Lecture Notes: Relativistic Quantum Mechanics
Lecture Notes: Relativistic Quantum Mechanics
1
Disclaimer
These lecture notes accompany the final-year undergraduate lecture course
on “Relativistic Quantum Mechanics”, consisting of 12 lectures, delivered
during the 2020/2021 academic year in an online format. The notes are
by no means original. Instead I shamelessly borrowed from a multitude of
resources, and tried to put material into a coherent form. I also try and
cite online links to books etc., where possible – I am aware that the library
holds some of them, also in electronic form.
The lectures mainly deal with second quantisation, a topic that has been
excellently covered in the literature and on the web. There are many truly
excellent textbooks on the topic, often named “Introduction to Quantum
field Theory” or similar, for example the books by Peskin & Schröder [1],
Griffiths [2], Schwartz, Zee, or Hatfield [3], in addition to a multitude of
freely available lectures notes on the web:
Mark Srednicki’s notes on Quantum Field Theory [4], which have since
been published as a book;
The lecture notes will be continuously updated over the course of the year -
please check the date on the front page to keep track of changes. When you
compare the notes with books you will realise that notation and conventions
differ between different resources. However, quite often these differences boil
down to trivial normalisations. I’ve tried, hopefully successfully, to be at
least self-consistent.
The notes are supplemented with worked examples and problems through-
out, and I cannot overemphasise how important it is to actually calculate
things on your own. Tougher, expert-level problems are identified with an
asterisk. They are outside the scope of examinable material and are solely
geared to helping interested students to develop a deeper understanding of
the subject and to contextualising the material in a wider perspective. I
have also added “extremely unbelievably hard” questions, indicated with
two asterisks. They cover material that is entirely beyond the scope of the
course, but may trigger some further reading and digging by students with
a soft spot for the abyss that is Quantum Field Theory.
Over the course of six weeks we will work through the analogue of 12 lectures
- I will try to highlight and explain crucial concepts in short movies with
me working through things on a white board - however, these movies are
i
Week Chapters Comments & Suggested Problems
1 1-3 Reminder of important concepts. Introduction
to Classical Field Theory in Lagrange formalism:
real and complex scalar fields and electrodynam-
ics. Conserved current and conserved charge.
2 4 Logic of 2nd quantisation and first example: real
scalar field theory. 2nd quantisation of complex
scalar theory. More on conserved current and
charges, this time in the quantum world.
3 5 Introducing the Dirac equation without quantisa-
tion: Linearising Klein-Gordon equation, spinors,
their properties, and γ matrices. 2nd quantisation
of the Dirac equation: using anti-commutators for
the quantisation conditions on fermions.
4 6 Free electrodynamics fields. Impact of gauge in-
variance: “over-quantising”. 2nd quantisation in
Coulomb and Lorentz gauge.
5 7 Time-ordered products are the Green’s functions
(propagators) of free theories.
6 8 Interacting field theories. A first stab at the S-
matrix and Wick’s theorem. This is extended
reading and will not be subject of the relativis-
tic quantum mechanics part of the exam.
by no means complete and they are mainly meant to structure your own,
self-driven learning. Below a table, Tab 1, of what material would have been
covered week-by-week.
ii
Contents
1 Introduction 1
2 Recapitulation 3
2.1 Natural Units . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.2 Some mathematics . . . . . . . . . . . . . . . . . . . . . . . . 3
2.3 Four-Vectors and Minkowski Space . . . . . . . . . . . . . . . 4
2.4 Lorentz Transformations . . . . . . . . . . . . . . . . . . . . . 7
2.5 Lagrange and Hamilton Formalism for Point Particles . . . . 10
2.6 First Quantisation of the Harmonic Oscillator . . . . . . . . . 12
2.7 Problems & Solutions . . . . . . . . . . . . . . . . . . . . . . 15
3 Classical Fields 31
3.1 One-Dimensional Lattice . . . . . . . . . . . . . . . . . . . . . 31
3.2 Scalar Fields: Real Scalars . . . . . . . . . . . . . . . . . . . . 35
3.3 Scalar Fields: Complex Scalars . . . . . . . . . . . . . . . . . 37
3.4 Vector Fields: Maxwell’s Equations . . . . . . . . . . . . . . . 40
3.5 Hamiltonian Formulation . . . . . . . . . . . . . . . . . . . . 43
3.6 Problems & Solutions . . . . . . . . . . . . . . . . . . . . . . 45
4 Second Quantisation 60
4.1 A How-To Guide . . . . . . . . . . . . . . . . . . . . . . . . . 60
4.2 Second Quantisation of the Real Scalar Field . . . . . . . . . 61
4.3 A Little Detour: Causal Structure of the Theory . . . . . . . 67
4.4 Second Quantisation of the Complex Scalar Field . . . . . . . 69
4.5 Problems & Solutions . . . . . . . . . . . . . . . . . . . . . . 72
5 Fermions 87
5.1 The Dirac Equation . . . . . . . . . . . . . . . . . . . . . . . 87
5.2 Second Quantisation . . . . . . . . . . . . . . . . . . . . . . . 94
5.3 Problems & Solutions . . . . . . . . . . . . . . . . . . . . . . 101
6 Electrodynamics 121
6.1 Gauge Invariance as Obstacle . . . . . . . . . . . . . . . . . . 121
6.2 Coulomb Gauge . . . . . . . . . . . . . . . . . . . . . . . . . . 124
6.3 Lorentz Gauge . . . . . . . . . . . . . . . . . . . . . . . . . . 128
6.4 Problems & Solutions . . . . . . . . . . . . . . . . . . . . . . 132
iii
8 Interacting Fields 164
8.1 Perturbative Expansion: Born Series . . . . . . . . . . . . . . 164
8.2 Interacting Field Theory: General Thoughts . . . . . . . . . . 165
8.3 Interacting Field Theory: λφ4 . . . . . . . . . . . . . . . . . . 168
iv
1 Introduction
In this course, “Relativistic Quantum Mechanics”, we combine Quantum
Mechanics with Special Relativity and develop a formalism to quantise fields
in a Lorentz-invariant way.
We will recapitulate the Lagrange and Hamilton formalism for the treat-
ment of classical point particles as well as the quantisation of the harmonic
oscillator through creation and annihilation operators. Building on the for-
mer, we will briefly analyse the Lagrange formalism and the derivation of
Euler-Lagrange equations of motion for a discrete system, before taking the
continuum limit, resulting in the Lagrange formulation of field dynamics.
We will analyse free real and complex scalar fields in this formalism, and
for the latter, we will find a symmetry – phase shifts of the fields – that
leaves the Lagrangian invariant. We will see that such invariances result in
conserved currents and charges. We will further exemplify the power of the
formalism by constructing a Lagrange density for the electromagnetic fields
and deriving Maxwell’s equations from it.
To quantise fields we will copy the steps known from single-particle systems,
in particular the harmonic oscillator, and adapt it to the case of fields. In
so doing we effectively replace the role of position and momentum of the
particle, and the corresponding operators, with the field and its conjugate
momentum. The resulting logic is to replace the functions describing fields
and their conjugate momenta with field and momentum operators, and to de-
mand suitable commutator relations for them. This is called second quanti-
sation. As a consequence of relativistic invariance, encoded in the quadratic
energy-momentum relation of E 2 − p2 = m2 , solutions with negative energy
become possible. Demanding a Hamiltonian with an energy spectrum that
is bounded from below, i.e. a physically meaningful ground state or vac-
uum, necessitates their interpretation as anti-particles. It also immediately
implies we have arrived at a multi-particle theory, because pairs of parti-
cles and anti-particles with short lifetimes can be produced. We will check,
by explicit calculation, that the resulting theory maintains causality at a
microscopic level, by asserting that commutators of causally disconnected
fields always vanish and that they therefore cannot impact onto each other.
After second quantisation of the simplest possible theory, a single free real
scalar field, we analyse the structure of a free complex scalar field. We will
recover the current and charge stemming from the phase invariance of the
Lagrangian and we will by explicit calculation show that the charge and
the Hamilton operators commute, making charge conservation of the theory
manifest.
After analysing the free scalar or Klein-Gordon fields we will turn our atten-
tion to the treatment of spin-1/2 particles in the celebrated Dirac equation.
We will analyse its structure and ingredients – γ matrices and spinors –
and their properties before second quantisation of the theory. Reflecting
1
the fermionic nature of the particles, we will use anti-commutators {·, ·}
instead of commutators [·, ·] for the quantisation conditions. Similar to the
case of the complex scalar field, also the Lagrangian for free spinors enjoys
invariance under phase transformations of the fields, and again this leads to
a conserved charge.
