gauge theories
that 1-loop beta functions are independent of the renormalization scheme.
g g
Since O10 (g) = − 11
12
n 2π 2
= − 11
3
n 8π 2
, the 1-loop beta function β(g) = 12 g 2 O10 (g) is
11 g3
β(g) = − n + O(g 5 ) .
3 16π 2
This is equation (3) on p. 388. In the non-abelian gauge theory of the strong interaction, we
have n = 3.
225
VII.4 Large N
1. Since the number of gluons only differs by one, it is generally argued that it does not make
any difference whether we choose to study the U (N ) theory or the SU (N ) theory. Discuss
how the gluon propagator in a U (N ) theory differs from the gluon propagator in an SU (N )
theory and decide which one is easier.
Solution:
Pdim G a a j
Let (Aµ )i j = a = 1 Aµ (T )i be the matrix-valued gluon field for G = U (N ) that acts
on the defining representation, meaning that the indices i, j run from 1 to N . Then:
where
d4 k e ik·x
Z
kµ kν
∆µν (x) = −ηµν + (1 − ξ) 2
(2π)4 k 2 + iε k + iε
is the usual abelian massless vector boson propagator in Rξ gauge. The new feature is just
the group theory factor
N 2
X
(T a )i j (T a )k` = C δi ` δk j (1)
a=1
226
is to take equation (1) (with C = 1 as per the previous discussion) and move the (N 2 )th
generator from the left-hand side to the right-hand side. In other words,
N 2 −1
X 1 j `
(T a )i j (T a )k` = δi ` δk j − δ δ (10 )
a=1
N i k
The previous discussion for U (N ) carries through exactly, except with the group theory factor
from equation (10 ) instead of (1). Therefore the SU (N ) gauge boson propagator is
j ` ` j 1 j `
h(Aµ )i (x)(Aν )k (0)iSU (N ) = ∆µν (x) δi δk − δi δk .
N
This too has a straightforward graphical interpretation, where we subtract the trace from
the U (N ) propagator:
i l 1 i l
j k N j k
Solution: 27
Define light cone coordinates p± = √12 (p0 ± p1 ), so that 2p+ p− = p20 − p21 = m2 , and fix
light cone gauge: A− = 0. The Lagrangian for (1+1)-dimensional QCD is then
where (A+ )i j = Aa+ (T a )i j is the matrix-valued gauge field. Treating x+ as the temporal
direction, we see that A+ has no dynamics and can be replaced by the 1D Coulomb potential.
0 1
The gamma matrices are defined by the Clifford algebra {γµ , γν } = 2 , so that
1 0
γ+2 = γ−2 = 0 and γ+ γ− +γ− γ+ = 2. The only interaction vertex in the theory is −igγ− (T a )i j ,
so the gamma matrices can be removed from the Feynman rules. For example, consider the
1-loop correction to the quark propagator due to single-gluon exchange:
Z
2 a a i dk− dk+ γ− [γ− (6 k−6 p)+ + γ+ (6 k−6 p)− + m] γ−
−g (T T ) j 2
(2π)2 [k− ][2(k − p)+ (k − p)− ]
Since γ−2 = 0 and γ− γ+ γ− = γ− (−γ− γ+ + 2) = 0 + 2γ− , only the part of the fermion
propagator proportional to γ+ contributes, and its contribution is simply a factor of 2. Thus
the Feynman rules can be taken as:
27
We thank G. ’t Hooft for helpful discussion.
227
−i
=
k −2
ik−
=
2 k + k −− m 2
= −i2g
The large-N approximation is to neglect non-planar diagrams and to consider loops due to
gluon exchange only. The quark self-energy Σ is related to the exact quark propagator S by
ik−
iS(k) =
2k+ k− − m2 − k− Σ(k)
and satisfies the implicit equation
= S
g2
Z
1
Σ(p− ) = dk− 2 sgn(k− + p− ) .
2π k−
R∞ R∞ R −µ
This integral diverges near k− → 0. To deal with this, replace −∞ dk− → µ dk− + −∞ dk−
with µ > 0 a positive IR regulator. The self-energy is
Z ∞ Z −µ
g2
1 1
Σ(p− ) = dk− 2 sgn(k− + p− ) + dk− 2 sgn(k− + p− )
2π µ k− −∞ k−
∞
g2
Z
1
= dk− 2 [sgn(p− + k− ) + sgn(p− − k− )] .
2π µ k−
228
d
Use dx
sgn(x) = 2δ(x) to evaluate the integral by parts:
g2 ∞
Z
d 1
Σ(p− ) = − dk− [sgn(p− + k− ) + sgn(p− − k− )]
2π µ dk− k−
" #
2 ∞ Z ∞
g 1 2
=− (sgn(p− + k− ) + sgn(p− − k− )) − dk− (δ(p− + k− ) − δ(p− − k− ))
2π k− k− = µ µ k−
g2
1 1
=− − (sgn(p− + µ) + sgn(p− − µ)) + 2sgn(p− )
2π µ |p− |
2
g 1 1
→+ sgn(p− ) −
π µ p−
where in the last line we have taken µ → 0+ in the numerator.
Thus we have solved for the quark self-energy. The denominator of the exact propagator
is therefore
g 2 |k− |
2k+ k− − − 1 − m2 .
π µ
In the limit µ → 0+ , the pole of the propagator is shifted to k+ → ∞. This indicates that
there is no physical single-quark state.
Next consider the implicit equation given by relating the nth and (n + 1)th ladder diagram:
m1 p p + k p
S S S
= k
S S S
m2 p−q p + k − q p−q
Here the blob, which we denote ψ(p, q), stands for an arbitrary process in which a quark-
antiquark pair emerges, the quark with mass m1 and momentum p, and the antiquark with
mass m2 and momentum q − p. The source-free part of this equation is:
i
iψ(p, q) = (−2g)2 (p− − q− )pi 2 g2
×
2(p+ − q+ )(p− − q− ) − (m22 − gπ ) − |p
πµ −
− q− |
dk− dk+ −i
Z
i
2 2 2
iψ(p + k, q) .
g g
2p+ p− − (m21 − π ) − πµ |p− | (2π)2 k−
After a series of manipulations, this can be put into the form of the eigenvalue equation28
Z 1
2 α1 α2 ϕ(y)
ρ ϕ(x) = + ϕ(x) − P dy
x 1−x 0 (y − x)2
28
The principal value integral is the average value of the integral across the pole:
Z
ϕ(x − i)
Z Z
ϕ(x) 1 ϕ(x + i)
P dx 2 = dx + dx .
x 2 (x + i)2 (x − i)2
229
R
where we have defined ϕ(p− , q) ≡ dp+ ψ(p, r) and the dimensionless variables x ≡ p− /r− , αi ≡
π
m2 − 1 and ρ2 ≡ gπ2 (2q+ q− ). Here ρ is the meson mass in units of g/π 1/2 .
g2 i
An approximate solution to the eigenvalue equation may be found by observing that the
dominant contribution to the integral on the right-hand side comes from y ≈ x, at which
point the denominator goes to zero. Using the trial function ϕ(x) = e iωx , we have
Z 1 Z ∞
e iωy e iωy
P dy 2
≈ P dy 2
= −π|ω| e iωx
0 (y − x) −∞ (y − x)
so that e iωx is an approximate solution with eigenvalue ρ2 ≈ π|ω|, where we have further
assumed that the quark and antiquark have equal masses (e.g., a meson made from a uū
pair rather than a ud¯ pair) and satisfy m2q ≈ g 2 /π, so that α1 = α2 ≈ 0. The appropriate
boundary conditions are ϕ(0) = ϕ(1) = 0, so that we are to take the linear combination
ϕω (x) ∝ e iωx − e−iωx ∝ sin(ωx) as the solution.
Defining ω = πn, where n is any positive integer, we thereby arrive at the meson spectrum
2
m2π ≡ gπ ρ2 given by
(m2π )n ≈ g 2 πn , n = 1, 2, 3, ...
The approximation becomes better for larger values of n. In terms of the quark mass mq , we
have
(m2π )n
≈ π2n .
m2q
For corrections to this lowest order approximation as well as for mesons built of quarks with
unequal masses, see the reference [G. ’t Hooft, “A two-dimensional model for mesons,” Nucl.
Phys. B75 (1974) 461-470]. For a path integral treatment as well as for the inclusion of
baryons in the spectrum, see E. Witten, “Baryons in the 1/N expansion,” Nucl. Phys. B160
(1979) 57-115.
3. Show that if we had chosen to calculate G(z) ≡ h(1/N )tr[1/(z − ϕ)]i, we would have
to connect the two open ends of the quark propagator. We see that figures VII.4.5b and d
lead to the same diagram. Complete the calculation of G(z) in this way.
If we want to calculate G(z) directly, then we can’t have any free indices dangling on the ran-
dom matrices ϕ. The trace takes the upper index on the incoming quark line and contracts
it with the lower index on the outgoing quark (which behaves like an incoming antiquark,
therefore in an antifundamental rep.) Therefore, we have really closed the loop by connecting
the two quark lines in the diagram from the text. We will calculate G(z) by making use of
a helpful recursion relation.
230
As we know from the text, we can write
∞
X 1 1
G(z) = tr ϕ2n
n=0
z 2n+1 N
where the matrix propagator gives a factor of 1/(N m2 ). To get the large N limit, the only
surviving terms are the planar graphs, which means that the trace, when computed as Wick
contractions, breaks into one trace per propagator (plus the original quark line trace). Thus,
we can write ∞
X 1
G(z) = Cn 2n+1
n=0
z
where Cn is the number of planar terms with n propagators. Let’s now derive a recursion
relation for the Cn , starting with the zero propagator case, C0 = 1. Let the notation [ϕ...ϕ]
denote the canonical Wick contraction, where the two fields at each end (ie, adjacent to
each bracket) are contracted. Consider the case with n propagators, or 2n matrices. (To
get planar graphs, there must be an even number of matrices between each contracted pair.)
Thus, the total number of Wick contractions for 2n fields is given by
(all for 2n) = [ϕ(all for 2n − 2)ϕ] + [ϕ(all for 2n − 4)ϕ][ϕϕ] + ...[ϕϕ](all for 2n − 2)
n−1
X
Cn = C0 Cn−1 + C1 Cn−2 + ... + C0 Cn−1 = Ck Cn−1−k .
k=0
In the second line, we have used the recursion relation and set n0 = n − 1. In the third
line, we noted that the second line was just the power series expansion of the product of two
series. Now we have the same quadratic equation as in the text, leading to
r !
m2 4
G(z) = z − z2 − 2 .
2 m
231
4. Suppose the random matrix ϕ is real symmetric rather than hermitean. Show that
the Feynman rules are more complicated. Calculate the density of eigenvalues. [Hint: The
double-line propagator can twist.]
Solution:
First, as clarification, this question is intended to follow the procedure on p. 397 with
the quadratic (Gaussian) potential V (ϕ) = 21 m2 ϕ2 .
In group theory language, the complication here is that there is no longer any distinction be-
tween up and down indices. For ϕ an N ×N hermitian matrix, meaning (ϕ† )i j ≡ (ϕj i )∗ = ϕi j ,
we know hϕi j ϕk` i ∝ δi ` δk j by matching upper and lower indices. But if ϕ is an N × N real
symmetric matrix, meaning ϕij = ϕji with all entries being real numbers, then matching
indices tells us hϕij ϕk` i = C1 δi` δjk + C2 δik δj` , with two a priori undetermined constants. The
first term corresponds to the hermitian case (just lower the indices ` and j from before), while
the second term involves a matrix transpose, which diagrammatically appears as a twist in
the double-line propagator.
The difference between the complex Hermitian case and the real symmetric case is in calcu-
lating the correlation function h(ϕ2 )ij i. The “propagator” hϕik ϕ`j i is given by the Gaussian
integral Z
1 1 2 2
dϕ e−N tr 2 m ϕ ϕik ϕ`j = C1 δij δk` + C2 δik δj`
Z
where we will now fix the constants C1 and C2 by taking special cases of the above expression.
First notice that trϕ2 = ϕij ϕji = ϕ211 + 2ϕ212 + ..., meaning that diagonal terms ϕ2ii (no
sum) come in with a factor of 1, while off-diagonal terms ϕ2i,j6=i come in with a factor of 2.
Now we will solve for the constants C1 and C2 in our matrix theory. First set i = k = j =
` = 1 to get
Z
1 1 2 2 1 1
dϕ e− 2 (N m )ϕ11 +... ϕ211 = = = C1 + C2 .
Z α α = N m2 N m2
232
1 1 1
Therefore C2 = N m2
− 2N m2
= 2N m2
. We arrive at the somewhat intuitively clear result
Z
1 1 2 ϕ2 1
hϕik ϕ`j i ≡ dϕ e−N tr 2 m ϕik ϕ`j = (δij δk` + δik δj` )
Z 2N m2
so that the previous factor 1/(N m2 ) is split evenly among the two tensors dictated by group
theory.
As in the text, the next step is to study the planar diagrams contributing to the n = 2
term and thereby arrive at a quadratic equation for G(z). The first half of the algorithm
(“repeat”) translates to the present case exactly as in the text, so that again we arrive at
equation (7) on p. 399:
1
G(z) = .
z − Σ(z)
Next we look at Figure VII.4.6b on p. 400 and observe that in our case we have two terms,
one which is depicted in the figure and another one for which the overarching propagator
twists. In other words, we have
1
Σ(z) = G(z)
2m2
with the extra factor of 1/2 arising just as in the above calculation of h(ϕ2 )ij i. Together
these result in the quadratic equation
1
G(z) = =⇒ G2 − 2m2 z G + 2m2 = 0
z − G(z)/(2m2 )
233
5. For hermitean random matrices ϕ, calculate
1 1 1 1 1 1 1 1
Gc (z, w) ≡ tr tr − tr tr
N z −ϕN w−ϕ N z−ϕ N w−ϕ
for V (ϕ) = 12 m2 ϕ2 using Feynman diagrams. [Note that this is a much simpler object to
study than the object we need to study in order to learn about localization (see exercise
VI.6.1).] Show that by taking suitable imaginary parts we can extract the correlation of the
density of eigenvalues with itself. For help, see E. Brézin and A. Zee, Phys. Rev. E51: p.
5442, 1995.
Solution:
∞
1 X D ϕ n ϕ m E
= 2 ∂z ∂w tr tr .
N n,m = 1
z w c
In terms of diagrams, each tr(ϕn ) is a quark loop, and the average of tr(ϕn )tr(ϕm ) tells
us that the two quark loops are to be attached to one another through gluon (double-line)
propagators.
The sum consists of two classes of diagrams, the first in which index contractions are made
between the two traces only, and the second in which we allow contractions within the same
trace.
For the first class, consider an individual diagram for a fixed value of n, and draw the z-type
quark loops inside the w-type quark loops. We have n different ways to attach the z-type
loops to the w-type loop (using a double-line propagator connecting from the z-diagram to
234
one of the loops in the w-diagram). Once this choice is made, there is only one way to attach
the rest of the z-loops to the w-loops without having the propagators cross. (Recall we are
working in the large-N limit, in which planar diagrams dominate.) Thus each such diagram
has a symmetry factor 1/n.
Given the gluon propagator on p. 398, each attachment of a gluon propagator contributes a
factor 1/(N m2 ), and each resulting closed loop contributes a factor of N . The factors of N
cancel, and we obtain for the sum of these diagrams:
∞ ∞ n
X 1 X 1 1 1
n
htr(ϕn )tr(ϕn )ic =
n=1
(zw) n=1
(zw) n m2
n
class 1
1
= − ln 1 − .
zwm2
Now consider the second class of diagrams, in which we allow contractions within the same
trace. In the large-N limit, these diagrams serve only to dress the bare propagator z1 to
1
the full propagator G(z) = z−Σ(z) (see p. 399). Therefore the full sum of diagrams can be
achieved by taking the result from class 1 and simply replacing 1/z with G(z) and 1/w with
G(w). We have
1 1
Gc (z, w) = − 2 ∂z ∂w ln 1 − 2 G(z)G(w)
N m
where (as on p. 400) !
r
m2 4
G(z) = z− z2 − 2 .
2 m
From this we may obtain the connected correlation between eigenvalues:
1
ρc (E, E 0 ) = − [Gc (+, +) + Gc (−, −) − Gc (+, −) − Gc (−, +)]
4π 2
where Gc (±, ±) ≡ limε,ε0 →0+ Gc (E ± iε, E 0 ± iε0 ). The result is
1 1 4 − m2 EE 0
ρc (E, E 0 ) = − .
4π 2 N 2 (E − E 0 )2 (4 − m2 E 2 )(4 − m2 E 02 )
p
235
6. Use the Faddeev-Popov method to calculate J in the Dyson gas approach.
Solution:
(There are no sums on i and j in the last equality.) So if we choose our gauge transformation
ϕg = g † Λg such that ϕg = Λ, or equivalently if we choose the gauge fixing function f (ϕ) =
ϕ − Λ, then the integral over the delta function is:
Z Z
Dg δ (f (ϕg )) = Dθ δ(−i[θ, Λ])
YZ
= d2 θij δ (2) ((λj − λi )θij ))
ij
Y 1
=
ij
(λj − λi )2
We wrote d2 θij for each matrix element θij because these generator matrices are complex,
∗
and we integrate over their real and imaginary parts (or equivalently, over θij and θij ). This
generates the exponent 2 in the determinant. Therefore, the Faddeev-Popov determinant
(Jacobian) is Y
J= (λj − λi )2
ij
236
7. For V (ϕ) = 12 m2 ϕ2 + gϕ4 , determine ρ(E). For m2 sufficiently negative (the double well
potential again) we expect the density of eigenvalues to split into two pieces. This is evident
from the Dyson gas picture. Find the critical value m2c . For m2 < m2c the assumption of
G(z) having only one cut used in the text fails. Show how to calculate ρ(E) in this regime.
Solution:
√ p
The function z 2 − a2 = (z − a)(z + a) has a single p branch cut of finite length along
the real axis from −a to +a. More generally, the function (z − a)(z − b) with b > a has a
single branchpcut of finite length along the real axis from a to b. From this it is clear that
the function (z − a)(z − b)(z + c)(z + d) , with a, b, c, d real and positive and b > a, d > c,
has two branch cuts of finite length along the real axis, one from a to b and one from −d to −c.
In particular,
p consider the case b = ka, c = p −a, d = kc = −ka for which the function
f (z) ≡ (z − a)(z − ka)(z + a)(z + ka) = [z 2 − a2 ][z 2 − (ka)2 ] contains two discon-
nected branch cuts of length ka, one from a to ka and the other from −ka to −a. We
know that the two cuts have the same length by the Z2 : z → −z symmetry of the potential
V (z), hence the length ka for both of them. The parameter a fixes how far away from the
origin the cuts begin. Together these constitute two unknown parameters.
Requiring the cubic and linear terms in G(z → ∞) to vanish results in two equations.
The requirement G(z → ∞) → (+1) z1 provides the third equation. The three equations for
the three unknowns B, C, D imply
1
B = 4g , C = −4g m2 , D = .
2g
237
Notice that the definition C = 12 a2 (k 2 + 1) implies that C > 0, so that this solution is only
valid for m2 < 0, as expected. Solving for a and k gives two solutions
s s
1 ∓8g 3/2 m2 − (4g 3/2 m2 )2 − 1
a± = −4g m2 ± √ , k± =
g 1 − (4g 3/2 m2 )2
where the ± signs are correlated, meaning that (a+ , k+ ) is one solution and (a− , k− ) is the
second solution. The solution (a− , k− ) is only valid when |m2 | ≥ 1/(4g 3/2 ), which defines the
critical value of m2 :
1
−m2c = 3/2
4g
For |m2 | < |m2c |, the two-cut solution ceases to be valid. Let us make sure this is compatible
with the solution for k− . Let q ≡ 4g 3/2 m2 < 0, so that
2 +2q − q 2 − 1
k− = .
1 − q2
2
The condition k− > 0 is satisfied if 1 − q 2 < 0, which again implies |m2 | > 1/(4g 3/2 ) with
2 2
m = −|m | < 0. Thus we have two solutions for the function G(z):
1
q
0
G+ (z) = V (z) − 4gz [z 2 − a2+ ][z 2 − (k+ a+ )2
2
1
q
0 2 2 2 2
G− (z) = V (z) − 4gz [z − a− ][z − (k− a− )
2
where (a+ , k+ ) and (a− , k− ) are given above as functions of the parameters m2 < 0 and g > 0.
The density of states ρ(E) = − π1 Im G(z) is given by
p
2gE p[a2+ − E 2 ][(k+ a+ )2 − E 2 ] for a+ ≤ E ≤ k+ a+
ρ(E) = 2gE [a2− − E 2 ][(k− a− )2 − E 2 ] for a− ≤ E ≤ k− a−
0 otherwise
We see that the density of states has split into two disconnected regions.
For a treatment involving orthogonal polynomials, see N. Deo, “Multiple Minima in Glassy
Random-Matrix Models,” J. Phys.: Condensed Matter, Vol. 12 No. 29, 24 July 2000.