We then turn our attention to the quantisation of electrodynamics and the
free electromagnetic fields. There, we will encounter an interesting problem:
the vector potential Aµ , on which we build the theory, naively speaking, has
four degrees of freedom in its four-components, but the physical field has
only two degrees of freedom, the well-known linear or circular polarisation
states of the photons, the quanta of electromagnetism. This necessitates the
imposition of additional conditions onto the theory, to correctly reflect its
physical content. In more formalised language, this problem is a result of the
gauge invariance of the underlying theory, electromagnetism, which results
in identical physical fields for different vector potentials. It will become clear
that the problem of the additional content will be fixed by fixing the gauge
of the theory, and we will see how this is shapes the additional conditions
we will impose on the theory.
Having quantised various free field theories and discussing some of their
properties, we will start with developing a framework to analyse their dy-
namical behaviour. To this end we will build on the concept of Green’s
functions and construct the Green’s functions of our quantised theories. It
will turn out that these “propagators” are the vacuum expectation values
of time-ordered products of the field operators.
2
2 Recapitulation
In this section we recapitulate important concepts from previous lectures
and properties of the objects we will use throughout the lecture. The aim
is not to explain in detail how things work or why, but to provide you with
a unified notation and nomenclature. If necessary, please, re-familiarise
yourselves with the concepts in this section.
If you feel you need to read up on
tensors and indices, please, take a look at the lecture notes of Dulle-
mond and Peeters [8]; also chapter 3 of Griffiths’ book [2] or chapter 7
of Goldstein’s book [9] may be helpful, although the latter keep factors
of c.
Lagrange and Hamilton formalism and related problems, take a look
at the classical textbooks of Goldstein [9] (chapters 1 and 2 for La-
grange formalism, chapter 8 for Hamilton formalism, chapter 3 for
central force problem, and chapter 6 for oscillations) and Landau and
Lifshitz [10] (chapter 1 for Lagrange formalism, chapter 7 for Hamil-
ton formalism, chapter 3 for motion in a potential and chapter 5 for
oscillations).
harmonic oscillator in Quantum Mechanics, creation and annihila-
tion operators, maybe you may want to check Sec. 2.3 in Sakurai’s
book [11]?
3
time ←→ length with c ≈ 0.3 · 109 m/s
momentum ←→ energy with c
mass ←→ energy with c2
time ←→ 1/energy with ~ ≈ 6.5 · 10−22 MeV s
length ←→ 1/energy with ~c ≈ 200 MeV fm
Z
f (x) = dk eikx f˜(k) . (2)
The extension to higher dimensions – for example for the Fourier transfor-
mation of three-vectors – is straightforward:
d3 x −ik·x
Z
˜
f (k) = e f (x)
(2π)3
Z
f (x) = d3 k eik·x f˜(k) . (3)
In addition, we have
Z
dx e−ix(k−q) = (2π)δ(k − q) (5)
4
Einstein Convention When not stated otherwise we will use Einstein’s
convention of summing over repeated indices, for example
Raising and Lowering Indices We have already seen, how the metric
tensor is used to raise or lower indices of four-vectors, e.g.,
For tensors with n indices, one metric tensor is necessary to raise or lower
one index. For example, for a tensor F µν of rank two, two metric tensors are
necessary to lower both indices. As an example, consider the field-strength
tensor of electromagnetism, given by
0 −Ex −Ey −Ez
Ex 0 −Bz By
F µν = . (14)
Ey Bz 0 −Bx
Ez −By Bx 0
5
Therefore,
0 0
Fµν = gµµ0 gν 0 ν F µ ν , (15)
x · p = xµ pµ = xµ pµ = x0 p0 − x · p = x0 p0 − x1 p1 − x2 p2 − x3 p3 (18)
6
∂
∂µ = = (∂/∂t, −∇) . (20)
∂xµ
Note that the derivatives have a relative sign in the spatial coordinates!
p2 = E 2 − p2 = m2 (21)
are linear transformations that connect four-vectors with each other. The
Λµν are usually divided into active transformations where the four-vector in
question is moved while the reference system is fixed, and passive transfor-
mations, where the four-vector is fixed, but the reference system is changed.
The difference between active and passive transformations is encoded in a
relative sign of the defining parameters.
In the context of this lecture, the idea of Lorentz transformations is gener-
alised such that they contain both boosts B µν and rotations Rµν , where the
former are defined by three velocities and the latter defined by three angles.
In fact, the rotation are the Galilei transformations, which are superseded
by the Lorentz transformations.
7
Boosts The (active) boosts B µν are defined by the three-velocity1 v; for
example for a boost long the z-axis with velocity v = vz
1 0 0 −v cosh η 0 0 − sinh η
0 1/γ 0 0 0 1 0 0
B µν (vz ) = γ
= ,
0 0 1/γ 0 0 0 1 0
−v 0 0 1 − sinh η 0 0 cosh η
(26)
where
1
cosh η = γ = √ (27)
1 − v2
is the Lorentz factor and η is the rapidity.
To construct the boost defined by a three-velocity v, B µν (v), it is advan-
tageous to realise that the spatial dimensions can be decomposed into one
component parallel to the boost-vector v, x0k and two perpendicular ones,
~x⊥ . With v = vn and xk = x · n, the transformations read
t0 = γ(t − x · v)
x0k = γ(xk − vt)
~x0⊥ = ~x⊥ , (28)
Rotations Similar to the boosts, the (active) rotations Rµν are defined by
three Euler angles; for example a rotation around the z-axis with angle θ is
mediated by
1 0 0 0
0 cos θ − sin θ 0
Rµν (θ) =
0 sin θ
. (31)
cos θ 0
0 0 0 1
1
Note that we express the velocity in natural units - in many books the velocity is
given as v = cβ with c the speed of light.
8
Invariance of Norm of Four-Vectors The Lorentz transformations
have been constructed such that the norm of a four vector is invariant under
a boost or rotation. To see how this works look at a four-vector x, boosted
with velocity v. The square of its norm is given by
2 2 2 02
||x0 || = x0 = t0 − ~x02 ⊥ − xk
h i
= γ 2 (1 − v 2 )t2 − (1 − v 2 )x2k − ~x2⊥ = x2 (32)
9
This implies that
0
(Λ−1 )µµ0 = (ΛT )µµ0 ) = Λµµ , (36)
Zt1
S(t1 , t0 ) = dtL(qi (t), q̇i (t), t) . (38)
t0
Zt1
δS = dt [L(qi + i , q̇i + ˙i ) − L(qi , q̇i )]
t0
Zt1 Zt1
∂L t1
∂L ∂L ∂L d ∂L
= dt i + ˙i = dt i − i + i , (39)
∂qi ∂ q̇i ∂qi dt ∂ q̇i ∂ q̇i t0
t0 t0
where in the last step the term with ˙i has been partially integrated.
∂L d ∂L
− = 0. (40)
∂qi dt ∂ q̇i
10
Canonical Momentum and Hamilton Function Introducing the canon-
ical momenta
∂L
pi = (41)
∂ q̇i
and expressing the generalised velocities through the canonical momenta pi
allows to construct the Hamilton function as
∂L
H(pi , qi ) = q̇i pi − L(qi , q̇i ) = q̇i − L(qi , q̇i ) = T + V , (42)
∂ q̇i
dpi ∂H
ṗi = =−
dt ∂qi
dqi ∂H
q̇i = =+ . (43)
dt ∂pi
∂f ∂g ∂f ∂g
{f, g} = − . (44)
∂qi ∂pi ∂pi ∂qi
anti-commutativity:
Jacobi identity:
{qi , qj } = {pi , pj } = 0
11
{qi , pj } = δij . (48)
Equations of motion can therefore be expressed as
∂H
ṗi = − = {pi , H}
∂qi
∂H
q̇i = + = {qi , H} . (49)
∂pi
The time evolution of any function f (pi , qi , t) can be evaluated using the
chain rule,
df ∂f ∂f ∂f ∂f
= q̇i + ṗi + = {f, H} + . (50)
dt ∂qi ∂pi ∂t ∂t
This translates into explicitly time-independent f are constant of motion, if
their Poisson bracket with the Hamilton function vanishes2 .
12
√
† 1 i
â = √ ω x̂ − √ p̂ . (53)
2 ω
Direct calculation shows the following commutation relations:
1 2 ω2 2 ω h i
Ĥ = p̂ + x̂ = −(â − ↠)2 + (â + ↠)2
2 2 4
ω † †
ω h †
i
†
† 1
= ââ + â â = â, â + 2â â = ω â â + (55)
2 2 2
N̂ = ↠â . (57)
13
squares of real numbers as eigenvalues, implies that there must be a smallest,
non-negative energy with a corresponding lowest-energy ground state of the
system. Denoting this state as “vacuum”, the only way to guarantee that
there are on lower energy eigenvalues is to demand that the annihilation
operators annihilate this state,
â |0i = 0 , (61)
and so on. Applying the number operator suggests that the vacuum contains
zero quanta, thereby justifying the notation of |0i and similarly that the first
excited state contains one quantum:
N̂ |0i = 0
h i
† † † †
N̂ |1i = â ââ |0i = â â, â + â â |0i = 1 · ↠|0i = 1 |1i .
†
(63)
This suggests that the number operator enjoys the eigenvalue equation
14
2.7 Problems & Solutions
1. Levi-Civita symbol
ijk ilm = δ jl δ km − δ jm δ kl
ijk ijl = 2δ kl
ijk ijk = 6 ,
Solution
15
(b) Levi-Civita in four dimensions:
The identities for the totally antisymmetric tensor in four dimen-
sions follow from using the same logic as before. The realitve
minus sign in front of the expressions is relatively easy to ex-
plain with the signs in the metric tensor, since for the spatial
0 0
components g ρρ = −δ ρρ .
(b) Repeat the exercise for two consecutive rotations around the z-
axis with angles θ1 and θ2 .
Solution
cosh(η1 + η2 ) 0 0 − sinh(η1 + η2 )
0 1 0 0
= 0 0 1 0 = B1 B2 .
− sinh(η1 + η2 ) 0 0 cosh(η1 + η2 )
16
and therefore, consecutively applying boost 2 after boost 1
1 0 0 0
1 0 0 0
0 cos θ2 − sin θ2 0 0 cos θ1 − sin θ1 0
R2 R1 = 0 sin θ2 cos θ2 0 0 sin θ1 cos θ1 0
0 0 0 1 0 0 0 1
1 0 0 0
0 cos(θ1 + θ2 ) − sin(θ1 + θ2 ) 0
= 0 sin(θ1 + θ2 ) cos(θ1 + θ2 ) 0 = R1 R2 .
0 0 0 1
to show that
[Λµν ]−1 = Λν µ
(b) what is the form of inverse Lorentz boosts and rotations along or
around the z-axis?
Solution
(a)
2
x2 = gµν xµ xν = x0 = gρσ x0ρ x0σ = gρσ Λρ α xα Λσβ xβ
=⇒ gµν = gρσ Λρ µ Λσν
=⇒ δµγ = gµν g νγ = gρσ g νγ Λρ µ Λσν
=⇒ δµγ = gρσ Λσγ Λρ µ
=⇒ δµγ = Iγµ = Λρ γ Λρ µ
Λ−1 = gΛT g
(b) Using the metric tensor to raise and lower the two indices, yields
Λν µ = gρν Λρ σ g µσ
− sinh u
1 0 0 0 cosh u 0 0
1 0 0 0
0 −1 0 0 0 1 0 0 0 −1 0 0
= 0 0 −1 0 0 0 1 0 0 0 −1 0
0 0 0 −1 − sinh u 0 0 cosh u 0 0 0 −1
1 0 0 0 cosh u 0 0 sinh u
0 −1 0 0 0 −1 0 0
= 0 0 −1 0 0 0 −1 0
0 0 0 −1 − sinh u 0 0 − cosh u
cosh u 0 0 sinh u
0 1 0 0
= 0 0 1 0
sinh u 0 0 cosh u
17
cosh(−u) 0 0 sinh(−u)
0 1 0 0
= ,
0 0 1 0
sinh(−u) 0 0 cosh(−u)
Λν µ = gρν Λρ σ g µσ
1 0 0 0 1 0 0 0
T 1 0 0 0
0 −1 0 0 0 cos θ sin θ 0 0 −1 0 0
= 0 0 −1 0 0 − sin θ cos θ 0 0 0 −1 0
0 0 0 −1 0 0 0 1 0 0 0 −1
1 0 0 0 1 0 0 0
0 −1 0 0 0 − cos θ sin θ 0
= 0 0 −1 0 0 − sin θ − cos θ 0
0 0 0 −1 0 0 0 −1
1 0 0 0
0 cos θ − sin θ 0
= 0 sin θ cos θ 0 ,
0 0 0 1
1 0 0 0
0 cos(−θ) sin(−θ) 0
=
0 − sin(−θ) cos(−θ) 0 ,
0 0 0 1
again, as expected.
where the M̂µν are the generators of the group. We obtain them
by considering infinitesimal transformations and comparing co-
efficients. Chooisng generators from three infinitesimals boosts
along the x, y, and z-axis and the three infinitesimals boosts
18
around the x, y, and z-axis we arrive at:
0 0 0 0 0 1 0 0
0 0 0 0 Bx = M̂01 = −i 1 0 0 0
Rx = M̂23 = i
0 0 0 −1 0 0 0 0
0 0 1 0 0 0 0 0
0 0 0 0 0 0 1 0
0 0 0 −1 By = M̂02 = −i 0 0 0 0
Ry = M̂13 = i
0 0 0 0 1 0 0 0
1 0 0 0 0 0 0 0
0 0 0 0 0 0 0 1
0 0 −1 0 Bz = M̂03 = −i 0 0 0 0
Rz = M̂12 = i
0 1 0 0 0 0 0 0
0 0 0 0 1 0 0 0
Solution
ων 0 µ0 + ωµ0 ν 0 = 0 −→ ων 0 µ0 = −ωµ0 ν 0 ,
i.e., anti-symmetric ω.
19
(b) Let us consider a number of cases, namely the commutator of two
two boosts, of two rotations, and of a boost and a rotation.
h i
M̂01 , M̂02 = ig00 M̂12 = iM̂12
h i
M̂12 , M̂13 = −ig11 M̂23 = iM̂23
h i
M̂01 , M̂12 = ig11 M̂02 = −iM̂02
20
The non-vanising commutators in contrast are given by
h i 1 1 1
± ±
X̂i , X̂j = ikl M̂kl ± iM̂i0 , jmn M̂mn ± iM̂j0
4 2 2
1 ikl jmn
n h i
= M̂kl, M̂mn
4 4
ijmn h i i h i h i
ikl
± M̂i0 , M̂mn ± M̂kl , M̂j0 − M̂i0 , M̂j0
2 2
i n ikl jmn
= gkn M̂lm + glm M̂kn − gkm M̂ln − gln M̂km
4 4
ijmn
± gin M̂0m + g0m M̂in − gim M̂0n − g0n M̂im
2
iikl
± gk0 M̂lj + glj M̂k0 − gkj M̂l0 − gl0 M̂kj
2 o
+ g00 M̂ij
i 1h
= − (δ lj δ im − δ lm δ ij )M̂lm + (δ in δ jk − δ ij δ kn )M̂kn
4 4
i
− (δ ln δ ij − δ lj δ in )M̂ln − (δ ij δ km − δ im δ kj )M̂km
ih i ih i
∓ ijm M̂0m − jin M̂0n ± ikj M̂k0 − ijl M̂l0 + M̂ij
2 2
i 1h i
= − M̂ji + M̂ji + M̂ji + M̂ji + M̂ij
4 4
ih i
∓ ijm M̂0m − jim M̂0m − imj M̂m0 + ijm M̂m0
2
i
i ih
= −M̂ji + M̂ij ∓ ijm M̂0m (1 + 1 + 1 + 1)
4 2
i
= M̂ij ± iijk M̂k0
2
±,
and direct comparison with the definition of X̂m
± iijm mrs
iijm X̂m = M̂rs ± iM̂m0
2 2
i h ijm rsm i i δir δjs − δis δjr
= M̂rs ± iM̂m0 = M̂rs ± iM̂m0
2 2 2 2
i 1 ih i
= M̂ij − M̂ji ± iM̂m0 = M̂ij ± iM̂m0 =
2 2 2
yields the desired result.
This proves that the six generators of the Lorentz group fac-
torise into two groups of three generators, where each group has
a commutator structure that is identical to the one enjoyed by the
21
generators of the angular momentum group, and where the gener-
ators of the two groups do commute. In other words, the Lorentz
group SO(3, 1) decomposes as SO(3, 1) = SU (2)⊗SU (2), hinting
at a deep connenction between the Lorentz group and spin.
5. ∗ Poincare transformations
The Poincare transformation U (Λ, a) is defined by the combination of
a Lorentz transformation, Λµν , and a shift in space-time, aµ , as
xµ → x0µ = Λµν xν + aµ .