238
8. Calculate the mass of the soliton (25).
N
mS = mF (25)
π
Solution:
After introducing the auxiliary field σ and integrating out the fermions the action is
Z
N
S[σ] = − d2 x 2 [σ(x, t)]2 − i 12 N tr ln(−∂ µ ∂µ − ∂x σ − σ)
2g
up to an overall additive constant. We know the shape of the time-independent soliton
2
configuration is given by σ(x) = m tanh(mx), where m = µe1−π/g , which satisfies the
differential equation ∂x σ + σ = m2 . Thus the entire tr ln(...) term is merely an additive
constant, and so the energy of the configuration is
Z ∞
N
dx 2 m2 C − tanh2 (mx)
M=
−∞ 2g
where we have written explicitly the additive constant C, which can be thought of as a
counterterm to set the vacuum energy to zero. Recall that the coupling constant g = g(µ)
is a function of the renormalization point µ. Choosing µ = m implies g 2 = π, so that the
energy is
Nm ∞
Z
du C − tanh2 u .
M=
2π −∞
For a formal derivation of the constant C, see A. Klein, “Bound states and solitons in the
Gross-Neveu model,” Phys. Rev. D, Vol. 14 No. 2, 15 Jul 1976. Here we will content
ourselves with observing that limu→∞ tanh u = 1, so that to get a finite answer we must have
C = 1. A physical way to see what is going on is to plot the function 1 − tanh2 u = sech2 u:
Sech2 u
0.25
0.20
0.15
0.10
0.05
u
-10 -5 5 10
We see that it is a peak localized atRthe origin, except that it asymptotes to zero only when
∞
C = 1. The value of the integral is −∞ du sech2 u = 2 tanh(∞) = 2, so we find
Nm
M= .
π
239
VII.5 Grand Unification
1. Write down the charge operator Q acting on 5, the defining representation ψ µ . Work out
the charge content of the 10 = ψ µν and identify the various fields contained therein.
Solution:
Recalling the definition in the text of the ψµ ∼ 5̄ in terms of the familiar fields at low energy
d¯α
ψµ = ν
e
you might worry about the charges of the lower two components of the ψ µ ∼ 5 worked out
above. When working out the generator of electric charge for the 5̄, be careful to note that
although the SU (2) generator T 3 is the same, the sign of the hypercharge generator flips
sign:
1 1
0 +3 ψ1 (+ 3 )ψ1
0
+ 13 ψ2 (+ 1 )ψ2
31
1
[Qψ]µ =
0 +
+ 3
ψ3 = (+ )ψ3
3
1 1
2
− 2
ψ4 (0)ψ4
− 12 − 12 ψ5 (−1)ψ5
So the electron field e = ψ5 indeed has charge −1, and the neutrino field ν = ψ4 is neutral.
Now for the 10. Given a representation R and its generators TR , and given a second repre-
sentation R0 and its generators TR0 , the generators of the product representation R ⊗ R0 are
given by TR⊗R0 = TR ⊗ IR0 + IR ⊗ TR0 . Explicitly in terms of components, if µ, ν, ... denote the
matrix indices of representation R and if a, b, ... denote the matrix indices of representation
R0 , then
(TR⊗R0 )µaνb = (TR )µ ν δ a b + δ µ ν (TR0 )a b .
If the representations R0 and R are the same, then we can symmetrize and antisymmetrize
the two upper indices (or the two lower indices). The generator of the symmetric product
representation R ⊗S R is
240
where the parentheses mean symmetrization in those indices:
1
M (µν) ≡ (M µν + M νµ ).
2
Similarly, the generator of the antisymmetric product representation R ⊗A R is
2
ψ 0αβ = (− )ψ αβ =⇒ ψ αβ = εαβγ ūγ
3
045
ψ = (+1)ψ 45 =⇒ ψ 45 = ē .
1 2
ψ 0α4 = (− + 1)ψ α4 = + ψ α4 =⇒ ψ α4 = uα
3 3
0α5 1 1
ψ = (− + 0)ψ = − ψ α5 =⇒ ψ α5 = dα .
α5
3 3
ν
Actually there is a slight mistake here. The lepton doublet `i ≡ is defined correctly
c e
dα
with a lower SU (2) index, since it is part of the 5̄ ψµ = . The usual Standard-Model
`i
u
convention for the quark doublet is qi ≡ with a lower SU (2) index just like for the
d
d
leptons. This implies q i = εij qj = , so that ψ α5 = dα as written, but ψ α4 = −uα with
−u
an extra minus sign arising from ε54 = −ε45 .
241
2. Show that for any grand unified theory, as long as it is based on a simple group, we have
at the unification scale P 2
T
sin θ = P 32
2
Q
where the sum is taken over all fermions.
Solution:
Consider the case for one fermion ψ that transforms under some representation of the simple
group G. Its gauge-covariant derivative in the electroweak subgroup SU (2) ⊗ U (1) is
√ 1
a
Dµ ψ = ∂µ − ig Aµ Ta + Bµ α Y ψ
2
As discussed on p. 410 of the text for the particular case α = 3/5, we fix α by demanding
that√the SU (2) and U (1) generators are normalized equally, or in other words tr[(T3 )2 ] =
tr[( α 12 Y )2 ]. This implies
tr[(T3 )2 ]
α=
tr[( 12 Y )2 ]
√ √
We define the weak mixing angle by tan θ ≡ g1 /g2 = α , so since tan θ = s/c = s/ 1 − s2 ,
where c ≡ cos θ and s ≡ sin θ, we have
α
s2 = .
1+α
The generator of electric charge is Q = T3 + 12 Y , so tr[( 12 Y )2 ] = tr(Q2 − 2QT3 + T32 ) and
therefore
tr(T32 )
s2 = .
2 tr[T3 (T3 − Q)] + tr(Q2 )
But T3 − Q = − 21 Y , and tr(T3 12 Y ) = 0. Therefore, for a single fermion, we have
tr(T32 )
s2 =
tr(Q2 )
In the case of multiple fermions, the currents in each direction of the Lie algebra get con-
tributions from eachPfermion
√ f that transforms (funder the group. In other words, we fix α
(f ) )
by f tr[(T3 )2 ] = f tr[( α 12 Y (f ) )2 ], where T3 and 12 Y (f ) are the generators appropriate
P
for the representation to which fermion f belongs. Everything carries through as before, and
we arrive at the result P (f ) 2
2 f tr[(T3 ) ]
s =P (f ) )2 ]
.
f tr[(Q
242
3. Check that the SU (3) ⊗ SU (2) ⊗ U (1) theory is anomaly-free. [Hint: The calculation is
more involved than in SU (5) since there are more independent generators. First show that
you only have to evaluate tr Y [Ta , Tb ] and tr Y 3 , with Ta and Y the generators of SU (2) and
U (1), respectively.]
Actually, we have to calculate slightly more than what is stated in the original problem.
In addition to denoting SU (2) generators by T , denote normalized SU (3) generators as t.
One thing we can note: since the charges are repeated in each generation of particles, we
only need to calculate the anomaly cancellation for one of them. Note that when we say “tr”
here, we mean to treat the gauge groups as tensor products (so the traces factorize) and to
put an additional minus sign on the right-handed field traces.
We start with anomalies involving SU (3). Since SU (3) couples equally to left and right
handed particles, the one with 3 ts vanishes automatically. With only one t, one of the
factors in the trace of generators is tr t = 0, so that always vanishes also. Now consider the
traces with 2 ts. We have tr(ta tb )trT c = 0 and tr(ta tb )trY = 12 δ ab trq Y . The trace now runs
only over the quarks because the leptons have t = 0. We have trq Y = 2 × ( 16 ) − 23 − (− 31 ) = 0.
Note that the factor of 2 in the first term is because the left-handed quarks are in an SU (2)
doublet, and that we have subtracted the right-handed ones.
Now take anomalies involving SU (2) generators. Since the fermions are all in doublets or
singlets, we can use the Pauli matrix anticommutator, which gives {T a , T b } = 12 δ ab . Thus,
the 3T anomaly is tr({T a , T b }T c ) = 12 δ ab tr(T c ) = 0. Similarly, tr(Y 2 T a ) = trY 2 trT c = 0.
The last case is with 2 T s, tr(Y T a T b ) = 12 δ ab trL Y , where the trace is now over left-handed
fields. This gives trL Y = 3( 16 ) − 12 = 0. Note the factor of 3 because the quarks are in an
SU (3) triplet.
243
4. Construct grand unified theories based on SU (6), SU (7), SU (8), ... , until you get tired
of the game. People used to get tenure doing this. [Hint: You would have to invent fermions
yet to be experimentally discovered.]
Solution:
This is of course an open-ended problem with many solutions. We will simply point to-
wards some of the literature on unified models based on SU (n) for n > 5:
• SU (9): Y. Fujimoto and P. Sodano, “SU(9) Grand Unified Theory,” Phys. Rev. D
Vol. 23 No. 7, 1 Apr. 1981
• SU (11): H. Georgi, “Towards a Grand Unified Theory of Flavor,” Nucl. Phys. B, 1979
• SU (15): P. Frampton and B. H. Lee, “SU(15) Grand Unification,” Phys. Rev. Lett.
Vol. 64 No. 6, 5 Feb 1990
244
VII.6 Protons Are Not Forever
1. Suppose there are F 0 new families of quarks and leptons with masses of order M 0 . Adopt-
ing the crude approximation described in exercise IV.8.2 of ignoring these families for µ below
M 0 and of treating M 0 as negligible for µ above M 0 , run the renormalization group flow and
discuss how various predictions, such as proton lifetime, are changed.
Solution:
For energies below the scale M 0 we still have only F “massless” fermion families, so equations
(1)-(3) on p. 414 become
0
1 1 1 M
= 0
+ (4F − 33) ln
αS (µ) αS (M ) 6π µ
2 2 0
0
sin θ(µ) sin θ(M ) 1 M
= 0
+ (4F − 22) ln
α(µ) α(M ) 6π µ
0
2 2
0
cos θ(µ) cos θ(M ) 1 20 M
= 0
+ F ln .
α(µ) α(M ) 6π 3 µ
Above the scale M 0 we excite the new degrees of freedom, so for M 0 < µ < MGUT , these
equations become exactly those on p. 414 except with F replaced by F + F 0 :
1 1 1 0 MGUT
= + [4(F + F ) − 33] ln
αS (µ) αS (MGUT ) 6π µ
sin2 θ(µ) sin2 θ(MGUT )
1 0 MGUT
= + [4(F + F ) − 22] ln
α(µ) α(MGUT ) 6π µ
2 2
cos θ(µ) cos θ(MGUT ) 1 20 MGUT
= + (F + F 0 ) ln .
α(µ) α(MGUT ) 6π 3 µ
As discussed in the text, the couplings are assumed to unify at the scale of grand unification
α(MGUT ) = αS (MGUT ) ≡ αGUT , and the angles are normalized by tan2 θ(MGUT ) = 3/5.
The best data we have is provided at the mass of the Z boson, MZ = 91.188 GeV. We
use the first set of equations to specify αS (M 0 ), α(M 0 ) and sin2 θ(M 0 ) in terms of the param-
eters evaluated at µ = MZ . We then use these values of the parameters at M 0 as inputs into
the second set of equations, which run up to MGUT . This results in
1 1 1 MGUT 1 0 MGUT
= − (4F − 33) ln − 4F ln
αS (MGUT ) αS (MZ ) 6π MZ 6π M0
sin2 θ(MGUT ) sin2 θ(MZ )
1 MGUT 1 0 MGUT
= − (4F − 22) ln − 4F ln
α(MGUT ) α(MZ ) 6π MZ 6π M0
cos2 θ(MGUT ) cos2 θ(MZ )
1 20 MGUT 1 20 0 MGUT
= − F ln − F ln .
α(MGUT ) α(MZ ) 6π 3 MZ 6π 3 M0
You can see what is happening: if F 0 = 0, then we recover the equations on p. 414 with
µ = MZ . The presence of the new families when F 0 6= 0 affects the running between the
245
scales M 0 and MGUT .
See W. J. Marciano, “Weak mixing angle and grand unified gauge theories,” Phys. Rev.
D, Vol. 20 No. 1, 1 July 1979 for running the renormalization group flow to compare SU (5)
GUT predictions with experiment.
2. Work out proton decay in detail. Derive relations between the following decay rates:
Γ(p → π 0 e+ ), Γ(p → π + ν̄), Γ(n → π − e+ ), and Γ(n → π 0 ν̄).
Solution:
We will use two-component spinor notation (see Appendix E). As explained on p. 407 in
the text, we write all fields as left-handed with the following transformation properties under
GSM ≡ SU (3)c ⊗ SU (2)W ⊗ U (1)Y :
u 1 2 ¯ 1 ν 1
q≡ ∼ (3, 2, ), ū ∼ (3̄, 1, − ), d ∼ (3̄, 1, ), ` ≡ ∼ (1, 2, − ), ē ∼ (1, 1, 1)
d 6 3 3 e 2
The indices α, β, γ run from 1 to 3 and denote the 3-representation of SU (3)c , and the indices
i, j run from 1 to 2 and denote the 2-representation of SU (2)W . The antisymmetric tensor
εαβγ is invariant under SU (3)c and the antisymmetric tensor εij is invariant under SU (2)W .
d¯α
The SU (5) fermion fields are (ψ5̄ )µ ≡ and (ψ10 )µν = −(ψ10 )νµ , whose components are
α `i
u
(ψ10 )αβ = εαβγ ūγ , (ψ10 )αi = qiα = and (ψ10 )ij = εij ē. As pointed out in problem VII.5.1,
dα
we have (ψ10 )αi = εij (ψ10 )αj so that (ψ10 )α4 = ε45 dα = +dα and (ψ10 )α5 = ε54 uα = −uα .
Let us now work out the new currents that arise from the SU (5) unified theory. The covariant
derivative acting on a 5 of SU (5) is
24
X
(Dψ5 )µ = ∂ψ5µ − ig X µν ψ5ν , X µν = X a (T5a )µν .
a=1
At this point we should note an annoying circumstance: Since we have already used µ =
1, ..., 5 as an index for SU (5), α = 1, 2, 3 as an index for SU (3), i = 1, 2 as an index for
SU (2) and a = 1, ..., 24 as an index for the adjoint representation of SU (5), we will denote
Lorentz vector indices by M = 0, 1, 2, 3, and Lorentz spinor indices by m = 1, 2 and ṁ = 1, 2
when we choose to display them explicitly.
246
The adjoint components X a are the same for the 5 and for the 5̄. The generators of the 5̄ are
T5̄a = −(T5a )∗ = −(T5a )T (since T5a are hermitian, we have (T5a )∗ = (T5a )T ). In components,
this means (T5̄a )µν = −[(T5a )T ]µν = −(T5a )ν µ . Therefore the covariant derivative of a 5̄ is
24
X
(Dψ5̄ )µ = ∂ψ5̄µ + ig X a (T5a )ν µ ψ5̄ν
a=1
This form for the current term is awkward because the SU (5) indices are not in matrix
multiplication order. To take care of this properly, we should display the Lorentz indices:
ψ5̄† µ X a (T5a )ν µ ψ5̄ν = XM
a
(ψ5̄† µ )ṁ σ̄ M ṁm (T5a )ν µ (ψ5̄ν )m
a †µ m
= − XM (ψ5̄ν )m σm
M a ν
ṁ (T5 ) µ (ψ5̄ )
247
Putting together the contributions from the 5̄ and the 10, we have
∆B6=0 α ¯ M †i ij †γ M β †i M
L = g(XM )i dα σ ` − ε εαβγ ū σ̄ qj + qα σ̄ ē + h.c.
ν
Expanding out the SU (2) index with `i = , we arrive at the interactions
e
Recall the electric charge of the down quark Q(d) = − 31 . Since Q(dν ¯ † ) = −Q(d) + Q(ν) =
1
+ 3 + 0, invariance under electromagnetism tells us that the gauge boson (XM )α4 has electric
charge − 31 . Since Q(de ¯ † ) = −Q(d) − Q(e) = + 1 + 1 = + 4 , the gauge boson (XM )α has
3 3 5
electric charge − 34 . Let us therefore write (X −1/3 )αM ≡ (XM )α4 and (X −4/3 )αM ≡ (XM )α5 and
rewrite the above interactions as
(J +1/3 )M ¯ M † †γ M β † M
α ≡ dα σ ν − εαβγ ū σ̄ d + uα σ̄ ē
As a consistency check, we should verify the electric charges of all of the terms in the currents.
We have Q(ū† d) = +Q(u)+Q(d) = + 32 − 31 = + 13 and Q(u† ē) = −Q(u)−Q(e) = − 23 +1 = + 13 ,
which both match Q(dν ¯ † ) = + 1 . Also Q(ū† u) = 2Q(u) = + 4 and Q(d† ē) = −Q(d) − Q(e) =
3 3
+ 13 + 1 = + 43 , which match Q(de ¯ † ) = + 4 as they must.
3
Suppose the gauge bosons X −1/3 and X −4/3 have the same GUT-scale mass MX ∼ 1016
GeV so that they can be integrated out at low energies. Performing this integration gives the
low-energy effective Lagrangian for baryon-number-violating processes such as proton decay:
g 2 +1/3 M −1/3 α
L∆B6 =0
)M + (J +4/3 )M −4/3 α
eff = 2
(J )α (J α (J )M .
MX
The products of currents in this Lagrangian are
J +1/3 J −1/3 = (d¯α σ M ν † − εαβγ ū†γ σ̄ M dβ + u†α σ̄ M ē)(νσM d¯†α − εαδ d†δ σ̄M ū + ē† σ̄M uα )
J +4/3 J −4/3 = (d¯α σ M e† + εαβγ ū†γ σ̄ M uβ + d† σ̄ M ē)(eσM d¯†α + εαδ u† σ̄M ū + ē† σ̄M dα )
α δ
For this problem, we are interested in only those interactions that contribute to proton decay
and neutron decay, so we are interested in operators of the schematic form ∼ qqq`. These
are:
J +1/3 J −1/3 = (d¯α σ M ν † )(−εαβγ d†β σ̄M ūγ ) + (−εαβγ ū†γ σ̄ M dβ )(ē† σ̄M uα ) + h.c.
and
J +4/3 J −4/3 = (d¯α σ M e† )(+εαβγ u†β σ̄M ūγ ) + (+εαβγ ū†γ σ̄ M uβ )(ē† σ̄M dα ) + h.c.
248
These can be simplified using the identities
(σ M )mṁ (σ̄M )ṅn = 2 δm
n ṅ
δṁ and (σ̄ M )ṁm (σ̄M )ṅn = 2 εmn εṁṅ
where we use the metric η = (+, −, −, −) and the convention ε12 = ε1̇2̇ = +1. Using these, we
have (ξσ M ν † )(ψ † σ̄M χ) = −2(ξχ)(ψ † ν † ) and (ξ † σ̄ M ν)(ψ † σ̄M χ) = +2(ξ † ψ † )(νχ) for any four
Weyl spinors ξ, ν, ψ and χ. (Relations like these are called Fierz identities.) So the above
products of currents simplify to
J +1/3 J −1/3 = +2 εαβγ (d¯α ūγ )(d†β ν † ) − 2 εαβγ (ū†γ ē† )(dβ uα )
J +4/3 J −4/3 = −2 εαβγ (d¯α ūγ )(u† e† ) + 2 εαβγ (ū†γ ē† )(uβ dα ) .
β
Therefore, the part of the baryon-number violating effective Lagrangian relevant for pro-
ton decay is
∆B6=0 i + 1 0 † † i + † 1 0 †
Leff =λ e p+ π n + 2 π p + 2ē p̄ −
√ π n̄ + 2 π p̄
√ + h.c.
f f
where we have defined the coupling
µ32
λ ≡ 2g
MX2
which has dimensions of mass.
Since all calculations in the text are done with 4-component Dirac spinors instead of 2-
component spinors, let us now rewrite L∆B6
eff
=0
in terms of Dirac spinors. Define
c ē p n
E ≡ † , P≡ , N ≡
e p̄† n̄†
29
See the addendum to this chapter of solutions.
249
where E c is the Dirac
field for the positron, or in other words the charge-conjugate of the
e
Dirac field E ≡ † for the electron. Since E c PL = ep and E c PR = ē† p̄† , we have:
ē
∆B6=0 c
i 0
Leff = λ E (PL + 2PR ) + π (PL − 2PR ) P + h.c.
2fπ
i
+λ√ π + E c (PL − 2PR )N + h.c.
2 fπ
where PL ≡ 12 (I − γ 5 ) and PR ≡ 1
(I + γ 5 ), and we have defined a modified pion decay
√ 2
constant fπ ≡ f / 2 .
We will first compute the rate for p → π 0 e+ . We see from the above Lagrangian that
one contribution to the amplitude arises from the cubic π 0 e+ p vertex
e+
p = − P L − 2P R
2f
0
There is also a contribution from the pion-nucleon interactions derived from the chiral La-
grangian:
gA
∂µ π 0 Pγ µ γ 5 P − N γ µ γ 5 N
Lπ0 pn =
2fπ
where gA is the axial vector coupling. The contributing diagram to the proton decay ampli-
tude is
p e+
where the solid dot indicates the iλE c (PL + 2PR )P vertex from L∆B6 eff
=0
. Let k0 , k1 , k2 be
the momenta of the proton, positron, and pion respectively. Then the second diagram con-
tributes:
gA 5
ū1 [iλ (PL + 2PR )] [iSp (k1 )] − 6 k2 γ u0
2fπ
λgA 6 k1 + mp
=+ ū1 (PL + 2PR ) 2 6 k2 γ 5 u0 .