(a) Determine the product, the unit and inverse of the resulting
group.
(b) Verify that
Solution
22
The unit element is obviously given by no Lorentz-transformation,
the unit matrix plus a zero shift, U1 = U (1, 0) and the inverse is
given by
U −1 (Λ, a) = U (Λ−1 , −Λ−1 a) .
To check this explicitly, show that
as expected.
(b)
see the previous problem. Using the transformation law for the
momentum above, and specify it for an infinitesimal Lorentz
transformation we therefore have
i µν i µν
1 + M̂µν ω P̂σ 1 − M̂µν ω = (δ µσ − ω µσ ) P̂µ
2 2
iω µν ω µν
M̂µν P̂σ − P̂σ M̂µν = − gνσ P̂µ − gνσ P̂ν
2 h i 2
M̂µν , P̂σ = gνσ P̂µ − gνσ P̂ν ,
where in going from the first to the second line we ignored terms
quadratic in ω and we explicitly anti-symmetrised the right-hand
side when lifting the Lorentz index of the ω µσ to reflect its prop-
erty.
23
(c) Start with
x0µ = U −1 (Λ, 0)U (Λ̃, 0)U (Λ, 0)xµ = (Λ−1 )µν Λ̃νρ Λρσ xσ
24
(a) write down the Lagrange function;
(b) derive and solve the Euler-Lagrange E.o.M.;
(c) construct the canonical momenta;
(d) find the Hamilton function;
(e) derive the Hamilton E.o.M.;
(f) try to directly infer constants of motion where possible.
Solution
(a)
m 2
(i) : L = ẋ
2
ml2 2
(ii) : L = θ̇ − mglθ2
2
m 2 2 2
(iii) : L = ṙ + r θ̇ − V (r)
2
d ∂L ∂L
0= −
dt ∂ q̇ ∂q
and therefore
(i) : 0 = mẍ
(ii) : 0 = ml2 θ̈ + mglθ
(iii) : 0 = m(r2 θ̈ + 2rṙθ̇)
∂V
0 = mr̈ − mrθ̇2 +
∂r
∂L
p=
∂ q̇
and therefore
(i) : px = mẋ
(ii) : pθ = ml2 θ̇
(iii) : pθ = mr2 θ̇
pr = mṙ
25
(d) Summing over all coordinates and momenta i
X
H= q̇i pi − L
i
(e) Using
∂H ∂H
ṗi = − and q̇i = +
∂qi ∂pi
we have
p
(i) : ṗ = 0 and ẋ =
m
pθ
(ii) : p˙θ = −mglθ and θ̇ =
ml2
p2θ ∂V pr
(iii) : p˙r = − 3
− and ṙ =
mr ∂r 2m
pθ
p˙θ = 0 and θ̇ = .
mr2
Solution
26
∂L ∂L ∂L dL d ∂L
= ∂t q̇ − q̈ + q̈ − = q̇ − L
∂ q̇ ∂ q̇ ∂ q̇ dt dt ∂ q̇
d ∂L
= q̇ − T + V ,
dt ∂ q̇
where Lagrange E.o.M. and L = T − V have been used with T and
V are the kinetic and potential energy. Using Euler’s theorem for
homogeneous functions,
∂L
q̇ = 2T
∂ q̇
if the kinetic energy is a quadratic function of generalised velocities q
(which is usually the case), and if the potential does - as usual - only
depend on generalised coordinates.
This shows that
d(T + V ) dE
= =0
dt dt
Alternatively:
dH ∂H dp ∂H dq ∂H
= + +
dt ∂p dt ∂q dt ∂t
∂H
= q̇ ṗ − ṗq̇ + (using Hamilton’s equations)
∂t
∂H
=
∂t
= 0 if H not explicitly dependent on t.
27
(c) Re-express the Hamilton operator first through the creation and
annihilation operators and then through the number operator.
Evaluate the commutators [N̂ , â] and [N̂ , ↠].
(d) Use the fact that â|0i = 0 to calculate the wave function of the
ground state in position space, i.e.
ψ0 (x) = hx|0i
To do so, you have to express the annihilation operator in position
space and suitably transform hx|â|0i.
(e) Speculate about the spectrum of the fermionic quantum har-
monic oscillator, given by the same Hamiltonian, but where the
creation and annihilation operators ↠and â anti-commute:
{â, ↠} = â↠+ ↠â = 1
{â, â} = {↠, ↠} = 0
Solution
(a) Momentum
∂L
p = = mẋ
∂ ẋ
and Hamilton function
1 2 mω 2 2
H = pẋ − L = p + x .
2m 2
(b) Hamilton operator
1 2 mω 2 2
Ĥ = p̂ + x̂ .
2m 2
With
r
mω i
â = x̂ + p̂
2 mω
r
† mω i
â = x̂ − p̂
2 mω
we find
mω i
[â, â] = + [x̂, p̂] + [p̂, x̂] =0
2 mω
mω i
[â, â] = − [x̂, p̂] + [p̂, x̂] =0
2 mω
h
†
i mω i i
â, â = − [x̂, p̂] − [p̂, x̂] = − 2 [x̂, p̂]
2 mω 2
= 1.
28
(c) Express p̂ and x̂ through â and ↠as
r
1 †
x̂ = + â + â
2mω
r
mω †
p̂ = − i â − â
2
and therefore
" 2 2 #
ω † †
Ĥ = − â − â + â + â
4
ω † † † 1 1
= ââ + â â = ω â â + = ω N̂ + .
2 2 2
Commutators:
[N̂ , â] = ↠ââ − â↠â = ↠ââ − ↠ââ − [â, ↠]â = − â
[N̂ , ↠] = ↠â↠− ↠↠â = ↠ââ − ↠â↠+ [â, ↠]↠= + ↠.
29
(e) Writing, in full analogy, the Hamiltonian as
1 † 1
Ĥ = ω N̂ − = ω â â −
2 2
â|0i = 0 −→ E0 = −ω/2
but because of
1 n † †o
↠↠= â , â = 0
2
the application of creation operator on the first excited state will
annihilate it:
30
3 Classical Fields
In this section we re-derive the Lagrange functions for classical fields. For a
more exhaustive explanation of how to make the transition from a discrete to
a continuous system, chapter 13 of Goldstein [9] may be helpful. There, you
will also find a good derivation of the Euler-Lagrange Equations of Motion
for fields. If you are mainly interested in using the formalism in the context
of the course, you may want to consult Sec. 2.2 of Peskin & Schröder [1].
Coupling the particles with springs with constants k yields the potential
energy
kX
V = (ξi+1 (t) − ξi (t))2 , (68)
2
i
31
Continuum Limit Going from discrete lattice distances to a continuum
can be understood as replacing the index i with a position x, ξi (t) → ξ(x, t),
and by taking the limit a → 0 for the lattice spacing. The ξ differences
become
ξi+1 (t) − ξi (t) ξ(x + a, t) − ξ(x, t) ∂ξ(x, t)
lim = lim = (71)
a→0 a a→0 a ∂x
Summation over i translates into an integral over x,
X
a → dx (72)
i
and the discrete Lagrange function of Eq. (69) turns into the Lagrangian
Z
1 ∂ξ
L= dx µξ˙2 − Y (73)
2 ∂x
for the continuous rod; from now on we suppress the arguments of the ξ.
Going back to the equation of motion, Eq. (69), and taking a closer look at
the second term in the limit of vanishing spacing a
ξi+1 − ξi ξi − ξi−1 ξ(x + a) − ξ(x) ξ(x) − ξ(x − a)
− −→ −
a2 a2 a2 a2
a→0 ∂ξ(x + a)/∂x − ∂ξ(x)/∂x ∂ 2 ξ(x)
−→ lim = ,
a→0 a ∂x2
(74)
it is clear that this is a second derivative, and the E.o.M. for the continuous
elastic rod therefore is given by
∂2ξ ∂2ξ
µ − Y = 0, (75)
∂t2 ∂x2
with longitudinal waves as solution.
which becomes the Lagrange function through integration over the (one-
dimensional) space,
Z
L = dxL(ξ, ∂ξ/∂t, ∂ξ/∂x, x, t) (77)
32
and the action S, as usual, by integrating the Lagrange function over time,
Zt1 Zx1
S(t0 , t1 ) = dt dx L . (78)
t0 x0
In the rest of the lecture course we will assume that Lagrange densities
depend on fields and their derivatives only and do not explicitly depend on
position or time, i.e.