2fπ k1 − m2p
Using γ 5 γ µ , we can move the 6 k1 to the left of PL + 2PR , picking up some minus signs. We
will neglect the electron mass me mp , and so ū1 6 k1 = ū1 me ≈ 0. In the denominator, we
also have k12 = m2e ≈ 0.
Furthermore, we have k2 = k0 − k1 , so that again we can move 6 k1 to the left and get
250
zero when acted on by ū1 . Finally, using PL γ 5 = −PL and PR γ 5 = +PR , we arrive at two
powers of mp in the numerator to cancel out the 1/m2p from the propagator and obtain:
λgA
− ū1 (PL − 2PR )u0
2fπ
for the total contribution to the amplitude from the second diagram.
This is exactly the same as the contribution from the π 0 e+ p vertex, up to a factor of gA . The
amplitude is
λ
iM = −(1 + gA ) ū1 (PL − 2PR )u0 .
2fπ
Taking the magnitude squared, averaging over the proton spins and summing over the
positron spins using the usual manipulations gives
X 5 2 m2p − m2π
1
2
|M|2 = λ (1 + gA ) 2
s0 ,s1
32fπ2 mp me
where we have set me ≈ 0 in the numerator but kept the mass of the neutral pion. The decay
rate is (p. 141)
d3 k1 d3 k2
Z X
4 4
Γ= 3 3
(2π) δ (k0 − k1 − k2 ) 1
2
|M|2
(2π) (ω1 /me ) (2π) 2ω2 s ,s 0 1
p
where ωi = |~pi |2 + m2i . The amplitude is a constant, and the 2-body phase space integral
is given by equation (40) on p. 142 suitably adjusted to account for the fermion factor ω1 /me :
2
d3 k1 d3 k2 4 mp − m2π
Z
δ (k0 − k1 − k2 ) ≈ πme
(ω1 /me ) 2ω2 m2p
where again we have taken me ≈ 0 wherever possible. The decay rate is therefore
5 λ2 (1 + gA )2 (m2p − m2π )2
Γ(p → e+ π 0 ) = .
128π fπ2 m3p
Actually this is still not quite complete, since the coefficients of the hadronic operators in the
effective Lagrangian are subject to renormalization group corrections. See F. Wilczek and
A. Zee, “Operator analysis of nucleon decay,” Phys. Rev. Lett. Vol. 43 No. 21, 19 Nov 1979.
Using the effective Lagrangian, the other decay rates p → ν̄π + , n → e+ π − and n → ν̄π 0
may be computed in exactly the same way. For general relations between the rates, see
problems VIII.3.3 and VIII.3.4.
251
3. Show that SU (5) conserves the combination B −L. For a challenge, invent a grand unified
theory that violates B − L.
Solution:
After ϕµ acquires a vacuum expectation value, the U (1)X symmetry gets broken, but a
subgroup remains unbroken. Recall that in terms of low-energy fields, the SU (5) fermions
d¯α
α
αβ αβγ α α u
are (ψ5̄ )µ ≡ for the 5̄, and for the 10: (ψ10 ) = ε ūγ , (ψ10 )i = qi =
`i dα
and (ψ10 )ij = εij ē. Again recalling problem VII.5.1, we have (ψ10 )αi = εij (ψ10 )αj so that
(ψ10 )α4 = ε45 dα = +dα and (ψ10 )α5 = ε54 uα = −uα .
We can thereby evaluate the generator 21 Y of hypercharge on the ψµ and ψ µν to find that the
combination X + 4(Y /2) generates a symmetry even after spontaneous symmetry breaking of
SU (5). The other conserved generators are Q, the generator of electric charge, and {T a }8a = 1 ,
the generators of color SU (3). By explicitly evaluating the generator X + 4(Y /2) on the
components of ψµ and ψ µν , we will find that X + 4(Y /2) is a multiple of B − L:
[X + 4( 12 Y )]d¯ = [−3 + 4(+ 13 )]d¯ = 5(− 13 )d¯ (← B = − 13 , L = 0 X)
[X + 4( 12 Y )]` = [−3 + 4(− 12 )]` = 5(−1)` (← B = 0 , L = +1 X)
[X + 4( 12 Y )]ū = [+1 + 4(− 23 )]ū = 5(− 13 )ū (← B = − 13 , L = 0 X)
[X + 4( 21 Y )]q = [+1 + 4(+ 16 )]q = 5(+ 13 )q (← B = + 13 , L = 0 X)
[X + 4( 12 Y )]ē = [+1 + 4(+1)]ē = 5(+1)ē (← B = 0 , L = −1 X)
Therefore, B − L = 15 [X + 4( 12 Y )].
252
is a generator of the SO(10) theory due to the presence of the right-handed gauge-singlet
neutrino which can be assigned a lepton number of −1.
One way to violate B − L in the SU (5) theory is to add another Higgs field, H, transforming
as H µν ∼ 5 ⊗A 5 = 10. The field H couples to fermions through the Yukawa interaction
where the repeated indices I, J label the different families31 . At this stage H could simply be
assigned a charge under X such that B − L remains conserved, but in the absence of further
restrictions there is a cubic scalar interaction:
L = µ H µν H ρσ ϕλ εµνρσλ
where µ is a coupling with dimensions of mass. The clash between these two terms violates
B − L by two units and thus implies (B − L)-violating processes such as n → µ− K + and
p → µ− K + π + .
Here we review the nonlinear sigma model used to parameterize low-energy QCD, also known
as the chiral Lagrangian. For more details, see M. Claudson and M. Wise, “Chiral Lagrangian
for deep mine physics,” Nucl. Phys. B195 (1982) 297-307 and O. Kaymakcalan, L. Chong-
Huah and K. C. Wali, “Chiral Lagrangian for proton decay,” Phys. Rev. D, Vol. 29 No. 9,
1 May 1984.
¯ s̄) are the two-component spinors for the up, down and
where q = (u, d, s) and q̄ = (ū, d,
µ
6 ≡ σ̄ Dµ is the gauge-covariant derivative for each quark field.
strange quarks, and D̄
This Lagrangian exhibits the global symmetry SU (3)L ⊗ SU (3)R , under which the quarks
transform as q ∼ (3, 1) and q ∼ (1, 3̄). We will use the indices A = 1, 2, 3 and A0 = 1, 2, 3 to
label the 3-dimensional representations of SU (3)L and SU (3)R respectively. We also choose
upper indices for the fundamental, 3, and lower indices for the anti-fundamental, 3̄. There is
also the global U (1)V symmetry (q, q̄ † ) → e−iθ (q, q̄ † ). The axial U (1)A : (q, q̄) → e−iθ (q, q̄) is
anomalous.
31
Since H µν = −H νµ , the coupling f must be antisymmetric: fIJ = −fJI . Thus this interaction couples
one generation of fermions to another. This is reminiscent of the interaction fab h+ εij `ai `bj present in the
Zee model of neutrino masses. See problem VIII.3.2.
253
As discussed in the text, the QCD vacuum is supposed to break the global symmetry
SU (3)L ⊗ SU (3)R down to the diagonal subgroup SU (3)V through the formation of a chiral
condensate:
hq A q̄A0 i = −v 3 δ AA0
where v is a positive parameter with dimensions of mass, and the minus sign arises from the
fact that fermions contribute negative energy to the vacuum.
We may map a spacetime point xµ to any point on the group manifold of SU (3)A by treating
the components of the meson octet as fields and exponentiating them to form a 3-by-3 special
unitary spacetime-dependent matrix:
Σ(x) ≡ e i2Π(x)/f
where f is the pion decay constant, f ≈ 139 MeV. Under a general SU (3)L ⊗ SU (3)R trans-
formation, the matrix Σ transforms as Σ → LΣR† , where L is an element of SU (3)L and
R is an element of SU (3)R . The diagonal subgroup SU (3)V is given by transformations for
254
which L = R.
It will prove convenient to define the “square root” of the matrix Σ as ξ(x) ≡ eiΠ(x)/f .
Under a general SU (3)L ⊗ SU (3)R transformation, the matrix ξ is defined to transform as
ξ → LξU † = U ξR†
where U is a 3-by-3 unitary matrix that is defined by the above transformation law and
thereby depends nonlinearly on the matrices L, R and Π. This ensures that Σ = ξ 2 trans-
forms as stated.
After seeing how the mesons arise from chiral symmetry breaking, we need the spectrum
of baryons at low energy.32 The quark fields transform under SU (3)L ⊗ SU (3)R as
u ū
q ≡ d ∼ (3, 1) and q̄ ≡ d¯ ∼ (1, 3̄) .
s s̄
Therefore q̄ † transforms as (1, 3), and
q̄ † q̄ † ∼ (1, 3) ⊗ (1, 3) = (1, 3̄A ) ⊕ (1, 6S ) .
A baryon is a color-singlet bound state of three quarks, so the color indices of the above
product of quarks are to be contracted with the 3-index antisymmetric tensor of SU (3)c .
The Lorentz-singlet part33 of q̄ † q̄ † forms a color anti-triplet
d¯ s̄
†β †γ
†β †γ
Qα ≡ 2 εαβγ q̄ q̄ = s̄†β ū†γ
1
ū†β d¯†γ
which transforms as (1, 3̄) under global SU (3)L ⊗ SU (3)R transformations. (Here an up-
per α = 1, 2, 3 denotes the fundamental of color and a lower index α denotes the anti-
fundamental.)
255
At low energy, the QCD vacuum breaks SU (3)L ⊗ SU (3)R → SU (3)V , under which (3, 3̄) →
3 ⊗ 3̄ = 1 ⊗ 8. We therefore find an octet of baryons34
1
√ Σ0 + √1 Λ0 Σ +
p
2 Σ− 6 − √12 Σ0 + √16 Λ0 n
B=
Ξ− Ξ0 − √26 Λ0
The next step is to write all possible interaction terms between baryons and mesons that
respect SU (3)L ⊗ SU (3)R symmetry and parity. The result is
The factor of 18 f 2 is so that the meson fields that comprise Σ = e i2Π/f have canonically
normalized kinetic terms, e.g. 12 ∂µ π 0 ∂ µ π 0 and ∂µ π + ∂ µ π − . The parameters F = 0.44 and
D = 0.81 are measured from semileptonic baryon decays.
Expanding LπB using ξ = I + iΠ/f + ... and the components of the baryon octet, we find
interactions between the protons, neutrons and pions:
256
√
and a modified pion decay constant fπ = f / 2 , so that the coupling of the neutral pion to
the proton and neutron can be written as
gA
∂µ π 0 Pγ µ γ 5 P − N γ µ γ 5 N
Lπ0 pn =
2fπ
gA
= ∂µ π 0 Ψγ µ γ 5 I3 Ψ ,
fπ
1
P +2 0
where Ψ ≡ is the nucleon doublet, and I3 ≡ is the third generator of
N 0 − 12
isospin.
Introducing nonzero quark masses results in nonzero masses for the meson octet and splits
the degeneracy among the masses in the baryon octet. For further details, see the references.
Solution:
The Clifford algebra for d = 2n + 1 is that of d = 2n with the addition of the γ FIVE from
d = 2n. For example, consider d = 3. The gamma matrices for d = 2 are just γ 1 = σ 1
and γ 2 = σ 2 as explained in the text. The chirality matrix γ FIVE ≡ −iγ 1 γ 2 = −iσ 1 σ 2 = σ 3
anticommutes with γ 1 and γ 2 , and it squares to 1. Therefore the Clifford algebra for d = 3 is
{γ i , γ j } = 2δ ij , γ i = σ i
which you already encountered way back in problem II.1.12, where you discovered that the
Dirac mass term in (2 + 1)-dimensional spacetime violates parity and time reversal. In gen-
eral, the γ FIVE for d = 2n anticommutes with the {γ i }2n i = 1 from d = 2n (since 2n is always
even) and squares to 1, and therefore will form a perfectly good (2n + 1)th gamma ma-
trix γ 2n+1 ≡ γ FIVE ≡ (−i)n γ 1 γ 2 ...γ 2n for d = 2n + 1. The point is that while the spinor
representation of SO(2n) is reducible into two chiral irreducible representations, the spinor
representation of SO(2n + 1) is not reducible.
Solution:
257
where µ, ν = 0, 1, 2, ..., d − 1. Let i, j = 1, 2, ..., d − 1 denote purely spatial indices. Then the
above can be written as
All we have to do is to take the Clifford algebra for d-dimensional Euclidean space and throw
in some factors of i to generate the appropriate minus signs in the metric. Immediately we
find
0 0 j j
γSO(d−1,1) = γSO(d) , γSO(d−1,1) = iγSO(d)
for j = 1, ..., d − 1. Note that we are using slightly different notation for the SO(d) gamma
matrices from the text. In the chapter, the vector indices run from 1 to d, whereas here our
indices run from 0 to d − 1. If you want to stick to the notation in the book, you can write
0 1 j j+1
γSO(d−1,1) = γSO(d) , γSO(d−1,1) = iγSO(d)
for j = 1, ..., d − 1.
3. Show that the Clifford algebra for d = 4k and for d = 4k + 2 have somewhat differ-
ent properties. (If you need help with this and the two preceding exercises, look up F.
Wilczek and A. Zee, Phys. Rev. D25: 553, 1982.)
Solution:
FIVE
The
Q chirality
matrix γ = σ3 ⊗...⊗σ3 acts on a chiral spinor |ε1 , ..., εn i as γ FIVE |ε1 , ..., εn i =
n
j = 1 εj |ε1 , ..., εn i. The “left-handed” chiral spinor |ε1 , ..., εn iL is defined by
γ FIVE |ε1 , ..., εn i = (−1)|ε1 , ..., εn i, and the “right-handed” chiral
Qspinor|ε1 , ..., εn iR is defined
n
by γ FIVE |ε1 , ..., εn iR = (+1)|ε1 , ..., εn iR . That is, the sign of j = 1 εj determines whether
the spinor is left-handed or right-handed.
Now, the charge conjugation matrix C = iσ2 ⊗ ... ⊗ iσ2 acts on the SO(2n) chiral spinor
|ε1 , ..., εn i as C|ε1 , ..., εn i = |(−ε1 ), ..., (−εn )i, which means that the chirality matrix γ FIVE
acts on a charge-conjugated spinor as
Let’s process what this means. Consider the case for which n is even, so that (−1)n = +1. For
that case, if γ FIVE |ε1 , ..., εn i = +γ FIVE C|ε1 , ..., εn i, then γ FIVE C|ε1 , ..., εn i = +C|ε1 , ..., εn i.
If γ FIVE |ε1 , ..., εn i = −γ FIVE C|ε1 , ..., εn i, then γ FIVE C|ε1 , ..., εn i = −C|ε1 , ..., εn i. That is,
|ε1 , ..., εn i and C|ε1 , ..., εn i have the same eigenvalue under γ FIVE ; the states |ε1 , ..., εn iL and
|ε1 , ..., εn iR are therefore self-conjugate (that is, not conjugate to each other) if n is even.
258
Now consider the case for which n is odd, meaning (−1)n = −1. Under that assump-
tion, we have the opposite situation from before. If γ FIVE |ε1 , ..., εn i = +γ FIVE C|ε1 , ..., εn i,
then γ FIVE C|ε1 , ..., εn i = −C|ε1 , ..., εn i. If γ FIVE |ε1 , ..., εn i = −γ FIVE C|ε1 , ..., εn i, then we
have γ FIVE C|ε1 , ..., εn i = +C|ε1 , ..., εn i. That is, |ε1 , ..., εn i and C|ε1 , ..., εn i have opposite
eigenvalues under γ FIVE ; the states |ε1 , ..., εn iL and |ε1 , ..., εn iR are therefore conjugate to
each other if n is odd.
As an important aside, note that the situation is reversed for SO(d − 1, 1), that is for
d-dimensional Minkowski spacetime rather than d-dimensional Euclidean spacetime. For
d = 4k (that is, d = 0 mod 4), the two chiral spinor representations are conjugate to each
other. For d = 2 + 4k (d = 2 mod 4), each chiral spinor is its own conjugate.
If you desire a second reference for this problem (and other properties of SO(2n) and
SO(2n − 1, 1) spinors), consult Volume II, Appendix B of J. Polchinski’s textbook on string
theory.
4. Discuss the Higgs sector of the SO(10). What do you need to give mass to the quarks
and leptons?
Solution:
Let ψa ∼ 2n−1 (with a = 1, ..., 2n−1 ) be the SO(2n) spinor that contains a single family
of matter fields. (We will specialize to n = 5 later.) The SO(2n) invariant tensors at our
disposal are the charge conjugation matrix C ab and its inverse Cab , and the gamma matrices
(γ µ )ab , where µ = 1, ..., 2n denotes the index for the vector (“defining”) representation of
SO(2n). Note that once we pick the convention for which the spinor ψ has one index down,
the rest of the index placements are fixed. For example, an SO(2n) transformation acts as
µν
ψ → e iωµν σ ψ = ψ + iωµν σ µν ψ + O(ω 2 ), so δψa = iωµν (σ µν )ab ψb . Since σ µν = 2i [γ µ , γ ν ], the
index placements must be (γ µ )ab . We can raise and lower indices using C and C −1 , which is
the whole point of defining such a matrix in the first place. The conventions we use are that
indices are raised and lowered by contracting with C or C −1 always on the second index:
ψ a ≡ C ab ψb and ψa ≡ Cab ψ b . This will introduce some minus signs, since in SO(2n) we have
the property
C T = (−1)n(n+1)/2 C (A11)
For a thorough discussion of this point as well as for other aspects of SO(2n), consult the
259
paper “Families from Spinors,” Phys. Rev. D25: 553, 1982 by F. Wilczek and A. Zee, as
referenced in problem VII.7.3. (We label the above equation as (A11) since that is the equa-
tion number in the reference.)
To write a mass term for the fermions contained within the spinor ψa , we need a term
quadratic in ψ that is invariant under SO(10) (and under the Lorentz group). For a single
family of fermions, we are restricted to terms of the form
where φµ1 ...µK is a Lorentz-scalar field that is completely antisymmetric in its SO(2n) vector
indices. (y is just a coupling constant.) We will suppress the Lorentz-spinor indices using
the conventions in Appendix E.
From the property (A11), raising and lowering a pair of up-and-down indices introduces the
sign (−1)n(n+1)/2 , meaning: ψ a ψa = (−1)n(n+1)/2 ψa ψ a , where we remind you that ψ a ≡ C ab ψb
and ψb ≡ Cbc ψ c . This means that raising and lowering two pairs of up-and-down indices can
always be done without changing a sign regardless of the value of n. In other words, for any
matrix Ma b , we have ψ a Ma b ψb = ψa M ab ψ b . So the above Lagrangian can equally well be
written as
L = −y ψ a Cab (γ µ1 ...γ µK )b c C cd ψd φµ1 ...µK + h.c.
The matrix sandwiched between the ψs is C −1 (γ µ1 ...γ µK )T C. We can use the property
C −1 (γ µ )T C = (−1)n γ µ (A10)
C −1 (γ µ1 ...γ µK )T C = C −1 (γ µK )T ...(γ µ1 )T C
= C −1 (γ µK )T CC −1 ...CC −1 (γ µ1 )T C
= (−1)Kn γ µK ...γ µ1
= (−1)Kn (−1)K−1 γ µ1 γ µK γ µK−1 ...γ µ2
= (−1)Kn (−1)K−1 (−1)K−2 γ µ1 γ µ2 γ µK γ µK−1 ...γ µ3
= ...(keep on anticommuting)...
PK−1
j µ1
= (−1)Kn (−1) j=1 γ ...γ µK
= (−1)Kn (−1)K(K−1)/2 γ µ1 ...γ µK
Therefore:
ψ a Cab (γ µ1 ...γ µK )b c C cd ψd = (−1)Kn+K(K−1)/2 ψ a (γ µ1 ...γ µK )ad ψd
Now recall what we said about contracting indices with the second index of the charge
conjugation matrix:
ψ a ≡ C ab ψb = ψb C ab = ψb (−1)n(n+1)/2 C ba .
260
Finally, after relabeling some dummy indices, we have the result
ψ a Cab (γ µ1 ...γ µK )b c C cd ψd = (−1)Kn+K(K−1)/2+n(n+1)/2 ψ a Cab (γ µ1 ...γ µK )b c C cd ψd
In other words, the mass term for ψ can only be nonzero when the quantity
K(K − 1) n(n + 1) (n + K)(n + K + 1)
Kn + + = −K
2 2 2
is an even number. (This is (A55) in the reference.) Since odd+odd = even and even+even
= even while odd+even = odd, this implies that K and 12 (n + K)(n + K + 1) are both even
or are both odd. Since odd×odd = odd, odd×even = even and even×even = even, we will
always have (n + K)(n + K + 1) = even. So the question is where 12 (n + K) is even or odd.
n+K
= even ≡ 2p =⇒ K = 4p − n
2
n+K
= odd ≡ 2q + 1 =⇒ K = 4q + 2 − n
2
where p and q are any integers, and of course we must have K > 0. The case at hand is
SO(10), which means n = 5. This implies
5+K
K = 4p − 5 = 3, 7, 11, ... = even
2
5+K
K = 4p + 2 − 5 = 1, 5, 9, ... = odd
2
Moreover, recall that SO(10) has the invariant tensor εµ1 ...µ10 . That means any number of
antisymmetrized indices greater than 5 is equivalent to a number less than 5. For example,
the definition φµ1 = εµ1 µ2 ...µ10 φµ2 ...µ9 is a statement invariant under SO(10), so K = 9 is
equivalent to K = 1. So the above is really
5+K
K = 4p − 5 = 3 = even
2
5+K
K = 4p + 2 − 5 = 1, 5 = odd
2
for SO(10). However, the first of these statements is a contradiction. Recall that we said
we must have 21 (5 + K)(6 + K) and K either both even or both odd. For K = 3, we have
1
2
(5 + 3)(6 + 3) = 12 × 8 × 9 = 36, which is even. Finally we have deduced that only the terms
for which K = 1 or 5 are allowed in the Lagrangian. (This is the conclusion reached below
(A56b) in the reference.) In other words, the masses for one family of fermions comes from
the Lagrangian
L = −y ψCγ µ ψ φµ − y 0 ψCγ µ1 ...γ µ5 ψ φ0µ1 ...µ5 + h.c.
where φ and φ0 are Lorentz-scalar fields, and y and y 0 are coupling constants. From looking
at the indices (and recalling that they are all implicitly antisymmetrized), we see the trans-
formation properties φµ ∼ 10 and φ0µ1 ...µ5 ∼ 10 ⊗A 10 ⊗A 10 ⊗A 10 ⊗A 10 = 126 of SO(10).
261
Note that for multiple families of fermions, the analysis becomes more complicated. The
mass terms have the form
X
L=− yF F 0 ψF Cγ µ1 ...γ µK ψF 0 φµ1 ...µK + h.c.
F,F 0
where F and F 0 run from 1 to the number of fermion families. Lorentz-scalars φ satisfying the
previous constraints would contribute terms for which yF F 0 = yF 0 F , and those not satisfying
those constraints would contribute terms for which yF F 0 = −yF 0 F .
5. The group SO(6) has 6(6 − 1)/2 = 15 generators. Notice that the group SU (4) also
has 42 − 1 = 15 generators. Substantiate your suspicion that SO(6) and SU (4) are isomor-
phic. Identify some low dimensional representations.
Solution:
γ FIVE ≡ −iγ 0 γ 1 γ 2 γ 3 γ 4 γ 5
FIVE
= +i(γ(4) ⊗ σ34 )(I(4) ⊗ σ1 σ2 )
FIVE
= +i(γ(4) ⊗ I)(I(4) ⊗ iσ3 )
FIVE
= −γ(4) ⊗ σ3
FIVE
−γ(4)
0(4)
= FIVE
0(4) +γ(4)
which commutes with σ µν . Thus if Ψ transforms as a 23 -dimensional spinor, the chiral spinors
ΨL ≡ 21 (1 − γ FIVE )Ψ and ΨR ≡ 12 (1 + γ FIVE )Ψ have 23−1 = 4 components and transform
irreducibly under SO(6).
Since the chiral spinors are 4-dimensional, the Lorentz generators σ µν acting on them should
split up into two sets of fifteen 4-by-4 traceless hermitian matrices. The generators of SU (4)
262
are precisely the set of fifteen 4-by-4 traceless hermitian matrices, so that constructing the ir-
reducible rotation generators acting on the chiral spinor would constitute a proof that SO(6)
is locally isomorphic to SU (4).
The rotation generators σ µν take the following forms in terms of 4-dimensional matrices.
For µ, ν = 0, 1, 2, 3 only, we have:
µ
σ µν = i 21 [γ µ , γ ν ] = i 12 [γ(4) ν
⊗ σ3 , γ(4) ⊗ σ3 ]
µ ν µ
= i 21 γ(4) γ(4) ⊗ σ32 − γ(4) ν
γ(4) ⊗ σ32
µ ν
= i 12 [γ(4) , γ(4) ]⊗I
µν
µν σ(4) 0(4)
= σ(4) ⊗ I = µν (µ, ν = 0, 1, 2, 3 only) .
0(4) σ(4)
and
µ5 µ 5 µ
σ = iγ γ = i ⊗ σ3 I(4) ⊗ σ2
γ(4)
µ
µ 0(4) γ(4)
= iγ(4) ⊗ (−iσ1 ) = µ (µ = 0, 1, 2, 3 only) .
γ(4) 0(4)
σ 45 = iγ 4 γ 5 = i I(4) ⊗ σ1 I(4) ⊗ σ2
−I(4) 0(4)
= iI(4) ⊗ (iσ3 ) = .
0(4) +I(4)
Let us now examine the chiral spinors in more detail. Consider the left-handed chiral spinor
ΨL ≡ PL Ψ, where
1 FIVE
1 FIVE (I + γ(4)
2 (4)
) 0(4) pR 0
PL ≡ 2 (I − γ )= 1 FIVE ≡ .
0(4) (I − γ(4)
2 (4)
) 0 pL
263
In terms of these, the SO(6) left-handed chiral spinor is
0
χ̄α̇
pR ψ
ΨL = 0 χ0β .
=
pL ψ
0
So if we want to extract the parts of the rotation generators σ µν that act only on these
components, then we may consider the matrices σLµν ≡ PL σ µν PL .
It is now helpful to write the Dirac matrices for (3 + 1)-dimensional spacetime in terms
of the Pauli matrices:
(σ µ )αβ̇
µ 0
γ(4) =
(σ̄ µ )α̇β 0
where numerically we have (σ µ )αβ̇ = (I, i~σ ) and (σ̄ µ )α̇β ≡ εαβ εα̇β̇ (σ µ )β β̇ = (I, −i~σ ).
For µ, ν = 0, 1, 2, 3, we have:
µν
µν σ(4) 0(4)
σ = µν
0(4) σ(4)
1 µ ν
i 2 (σ σ̄ − σ ν σ̄ µ ) 0(2) 0(2) 0(2)
0(2) i 12 (σ̄ µ σ ν − σ̄ ν σ µ ) 0(2) 0(2)
= .
0(2) 0(2) i 21 (σ µ σ̄ ν − σ ν σ̄ µ ) 0(2)
1 µ ν ν µ
0(2) 0(2) 0(2) i 2 (σ̄ σ − σ̄ σ )
Therefore, the middle 4-by-4 block of σLµν (for µ, ν = 0, 1, 2, 3 only) is given by the matrix
1 µ ν
µν i 2 (σ̄ σ − σ̄ ν σ µ ) 0(2)
W = .
0(2) i 21 (σ µ σ̄ ν − σ ν σ̄ µ )
The middle 4-by-4 block of σLµ4 (for µ = 0, 1, 2, 3 only) is given by the matrix
µ 0(2) +iσ̄ µ
X =
−iσ µ 0(2)
The 12 (4 × 3) = 6 matrices W µν , the 4 matrices X µ , the 4 matrices Y µ and the matrix Z con-
stitute 6+4+4+1 = 15 traceless hermitian 4-by-4 matrices. These therefore generate SU (4),
264
and we have proven that SO(6) and SU (4) are locally isomorphic.
We also have ψAB ∼ 4 ⊗ 4̄ = 1 ⊕ 15, where the singlet is the trace tr ψ = δBA ψAB and the 15
is comprised of the traceless hermitian matrices displayed previously, or in other words the
adjoint representation of SU (4). Also ψ [µν] ∼ 6 ⊗A 6 ∼ 6×5
2
= 15 of SO(6), which is the set
of traceless antisymmetric 6-by-6 matrices and therefore the adjoint of SO(6). Indeed, the
adjoint of SO(6) is equivalent to the adjoint of SU (4), as we have just proven by showing
that the two groups are locally isomorphic.
6. Show that (unfortunately) the number of families we get in SO(18) depends on which
subgroup of SO(8) we take to be hypercolor.
Solution:
Let us use the notation SR to denote the “right-handed” spinor S, by which we mean the
chiral spinor with eigenvalue +1 under the γ FIVE matrix, and SL to denote the chiral spinor
with eigenvalue −1 under the γ FIVE matrix. As shown in the reference for problem VII.7.3
(equations A20a, b), the chiral spinors 2n+m−1
R and 2n+m−1
R of SO(2n + 2m) decompose under
restriction to the subgroup SO(2n) ⊗ SO(2m) as
2n+m−1
R → (2n−1 m−1
R , 2R ) ⊕ (2n−1 m−1
L , 2L )
n+m−1 n−1 m−1 n−1 m−1
2L → (2R , 2L ) ⊕ (2L , 2R )
265
for the restriction of SO(18) to the subgroup SO(10) ⊗ SO(8). Let us choose the 256R as the
spinor in which to place all of the fermions. We want a subgroup GHC of SO(8) such that
8L decomposes into a bunch of stuff that is confined, meaning non-singlets of GHC , while 8R
decomposes into a bunch of free particles that do not participate in these strong “hypercolor”
interactions, meaning we want 8R to yield singlets of GHC .
In any case, we see that the restriction SO(8) → SU (4) exhibits the desired properties:
8R contains SU (4)-singlets, while 8L does not. At this stage, we see that the SO(10) ⊗ SU (4)
spinors (16R , [0])1 and (16R , [0])2 (where “1” and “2” are merely labels) constitute two
copies of the spinor S + of equation (14) on p. 425 of the text – that is, we have discov-
ered exactly two families of fermions, which upon the decomposition SO(10) → SU (5) →
SU (3)c ⊗ SU (2)W ⊗ U (1)Y become two families of the low-energy matter fields of the Stan-
dard Model.
Since in fact we observe three families, the theory as it stands with GHC = SU (4) is wrong, or
incomplete. We must break SU (4) down further. Let ψA transform as a 4 of SU (4). Suppose
we decide to restrict SU (4) to the subgroup SU (2) ⊗ SU (2), meaning
4×4 special unitary matrix → 2×2 special unitary matrix O
O 2×2 special unitary matrix
266
where the old guys “1” and “2” appeared at SO(8) → SO(4), while the new guys “3” and
“4” appeared at SO(4) → SU (2) ⊗ SU (2), comprise four families of SO(10) spinors. With
GHC = SU (2) ⊗ SU (2) ⊂ SO(8), we predict the three observed fermion families plus an
additional as of yet unobserved fourth generation.
At this point we see how the game is played and simply quote from section III of the reference:
“...the number of V − A families is ‘predicted’ to be two, three, four, or five, respectively,
according to whether SO(8) is broken down to SO(6), SO(5), SO(4), or SO(3). [These
orthogonal subgroups are not embedded in SO(8) in the obvious way, however.]” For further
details, consult the literature.
7. If you want to grow up to be a string theorist, you need to be familiar with the Dirac
equation in various dimensions but especially in 10. As a warm up, study the Dirac equation
in 2-dimensional spacetime. Then proceed to study the Dirac equation in 10-dimensional
spacetime.
Solution:
The salient feature will be whether the solutions to the Dirac equation can be classified
as Weyl, Majorana, or both.36 In general, let Ψ denote a Dirac spinor in any dimension.
Then a Weyl spinor is one that satisfies Ψ = ΨL or R , where L, R denote chiral projections,
and a Majorana spinor is one that satisfies Ψ = Ψc , where c denotes the charge conjugation
operation in whichever dimension is relevant.
which satisfy the Clifford algebra {γ µ , γ ν } = 2η µν with η = diag(−, +). The Dirac operator
i6 ∂ ≡ iγ µ ∂µ is diagonal:
∂0 − ∂1 0
i6 ∂ = −
0 ∂0 + ∂1
ψ
and therefore, writing Ψ = , we see that the two spinors ψ, ψ̃ do not mix under the
ψ̃
Dirac equation and therefore can be taken as independent degrees of freedom.
267
γ FIVE ΨL = (−1)ΨL and γ FIVE ΨR = (+1)ΨR .
Furthermore, since it is possible to choose a basis in which the Dirac operator is purely
real (as we have done), it is consistent to impose that its solutions are real: ψ ∗ = ψ, ψ̃ ∗ = ψ̃,
or in other words Ψ∗ = Ψ.
Thus, if we define the charge-conjugated Dirac field Ψc ≡ Ψ∗ as well as the left-chiral Dirac
field ΨL ≡ 21 (I − γ FIVE )Ψ, then we find it is consistent to impose the condition Ψ = Ψc = ΨL .
A spinor satisfying this condition is called a Majorana-Weyl spinor. (We can also use ΨR
instead of ΨL .)
Before jumping to d = 9 + 1, let us compare this briefly with what we know from d = 3 + 1.
We learned in Chapter II (and as explained in Appendix E), a Dirac spinor Ψ in (3 + 1)-
dimensional spacetime transforms reducibly as 4Dirac = 2L ⊕ 2R :
ψα
Ψ=
χ̄α̇
However, it is possible to impose the condition χ̄† = ψ. In other words, if we define the
conjugate spinor †
c (χ̄ )α
Ψ ≡
(ψ † )α̇
then it is possible to impose the Majorana condition Ψ = Ψc .
268
c ψ ψ
Ψ=Ψ = † , then ΨL = 6= Ψ. In d = 3 + 1, a Weyl spinor cannot be Majorana,
ψ 0
and a Majorana spinor cannot be Weyl.
Now consider d = 9 + 1. As explained in the text, the gamma matrices can be constructed
iteratively starting from d = 1 + 1:
0 0 1 1 0 1
γ(d=2) = , γ(d=2) =
−1 0 1 0
µ = 0,...,7 µ −1 0 8 0 1 9 0 −i
γ(d=10) = γ(d=8) ⊗ , γ(d=10) = I16 ⊗ , γ(d=10) = I16 ⊗
0 1 1 0 i 0
where I16 is the 16 = 24 -dimensional identity matrix. (Thegamma d/2
matrices have size 2 ,
0 1
which for d = 10 is 25 = 32. For instance, the matrix I16 ⊗ is indeed 32-by-32.)
1 0
Define the chiral projection matrix γ FIVE ≡ γ 0 γ 1 γ 2 ...γ 9 . The chiral spinor ΨL ≡ 12 (1−γ FIVE )Ψ
satisfies γ FIVE ΨL = (−1)ΨL , and the chiral spinor ΨR ≡ 12 (1 − γ FIVE )Ψ satisfies γ FIVE ΨR =
(+1)ΨR .
The question now is whether these irreducible chiral representations are self conjugate or
conjugate to each other. Under Lorentz transformations, the Dirac spinor Ψ transformas as
µν
Ψ → e iωµν σ Ψ = (I + iωµν σ µν + ...)Ψ
B −1 Ψ∗ → B −1 (I + iωµν σ µν )∗ Ψ∗
= B −1 (I − iωµν (σ µν )∗ )Ψ∗
= (I + iωµν σ µν )B −1 Ψ∗
which is identical to the transformation property of Ψ. Thus the (reducible) Dirac represen-
tation Ψ is self-conjugate. The question is whether its irreducible chiral components are also
self-conjugate.
269
Ψc ≡ C Ψ̄T = BΨ∗ denotes the charge conjugate spinor. We have therefore learned that, just
like d = 1 + 1 but unlike d = 3 + 1, it is consistent to impose the condition Ψ = ΨL = Ψc
in d = 9 + 1 dimensions. Again, a spinor satisfying this condition is called a Majorana-Weyl
spinor.
270
VIII Gravity and Beyond
VIII.1 Gravity as a Field Theory and the Kaluza-Klein Picture
1. Work out T µν for a scalar field. Draw the Feynman diagram for the contribution of
one-graviton exchange to the scattering of two scalar mesons. Calculate the amplitude and
extract the interaction energy between two mesons sitting at rest, thus deriving Newton’s
law of gravity.
Solution:
For this and the next problem, we will need to vary the determinant of the metric.
−1 δg)
det(g + δg) = det g det(1 + g −1 δg) = det g etr ln(1+g
−1
= det g etr(g δg) = det g 1 + tr(g −1 δg)
So defining as usual δ det g = det(g + δg) − det g, we get δ det g = det g tr(g −1 δg). It will
be convenient to vary with respect to g −1 rather than g, so to relate the two consider the
definition gg −1 = 1. Varying this yields δgg −1 + gδg −1 = 0 =⇒ δg = −g δg −1 g. Therefore
we have
δ det g = − det g tr(δg −1 g)
√
We actually need δ − det g , which is:
p 1 1p
δ − det g = √ δ det g = − − det g tr(δg −1 g)
2 − det g 2
Just to clarify, tr(δg −1 g) = δg µν gµν . Now we are ready for the problem. The action for a
real scalar field φ is
Z p 1
S = d4 x − det g L , L = g µν ∂µ φ ∂ν φ − V (φ) .
2
1
Tµν = ∂µ φ ∂ν φ − gµν g ρσ ∂ρ φ ∂σ φ − m2 φ2 .
2
271
Now we need the weak field action for gravity coupling to matter. This is given in equation
(10) on p. 437, which we repeat below:
Z
4 1 1 µν λ 1 µ λ ν µν
S= dx ∂λ h ∂ hµν − 2 ∂λ hµ ∂ hν − h Tµν
2 32πG
√ √
Define MP ≡ 1/ 16πG and rescale the graviton as hµν → M1P 2 hµν to obtain
Z
4 1 µν λ 1 µ λ ν
1 µν
S = d x 2 ∂λ h ∂ hµν − 2 ∂λ hµ ∂ hν − √ h Tµν
2 MP
k’
which is equal to
−i µ 0ν
k k + k ν k 0µ − η µν (k · k 0 + m2 )
√
2 MP
where both momenta flow into the vertex.
272
p1 p 1’
p2 p 2’
−i 0 0 −i 0 0
−p1 p10 − p~1 · p~10 − m21 [i∆00,00 (q)] √ −p2 p20 − p~2 · p~20 − m22
=√
2 MP 2 MP
−i
= 2 2 2 (p01 p010 + p~1 · p~10 + m21 )(p02 p020 + p~2 · p~20 + m22 )
2 MP q
where q = p1 − p10 = p20 − p2 is the momentum transfer. The minus signs on the momenta in
the second line come from the fact that in the diagram, one momentum flows into the vertex
while the other flows out. In the definition of the vertex given above, both momenta flow
into the vertex.
If the two incoming particles are at rest, then pµ1 = (m1 , ~0) and pµ2 = (m2 , ~0), which im-
plies (p01 p010 + p~1 · p~10 + m21 )(p02 p020 + p~2 · p~20 + m22 ) = (2m21 )(2m22 ) = 4m21 m22 . In the denominator
we have
The amplitude is
m21 m22
M= .
MP2 |~p10 |2
Scattering from a potential in nonrelativistic quantum mechanics gives37
d3 q i~q·~x
Z
1
V (~x ) = − e M(~q )
4m1 m2 (2π)3
where ~q = p~1 − p~10 = −~p10 is the transfer of 3-momentum. Since
d3 q i~q·~x 1
Z
1
3
e 2
=
(2π) |~q | 4π|~x |
37
See Section 4.5.7 of http://arxiv.org/abs/0812.1594 and the textbook M. Maggiore, A Modern Introduc-
tion to Quantum Field Theory (Oxford University Press, Oxford, UK, 2005) pp. 167-170 .
273
we have
m21 m22
1 1 m1 m2
V (~x ) = − =− .
4m1 m2 MP2 4π|~x | 16πMP2 |~x |
The Planck mass is defined in terms of Newton’s constant as MP2 = 1/(16πG), so we get
Gm1 m2
V (~x ) = −
|~x |
Note that various conventions exist for defining the Planck mass MP in terms of Newton’s
constant G, and so various factors of 2 can appear between references. For a discussion of
graviton scattering in various contexts, see K. Hinterbichler, “Theoretical aspects of massive
gravity,” arXiv:1105.3735v1 [hep-th].
Solution:
Yang-Mills 1
Tαβ = g µν Fαµ
a a
Fβν a
− gαβ Fµν F aµν
4
Note that since γ αβ g µν Fαµ
a a
Fβν a
= Fµν F aµν and g αβ gαβ = 4, this stress tensor is in fact trace-
less.
274
3. Show that if hµν does not satisfy the harmonic gauge, we can always make a gauge trans-
formation with εν determined by ∂ 2 εν = ∂µ hµν − 12 ∂ν hλλ so that it does. All of this should be
conceptually familiar from your study of electromagnetism.
Solution:
1 1
∂µ hµν − ∂ν hλλ = ∂µ (h0µ µ µ 0λ λ
ν + ∂ εν + ∂ν ε ) − ∂ν (hλ + 2∂λ ε )
2 2
0µ 1 0λ
= ∂µ hν − ∂ν hλ + ∂ εν + ∂µ ∂ν εµ − ∂ν ∂λ ελ
2
2
1
0µ
= ∂µ hν − ∂ν h0λ 2
λ + ∂ εν
2
We can choose ∂ 2 εν = ∂µ hµν − 12 ∂ν hλλ to make h0 satisfy the harmonic gauge.
Solution:
Suppose we have already imposed the condition k µ hµν = 12 kν hλλ as indicated, so that we
have 6 degrees of freedom. Consider the gauge transformation hµν = h0µν − kµ εν − kν εµ .