∂ξ ∂ξ
L = L ξ, , . (79)
∂t ∂x
33
and, finally, the equations of motion
∂ ∂L ∂ ∂L ∂L
+ − = 0. (84)
∂t ∂ ∂ξ ∂x ∂ ∂ξ ∂ξ
∂t ∂x
A simple calculation will show that the integral over the space-time volume
is boost and hence Lorentz-invariant. In a similar way, the two derivative
terms in the Lagrange density in Eq. (79) will be amalgamated such that
the Lorentz-invariant Lagrange density is given by
L = L (ξ, ∂µ ξ) . (86)
There is one big caveat, however. This Lagrange density must be a Lorentz-
scalar; pictorially speaking, all indices must be contracted. This implies that
terms of the type ∂µ ξ must come at least in squares, like, e.g. (∂µ ξ)(∂ µ ξ)
such that the two Lorentz-indices are contracted off.
To obtain Euler-Lagrange equations of motion from the action
µ
Zx1
S= d4 x L (ξ, ∂µ ξ) , (87)
xµ
0
steps similar to the one before will be necessary. In particular, we will now
demand that the virtual variations of the field vanish on the surface of the
d4 x-integration, leading to
µ
Zx1
dS 4 ∂L ∂ξ ∂L ∂(∂µ ξ)
0≡ = d x +
dα ∂ξ ∂α ∂(∂µ ξ) ∂α
xµ
0
34
µ
Zx1
4 ∂ξ ∂L ∂L ∂L ∂ξ
= d x − ∂µ + ∂µ ∂µ . (88)
∂α ∂ξ ∂(∂µ ξ) ∂(∂µ ξ) ∂α
xµ
0
∂L ∂L
∂µ − = 0. (89)
∂(∂µ ξ) ∂ξ
∂2
− ∇2 + m2 φ(x) = ∂µ ∂ µ + m2 φ(x) = 0 .
(91)
∂t2
35
A few comments are in order here:
1. In Eq. (142) we have directly used the continuum limit. This neces-
sitates the integration over all momenta instead of a summation over
a discrete set of eigenvalues for the momentum. The latter would be
the case for example when second quantising on a lattice with lattice
spacing a, where the eigenvalues for the momentum are discrete and
behave like kn = n/a.
2. The measure of integration, that sums over the different wave vector,
should better be Lorentz-invariant. It is not trivial to see immedi-
ately that d3 k/(2k0 ) fulfils this criterion. To realise that this is indeed
the case, let us start with a manifestly Lorentz-invariant integration
measure,
d4 k d3 k
Z Z
2 2
dk0 δ(k02 − k 2 − m2 )Θ(k0 )
4
(2π)δ(k − m )Θ(k 0 ) = 3
(2π) (2π)
d3 k
Z
X 1
= Θ(k0 )
(2π)3 √ 2k0
k0 =± k2 +m2
d3 k
Z
= , (94)
(2π)3 (2k0 )
which replaces the integral over the δ-function of a function f (x) with
an integral over a sum of its zeroes xi (given by f (xi ) = 0), normalised
by the first derivative of the function at the zero.
1 m2 2
L(∂µ φ, φ) = (∂µ φ)(∂ µ φ) − φ . (96)
2 2
Note that, wherever the dependence is self-evident, we will ignore the ar-
guments of the fields from now on. To see this, let us plug this Lagrange
36
density into Eq. (89), with the obvious replacement ξ → φ.
∂L ∂L
0 = ∂µ −
∂(∂µ φ) ∂φ
2
∂[ 12 (∂µ φ)(∂ µ φ)] ∂[ m2 φ2 ]
= ∂µ − , (97)
∂(∂µ φ) ∂φ
where we have replaced the Lagrange density in the first line with the rel-
evant parts of Eq. (96) in the second one. The first expression looks a bit
tricky and, naively, it seems as if derivation w.r.t. ∂µ φ would only deliver
1 µ
2 ∂ φ – this however is wrong, and it is easy to see why. Rewriting this part
component by component we would arrive at terms like
1 ∂ ∂ φ̇2 1 ∂ φ̇ ∂2φ
=2· = 2 (98)
2 ∂t ∂ φ̇ 2 ∂t ∂t
and similar for the spatial components. Another way to see this is to rewrite
the Lorentz-scalar of the derivatives with other indices – replacing the µ’s
with ν’s in the Lagrangian (it doesn’t matter, they get contracted anyway,
so I can sum over µ’s, ν’s or any other symbol I chose as Lorentz index)
νρ
∂ 1 ν ∂ g
(∂ φ)(∂ν φ) = (∂ν φ)(∂ρ φ)
∂(∂µ φ) 2 ∂(∂µ φ) 2
g νρ
∂(∂ρ φ) ∂(∂ν φ)
= (∂ν φ) + (∂ρ φ)
2 ∂(∂µ φ) ∂(∂µ φ)
νρ
g
(∂ν φ)δρµ + (∂ρ φ)δνµ = ∂ µ φ
= (99)
2
Taking into account of this insight, we ultimately arrive at
0 = ∂µ ∂ µ φ + m2 φ , (100)
as requested.
If both masses are equal, m1 = m2 , the two real fields can be re-arranged
into one complex one,
φ1 + iφ2 φ1 − iφ2
φ= √ and φ∗ = √ , (102)
2 2
37
or
φ + φ∗ −i(φ − φ∗ )
φ1 = √ and φ2 = √ . (103)
2 2
The Lagrange density for the free complex scalar field then becomes
It is important to stress here that while the fields φ and φ∗ are connected
through complex conjugation, they still encode two independent degrees
of freedom and therefore must be treated as independent quantities when
analysing the structure of the Lagrange density, or deriving E.o.M..
Equations of Motion The E.o.M. are obtained in the now familiar fash-
ion as
∂L ∂L
0 = ∂µ − = ∂µ ∂ µ φ∗ + m2 φ∗
∂(∂µ φ) ∂φ
∂L ∂L
0 = ∂µ − = ∂µ ∂ µ φ + m2 φ . (105)
∂(∂µ φ∗ ) ∂φ∗
Note that, as we have two independent degrees of freedom (the two fields),
we have two E.o.M., obtained by differentiating the Lagrangian with respect
to each of the two fields.
Clearly, the Lagrangian and therefore the action are invariant under this set
of transformations.
0 ≡ δS
Z
4 ∂L ∂L ∂L ∗ ∂L ∗
= d x δ(∂µ φ) + δφ + δ(∂µ φ ) + δφ (108)
∂(∂µ φ) ∂φ ∂(∂µ φ∗ ) ∂φ∗
38
Realising that, for example,
δφ = φ0 − φ = (eiθ − 1)φ =⇒ ∂µ (δφ) = δ(∂µ φ) (109)
and using the by now familiar trick of integrating by parts, we arrive at
Z (
∂L ∂L ∂L ∂L
4
δS = d x iθφ − ∂µ − iθφ∗ − ∂ µ
∂φ ∂(∂µ φ) ∂φ∗ ∂(∂µ φ∗ )
)
∂L ∂L ∗
+ iθ∂µ φ− φ . (110)
∂(∂µ φ) ∂(∂µ φ∗ )
The first line of the result above equals 0, by virtue of the E.o.M. for both
φ and φ∗ , and in order for the second line to integrate to 0 we must have
∂L ∗ ∂L ∗ µ µ ∗
0 ≡ ∂µ φ − φ = ∂µ φ (∂ φ) − (∂ φ )φ . (111)
∂(∂µ φ∗ ) ∂(∂µ φ)
This implies the existence of a conserved current, i.e.
∂µ j µ = 0 (112)
←
→
µ ∗ µ µ ∗
j = φ (∂ φ) − (∂ φ )φ ≡ φ∗ ∂ µ φ . (113)
Conserved Charge The current from Eq. (113) implies the existence
of a conserved charge Q with dQ/dt = 0, constructed by integrating the
temporal component over three-dimensional space,
Z
Q = d3 x j 0 . (115)
39
3.4 Vector Fields: Maxwell’s Equations
A Little Game of Symmetry Assume you want to introduce two differ-
ent three-vector fields. From a (classical) symmetry point of view, they can
be distinguished through parity, i.e. one of them is parity-odd – a “proper”
vector – while the other one is parity-even – an axial-vector. We call the
parity odd fields (or 1− in spin-parity notation) E, and the parity even ones
(or 1+ ) B. Now let us assume that you only want to allow first derivatives
of the fields, ∂t and ∇ and scalar and pseudo-scalar charge densities ρE,B
and corresponding currents j E,B . Then you can sort resulting quantities by
spin and parity as in Table 3.
scalars 0+ ∇ · E, ρE
pseudo-scalars 0− ∇ · B, ρB
vectors 1− ∂t E, ∇ × B, j E
axial-vectors 1− ∇ × E, ∂t B, j B
∇ · E = 4πρE ∇·B = 0
(118)
∇ × B − ∂t E = 4πj E ∇ × E + ∂t B = 0
Note that we absorbed the usual factors of 0 and µ0 into the definition of
the charge and current, and we have used natural units with c = 1.
40
The Vector Potential The left column in Eq.(118) suggest to use a scalar
potential Φ, which we denote as A0 , and a vector potential A and write
Of course this now forms a four-vector potential Aµ = (A0 , A), and we will
continue the analysis of electrodynamics mainly based on this object.
Aµ → A0µ = Aµ − ∂ µ Λ , (120)
and therefore
E → E 0 = − ∇(A0 − ∂ 0 Λ) − ∂t (A + ∇ · Λ) = −∇A0 − ∂t A = E
B → B 0 = ∇ × (A + ∇ · Λ) = ∇ × A = B , (121)
where we have used that rot·grad of a scalar function vanishes. This suggest
that it would be beneficial to express the theory in terms of gauge invariant
quantities made from Aµ , to directly encode this symmetry.
41
Lagrange Density in Terms of the Fields There are various ways
to express the Lagrange density; a version probably familiar from previous
lectures expresses it through the electric and magnetic fields and reads
E2 − B2
L= − ρφ + j · A . (125)
8π
The E.o.M. are obtained in terms of the potential φ and A, using the fact
that the electromagnetic fields are expressed through their derivatives. This
also fixes the two homogeneous Maxwell equations, i.e. the right column of
Esq. (118). This also implies that we are left with the task to check if the
Lagrange density above yields the correct inhomogenous equations – the left
column of Eq. (118).
For example, for φ we have:
∂L
= −ρ
∂φ
∂L Ek ∂Ek Ek
= = − , (126)
∂(∂φ/∂xk ) 4π ∂(∂φ/∂xk ) 4π
where
∂Ek
= −1 (127)
∂(∂φ/∂xk )
follows directly from Eq. (121). Assembling all parts, and making the sum-
mation over repeated indices explicit therefore yields Gauss’ law,
X ∂ ∂L
∂L ∇·E
− = − + ρ = 0. (128)
∂xk ∂(∂φ/∂xk ) ∂φ 4π
k
where Eq. (121) has again been used, noting that, expressed in component
notation
B = ∇ × A ←→ Bk = ijk ∂i Aj . (130)
42
specialising i = 1 we are left with Ampere’s law,
1 ∂B3 ∂B2 1 ∂E1
− − − j1 = 0 , (131)
4π ∂x2 ∂x3 4π ∂t
or, in vector form,
∂E
∇×B− = 4πj . (132)
∂t
1
L = − F µν Fµν − 4πj µ Aµ . (133)
4
For the “source” term j µ Aµ , which couples the potentials to charge and cur-
rent densities, we have assumed the so-called “minimal coupling”, typically
of the form source · fields, in a Lorentz-invariant way. This form also fixes
the gauge transformation of the four-vector current j µ . The (E 2 − B 2 )-term
is replaced by a product of field-strength tensors, by realising that
F µν Fµν = −F νµ Fµν
0 −Ex −Ey −Ez 0 −Ex −Ey −Ez
Ex 0 −Bz By Ex 0 Bz −By
= Tr
Ey Bz 0 −Bx Ey −Bz 0 Bx
Ez −By Bx 0 Ez By −Bx 0
2
−E • • •
• −Ex + Bz2 + By2
2 • •
= Tr
• 2 2 2
• −Ey + Bz + Bx •
• • • 2 2 2
−Ez + By + Bx
= − 2(E 2 − B 2 ) .
(134)
∂L(φi , ∂µ φi )
πi (x) = (135)
∂(∂t φi )
43
and a Hamilton density is constructed as
X
H= πi φ̇i − L . (136)
i
44
3.6 Problems & Solutions
1. General Solutions for the Klein-Gordon Equation
Consider a real scalar field, given by thr Klein-Gordon Lagrangian,
Eq. (96).
(a) Proof that the solutions to its Equation of Motion, Eq. (91), are
given by the expression in Eq. (93).
(b) Calculate the Hamiltonian and momentum for a free scalar field
using their definitions,
Z
1
d3 x (∂t φ)2 + (∇φ)2 + m2 φ2
H =
2
Z
P = − d3 x [(∂t φ)(∇φ)] .
Solution
(a) Inserting the solution for the Klein-Gordon equation from Eq. (93)
into the E.o.M. yields
d3 k
Z
2 −ik·x ∗ ik·x
+m a(k)e + a (k)e
(2π)3 2k0
d3 k
Z
2 −ik·x ∗ 2 ik·x
= a(k)( + m )e + a (k)( + m )e
(2π)3 2k0
d3 k
Z
2 2 −ik·x ∗ ik·x
= (−k + m ) a(k)e + a (k)e
(2π)3 2k0
d3 k
Z
2 2 2 −ik·x ∗ ik·x
= (−k0 + k + m ) a(k)e + a (k)e =0
(2π)3 2k0
+ a(k)a∗ (q)e−i(k−q)·x k0 q0 − k · q + m2
+ a∗ (k)a(q)e+i(k−q)·x k0 q0 − k · q + m2
45
∗ ∗ +i(k+q)·x 2
+ a (k)a (q)e −k0 q0 − k · q + m
d3 k d3 q
Z
1
=
2 (2π)3 2k0 (2π)3 2q0
a(k)a(q)(2π)3 δ 3 (k + q)e−i(k0 +q0 )x0 −k0 q0 − k · q + m2
∗ ∗ 3 3 +i(k0 +q0 )x0 2
+ a (k)a (q)(2π) δ (k + q)e −k0 q0 − k · q + m
d3 k
Z
1
=
2 (2π)3 (2k0 )2
a(k)a(−k)e−i(k0 +k0 )x0 −k02 + k 2 + m2
∗ ∗ +i(k0 +k0 )x0
−k02 2 2
+ a (k)a (−k)e +k +m
d3 k
Z
1 2 ∗ ∗
= 2k a(k)a (k) + a(k)a (k)
2 (2π)3 (2k0 )2 0
d3 k
Z
= k0 a(k)a∗ (k)
(2π)3 (2k0 )
and
Z
P =− d3 x [(∂t φ)(∇φ)]
d3 k d3 q
Z
3
= − d x
(2π)3 2k0 (2π)3 2q0
a(k)a(q)e−i(k+q)·x (k0 q) + a(k)a∗ (q)e−i(k−q)·x (−k0 q)
+ a∗ (k)a(q)e+i(k−q)·x (−k0 q) + a∗ (k)a∗ (q)e+i(k+q)·x (k0 q)
d3 k d3 q
Z
= − k0 q
(2π)3 2k0 (2π)3 2q0
a(k)a(q)(2π)3 δ 3 (k + q)e−i(k0 +q0 )x0
46
− a(k)a∗ (q)(2π)3 δ 3 (k − q)e−i(k0 −q0 )x0
d3 k
Z
= k0 k
(2π)3 (2k0 )2
a(k)a(−k)e−i(k0 +k0 )x0 + a(k)a∗ (k)e−i(k0 −k0 )x0
∗ +i(k0 −k0 )x0 ∗ ∗ +i(k0 +k0 )x0
+ a (k)a(k)e + a (k)a (−k)e
d3 k
Z
= k0 k a(k)a∗ (k) + a∗ (k)a(k)
(2π)3 (2k0 )2
d3 k
Z
= k a(k)a∗ (k)
(2π)3 2k0
47
Solution
∂t φ + i/m∂t2 φ
i∂χ i
= = Hχ
∂t 2 ∂t φ − i/m∂t2 φ
because then
∇2
φ m φ + i/m∂t φ
Hχ = − +
2m −φ 2 −φ + i/m∂t φ
∇2
m i i 1 2
−
2m φ + 2 φ + ∂t φ ∂ φ− ∂t φ
2 2 t 2m
= =
∇2 m i i 1 2
+ φ− φ+ ∂t φ ∂t φ + ∂t φ
2m 2 2 2 2m
This gives rise to two two identical equations, namely the Klein-
Gordon E.o.M.,
∂t2 ψ − ∇2 φ + m2 φ = 0 .