Putting this into the above condition gives
1
k µ (h0µν − kµ εν − kν εµ ) = kν (h0λλ − 2k λ ελ )
2
1
k µ h0µν − k 2 εν − kν k µ εµ = kν h0λλ − kν k λ ελ
2
1
k µ h0µν = kν h0λλ
2
since k 2 = 0. The function h0µν also satisfies the harmonic gauge for arbitrary εµ , so we
have not completely used up our gauge freedom. Indeed, we still have 4 functions to choose,
namely {εµ }3µ=0 , which cuts the number of physical degrees of freedom from 6 to 2. Thus the
graviton has 2 physical degrees of freedom.
275
5. The Kaluza-Klein result that we argued by symmetry considerations can of course be
derived explicitly. Let me sketch the calculation for you. Consider the metric
where θ denotes an angular variable 0 ≤ θ < 2π. With Aµ = 0, this is just the metric of a
curved spacetime, which has a circle of radius a attached at every point. The transformation
θ → θ + Λ(x) leaves ds invariant provided that we also transform Aµ (x) → Aµ (x) − ∂µ Λ(x).
Calculate the 5-dimensional scalar curvature R5 and show that R5 = R4 − 41 a2 Fµν F µν . Except
for the precise coefficient 14 this result follows entirely from symmetry considerations and from
the fact that R5 involves two derivatives on the 5-dimensional metric, as explained in the
text. After some suitable rescaling this is the usual action for gravity plus electromagnetism.
Note that the 5-dimensional metric has the explicit form
5 gµν − a2 Aµ Aν −a2 Aµ
gAB = .
−a2 Aν −a2
Solution:
We will compute the result using differential forms. First, find the veilbein 1-forms em =
em
M dx
M
= em µ m m
µ dx + eθ dθ, then find the connection 1-forms ω n from de
m
= −ω mn en , then
compute the curvature 2-form Rmn = dω mn + ω mp ω pn . From the curvature 2-form, extract the
components via Rmn = 21 Rmnrs er es . Finally, compute the curvature scalar R ≡ η ns Rmnms .
The 5D metric is
ds2 = GM N dxM dxN = (gµν − a2 Aµ Aν )dxµ dxν − a2 dθ2 − aAµ (dxµ dθ + dθdxµ )
we have:
Gθθ = −a2 =⇒ eθθ = a, eaθ = 0
Gµθ = −aAµ =⇒ eθµ = aAµ
Gµν = gµν − a2 Aµ Aν =⇒ eaµ = ea(4)µ
Now define the 1-forms em ≡ em M
M dx . This implies
276
where ea(4) are the 4D veilbeins defined by gµν = ea(4) ηab eb(4) , and A ≡ Aµ dxµ . Recall that the
point of using forms notation is to be able to expand in any basis we choose, not necessarily
the coordinate basis dxµ . Thus we define the component fields Aa via A ≡ Aa ea(4) , and for
later convenience the partial derivatives ∂a via dxµ ∂µ ≡ ea(4) ∂a .
Now we need to find the connection 1-forms ω mn , which are defined by the equation dem +
ω mn en = 0. Since eθ = aA, we have38 deθ = d(dθ + aA) = 0 + aF = a 21 Fab ea(4) eb(4) . Since
ω θθ = 0, the definition deθ = −ω θm em = −ω θa ea(4) implies
1
ω θa = + a Fab eb(4)
2
where we have been careful to keep track of the order of the 1-forms ea eb when extracting
the components Fab = −Fba , hence the + sign in the above.
1
ω ac = ω(4)c
a
− a ηθθ F ac eθ .
2
Now we are ready to define the curvature 2-form, Rmn = dω mn + (ω 2 )mn . First compute
Rab = dω ab + (ω 2 )ab . We have dω ab = dω(4)b a
− 12 a ηθθ d(F ab eθ ), and d(Fab eθ ) = ∂c Fab ec(4) eθ +
Fab (aF ) = ∂c Fab ec(4) eθ + 12 aFab Fcd ec(4) ed(4) . To summarize:
1 1
dω ab = dω(4)b
a
− a ηθθ ∂c F ab ec(4) eθ − a2 ηθθ F ab Fcd ec(4) ed(4) .
2 4
Next we need (ω 2 )ab = ω am ω mb = ω ac ω c b + ω aθ ω θb . Multiplying the terms out gives
1 1
(ω 2 )ab = (ω(4)
2 a
) b + a ηθθ (F ac ω(4)b
c a
− ω(4)c F cb )eθ − a2 F ac Fbd ec(4) ed(4) .
2 4
Therefore, the Rab part of the curvature 2-form is
1 1
Rab = R(4)b
a
− a2 (F ac Fbd + ηθθ F ab Fcd )ec(4) ed(4) + a ηθθ (F ac ω(4)b
c a
− ω(4)c F cb − ∂c F ab ec(4) )eθ .
4 2
38
To make sure the meaning of the last equality here is clear: In the coordinate basis, we have F =
1 µ ν
2 Fµν dx dx , so that F is a 2-form and the components Fµν are just numbers. We use the definition of the
veilbein 1-forms ea ≡ eaµ dxµ to define the c-number components Fab via Fµν ≡ eaµ ebν Fab . So while ω ab is a
collection of 1-forms, F ab ≡ η ac Fcb is not a collection of forms but instead is just a collection of numbers.
39
We will keep the explicit factor of ηθθ present to show that it cancels out of the final result, and hence
that the resulting action is independent of whether we use the metric convention η = (+, −, −, −, −) or the
one more commonly used in the gravitational literature, η = (−, +, +, +, +).
277
A few comments are in order at this point. First, to compute the curvature scalar we will not
need coefficients of the 2-form ec(4) eθ , so we will drop them from now on. Second, look at the
terms in parentheses multiplying ec(4) ed(4) . Since Fcd = −Fdc , the second term in parentheses is
manifestly antisymmetric in (cd), but the first term is not. Thus to extract the components
Rabcd from the 2-form Rab = 21 Rabcd ec ed , we must write F ac Fbd ec ed = 12 (F ac Fbd − F ad Fbc )ec ed .
Therefore, we have
1 1
Rabcd = (R(4) )abcd − a2 (F ac Fbd − F ad Fbc ) − a2 ηθθ F ab Fcd .
4 2
m
Finally, let us make sure the dimensions are correct. The e have dimensions of length (recall
em = em µ m θ
µ dx with eµ dimensionless), so writing e = aA implies A is dimensionless. Since
A = Aa ea , we find Aa has dimensions of inverse length. Since ea ∂a = dxµ ∂µ , the derivatives
∂a have dimensions of inverse length (just like ∂µ ), so Fab = ∂a Ab − ∂b Aa has dimensions of
1/length2 . Thus a2 F ac Fbd has dimensions of 1/length2 , as it should be to match Rabcd . Note
that the operator d = ea ∂a is dimensionless, so that the 1-forms ω mn defined by dem = −ω mn en
are also dimensionless.
When simplifying this expression, it is important to keep track of whether a particular term
b
is a form or a number. In particular, Fbc is a number, ω(4)a is a 1-form, and ec(4) is a 1-form.
Therefore:
1 1
(ω 2 )θ a = ( a Fbc ec(4) )(ω(4)a
b
− a ηθθ F ba eθ )
2 2
1 1 2
= − a Fbc ω(4)a e(4) − a ηθθ Fbc F ba ec(4) eθ
b c
2 4
where in the first term, the minus sign was generated from moving the 1-form ec(4) to the
b
right of the 1-form ω(4)a . In contrast, no minus sign was generated in the second term when
moving the 1-form e(4) to the right of the number F ba . Putting it all together, we have40
c
1 1 1 1
Rθ a = a ∂c Fab ec(4) eb(4) − a Fab ω(4)c
b
ec(4) − a Fbc ω(4)a
b
ec(4) − a2 ηθθ Fbc F ba ec(4) eθ .
2 2 2 4
40
Let us again pause to check the dimensions. Fab has dimensions of 1/length2 , and ec(4) has dimensions of
length. a has dimensions of length, and ∂c has dimensions of inverse length. So the first term is dimensionless.
b
Since ω(4)c is dimensionless, the second and third terms are also dimensionless. The fourth term has two
powers of length from a2 , one power of length from each of ec(4) and eθ , which are indeed canceled by four
powers of inverse length from the product Fbc F ba .
278
To get the curvature scalar, we need only the components Rmamc = Rb abc +Rθ aθc , so we need to
extract Rθ aθb from the above expression for Rθ a . Since Rθ a = 12 Rθ amn em en = 12 Rθ abc eb(4) ec(4) +
2 × 12 Rθ aθc eθ ec(4) , we compare “Rθ a = − 41 ηθθ Fbc F ba ec(4) eθ + ... = + 14 ηθθ Fbc F ba eθ ec(4) + ...” with
“Rθ a = Rθ aθc eθ ec(4) + ...” to get
1
Rθ aθc = ηθθ Fbc F ba .
4
Next, recall the previously obtained result
1 1
Rabcd = (R(4) )abcd − a2 (F ac Fbd − F ad Fbc ) − a2 ηθθ F ab Fcd
4 2
and contract with δ c a to get
1 1
Rabad = (R(4) )abad − a2 (0 − F ad Fba ) − a2 ηθθ F ab Fad
4 2
1 2 a 1 2
= (R(4) ) bad + a F d (−Fab ) − a ηθθ F ab Fad
a
4 2
1 2
= (R(4) ) bad − a (1 + 2ηθθ )Fab F ad .
a
4
Since we already found Rθ bθd = 14 ηθθ Fab F ad , we arrive at the following expression for Rbd ≡
Rmbmd = Rabad + Rθ bθd :
1
Rbd = (R(4) )bd − a2 (1 + ηθθ )Fab F ad .
4
We use the calligraphic font to denote Rbd ≡ Rmbmd to avoid potential confusion with the
previously defined 2-form Rac , which can display a lowered index via Rac = ηab Rb c . To
belabor the point, the object Rab is a collection of ordinary numbers, while the object Rab is
a collection of 2-forms.
To complete the calculation we need the quantity Rθθ ≡ Rmθmθ = Raθaθ = η ab Rbθaθ =
η ab Rθbθa = η ab ηθθ Rθ bθa = η ab ηθθ ( 41 a2 ηθθ Fcb F ca ) = 41 a2 Fcb F cb .
Solution:
Indices:
M = 0, 1, 2, ..., D ← full spacetime, x̄M
µ = 0, 1, 2, 3 ← non-compact part, xµ
m = 4, 5, ..., D ← compact part, θm
a = 1, 2, ..., K ← isometry group of compact part; generated by ξam (θ)
µ̂ ν̂
T eµ (x) −Bµm̂ (x, θ) ηµ̂ν̂ 0 eν (x) 0
GM N (x, θ) = (EηE )M N =
0 em̂
m (θ) 0 −δm̂n̂ −Bνn̂ (x, θ) en̂n (θ)
eµ̂µ (x)ηµ̂ν̂ eν̂ν (x) − Bµm̂ (x, θ)δm̂n̂ Bνn̂ (x, θ) Bµm̂ (x, θ)δm̂n̂ en̂n (θ)
=
em̂ n̂
m (θ)δm̂n̂ Bν (x, θ) −em̂ n̂
m (θ)δm̂n̂ en (θ)
gµν (x) − Bµm (x, θ)gmn (θ)Bνn (x, θ) Bµm (x, θ)gmn (θ)
=
gmn (θ)Bνn (x, θ) −gmn (θ)
∂ x̄N
Under the transformation x̄M → x̄0M , we have EM
M̂ 0M̂
(x̄) → EM (x̄0 ) = E M̂ (x̄).
∂ x̄0M N
Rearrange
the primes:
M̂ ∂ x̄0N 0M̂ 0 ∂x0ν 0M̂ 0 ∂θ0n 0M̂ 0
EM (x̄) = E (x̄ ) = E (x̄ ) + E (x̄ ) .
∂ x̄M N ∂ x̄M ν ∂ x̄M n
Look at Eµm̂ = −Bµm̂ , with x0µ = xµ and θ0m = θm + ξam (θ)εa (x):
∂x0ν 0m̂ 0 0 ∂θ0n 0m̂ 0
−Bµm̂ (x, θ) = − B (x , θ ) + e (θ )
∂xµ ν ∂xµ n
= −Bµ0m̂ (x, θ0 ) + ξan (θ)∂µ εa (x)en0m̂ (θ0 ) .
280
Write Bµm (x, θ) ≡ ξam (θ)Aaµ (x) to get:
Be mindful of the primes on ξa0m̂ (θ0 ) on the left-hand side: we have to ask how a Killing
vector transforms under an infinitesimal transformation of the compact coordinates. For
0m
θm → θ0m = (Λ−1 )mn θn , with (Λ−1 )mn = ∂θ
∂θn
= δnm + ∂n ξam (θ)εa (x), we have:
For further discussion, see A. Salam, J. Strathdee, “On Kaluza-Klein theory,” Ann. Phys.
141, 316-352 (1982) and F. Cianfrani, G. Montani, “Non Abelian gauge symmetries induced
by the unobservability of extra-dimensions in a Kaluza-Klein approach,” Mod.Phys.Lett.
A21 (2006) 265-274 (arXiv:gr-qc/0511100v1).
281
8. The veilbeins for a spacetime with Minkowski metric is defined by gµν (x) = eaµ (x)ηab ebν (x),
where the Minkowski metric ηab replaces the Euclidean metric δab . The indices a and b are
to be contracted with ηab . For example, Rab = dω ab + ω ac ηcd ω db . Show that everything goes
through as expected.
Solution:
We will follow the discussion beginning on p. 443 replacing δab with ηab . We begin with
For your convenience, we will label the equations as in the chapter, with the subscript M to
denote “Minkowski.” Consider the “Minkowskian 2-sphere” defined by the line element
From the metric (gtt = 1, gϕϕ = sin2 t) we can read off e1t = 1 and e2ϕ = sin t, with all other
components zero. (Remember that ηab = diag(+1, −1) in this case!) We define the 1-forms
ea = eaµ dxµ , so that e1 = dt , e2 = sin t dϕ. Now define the curvature 1-form
dea = −ω ab eb (17M )
where now we are careful to use upper and lower indices because we have a nontrivial norm
on the spacetime. In our baby example, we have de1 = 0 and de2 = cos t dt dϕ, so as in the
text the connection has only one nonvanishing component ω 12 = −ω 21 = − cos t dϕ.
In the Euclidean case, the rotation eaµ (x) = Oab (x)e0bµ (x) leaves the metric invariant, that
is gµν (x) = eaµ (x)δab ebν (x) = e0a 0b T
µ (x)δab eν (x) if O O = 1. This time, under the same transfor-
a a 0b
mation eµ (x) = O b (x)eµ (x), we have (suppressing the spacetime arguments)
gµν = eaµ ηab ebν = Oac ecµ ηab Ob d edν = ecµ (OT )c a ηab Ob d edν = ecµ ηcd edν
only when OT η O = η. This is precisely the definition of the Lorentz transformation. That
is, the rotation eaµ = Oab eb leaves the curved metric gµν invariant if O leaves the flat metric
ηab invariant. Everything else up through equation (20) on p. 444 works out in a formally
identical manner, including the curvature of +1 in our Minkowskian 2-sphere.
282
VIII.3 Effective Field Theory
1. Consider
1
(∂ϕ1 )2 + (∂ϕ2 )2 − λ(ϕ41 + ϕ42 ) − g ϕ21 ϕ22
L=
2
We have taken the O(2) theory from chapter I.10 and broken the symmetry explicitly. Work
out the renormalization group flow in the (λg)-plane and draw your own conclusions.
Solution:
First let us make a few changes in notation. Normalize the couplings such that we won’t
have to worry about unnecessary numerical factors in the Feynman rules, meaning λ → 4!1 λ
and g → 212 g. To eliminate notational clutter, write ϕ1 ≡ ϕ and χ ≡ ϕ2 . For this problem
we will use dimensional regularization and therefore separate the mass parameter µ̃ from the
couplings: λ → λµ̃ε and g → g µ̃ε . The Lagrangian is
1 1 1
L = Z (∂ϕ)2 + (∂χ)2 − Zλ λµ̃ε (ϕ4 + χ4 ) − Zg g µ̃ε ϕ2 χ2
2 24 4
1 1 1
(∂ϕ0 )2 + (∂χ0 )2 − λ0 (ϕ40 + χ40 ) − g0 ϕ20 χ20
=
2 24 4
In the first line we have included the renormalizing Z-factors, or equivalently the counterterms
Ai ≡ Zi − 1. In the second line we have written the Lagrangian in terms of bare fields and
parameters, denoted by the subscript 0. For future reference, we will need the relationships
between the bare parameters and the renormalized parameters:
λ0 = Z −2 Zλ λµ̃ε and g0 = Z −2 Zg g µ̃ε
The Feynman rules are as follows. We use a solid line for the ϕ propagator, and we use a
dashed line for the χ propagator. Since we assume these fields have the same mass (namely
m2 = 0), both propagators are equal, so that (dashed line) = (solid line) = i∆(k), where
∆(k) ≡ 1/(k 2 − m2 ) = 1/k 2 . A vertex connecting 4 solid lines equals a vertex connecting 4
dashed lines, both of which equal −iZλ λµ̃ε . A vertex connecting 2 solid lines and 2 dashed
lines equals −iZg g µ̃ε .
+ + + + (higher order)
(4) (5) (6)
283
We use conventions for which all momenta flow into the vertex. We also use the notational
convention for which the diagram is just the picture as written, and the symmetry factor (in
this case 21 ) is written explicitly multiplying the picture (as shown above). The first 1-loop
diagram is:
dd `
Z
ε 2
(1) = (−iZλ λµ̃ ) i∆(`)i∆(` − k1 − k4 )
(2π)d
dd `
Z
ε 2 1 1
= (λµ̃ )
(2π) (` − m ) [(` − k1 − k4 )2 − m2 ]
d 2 2
Z 1
dd `
Z
ε 2 1
= (λµ̃ ) dx
(2π) 0d
{x(` − m ) + (1 − x)[(` − k1 − k4 )2 − m2 ]}2
2 2
i ε Z 1 4π ε/2
ε 2
(1) = (λµ̃ ) Γ dx
(4π)2 2 0 D4
ε 1 Z ε/2
ε 2 i 4π
(2) = (λµ̃ ) Γ dx
(4π)2 2 0 D2
ε 1 Z ε/2
ε 2 i 4π
(3) = (λµ̃ ) Γ dx
(4π)2 2 0 D3
Moreover, the other three diagrams (containing dashed lines) are numerically identical to the
three diagrams above, except with the replacement λ → g. Therefore, at 1-loop order we
have
( Z 1 " #)
2 2 ε/2 2 ε/2 2 ε/2
λ g ε 4π µ̃ 4π µ̃ 4π µ̃
Vλ = −λµ̃ε Zλ − 1+ 2 Γ dx + +
2(4π)2 λ 2 0 D2 D3 D4
ε 2
C ε/2 = + ln Ce−γ + O(ε) .
Γ
2 ε
284
Now let µ2 ≡ 4πe−γ µ̃2 . Using the above expansion and taking ε → 0+ wherever possible
gives
Z 1 2
g2
2 2
λ 6 µ µ µ
Vλ = −λ Zλ − 2
1+ 2 + dx ln + ln + ln
2(4π) λ ε 0 D2 D3 D4
We choose the “modified minimal subtraction” renormalization scheme, for which the coun-
terterm Aλ = Zλ − 1 is chosen purely to cancel the 1/ε pole and nothing more. To this order,
we therefore have
g2 1
3
Aλ = λ 1+ 2 .
(4π)2 λ ε
Now we need to renormalize the ϕ2 χ2 vertex. At 1-loop, we have
iV g = + 1 +
2
+ + + (higher order)
Fortunately, we have already done all of the computational work. The two s-channel diagrams
are numerically equal to the s-channel diagram from the ϕ4 vertex, except with the replace-
ment λ2 → λg. The t-channel and u-channel diagrams are numerically equal to the t- and
u- channel diagrams from the ϕ4 vertex, except with the replacement λ2 → g 2 . Immediately
we get the result
1 2
Vg = −g Zg − (λ + 2g) + (finite)
(4π)2 ε
so that choosing the counterterm Ag ≡ Zg − 1 purely to cancel the divergent piece gives
2 1
Ag = 2
(λ + 2g) .
(4π) ε
Now we compute the beta functions. Define H(λ, g) ≡ ln(Z −2 Zλ ) and G(λ, g) ≡ ln(Z −2 Zg ).