∇2
1 1 1 0
H = − +m
2m −1 −1 0 −1
2
p2
2 2
∇ ∇ p
− 2m +m − 2m + m
= = 2m 2m
∇2 ∇2 p2 p2
2m 2m − m − 2m − 2m − m
48
and therefore
∇2 ∇2
(m + T )χ+ = − − m χ+ − χ−
2m2 2m
∇2
∇
(m + T )χ− = − m χ− + χ+ .
2m 2m
∇2 ∇2
χ− = χ+ · ≈ χ + ·
2m(2m + T ) − ∇2 4m2
∇2 ∇2
T χ+ = − 1+ χ+
2m 4m2
1 m2 2 λ 4
L = (∂µ φ)(∂ µ φ) − φ − φ ;
2 2 4!
m2 λ
L = −(∂µ Aν )(∂ν Aµ ) + Aµ Aµ + (∂µ Aµ )2 ;
2 2
1 m2
L = − Fµν F µν + Aµ Aµ ;
4 2
49
Solution
∂Fµν ∂(∂µ Aν − ∂ν Aµ )
= = gµρ gνσ − gνρ gµσ
∂(∂ρ Aσ ) ∂(∂ρ Aσ )
and therefore
1 ∂Fµν F µν 1 ∂Fµν 1
= − F µν = − F µν gµρ gνσ − gνρ gµσ
−
4 ∂(∂ρ Aσ ) 2 ∂(∂ρ Aσ ) 2
1
= − (F ρσ − F σρ ) = −F ρσ = F σρ
2
and
1 ∂Fµν F µν
− ∂ρ = −∂ρ F ρσ = Aσ − ∂ σ (∂ · A) .
4 ∂(∂ρ Aσ )
1 ∂(m2 A2 )
= m 2 Aρ
2 ∂Aρ
λ ∂(∂ · A)2
∂ρ = λgµρ ∂ρ g σµ (∂ · A) = λ∂ σ (∂ · A)
2 ∂(∂ρ Aσ )
∂[−(∂µ Aν )(∂ν Aµ )] µ ν
ν ∂(∂ν A ) µ ∂(∂µ A )
∂ρ = −∂ρ (∂µ A ) + (∂ν A )
∂(∂ρ Aσ ) ∂(∂ρ Aσ ) ∂(∂ρ Aσ )
= −2∂ρ ∂ σ Aρ = −2∂ σ (∂ · A)
(λ − 2)∂ σ (∂ · A) − m2 Aσ = 0
(c) Previous results mean that we only have to put terms together
and arrive at
∂σ F σρ + m2 Aρ = [gσρ ( + m2 ) − ∂σ ∂ ρ ]Aσ = 0 .
50
(d) Here we have three active fields and arrive at:
1 m2
L = − Vµν V µν + Vµ V µ ,
4 2
where the field strength tensor Vµν assumes the usual form
Vµν = ∂µ Vν − ∂ν Vµ .
∂µ V µ = 0
Solution
∂σ (∂ρ V ρσ + mV σ ) = m∂ · V = 0 .
51
(c) Assuming the momentum being oriented along the z-axis, we have
p
k µ = (ω, 0, 0, κ) with ω = κ2 + m2
0 0 −ω
F µν = ∂ µ Aν − ∂ ν Aµ
is the field strength tensor for the vector potential Aµ . The additional
term 1/2(∂ · A)2 is also known as “gaug-fixing” term, and in this case
corresponds to the Lorentz gauge. We will come back to this in Section
6 of the notes.
Show that the Euler-Lagange E.o.M. lead directly to a wave equation
of the form
Aµ = ∂ν ∂ ν Aµ = 4πj µ .
Solution
52
where in the last step we have used that the first and last term are
identical when swapping λ and κ in the last term – which is allowed, as
both are just repeated indices summed over from 0 to 3, and similarly
for the second and third term.
Using the fact that
∂(∂ ρ Aσ )
= g ρξ g σχ
∂(∂ ξ Aχ )
we can write
∂ 1 κλ 1 ∂ h
κ λ κ λ
i
− F Fκλ = − ∂ A ∂ κ A λ − ∂ A ∂ λ Aκ
∂(∂ ν Aµ ) 4 2 ∂(∂ ν Aµ )
∂(∂ κ Aλ )
1 ∂(∂κ Aλ )
= − ∂ κ Aλ + ∂ κ Aλ
2 ∂(∂ ν Aµ ) ∂(∂ ν Aµ )
∂(∂ λ Aκ )
κ λ ∂(∂λ Aκ )
−∂κ Aλ − ∂ A
∂(∂ ν Aµ ) ∂(∂ ν Aµ )
1h i
= − ∂κ Aλ g κν g λµ + ∂ κ Aλ gκν gλµ − ∂κ Aλ g λν g κµ − ∂ κ Aλ gλν gκµ
2
= ∂µ Aν − ∂ν Aµ
In addition
∂ 1 κ 2
− (∂κ A ) = −∂κ Aκ gλν g λµ = −∂κ Aκ gµν .
∂(∂ ν Aµ ) 2
Therefore
∂L
∂ν = ∂ ν ∂µ Aν − Aµ − ∂µ (∂ · A) = −Aν .
∂(∂ ν Aµ )
Combining this with
∂L
= −4πjµ
∂Aµ
yields the desired result.
6. Free Schrödinger field The Lagrangian of the free Schrödinger Field
is given by
i ∗ 1
L = (φ ∂t φ − φ∂t φ∗ ) − (∇φ∗ ) · (∇φ)
2 2
(a) write down the equations of motion for the Schrödinger field.
(b) show that the conserved current is given by
i ∗
j 0 = φ∗ φ , j= (φ ∇φ − φ∇φ∗ ) ,
2
i.e. show that ∂µ j µ = 0
53
Solution
(a)
∂L ∂L ∂L
0 = ∂t − ∇ − = −i∂t φ∗ + 1 2 ∗
2∇ φ
∂(∂t φ) ∂(∇φ) ∂φ
∂L ∂L ∂L 1 2
0 = ∂t − ∇ − = +i∂t φ + 2∇ φ .
∂(∂t φ∗ ) ∂(∇φ∗ ) ∂φ∗
(b)
i
∂µ j µ = ∂t (φ∗ φ) − ∇(φ∗ ∇φ − φ∇φ∗ )
2
= φ∗ (∂t φ) + φ(∂t φ∗ )
i ∗ ∗ 2 ∗ 2 ∗
− (∇φ )(∇φ) + φ ∇ φ − (∇φ)(∇φ ) − φ∇ φ
2
∗ i 2 i 2
=φ ∂t − ∇ φ + φ ∂t + ∇ φ∗ ,
2 2
and both terms vanish under the E.o.M. above.
+∞ Z+L
m2 2
Z
1 µ
S= dt dx (∂µ φ)(∂ φ) − φ .
−∞ 2 2
−L
Find the Equation of Motion for the field φ, and discuss the importance
of the boundary terms.
Solution
54
Z +∞ Z+L
dx (∂t φ)∂t (δφ) − (∂x φ)∂x (δφ) − m2 φδφ
= dt
−∞
−L
Z+L +∞
t=+∞ Z
dt (∂t2 φ)δφ
= dx (∂t φ)δφ −
t=−∞
−L −∞
+∞ x=+L Z+L
Z
dx (∂x2 φ)δφ
− dt (∂x φ)δφ −
x=−L
−∞ −L
+∞
Z Z+L
− dt dx m2 φδφ
−∞ −L
Z+L t=+∞ Z+∞ Z+L
dx ( + m2 )φδφ ,
= dx (∂t φ)δφ − dt
t=−∞
−L −∞ −L
where we have assumed, as usual, that the variations of the field vanish
for t → ±∞, δφ(t → ±∞, x) = 0.
This leaves us with the equation of motion
( + m2 )φ = 0 .
This however holds true only either if the (Dirichlet) boundary condi-
tions
δφ(t, , x = ±L) = 0
or if the (Neumann) boundary conditions
∂x φ(t, , x = ±L) = 0
are fulfilled. The latter are better suited for the solution of our prob-
lem, since they are formulated as conditions on the field φ or its deriva-
tive and not on its – in principle arbitrary – variation δφ.
8. Symmetry and Conserved Current
Consider the Lagrangian density for two real scalars φ1, 2 given by
2
m2 2
1 µ µ 2 λ 2 2
L= (∂µ φ1 )(∂ φ1 ) + (∂µ φ2 )(∂ φ2 ) − φ1 + φ2 − φ + φ2 .
2 2 4! 1
Show that it invariant under the transformation
55
Solution
9. ∗ A SU (2) Symmetry
Consider a doublet of conplex scalars
Φ= φ1 , φ 2
with dynamics defined by the free Lagrangian
L = (∂µ Φ)† (∂ µ Φ) − m2 Φ† Φ .