Taking the logarithm of the relationships between the bare and renormalized couplings gives
The bare parameters are independent of the unphysical parameter µ, so differentiating the
above with respect to ln µ gives
dH dλ dG dg
0=λ + + ελ and 0 = g + + εg .
d ln µ d ln µ d ln µ d ln µ
285
Upon using the chain rule and rearranging a bit, these become
∂H dλ ∂H dg
0= λ +1 +λ + ελ (1)
∂λ d ln µ ∂g d ln µ
and
∂G dg ∂G dλ
0= g +1 +g + εg (2)
∂g d ln µ ∂λ d ln µ
The renormalizing Z-factors in the MS-bar renormalization scheme to this order are
Z=0
g2 1
3
Zλ = 1 + λ 1+ 2
(4π)2 λ ε
2 1
Zg = 1 + 2
(λ + 2g)
(4π) ε
so the functions H and G are (after expanding ln(1 + x) = x + O(x2 ))
g2 1 g2 1
3 ∂H 3 ∂H 6g 1
H(λ, g) = λ 1 + =⇒ = 1 − and =
(4π)2 λ2 ε ∂λ (4π)2 λ2 ε ∂g (4π)2 λ ε
2 1 ∂G 2 1 ∂G 4 1
G(λ, g) = 2
(λ + 2g) =⇒ = 2
and =
(4π) ε ∂λ (4π) ε ∂g (2π)2 ε
With these, the renormalization group equations (1) and (2) become
g2 1
3 dλ 6 1 dg
0= λ 1 − + 1 + g + ελ (10 )
(4π)2 λ2 ε d ln µ (4π)2 ε d ln µ
4 1 dg 2 1 dλ
0= 2
g +1 + 2
g + εg (20 )
(4π) ε d ln µ (4π) ε d ln µ
Now as per the usual procedure with dimensional regularization, we write
dλ dg
= −ελ + βλ and = −εg + βg
d ln µ d ln µ
and demand that βλ and βg be finite in the limit ε → 0+ . Putting these into equation (10 )
implies
3
0 = βλ − 2
(λ2 + g 2 ) + (things we insist sum to zero)
(4π)
0
while equation (2 ) implies
2
0 = βg − (λ + 2g)g + (things we insist sum to zero) .
(4π)2
Thus we have the 1-loop beta functions
dλ 3
βλ = =+ 2
(λ2 + g 2 ) (100 )
d ln µ (finite) (4π)
dg 2
βg = =+ (λ + 2g)g (200 )
d ln µ (finite) (4π)2
286
Before discussing the renormalization group dynamics in the (λ, g)-plane, let us perform a
check. There should be a value of the parameter g for which we recover an O(2) symmetric
theory. Consider the 2-component vector φ ~ = (φ1 , φ2 )T . An O(2)-invariant theory would
have the interaction Lagrangian
1 ~ ~ 2 1
L = − λ0 (φ · φ ) = − λ0 φ41 + φ42 + 2φ21 φ22
8 8
with some coupling λ0 . Compare this to our interaction Lagrangian
1 1 1
λ(ϕ4 + χ4 ) − g ϕ2 χ2 = − λ(ϕ4 + χ4 ) + 3! g ϕ2 χ2 .
L=−
4! 4 4!
We recover the O(2) symmetric theory only for the precise value 3! g = 2λ =⇒ g = 31 λ. If
we plug this particular value into the beta functions, we should recover a single beta function
for the coupling λ. If g = 13 λ, then g 2 = 91 λ2 , so that βλ = (4π)
1 10 2 2
2 3 λ . Also, βg = (4π)2 (1 +
2 1 1 10 2 1
)( )λ2 = (4π)
3 3 2 9 λ . Using the chain rule, we have dg/d ln µ = (dg/dλ)(dλ/d ln µ) = 3 βλ . So
1 10 2
βλ = (4π) 2 3 λ , which is exactly what we got before.
Now let us study the dynamics implied by (100 ) and (200 ). Equation (100 ) implies that
dλ/d ln µ > 0, so the coupling λ increases in magnitude as the parameter µ increases, irre-
spective of the signs of λ and g. Equation (200 ) is more interesting. Since d ln g/d ln µ ∝ λ+2g,
the relative signs of λ and g change the running of g qualitatively: if λ + 2g < 0, then the
coupling g decreases in magnitude as µ increases. At low energy there is an O(2) symmetric
fixed point.
2. Assuming the nonexistence of the right handed neutrino field νR (i.e., assuming the
minimal particle content of the standard model) write down all SU (2) ⊗ U (1) invariant terms
that violate lepton number L by 2 and hence construct an effective field theory of the neu-
trino mass. Of course, by constructing a specific theory one can be much more predictive.
Out of the product lL lL we can form a Lorentz scalar transforming as either a singlet or
triplet under SU (2). Take the singlet case and construct a theory. [Hint: For help, see A.
Zee, Phys. Lett. 93B: p. 389, 1980.]
Solution:
We use two-component spinor notation for the fermion fields. The neutrinos reside in SU (2)
doublets
ν
(`a )i = a ∼ (2, − 12 ) of SU (2) ⊗ U (1)
ea
where i = 1, 2 labels the fundamental of SU (2), and a = e, µ, τ labels the flavor. From this
field we can form the SU (2) singlet
287
which has hypercharge 2(− 12 ) = −1. Note that this singlet is antisymmetric in the flavor
indices. If we introduce a complex scalar field h+ with hypercharge +1 that is a singlet under
SU (2), then we can form the interaction terms
where the sums over a, b = e, µ, τ are implied, and due to fermi statistics41 only the antisym-
metric part of f contributes.
At this stage, we may simply assign two units of lepton number to the field h+ , and so
this interaction does not generate Majorana neutrino masses on its own. Let us now aug-
ment the standard model with additional Higgs doublets:
0
ϕI
(ϕI )i = ∼ (2, − 12 )
ϕ−
I
where again i = 1, 2 labels the fundamental of SU (2), while I = 1, 2, ..., nH labels the flavor
of Higgs. This allows for the dimension-3 interaction
where MIJ is an antisymmetric matrix of size nH ×nH whose entries have dimensions of mass.
I
If the Higgs fields give charged lepton masses in the usual way, LYuk = −yab ϕI · `a ēb + h.c.,
then they must be assigned lepton number zero. Thus, the clash between Lhϕϕ and Lh``
breaks lepton number by 2 and thereby generates neutrino masses at 1-loop.
For simplicity, let us assume that only ϕ1 gives mass to the charged leptons, and that nH = 2.
Then neutrino masses are generated by the diagram
+
h− 2
l
−
h+ 2
p e
p
e e
l+p
fe
41
Recall that νe = eν for two Grassmann-valued spinors να and eα contracted with the antisymmetric
tensor εαβ . See appendix E.
288
The diagram is42
d4 `
Z
∗
iAµe (p) = ve (p)(+iye )[iSe (` + p)](+ifµe )uµ (p)[i∆ϕ±2 (`)](+iM12 hϕ01 i∗ )[i∆h (`)]
(2π)4
∗
= fµe M12 m2e ve (p)uµ (p) I(p)
where
d4 `
Z
1 1 1
I(p) = − .
(2π) (` + p) − me ` − Mϕ± ` − Mh2
4 2 2 2 2 2
2
We want the mass term, so take p → 0. Also, me Mϕ±2 , Mh so take me → 0 inside the
integral. Define I0 ≡ limme →0 I(p = 0). Let M> denote the greater of Mϕ±2 and Mh , and M<
the lesser of the two.
d4 ` 1
Z
1 1
I0 = −
(2π) ` ` − M> ` − M<2
4 2 2 2 2
Z 1 3
d4 `
Z
X 1
= −Γ(3) dx1 dx2 dx3 δ( xi − 1)
0 i=1
(2π)4 [x1 `2 + x2 (`2 − M>2 ) + x3 (`2 − M<2 )]3
Z 1 Z 1−x2
d4 `
Z
1
= −Γ(3) dx2 dx3 , D = x2 M>2 + x3 M<2
0 0 (2π) (` − D)3
4 2
Z 1 Z 1−x2
i 1 1
= +Γ(3) dx2 dx3 2
0 0 (4π) Γ(3) D
Z 1 Z 1−x2
i 1 1 M<2
= dx 2 dx 3 , r ≡ <1
(4π)2 M>2 0 0 x2 + rx3 M>2
2
i 1 M>
≈ 2 2
ln
(4π) M> M<2
where we have taken M>2 M<2 . We have
M>2
∗ 1 1
(mν )eµ = Aµe (0) + Aeµ (0) ≈ feµ (m2µ − m2e )M12 ln
(4π) M>2
2 M<2
where we have used fµe = −feµ .
42
We are using the two-component spinor notation, in which the Fourier expansion of the free neutrino
fields is
XZ
ν(x) = (dp)[bs (~ p ) e−ip·x + d†s (~
p )us (~ p ) e ip·x ] .
p )vs (~
s
Therefore:
Z XZ Z
4
d x νe (x)νµ (x) = (dpe ) (dpµ )(2π)4 δ 4 (pµ − pe ) vse (~ pµ )d†se (~
pe ) usµ (~ pe ) bsµ (~
pµ ) + ... .
se ,sµ
289
The distinctive prediction of this model is that the diagonal entries of mν are zero in the
basis for which the charged lepton mass matrix is diagonal. Present neutrino oscillation data
rule out such a mass matrix in the absence of further structure.
See the reference along with D. Chang and A. Zee, “Radiatively Induced Neutrino Majorana
Masses and Oscillation,” arXiv:hep-ph/9912380v1 and R. A. Porto and A. Zee, “Neutrino
Mixing and the Private Higgs,” arXiv:0807.0612v1 [hep-ph] for further discussion.
3. Let A, B, C, D denote four spin-1/2 fields and label their handedness by a subscript:
γ 5 Ah = hAh with h = ±1. Thus, A+ is right handed, A− is left handed, and so on. Show
that
1
(Ah Bh )(C−h D−h ) = − (Ah γ µ D−h )(C−h γµ Bh )
2
This is an example of a broad class of identities known as Fierz identities (some of which
we will need in discussing supersymmetry.) Argue that if proton decay proceeds in low-
est order from the exchange of a vector particle then only the terms (˜lL CqL )(uR CdR ) and
(eR CuR )(q̃L CqL ) are allowed in the Lagrangian.
Solution:
In the first line we have written the current in 2-component notation, and in the second line
we have used Dirac notation. Here i = 1, 2 denotes the fundamental of SU (2), and α = 1, 2, 3
denotes the fundamental or anti-fundamental of SU (3) depending on the index height.
Integrating out the vector bosons that couple to these currents results in the effective La-
grangian J † J. The operators of the form qqq` contained in this Lagrangian are:43
εαβγ (ū† σ̄ µ εij qjβ )(`i σµ d¯†α ) = −2εαβγ εij (`i qjβ )(ū†γ d¯†α ) = −2εαβγ (νdβ − euβ )(ū†γ d¯†α )
= −2εαβγ (NRc DLβ − ERc ULβ )(ULcγ DR
α
)
εαβγ (ū†γ σ̄ µ εij qjβ )(ē† σ̄µ qiα ) = +2εαβγ εij (qiα qjβ )(ū†γ ē† ) = +4εαβγ (uα dβ )(ū†γ ē† )
= +4εαβγ (URcα DLβ )(ULcγ ER )
These are the first two effective operators given at the top of p. 456.
The other possibility consistent with SU (3) ⊗ SU (2) ⊗ U (1) invariance and ordinary matter
fields is that proton decay is mediated through vector bosons with electric charges 32 and − 13 .
These would couple to the current
The i = 1 component has electric charge Q(ūν † ) = −Q(u) − Q(ν) = − 23 , and the i = 2
component has electric charge Q(ūe† ) = −Q(u) − Q(e) = − 23 − (−1) = + 31 . The operator
contained in J 0† J 0 relevant for proton decay is:
εαβγ (εij qβ†j σ̄ µ d¯γ )(ūα σµ `†i ) = −2εαβγ εij (ūα d¯γ )(qβ†j `†i )
43
To translate
between
2-component
spinors and Dirac spinors, consider
the following example: E =
e c ē c 0 c u c U = eu.
=⇒ E = † =⇒ ER = † =⇒ ER = (e, 0). Then UL = =⇒ ER L
ē† e e 0
291
4. Given the conclusion of the previous exercise show that the decay rate for the processes
p → π + + ν̄, p → π 0 + e+ , n → π 0 + ν̄, and n → π − + e+ are proportional to each other,
with the proportionality factors determined by a single unknown constant [the ratio of the
coefficients of (˜lL CqL )(uR CdR ) and (eR CuR )(q̃L CqL )].
For help on these last three exercises see S. Weinberg, Phys. Rev. Lett. 43:1566, 1979;
F. Wilczek and A. Zee, ibid. p. 1571; H. A. Weldon and A. Zee, Nucl. Phys. B173: 269, 1980.
Solution:
The hadronic part of the first operator transforms as a doublet under isospin. Since q̃L CqL =
2d˜L CuL and parity acts on a Dirac spinor ψ as ψL → iγ 0 ψR , the hadronic part of the second
operator is just the parity transform of the first component of the hadronic part of the first
operator. Let the coefficients of the two operators be C1,2 , respectively, let us denote their
ratio by C2 /C1 = r. Then the transformation properties under isospin imply the relations
As we saw in problem VII.6.2, at tree level in SU (5) unification we have r = 2. For a discus-
sion of renormalization group corrections to r, see F. Wilczek and A. Zee, “Operator analysis
of nucleon decay,” Phys. Rev. Lett. Vol. 43 No. 21, 19 Nov 1979.
5. Imagine a mythical (and presumably impossible) race of physicists who only understand
physics at energies less than the electron mass me . They manage to write down the effective
field theory for the one particle they know, the photon,
1 1
L = − Fµν F µν + 4 [a(Fµν F µν )2 + b(Fµν F̃ µν )2 ] + ...
4 me
with F̃ µν = 21 εµνρσ Fρσ the dual field strength as usual and a and b two dimensionless constants
presumably of order unity.
a) Show that L respects charge conjugation (A → −A in this context), parity, and time
reversal, (and of course gauge invariance).
Solution:
Since Fµν = ∂µ Aν − ∂ν Aµ , each term in the Lagrangian has two powers of A and is
thereby invariant under A → −A. The term Fµν F µν is invariant under parity and time
reversal, as usual. The term Fµν F̃ µν = 12 εµνρσ Fµν Fρσ involves the antisymmetric tensor
εµνρσ and therefore breaks parity and time reversal. However, it appears squared in the
effective Lagrangian depicted above, so the Lagrangian preserves parity and time reversal.
Since Fµν is gauge invariant, the Lagrangian is also gauge invariant.
292
b) Draw the Feynman diagrams that give rise to the two dimension 8 terms shown. The
coefficients a and b were calculated by Euler and Kockel in 1935 and by Heisenberg and
Euler in 1936, quite a feat since they did not know about Feynman diagrams and any of
the modern quantum field theory setup.
Solution:
The electron propagator at zero momentum is m1e , so we will need four electron prop-
agators to generate the dimension-8 terms. We can draw the “box” diagram with four
external photons:
k1 k4
p+k1
p
p+k1+ k4
p−k 2
k2 k3
This is similar to the diagram in problem III.8.2, except now the internal lines are fermion
propagators and the external lines are spin-1 bosons rather than spin-0. To this diagram
(call it M1234 ) we add the crossed diagrams:
M = M1234 + M1243 + M1324 .
Recall also that the closed fermion loop generates a factor of (−1) in the amplitude.
we have
(Fµν F̃ µν )2 = Fµν F νρ Fρσ F σµ
293
In momentum space, this corresponds to an expansion in powers of ki /m.
1 7 e4
a = x and b = x where x ≡ .
4 16 360π 2
c) Explain why dimension 6 terms are absent in L. [Hint: One possible term is ∂λ Fµν ∂ λ F µν .]
Solution:
Dimension-6 terms can be removed by a field redefinition. See the appendix on p. 458.
d) Our mythical physicists do not know about the electron, but they are getting excited.
They are going to start doing photon-photon scattering experiments with a machine called
LPC that could produce photons with energy greater than me . Discuss what they will
see. Apply unitarity and the Cutkosky rules.
Solution:
If we cut the box diagram down the middle we find a process in which two photons
annihilate √
into an electron-positron pair. When the center-of-mass energy reaches the
threshold s = 2me , then the process γγ → e+ e− will occur. See problem III.8.3 for
discussion of a similar process.
45
arXiv:hep-th/0406216v1
294
6. Use the effective field theory approach to show that the scattering cross section of light
on an electrically neutral spin 1/2 particle (such as the neutron) goes like σ ∝ ω 2 to leading
order, not ω 4 . Argue further that the constant of proportionality can be fixed in terms of the
magnetic moment µ of the particle. [Historical note: This result was first obtained in 1954
by F. Low (Phys. Rev. 96: 1428) and by M. Gell-Mann and Murph L. Goldberger (Phys.
Rev. 96: 1433) using much more elaborate arguments.]
Solution:
The operator of lowest dimension we can write down is Fµν ψ̄σ µν ψ, which is dimension
[F ] + 2[ψ] = 2 + 2(3/2) = 5. This has only one factor of Fµν , meaning one derivative
on Aµ . The amplitude gets one power of frequency ω, so the cross section goes like ω 2 .
1 12 1 σ3 0
Since the spin operator is Sz = 2 σ = 2 and a magnetic field pointing in the
0 σ3
z-direction is B = 21 ε3ij Fij = F12 , the above operator is B ψ̄Sz ψ. The magnetic moment
µ is defined by the response of the spin to an external magnetic field, or in other words
µ = he|ψ̄Sz ψ|ei, where |ei is a state with one electron.
VIII.4 Supersymmetry
1. Construct the Wess-Zumino Lagrangian by the trial and error approach.
Solution:
With the benefit of hindsight, let us write only the terms we know we need: L = Lϕ + Lψ +
Lϕψ + LF , where
1
Lϕ = ∂µ ϕ† ∂ µ ϕ − ( µ2 ϕ2 + h.c.)
2
1 † µ 1
Lψ = iψ σ̄ ∂µ ψ − m ψψ + h.c.
2 2
Lϕψ = g ϕψψ + h.c.
LF = F † F − (κ F ϕ + λ F ϕ2 + h.c.)
We will for simplicity also assume that the couplings are all real, even though this is not
necessarily true in general. (Note that we have written 21 iψ † σ̄ µ ∂µ ψ + h.c. = iψ † σ̄ µ ∂µ ψ up to
a total derivative.)
295
The infinitesimal transformations have the form:
1
δLψ = iψ † σ̄ µ ∂µ (δψ) − m ψδψ + h.c.
2
1 † µβ̇α
= iψβ̇ σ̄ ∂µ (a σαν α̇ ξ †α̇ ∂ν ϕ + b ξα F ) − m ψ α (a σαµα̇ ξ †α̇ ∂µ ϕ + b ξα F ) + h.c.
2
1 1
= ia ψβ̇† σ̄ µβ̇α σαν α̇ ξ †α̇ ∂µ ∂ν ϕ + ib ψ † σ̄ µ ξ∂µ F − ma ψσ µ ξ † ∂µ ϕ − mb ψξF + h.c.
2 2
1
= ia ψβ̇† (2η µν δα̇β̇ )ξ †α̇ ∂µ ∂ν ϕ + 0 − ma ψσ µ ξ † ∂µ ϕ − mb ψξ F + h.c.
2
= ia ψ † ξ † ∂ 2 ϕ − ma ψσ µ ξ † ∂µ ϕ − mb ψξ F + h.c.
The 0 came from ∂µ F = 0, since F has no kinetic term and hence no dynamics. Since
iψ † ξ † ∂ 2 ϕ + h.c. = −i∂ µ ψ † ξ † ∂µ ϕ + h.c. = +iξ∂ µ ψ∂µ ϕ† + h.c. up to a total derivative, we can
write δLψ as
δLψ = +ia ξ∂ µ ψ∂µ ϕ† − ma ψσ µ ξ † ∂µ ϕ − mb ψξ F + h.c.
The first term in δLϕ cancels the first term in δLψ if a = +i. (When verifying this, recall
that ξψ ≡ ξ α ψα = −ψα ξ α = +ψ α ξα ≡ ψξ.)
296
Comparing the terms in δLϕψ and δLF , we want gb = λ and ig = −λc, where the minus sign
arises from ϕ2 ∂µ ψσ µ ξ † = −2ϕ ∂µ ϕ ψσ µ ξ † up to a total derivative. So we have
λ g
b= and c = −i .
g λ
At this point we should collect what we know so far. We have
g λ
δL = −µ2 ϕ ξψ − i( κ + m) ψσ µ ξ † ∂µ ϕ − (κ + m )ψξ F
λ g
We therefore require µ2 = 0 and λ = −κg/m. After rescaling F → mF/κ and then setting
κ = m, we obtain
† µ † µ † 1
L = ∂µ ϕ ∂ ϕ + iψ σ̄ ∂µ ψ + F F + − mψψ + gϕψψ − (m − gϕ)ϕF + h.c.
2
Upon the rephasing ψ → iψ and letting g → −g, we obtain the Lagrangian of equation (12)
on p. 467.
2. In general there may be N supercharges QIα , with I = 1, ..., N . Show that we can
have {QIα , QJβ } = εαβ Z IJ , where Z IJ denotes c-numbers known as central charges.
Solution:
Generalizing from p. 463, we have {QIα , QJβ } = Z IJ εαβ for some ordinary commuting num-
bers Z IJ . For the case N = 1, the fact that the left-hand side is symmetric in αβ while the
right-hand side is antisymmetric in αβ forces Z = 0. If N > 1 then we have more indices
to play with; if the left-hand side is symmetric under the exchange (α, I) ↔ (β, J), then the
right-hand side can also be symmetric and nonzero if Z IJ = −Z JI . So Z is an antisymmetric
matrix whose components are ordinary c-numbers.
3. From the fact that we do not know how to write consistent quantum field theories with
fields having spin greater than 2 show that the N in the previous exercise cannot exceed 8.
Theories with N = 8 supersymmetry are said to be maximally supersymmetric. Show that
if we do not want to include gravity, N cannot be greater than 4. Supersymmetric N = 4
Yang-Mills theory has many remarkable properties.
Solution:
To set the notation, consider N = 1 supersymmetry. Let |si denote a state of spin-s.
The generator Q raises the spin by +1/2, so that Q|si = |s + 21 i. Moreover, since Q is a
297
Grassmann generator we have Q2 = 0. Thus, for example, starting with the state s = 0 we
find the state Q|0i = | 21 i, and similarly Q† |0i = | − 12 i. Thus we find the chiral supermultiplet
containing the states s = 0 and s = ±1/2.
Now generalize to the case N > 1. The supersymmetry generators QI obtain an in-
dex I = 1, 2, ..., N . Each QI increases the spin by 1/2. Starting with a state |si, the
highest spin we can obtain using the QI is given by the action of all of the QI together:
Q1 Q2 ...QN |si = |s + 12 N i.
If we begin from s = −1 and we want to restrict to spin-0, spin-1/2 and spin-1 states
only, then we require −1 + 21 N ≤ 1 =⇒ N ≤ 4.
If we begin from s = −2 and demand that no states higher than spin-2 exist, then we
require −2 + 12 N ≤ 2 =⇒ N ≤ 8.
Solution:
∂ ∂
β
θα = εαγ β θγ = εαγ δβγ = εαβ .
∂θ ∂θ
Alternatively, Z
∂ ∂
β
θα = εαγ β θγ = εαγ dθβ θγ = εαγ δβγ = εαβ .
∂θ ∂θ
5. Work out δϕ, δψ, and δF precisely by computing δΦ = i(ξ α Qα + ξ¯α̇ Q̄α̇ )Φ.
Solution:
So when computing the action of the supercharges on Φ, we need to extract the term without
any θ or θ̄ to get δϕ, the term of order θ to get δψ, and the term of order (θθ) to get δF .
298
The supercharges are
∂
Qα = + − iσαµα̇ θ̄α̇ ∂µ
∂θα
∂
Q̄α̇ = − α̇ + iθα σαµα̇ ∂µ
∂ θ̄
Direct computation yields
√ √ √
Qα Φ = 2 ψα + i 2 (θσ µ θ̄)∂µ ψα + 2 (θ̄θ̄)θα (θ∂ 2 ψ) − iσαµα̇ θ̄α̇ (θθ)∂µ F − 2θα F
√
Q̄α̇ Φ = −σ̄ µα̇α θα ∂µ ϕ − 2 2 i σ̄ µα̇α θα (θ∂µ ψ)
εαβ ξ¯α̇ σ̄ µα̇β = εαβ εα̇β̇ ξ¯β̇ σ̄ µα̇β = −εαβ εβ̇ α̇ ξ¯β̇ σ̄ µα̇β = −ξ¯β̇ σαµβ̇
√
to get δψα = + √12 ∂µ ϕσαµα̇ ξ¯α̇ + i 2 ξα F , which is the form given under equation (9) on p.
465.
6. For any polynomial W (Φ) show that [W (Φ)]F = F [dW (ϕ)/dϕ]+ terms not involving
F . Show that for the theory (11) the potential energy is given by V (ϕ† , ϕ) = |∂W (ϕ)/∂ϕ|2 .
Solution:
299
√
numerical values are unimportant for this problem. Let p = (θθ)F and q = ϕ + 2 θψ in the
binomial expansion:
n
n
X n! √
Φ = [(θθ)F ] j (ϕ + 2 θψ)n−j
j =0
j!(n − j)!
We could continue to simplify the first term, but we are asked to consider only the terms
that involve F . So as far as this problem is concerned, we are free to write
D D D
X
n
X
n−1 d X d
W (Φ) = Cn Φ = Cn nF ϕ + ... = F Cn ϕn = F W (ϕ) + ...
n=0 n=0
dϕ n = 0 dϕ
where the (...) does not involve the field F . Now integrate out the auxiliary fields {Fa }3a = 1
using the Gaussian identity
Z
† † † †
R 4 R 4
DF DF † e i d x(F F +J F +F J) = e−i d x J (x)J(x)
where J † = dϕ
d
W (ϕ) and J = (J † )† = [ dϕ
d
W (ϕ)]∗ . After integrating out the F field, the
Lagrangian gets the term
∗ 2
† dW (ϕ) dW (ϕ) dW (ϕ)
LW = −J J = − =−
dϕ dϕ dϕ
Solution:
300
Let Ξ(x, θ) = ξ(x) + ... and {Φi (x, θ) = ϕi (x) + ...}ni= 1 be chiral superfields, and consider a
superpotential of the form46
Xn
W (Ξ, Φ) = Φi fi (Ξ)
i=1
Supersymmetry is spontaneously broken at tree level if there exist values of the scalar com-
ponents ξ(x) and ϕi (x) such that the potential is zero. From the previous problem, we know
that this is equivalent to the conditions
n
∂W (ξ, ϕ) X ∂W (ξ, ϕ)
= ϕi fi0 (ξ) = 0 and = fi (ξ) = 0 .
∂ξ i=1
∂ϕi
The first condition is satisfied by the field values ϕi = 0. The second equation, however,
imposes n conditions on a function of just one variable, ξ. If n ≥ 2, then these conditions
cannot be satisfied and supersymmetry is spontaneously broken.
For example, take n = 2 and consider the functions f1 (ξ) = ξ − λ and f2 (ξ) = ξ 2 , with
λ an arbitrary nonzero constant. Then ξ = λ sets f1 (ξ) = 0 but f2 (ξ) 6= 0, while ξ = 0 sets
f2 (ξ) = 0 but f1 (ξ) 6= 0.
8. If we can construct supersymmetric quantum field theory, surely we can construct super-
symmetric quantum mechanics. Indeed, consider Q1 ≡ 12 [σ1 P + σ2 W (x)] and Q2 ≡ 12 [σ2 P −
σ1 W (x)], where the momentum operator P = −i(d/dx) as usual. Define Q ≡ Q1 + iQ2 .
Study the properties of the Hamiltonian H defined by {Q, Q† } = 2H.
Solution:
46
Here we follow Section 26.5 in Weinberg, Volume III. The original paper is L. O’Raifeartaigh, “Sponta-
neous symmetry breaking for chiral scalar superfields,” Nucl. Phys. B96 (1975) 331-352.
301
then by direct computation we find Ṽ (x) = W (x)2 + √1
2
W 0 (x). The reason for defining this
new Hamiltonian is the following.
The underlying reason for this is that the two Hamiltonians belong to a supersymmetry
algebra. The operators
0 h†
H 0 0 0 †
H≡ , Q≡ , Q ≡
0 H̃ h 0 0 0
satisfy the algebra [H, Q] = [H, Q† ] = {Q, Q} = {Q† , Q† } = 0 and {Q, Q† } = H.
An important property of this Hamiltonian H is that the energy of any state is non-negative.
A state |invi invariant under supersymmetry satisfies Qi |invi = 0. A state |noti that is
not invariant under supersymmetry satisfies Qi |noti 6= 0. Since H is the sum of squares of
Hermitian operators, its eigenvalues are non-negative, so Qi |noti > 0. Since all states besides
|invi have energy greater than zero, the state |invi is the vacuum state. This tells us that if
a supersymmetric state exists, it must be the vacuum state of the theory.
This has implications for trying to break supersymmetry spontaneously. For further details,
consult Witten, “Dynamical Breaking of Supersymmetry,” Nucl. Phys. B185 (1981) 513-554.
IX Part N
IX.1 N.2 Gluon Scattering in Pure Yang-Mills Theory
1. Work out the two polarization vectors for general µ and µ̃ for a gluon moving along the
third direction.
Solution:
Let pµ = E(1, 0, 0, 1) be the momentum of a gluon moving along the third spatial direc-
tion. The 2-by-2 matrix representation of pµ is
µ 0 3 0 0
pαα̇ = σαα̇ pµ = E(σαα̇ − σαα̇ ) = 2E .
0 1
Therefore we have
√ √
0 0
λα = 2E , λ̃α̇ = 2E
1 1
302
as on p. 487. The polarization vectors (in 2-by-2 matrix notation) are given on p. 489:
µα λ̃α̇ λα µ̃α̇
+
αα̇ = , −
αα̇ =
hµλi [λµ]
where hµλi ≡ εαβ µα λβ and [λµ] ≡ εα̇β̇ λ̃α̇ µ̃β̇ . Using the explicit forms for λ and λ̃, we have
√
hµλi = ε12 µ1 λ2 + 0 = 2E µ1
and √
[λµ] = 0 + ε2̇1̇ λ̃2̇ µ̃1̇ = 2E µ̃1̇ .
The numerators are
√ √
µ1 0 µ1
µα λ̃α̇ = (0, 2E ) = 2E
µ2 0 µ2
and
√
0 0 0
λα µ̃α̇ = √ (µ̃1̇ , µ̃2̇ ) = 2E .
2E µ̃1̇ µ̃2̇
Therefore !
0 0
0 1
+
αα̇ = µ2 , −
αα̇ = µ̃2̇ .
0 µ1 1 µ̃1̇
3. Show that the result in (17) satisfies the reflection identity A(1− , 2− , 3+ , 4+ ) = A(4+ , 3+ , 2− , 1− ).
h12i4 p1 · p2
A(1− , 2− , 3+ , 4+ ) = = (17)
h12ih23ih34ih41i p2 · p3
Solution:
hcdi4
A(a+ , b+ , c− , d− ) = .
habihbcihcdihdai
Let a = 4, b = 3, c = 2, d = 1 to get
4. Show that the “all plus” and “all minus” cubic Yang-Mills vertices (see appendix 2)
vanish. [Hint: Choose the µ spinors wisely.]
Solution:
303
The color-stripped all-plus cubic amplitude is (see equation (23) on p. 495)
A(1+ , 2+ , 3+ ) = (+ + + + + + + + +
1 · 2 )(3 · p1 ) + (2 · 3 )(1 · p2 ) + (3 · 1 )(2 · p3 ) .
If µ1 = µ2 = µ3 , then all + + + + +
i · j = 0. Therefore A(1 , 2 , 3 ) = 0. Flipping all three helicities
from + to − will make each term in the amplitude have a factor of − −
i · j ∝ [µi µj ]. If we
choose all µ̃i equal, then A(1− , 2− , 3− ) = 0.
5. Why doesn’t the argument in the text that A(− + +... + +) vanish apply to A(− + +)?
Solution:
For the 3-point amplitude, either all of the hiji = 0 or all of the [ij] = 0 (see equation
(17) on p. 508). Therefore there are zeros in the denominator of the amplitude, so that
choosing the µ spinors wisely yields an indeterminate zero over zero situation rather than
just zero. Indeed, in appendix 2 on p. 496 we see the µ factors cancel out explicitly, leaving
behind the vertex (25) or (26).
6. Insert the expansion for the cubic vertex into (21) and derive M (W1+ , W2+ , Z3− ).
Z
α
M (Wi ) = d2 λi e iµ̃i λiα M (λi , λ̃i ) (21a)
Z
α̇
M (Zi ) = d2 λ̃i e iµi λ̃iα̇ M (λi , λ̃i ) (21b)
Solution:
Recall that the amplitudes M are related to the amplitudes A by including the momentum-
conserving delta function:
n
!
X
M (1, ..., n) = A(1, ..., n) δ (4) λi λ̃i .
i=1
304
We can write a fourier representation of the delta function
n
! Z
(4)
X
(i) (i) 1 4 ixαα̇ i λiα λ̃iα̇
P
δ λ λ̃ = d x e
i=1
(2π)4
and perform the relevant twistor transforms, using W for + and Z for −. The (2π)4 will
cancel from the resulting fourier transforms, so from now on we will simply drop the factors
of 2π. We have:
Z 3 Z
i(µ̃1 λ1 +µ̃2 λ2 +µ3 λ̃3 ) [12] αα̇
P
+ + − 2 2 2
M (W1 , W2 , Z3 ) = d λ1 d λ2 d λ̃3 e d4 x e ix i λiα λ̃iα̇
[23][31]
e i(µ3 +xλ3 )λ̃3
Z Z Z Z
3 4 2 i(µ̃1 +xλ̃1 )λ1 2 i(µ̃2 +xλ̃2 )λ2
= [12] dx d λ1 e d λ2 e d2 λ̃3
[23][31]
e i(µ3 +xλ3 )λ̃3
Z Z
= [12]3 d4 x δ 2 (µ̃1 + xλ̃1 ) δ 2 (µ̃2 + xλ̃2 ) d2 λ̃3 .
[23][31]
Since two spinors span the 2-dimensional spinor space, we can write a third spinor as a linear
combination of the first two: λ̃3 = a1 λ̃1 +a2 λ̃2 . The goal is to perform a change of integration
variables from d2 λ̃3 to da1 da2 .
The reader seeing this calculation for the first time may be distracted by the indices and
tildes, but in fact the change of variables is completely straightforward. Let
Thus we find d2 λ̃3 = |[12]|da1 da2 = [12]sgn([12])da1 da2 . We also have [13] = a2 [12] and
[32] = a1 [12] from writing the linear dependence equation as |3] = a1 |1]+a2 |2] and contracting
with the appropriate spinors. Therefore, the amplitude is
Z Z Z
+ + − 2 4 da1 i(µ3 +xλ3 )λ̃1 a1 da2 i(µ3 +xλ3 )λ̃2 a2
M (W1 , W2 , Z3 ) = [12] sgn([12]) d x e e
a1 a2
× δ 2 (µ̃1 + xλ̃1 )δ 2 (µ̃2 + xλ̃2 )
1
Using the property δ(ax) = |a| δ(x), the delta functions bring two powers of [12] in the
denominator, which cancels the factor of [12]2 in the numerator. In the exponentials, the
305
delta functions set xλ̃1 = −µ̃1 and = −µ̃2 . We also recognize the integral representation
R xλ̃2ixy
of the sign function as sgn(x) = dy y
e . We have:
M (W1+ , W2+ , Z3− ) = sgn([12]) sgn(µ3 λ̃1 − µ̃1 λ3 ) sgn(µ3 λ̃2 − µ̃2 λ3 )
h13i4
A(1− , 2+ , 3− , 4+ ) = (19)
h12ih23ih34ih41i
Solution:
[13]4
A(1+ , 2− , 3+ , 4− ) =
[12][23][34][41]
and the inverse transforms for W = (µ̃, λ̃) and Z = (λ, µ) are
Z Z 2
2 iµ̃α λiα d µ̃i −iµ̃αi λiα
M (Wi ) = d λi e i M (λi , λ̃i ) =⇒ M (λi , λ̃i ) = e M (Wi )
(2π)2
Z Z 2
2 iµα̇ λ̃iα̇ d µi −iµα̇i λ̃iα̇
M (Zi ) = d λ̃i e i M (λi , λ̃i ) =⇒ M (λi , λ̃i ) = e M (Zi )
(2π)2
306
Inverse twistor transforming gives
Z 2 Z 2 Z 2 Z 2
+ − + − d µ̃1 −iµ̃1 λ1 d µ2 −iµ2 λ̃2 d µ̃3 −µ̃3 λ3 d µ4 −iµ4 λ̃4
M (1 , 2 , 3 , 4 ) = e e e e
(2π)2 (2π)2 (2π)2 (2π)2
Z
da1 da2 da3 da4 i[a1 (µ̃1 λ2 −λ̃1 µ2 )+a2 (λ2 µ̃3 −µ2 λ̃3 )+a3 (µ̃3 λ4 −λ̃3 µ4 )+a4 (λ4 µ̃1 −µ4 λ̃1 )]
× e
a1 a2 a3 a4
Z Z 2 Z 2
da1 da4 d µ̃1 i(−λ1 +a1 λ2 +a4 λ4 )µ̃1 d µ2 −i(λ̃2 +a1 λ̃1 +a2 λ̃3 )µ2
= ... 2
e e
a1 a4 (2π) (2π)2
Z 2 Z 2
d µ̃3 i(−λ3 +a2 λ2 +a3 λ4 )µ̃3 d µ4 −i(λ̃4 +a3 λ̃3 +a4 λ̃1 )µ4
× 2
e e
(2π) (2π)2
Z
da1 da4 2
= ... δ (−λ1 + a1 λ2 + a4 λ4 ) δ 2 (λ̃2 + a1 λ̃1 + a2 λ̃3 )
a1 a4
× δ (−λ3 + a2 λ2 + a3 λ4 ) δ 2 (λ̃4 + a3 λ̃3 + a4 λ̃1 ) .
2
Now
R we Rneed the “inverse”
R 2 of the steps in the previous problem, namely to turn an integral
dai /ai daj /aj into d λ.
The first delta function sets λ1 = a1 λ2 + a4 λ4 , or in other words |1i = a1 |2i + a4 |4i. Multi-
h21i
plying on the left with h2| gives h21i = a4 h24i =⇒ a4 = h24i . Instead multiplying on the
left with h4| gives h41i = a1 h42i =⇒ a1 = h41ih42i
. As in N.2.6, we have d2 λ1 = |h24i| da1 da4
and therefore
da1 da4 |h24i| 2
= d λ1 .
a1 a4 h12ih41i
We can therefore perform the integrals over a1 and a4 to get
da2 da3 |h24i| 2 h41i
Z
+ − + −
M (1 , 2 , 3 , 4 ) = δ (λ̃2 + λ̃1 + a2 λ̃3 )
a2 a3 h12ih41i h42i
h21i
× δ 2 (−λ3 + a2 λ2 + a3 λ4 ) δ 2 (λ̃4 + a3 λ̃3 + λ̃1 ) .
h24i
Now repeat the previous procedure to perform the integral over a2 and a3 . The middle delta
function in the above expression sets λ3 = a2 λ2 + a3 λ4 , or in other words |3i = a2 |2i + a3 |4i.
h23i
Multiplying on the left with h2| implies h23i = a3 h24i =⇒ a3 = h24i . Multiplying with h4i
h43i
implies h43i = a2 h42i =⇒ a2 = h42i
. Using d2 λ3 = |h24i| da2 da3 , we get
da2 da3 |h24i| 2
= d λ3 .
a2 a3 h23ih34i
Therefore, we have
+ − + − |h24i| |h24i| h41i h43i h23i h21i
M (1 , 2 , 3 , 4 ) = δ 2 (λ̃2 + λ̃1 + λ̃3 ) δ 2 (λ̃4 + λ̃3 + λ̃1 )
h12ih41i h23ih34i h42i h42i h24i h24i
h24i4
= δ 2 (h24iλ̃2 −h41iλ̃1 −h43iλ̃3 ) δ 2 (h24iλ̃4 +h23iλ̃3 +h21iλ̃1 )
h12ih23ih34ih41i
307
The hope is that the two delta functions recollect into δ 4 ( 4i = 1 pi ). The equation 4i = 1 pi = 0
P P
is written in terms of spinors as λ1 λ̃1 + λ2 λ̃2 + λ3 λ̃3 + λ4 λ̃4 = 0. Contracting with λ4 implies
h41iλ̃1 + h42iλ̃2 + h43iλ̃3 = 0, which is the condition mandated by the first delta function
above (h42i = −h24i). Contracting with λ2 implies h21iλ̃1 + h23iλ̃3 + h24iλ̃4 = 0, which is
the condition mandated by the second delta function. We therefore arrive at the result
4
+ − + − h24i4 4
X
M (1 , 2 , 3 , 4 ) = δ ( pi ) .
h12ih23ih34ih41i i=1
Recall that we wanted to show that the prefactor A ≡ A(1+ , 2− , 3+ , 4− ) of the delta function
is equal to
[13]4
.
[12][23][34][41]
We will now show that the two forms are indeed equivalent. We will make repeated use of
momentum conservation written in the form
Since (p1 +p2 )2 = (p3 +p4 )2 , the term in parentheses equals 1. Since [14][32] = (−[41])(−[23]) =
[41][23], we have
[13]4
A=
[12][23][34][41]
which is the desired result. Note that had we chosen to integrate over the delta functions in
a different order we could have arrived at this form directly. It is however worth showing the
relationship between the forms with angle and square brackets.
308
8. Show that SL(4, R) is locally isomorphic to the conformal group. [Hint: Identify the
15 = 42 − 1 generators of the conformal group (3 rotations J i , 3 boosts K i , 1 dilation D,
4 translations P µ , and 4 conformal transformations K µ ) with the 15 traceless real 4 by 4
matrices.]
Solution:
Our strategy will be to write down all possible traceless real 4-by-4 matrices and identify
them with the conformal generators based on their commutation relations. The 15 generators
M µν , P µ , D, K µ of the conformal group47 satisfy the commutation relations:
with all others zero. The generators of rotations are J i = 21 εijk Mjk , and the generators of
boosts are B i = M 0i . (We will use B i to denote the boosts so as not to confuse them with
the spatial components of the special conformal generators.)
For two real spinors λα and µα̇ , we have learned in the text that conformal transformations
act naturally on the 4-dimensional column vector
λα
Z=
µα̇
which transforms under the 4-dimensional representation of SL(4, R). The lie algebra of
SL(4, R) consists of the 15 real traceless 4-by-4 matrices, which we now attempt to construct.
From learning about spinor representations of the Lorentz group back in Chapter II, we
know how the rotations J i and the boosts B i act on the vector Z:
1 i 1 i
i 2
σ 0 i −i 2 σ 0
J = , B = .
0 12 σ i 0 +i 12 σ i
Here we are working in the signature η = (−, +, +, +) of SO(3, 1). These generators can be
repackaged into the relativistic notation
1 µν β
µν 2
(σ )α 0
M =
0 − 21 (σ̄ µν )α̇β̇
47
Acting on scalar fields, these generators take the form:
Mµν = i(xµ ∂ν − xν ∂µ )
Pµ = −i∂µ , D = ixµ ∂µ
Kµ = −i(x2 ∂µ − 2xµ xν ∂ν ) .
In addition to these, spinors and vectors also transform with the appropriate matrices. These matrices are
what we are trying to identify.
309
where σ µν ≡ + 12 i(σ µ σ̄ ν − σ ν σ̄ µ ) and σ̄ µν ≡ − 12 i(σ̄ µ σ ν − σ̄ ν σ µ ), and we have defined the
4-vector
satisfies
det x = x20 − x21 − x22 − x23 = −η µν xµ xν
with the SO(3, 1) metric η µν = diag(−1, +1, +1, +1). Note that the components of the
matrices σ µν and σ̄ µν are numerically equal to:
σ ij = +εijk σ k , σ 0i = −iσ i
σ̄ ij = −εijk σ k , σ̄ 0i = −iσ i
From looking at the index structure of these tensors as well as that of the spinors in Z = (λ, µ),
the only other traceless 4-by-4 matrices we can write down are of the forms
1 µ 1 β
µ 0 σ
2 αβ̇
δ
2 a
0
X± ≡ and Y ≡
± 1 σ̄ µα̇β 0 0 − 12 δ α̇β̇
2
up to an overall constant for each. By direct computation, we find [Y, X±µ ] = X∓µ , and
therefore:
We also find
If we identify D = iY, P µ = −i(X+µ + X−µ ) and K µ = −i(X+µ − X−µ ), then we get the correct
commutation relations for the conformal group. This completes the problem.
310
IX.2 N.3 Subterranean Connections in Gauge Theories
1. Show that the structure of Lie algebra (21) emerges naturally.
fabe fcde + face fbde + fade fbce = 0 (21)
Solution:
First note that equation (21) on p. 509 has a sign error on the second term. The Jacobi
identity should read
fabe fcde + fcae fbde + fade fbce = 0 .
Recall that the recursion scheme is to separate the diagram into sets L and R, with r L
and s R. If (r, s) = (1, 4), we get one separation with a pole in s = (p1 + p2 )2 = (p1 + p3 )2 ,
and one with a pole in t = (p1 + p3 )2 = (p2 + p4 )2 :
−M(1a , 2b , 3c , 4d ) = fabe fecd A(1, 2, 3, 4) + face febd A(1, 3, 2, 4)
If (r, s) = (1, 3), then we get one pole in s and one pole in u = (p1 + p4 )2 = (p2 + p3 )2 :
−M(1a , 2b , 3c , 4d ) = fabe fedc A(1, 2, 4, 3) + fade febc A(1, 4, 2, 3)
If (r, s) = (1, 2), then we get one pole in t and one pole in u:
−M(1a , 2b , 3c , 4d ) = face fedb A(1, 3, 4, 2) + fade fecb A(1, 4, 3, 2)
To make sense of these, we should relate the various color-stripped amplitudes to each other.
Since the amplitudes with all +, all −, or 3 + or 3 − are zero, we can without loss of
generality consider the amplitude with two + helicities and two − helicities. From equation
(19) on p. 493 and the surrounding discussion, we know how to compute the amplitude for
any ordering of the + and − helicities, so without loss of generality we choose (1− , 2− , 3+ , 4+ ).
Thus consider the amplitude
h12i4
A(1− , 2− , 3+ , 4+ ) =
h12ih23ih34ih41i
from equation (17) on p. 492. The goal P
is to relate all of the other As to this one. We will
repeatedly use momentum conservation 4i = 1 pi = 0 in the form
|1i[1| + |2i[2| + |3i[3| + |4i[4| = 0 . (∗)
Consider the amplitude A(1− , 3+ , 2− , 4+ ). From relabeling equation (19) on p. 493, we know
h12i4 h12i4
− + − + −h12ih34i
A(1 , 3 , 2 , 4 ) = = .
h13ih32ih24ih41i h12ih23ih34ih41i h13ih24i
Multiply the momentum conservation equation (∗) by h1| on the left and by |4] on the right
to get
h12i [34]
h12i[24] + h13i[34] = 0 =⇒ =− .
h13i [24]
311
Thus
−h12ih34i +[34]h34i s
= = .
h13ih24i [24]h24i t
Therefore, we have
s
A(1− , 3+ , 2− , 4+ ) = A(1− , 2− , 3+ , 4+ ) .
t
Next, consider
h12i4 h12i4
− − + + −h23ih41i
A(1 , 2 , 4 , 3 ) = =
h12ih24ih43ih31i h12ih23ih34ih41i h24ih31i
Multiply (∗) by h2| and |1] to get h23i[31] + h24i[41] = 0 =⇒ h23i/h24i = −[41]/[31] and
thus
−h23ih41i [41]h41i u
= = .
h24ih31i [31]h31i t
We have
u
A(1− , 2− , 4+ , 3+ ) = A(1− , 2− , 3+ , 4+ ) .
t
Next, consider
h12i4 h12i4
− + − + −h12ih34i
A(1 , 4 , 2 , 3 ) = = .
h14ih42ih23ih31i h12ih23ih34ih41i h42ih31i
Multiply (∗) by h2| and [3] to get h21i[13] + h24i[43] = 0 =⇒ h12i/h42i = −[34]/[31] and
thus
−h12ih34i +[34]h34i s
= = .
h42ih31i [31]h31i t
We have
s
A(1− , 4+ , 2− , 3+ ) = A(1− , 2− , 3+ , 4+ ) .
t
Two more to go. The next one is
h12i4 h12i4
− + + − −h23ih41i
A(1 , 3 , 4 , 2 ) = = .
h13ih34ih42ih21i h12ih23ih34ih41i h13ih42i
Multiply (∗) by h3| and |4] to get h31i[14] + h32i[24] = 0 =⇒ h23i/h13i = −[41]/[42] and
thus
−h23ih41i +[41]h41i u
= = .
h13ih42i [42]h42i t
We have
u
A(1− , 3+ , 4+ , 2− ) = A(1− , 2− , 3+ , 4+ ) .
t
Fortunately, the last one is easy:
h12i4 h12i4
A(1− , 4+ , 3+ , 2− ) = = = A(1− , 2− , 3+ , 4+ ) .
h14ih43ih32ih21i h12ih23ih34ih41i
312
We therefore arrive at the amplitude
− + +
s
−M(1−
a , 2b , 3c , 4d ) = f f
abe ecd + f − − + +
ace ebd A(1 , 2 , 3 , 4 )
f
u t
s
= fabe fedc + fade febc A(1− , 2− , 3+ , 4+ )
ut t
= face fedb + fade fecb A(1− , 2− , 3+ , 4+ ) .
t
The equalities reflect the physical necessity for the amplitude to be the same regardless of
which (r, s) we choose for the recursion formula. Subtracting and multiplying by t gives the
consistency conditions
tfabe fecd + s(face febd − fade febc ) − ufabe fedc = 0
t(fabe fecd − fade fecb ) + sface febd − uface fedb = 0
Subtract (2) from (1) to get
tfade fecb − sfade febc + u(face fedb − fabe fedc ) = 0 .
Since fecb = −febc and since s + t + u = 0 =⇒ t + s = −u, we get
u(fade febc + face fedb − fabe fedc ) = 0 .
u 6= 0 and fedc = −fecd implies
fade febc + face fedb + fabe fecd = 0 .
This is the Jacobi identity.
Particle physics experimentalists are fond of saying that yesterday’s spectacular discovery
is today’s calibration and tomorrow’s annoying background. The canonical example is the
Nobel-winning discovery of the CP-violating decay of the KL meson into two pions. In the-
oretical physics, yesterday’s discovery is today’s homework exercise and tomorrow’s trivium.
X RL X ML (zL )MR (zL )
M(0) = − =− (3)
L,h
zL L,h
PL (0)2
Solution:
We are interested in the color-ordered amplitude A(1− , 2− , 3+ , 4+ ) with legs 1 and 2 com-
plexified. There is only one way to split up the amplitude into 3-point vertices for which legs
1 and 2 do not belong to the same partition while maintaining the color ordering:
313
1 2
1 2
K K
4 3
4 3
[12]3 h12i3
A(1+ , 2+ , 3+ ) = , A(1− , 2− , 3+ ) = .
[23][31] h23ih31i
When the internal momentum K ≡ |Ki[K| flows out of the vertex, we can effect the replace-
ment K → −K by replacing its spinors as |Ki → i|Ki, |K] → i|K]. We have:
" ! ! ! !#
3 3 3 3
1 [ K̂4] (ih K̂ 2̂i) h1̂K̂i (i[3 K̂])
A(1− , 2− , 3+ , 4+ ) = − +
2h23i[23] [41̂][1̂K̂] h2̂3i(ih3K̂i) hK̂4ih41̂i (i[K̂ 2̂])[2̂3]
" #
1 [K̂4]3 hK̂ 2̂i3 h1̂K̂i3 [3K̂]3
=+ + .
2h23i[23] [41̂][1̂K̂] h2̂3ih3K̂i hK̂4ih41̂i [K̂ 2̂][2̂3]
Here we choose (r, s) = (1, 2) and thereby complexify the momenta as:
with |1i and |2] unchanged. This effects the shift p̂1 = p1 +zL q , p̂2 = p2 −zL q with q = |1i[2|.
Here we have removed the hats from |1i and |2] since they are not complexified. Now
use the hatted version of equation (A):
314
Therefore:
[K̂4]3 hK̂ 2̂i3 (h21i)[41̂])3 h21i3 [41̂]
= = .
[41̂][1̂K̂]h2̂3ih3K̂i [41̂]h2̂3i(h34i[41̂]) h2̂3ih34i
[1̂4] [34]
=−
h2̂3i h21i
and therefore
h21i3 [41̂] h21i2 [34] h21i3
[34]
= = .
h2̂3ih34i h34i h23ih34i h21i[23]
Multiplying momentum conservation (unhatted) by h1|...|3] implies that the term in paren-
theses equals −h41i, so all together we have48
Now let us simplify the second term. Go back to equation (A) and obtain:
Therefore:
h41i[12]
h41i[12] + h43i[32] = 0 =⇒ = −h34i
[23]
315
This is the second half of the correct answer. Adding the two terms cancels the overall factor
of 2, and we arrive at the result
h12i4
A(1− , 2− , 3+ , 4+ ) = .
h12ih23ih34ih41i
3. Using the explicit forms given for A(1− , 2− , 3+ , 4+ ) and A(1− , 2+ , 3− , 4+ ) in the preceding
chapter, check the estimated large z behavior in (13-15).
1
M−+ (z) → (13)
z
−− 1
M (z) → (14)
z
M+− (z) → z 3 (15)
Solution:
h12i4 h13i4
A(1− , 2− , 3+ , 4+ ) = , A(1− , 2+ , 3− , 4+ ) =
h12ih23ih34ih41i h12ih23ih34ih41i
and the physical amplitude specified by helicity labels for particles r and s is Mhr hs (z) ≡
[hr r (z)]µ Mµν (z)[hs s (z)]ν . The spinorial version of the deformations pr (z) = pr + zq and
ps (z) = ps − zq are given on p. 501, with pi = λi λ̃i :
and q = λr λ̃s . Choose (r, s) = (1, 2). Then the only z-dependence in the amplitude
A(1− , 2+ , 3− , 4+ ) potentially comes from h12i and h23i. But h12i(z) = h12i − zh11i = h12i,
so really the only dependence on z comes from h23i(z) = h23i − zh13i ∼ −zh13i. Therefore
we find M−+ (z) ∼ 1/z.
The case for A(1− , 2− , 3+ , 4+ ) is identical, with the only z-dependence coming from a factor
of h23 in the denominator, leading immediately to M−− (z) ∼ 1/z.
Now choose (r, s) = (2, 3). Then h13i(z) = h13i − zh12i ∼ z, leading to a factor of z 4
in the numerator. This is canceled partially by a factor h34i(z) = h34i − zh14i ∼ z in the
denominator, leading to M+− (z) ∼ z 3 .
316
4. Worry about the sloppy handling of factors of 2 in appendix 1. [Hint: The final result
is correct because the polarization vectors in (5-6) are normalized to |ε|2 = 2 for convenience.]
Solution:
Let us be pedantic about distinguishing among pµ , pαα̇ and pα̇α . Given the Lorentz 4-vector
pµ , define 6 pαα̇ ≡ pµ σαµα̇ and 6 p̄α̇α ≡ pµ σ̄ µα̇α , where σ̄ µα̇α ≡ εαβ εα̇β̇ σβµβ̇ . Then for lightlike mo-
menta p2 ≡ pµ pµ = 0, we define 6 pαα̇ ≡ λα λ̃α̇ . Then λα λ̃α̇ = εαβ εα̇β̇ λβ λ̃β̇ = εαβ εα̇β̇ 6 pβ β̇ =6 p̄α̇α .
Therefore:
hλλ0 i[λλ0 ] = λα λ0α λ̃α̇ λ̃0α̇ =6 pαα̇ 6 p̄α̇α = pµ pν tr(σ µ σ̄ ν ) = 2 pµ pµ .
So the top of page 511 should read PL (0)2 ≡ PL (0)µ PL (0)µ = 2 p1µ pµ2 = h12i[12]. Thus the
result (25) is correct.
317
X Appendix E: Dotted and Undotted Indices
1. Show that ησ µν ψ = −ψσ µν η and χ̄σ̄ µ ψ = −ψσ µ χ̄.
= −ψ β σβµβ̇ χ̄β̇
= −ψσ µ χ̄ X
4 ησ µν ψ = ησ µ σ̄ ν ψ − (µ ↔ ν) = η α σαµβ̇ σ̄ ν β̇β ψβ − (µ ↔ ν)
= −ψβ σαµβ̇ σ̄ ν β̇β η α − (µ ↔ ν)
= −ψβ [(εαγ εβ̇ γ̇ σ̄ µγ̇γ )(εβ̇ δ̇ εβδ σδνδ̇ )]η α − (µ ↔ ν)
= −ψβ [εαγ εβδ (−δγ̇δ̇ )σ̄ µγ̇γ σδνδ̇ ]η α − (µ ↔ ν) [ note: εβ̇ γ̇ εβ̇ δ̇ = −εγ̇ β̇ εβ̇ δ̇ = −δγ̇δ̇ ]
= −ψβ (−εαγ εβδ σ̄ µγ̇γ σδνδ̇ )η α − (µ ↔ ν)
= +(εβδ ψβ )(σ̄ µγ̇γ σδνδ̇ )(εαγ η α ) − (µ ↔ ν)
= (−ψ δ )(σ̄ µγ̇γ σδνδ̇ )(−ηγ ) − (µ ↔ ν)
= +ψσ ν σ̄ µ η − (µ ↔ ν) [ put in matrix multiplication order ]
= −[ψσ µ σ̄ ν η − (µ ↔ ν)]
= −4 ψσ µν η =⇒ ησ µν ψ = −ψσ µν η X
¯ = − 1 (θσ µ ξ)(
2. Show that (θϕ)(χ̄ξ) ¯ χ̄σ̄µ ϕ).
2
Solution:
First we need
σαµα̇ σ̄µβ̇β = Cδαβ δα̇β̇
which we know by SU (2) ⊗ SU (2) invariance. Choose α = β = α̇ = β̇ = 1 to get
µ 11
C = σ11 σ̄µ = (σ 0 )11 (σ̄ 0 )11 − (σ 3 )11 (σ̄ 3 )11 = 1 − (−1) = 2
Therefore
σαµα̇ σ̄µβ̇β = 2 δαβ δα̇β̇ .
318
Now compute:
¯ χ̄σ̄µ ϕ) = θα σ µ ξ¯α̇ χ̄ σ̄ β̇β ϕβ
(θσ µ ξ)( αα̇ β̇ µ
3. Show that θα θβ = 12 (θθ)δ αβ . [Hint: simply evaluate the two sides for all possible cases.]
Solution:
The indices have to match up by SU (2) invariance, and then we can fix the constant.
θα θβ = C(θθ)δ αβ
Contract both sides with δ βα and note that δ αα = 2 to get
1
(θθ) = C(θθ)δ αα = 2C(θθ) =⇒ C = .
2
Alternatively, we can follow the hint and just check. α = 1, β = 2 implies 0 for the right-hand
side, and
θ1 θ2 = ε12 θ2 θ2 = 0
for the left-hand side, since θ22 = 0 since the square of any Grassmann number is zero. For
α = β = 1, we have 21 (θθ) = 12 (θ1 θ1 + θ2 θ2 ), and θ2 θ2 = ε21 θ1 ε21 θ1 = −(ε21 )2 θ1 θ1 = −θ1 θ1 =
+θ1 θ1 , so 21 (θθ) = θ1 θ1 , which matches the left-hand side θα θβ for α = β = 1.
Here we collect a few relations between two-component spinors and four-component spinors
for use in calculating Feynman diagrams using two-component notation. For an extensive
review, consult arXiv:0812.1594v5 [hep-ph].
eα (x)
Take the Dirac field for the electron for definiteness: E(x) ≡ . The mode ex-
ē†α̇ (x)
pansions for the fields eα (x) and ēα (x) are:
XZ d3 p −ip·x † +ip·x
eα (x) = b(~
p , s)u α (~
p , s) e + d (~
p , s)vα (~p , s) e
s
(2π)3 2ωp
XZ d3 p
d(~p, s)uα (~p, s) e−ip·x + b† (~p, s)vα (~p, s) e+ip·x .
ēα (x) = 3
s
(2π) 2ωp
319
The creation operators b† and d† create an electron e− and a positron e+ respectively.
The wave functions u(~p, s) and v(~p, s) are related to the 4-component wave functions U (~p, s)
and V (~p, s) as
uα (~p, s) vα (~p, s)
U (~p, s) = , V (~p, s) =
v †α̇ (~p, s) u†α̇ (~p, s)
P
The
P usual spin sum relations (see p. 110 of the text) s U (~ p, s)Ū (~p, s) =6 p + m and
p, s)V̄ (~p, s) =6 p − m, along with the gamma matrices in the Weyl basis (see appendix
s V (~
E)
µ 0 σµ
γ =
σ̄ µ 0
imply the spin sum relations
X uα v β uα u†
! !
β
β̇
m δ α 6 p αβ̇
†α̇ † = ¯ α̇β
†α̇ β
v v v uβ̇ 6 p m δ α̇β̇
s
X vα uβ vα v †
! !
β
β̇
−m δ α 6 p αβ̇
†α̇ † = ¯ α̇β
†α̇ β
u u u v 6 p −m δ α̇β̇
s β̇
α̇β
where we have defined 6 pαβ̇ ≡ σαµβ̇ pµ and 6 ¯p ≡ σ̄ µα̇β pµ . Since tr(σ µ σ̄ ν ) = 2η µν [with
β̇α
η = (+, −, −, −)] we have 6 pαβ̇ 6 p¯0 = 2p · p 0 .
An incoming electron state with momentum p~ and spin s is |e− (~p, s)i = (2π)3 2ωp b† (~p, s)|0i.
This implies the following rules for the external state wave function for an incoming electron:
h0|e(x)|e− (~p, s)i = u(~p, s) e−ip·x , h0|ē(x)|e− (~p, s)i = 0
h0|e† (x)|e− (~p, s)i = 0 , h0|ē† (x)|e− (~p, s)i = v † (~p, s) e−ip·x .
An outgoing electron state with momentum p~ and spin s is he− (~p, s)| = h0|b(~p, s)(2π)3 2ωp .
This implies the following rules for the external state wave function for an outgoing electron:
he− (~p, s)|e(x)|0i = 0 , he− (~p, s)|ē(x)|0i = v(~p, s) e+ip·x
he− (~p, s)|e† (x)|0i = u† (~p, s) e+ip·x , he− (~p, s)|ē† (x)|0i = 0 .
An incoming positron state with momentum p~ and spin s is |e+ (~p, s)i = (2π)3 2ωp d† (~p, s)|0i.
This implies the following rules for the external state wave function for an incoming positron:
h0|e(x)|e+ (~p, s)i = 0 , h0|ē(x)|e+ (~p, s)i = u(~p, s) e−ip·x
h0|e† (x)|e+ (~p, s)i = v † (~p, s) e−ip·x , h0|ē† (x)|e+ (~p, s)i = 0 .
An outgoing positron state with momentum p~ and spin s is he+ (~p, s)| = h0|d(~p, s)(2π)3 2ωp .
This implies the following rules for the external state wave function for an incoming positron:
he+ (~p, s)|e(x)|0i = v(~p, s) e+ip·x , he+ (~p, s)|ē(x)|0i = 0
he+ (~p, s)|e† (x)|0i = 0 , he+ (~p, s)|ē† (x)|0i = u† (~p, s) e+ip·x .
320