(a) Show that this Lagrangian is invariant under the three-parameter
transformations
† i
Φ → Φ = exp θa σa Φ ,
2
with the three constant real angles θ1, 2, 3 and where the σa are
the three Pauli matrices,
0 1 0 −i 1 0
σ1 = , σ2 = , σ3 =
1 0 i 0 −1 0
which enjoy the commutation relation
[σi , σj ] = iijk σk
56
Solution
(a) Because the angles are constant, the derivatives do not act on
them, and we only have to evaluate terms of the form
iθa † iθb
exp − σa exp + σb = exp X Y = exp Z ,
2 2
where Z is given by the Baker–Hausdorff formula as
1 1 1
Z = X + Y + [X, Y ] + [X, [X, Y ]] − [X, [X, Y ]] + . . . .
2 12 12
Making the summation over a and b explicit and using that σa† =
σa , the commutator is given by
3 3
X θa θ b X iabc θa θb σc
[X, Y ] = [σa , σb ] = = 0,
4 4
a,b=1 a,b=1
∂L
T µν = ∂ ν φα − g µν L .
∂(∂µ φα )
∂µ T µν = 0 .
(d) Show that for the free real scalar field, the energy-momentum
tensor is symmetric, T µν = T νµ .
(e) Verify that the interpretations of T 00 and T 0i from questions (a)
and (b) are correct for the free electromagentic field.
57
Solution
(a)
∂L
T 00 = φ̇α − L = πα φ̇α − L = H ,
∂ φ̇α
the Hamiltonian or energy density of the system.
(b)
∂L j
T 0j = ∂ φα − g 0j L = πα ∂ j φα ,
∂ φ̇α
the components of the three-momentum density of the system.
(c) Direct calculation yields
µν ∂L ν µν
∂µ T = ∂µ ∂ φα − g L
∂(∂µ φα )
∂L ∂L
= ∂µ · ∂ ν φα + · ∂µ ∂ ν φα − ∂ ν L
∂(∂µ φα ) ∂(∂µ φα )
∂L ∂L
= ∂µ · ∂ ν φα + · ∂µ ∂ ν φα
∂(∂µ φα ) ∂(∂µ φα )
∂L ∂L
− · ∂ ν φα − · ∂ ν ∂µ φα
∂φα ∂(∂µ φα )
∂L ∂L
= ∂µ − ∂ ν φα ≡ 0 ,
∂(∂µ φα ) ∂φα
as demanded.
(d) Direct calculation:
∂L
T µν = ∂ ν φ − g µν L
∂(∂µ φ)
g µν
= − (∂ µ φ)(∂ ν φ) (∂µ φ)(∂ µ φ) − m2 φ2 = T νµ ,
2
as expected. However, there may be more complicated Lagrangians,
where the energy-momentum tensor is not symmetric.
(e) The free field Lagrangian is given by
1 1
L = − Fµν F µν = − (∂µ Aν − ∂ν Aµ )(∂ µ Aν − ∂ ν Aµ ) .
4 4
Energy density: obviously there are no term Ȧ0 due to the sym-
metry of the F µν , this means in the summing over field compo-
nents we can concentrate on the Ȧi
∂L E2 − B2 E2 + B2
T 00 = Ȧi − L = E 2 − = ,
∂ Ȧi 2 2
58
the energy density of the field, as expected. This, of course, was
clear from the beginning, since we already explicitly calculated
T 00 = H in (a).
Momentum density: the momenta conjugate to Aµ , πµ are given
by
∂L π0 = 0
πµ = =
∂ Ȧµ πi = Ei
T 0j = Ei ∂ j Ai = (E × B)j ,
The last term in our expression for T 01 , ∇·(A1 E), is a total derivative.
This means that, when we integrate over all space, this term will vanish
and we are therefore free to ignore it.
59
4 Second Quantisation
At the beginning of this part of the course you may have read the title and
asked yourself: What does second quantisation actually mean? Haven’t we
already quantised the theory? The answer is that in “first quantisation”
we quantised the position and momentum of point particles. This led to
important properties related to our ability to measure them – the uncer-
tainty principle – and to a crucial reassessment of the inner working of the
world around us, replacing Laplace’s demon of deterministic physics with
a determinism of probabilities. So while in this first quantisation we re-
placed the real numbers x and p with operators x̂ and p̂ and constructed
wave functions for the emerging states, in “second quantisation” we quantise
something else, namely the fields. Consequently, x and p become “ordinary”
numbers again, which serve as arguments of the field operators φ̂ and π̂.
This step necessitates the introduction of a new state. While, formally
speaking, the states of Quantum Mechanics constitute a Hilbert space, the
field operators act on objects in a more complicated Fock space, which is not
labelled by eigen-positions or momenta, but by the number of field quanta
of a given momentum. We will, however, not discuss the properties of these
vector spaces in the lecture.
Simply put: while first quantisation quantised the point dynamics of Clas-
sical Mechanics, leading to Quantum Mechanics, second quantisation pro-
duces a Quantum Field Theory.
If you like to do some additional reading, I would recommend to take a closer
look at Chapter 3 of Hatfield’s book [3], in particular sections 3.1-3.4 or at
Sections 2.3-2.4 in Peskin & Schroeder [1].
60
How-to: Second Quantisation
π = ∂L/∂(∂t φ) = ∂L/∂ φ̇
1 m2 2
L(∂µ φ, φ) = (∂µ φ)(∂ µ φ) − φ ,
2 2
the conjugate momentum reads
∂L
π= = φ̇ . (139)
∂ φ̇
61
Hamiltonian The Hamilton function therefore is given by
Z Z
3 3 1 2 2 2 2
H = d x π φ̇ − L = d x π + (∇φ) + m φ . (140)
2
Field Operators and Commutators Promoting the field and its con-
jugate momentum to operators, φ(x) → φ̂(x) and π(x) → π̂(x), we demand
the equal-time commutators,
h i
φ̂(t, x), π̂(t, y) = iδ 3 (x − y)
h i
φ̂(t, x), φ̂(t, y) = π̂(t, x), π̂(t, y) = 0 (141)
d3 k
Z h i
φ̂(x) = â(k) e−ik·x + ↠(k) eik·x
(2π)3 2k0
d3 k
Z h i
π̂(x) = −ik0 â(k) e−ik·x + ik0 ↠(k) eik·x , (142)
(2π)3 2k0
62
a way that the annihilation operator â drops out. Multiplying, inside the
integral, the expression for φ̂ with k0 and π̂ with i and adding the expression
for π̂ we arrive at
d3 k 1 h
h i Z i
† ik·x
“ k0 φ̂(x) − iπ̂(x) ” = 2ik 0 â (k)e
(2π)3 2k0
d3 k †
Z
= â (k)e−ik·x eik0 x0 , (143)
(2π)3
which looks suspiciously like the Fourier transform of ↠times a factor.
Therefore, Fourier-back-transforming yields
d3 q †
Z h i Z Z
d3 xeik·x k0 φ̂(x) − iπ̂(x) = d3 xeik·x â (q)e−iq·x eik0 x0
(2π)3
d3 q 3
Z
= δ (k − q)↠(q)eik0 x0 = ↠(k)eik0 x0 .
(2π)3
(144)
63
relative sign in front of both momentum operators and in both exponential
factors,
h i
↠(k), ↠(q) = 0 . (147)
0
+â(k)â (k ) +k0 k0 kk + m e−i(k−k )·x
† 0 0 0 2
64
† 0 0 0 0
+â (k)â(k ) +k0 k0 kk + m e+i(k−k )·x
2
† † 0 0 0 2 +i(k+k0 )·x
+â (k)â (k ) −k0 k0 − kk + m e
d3 k d3 k 0
Z
1
=
2 (2π)3 2k0 (2π)3 2k00
3 3 0 −i(k0 +k00 )x0 0 0 0 2
(2π) δ (+k + k ) e â(k)â(k ) −k0 k0 − kk + m
3 3 0 −i(k0 −k00 )x0 † 0 0 0 2
+ (2π) δ (+k − k ) e â(k)â (k ) +k0 k0 + kk + m
3 3 0 +i(k0 −k00 )x0 † 0 0 0 2
+ (2π) δ (−k + k ) e â (k)â(k ) +k0 k0 + kk + m
3 3 0 +i(k0 +k00 )x0 † † 0 0 0 2
+ (2π) δ (−k − k ) e â (k)â (k ) −k0 k0 − kk + m
d3 k
Z
1 −2ik0 x0 2 2 2
= â(k)â(−k)e −k0 + k + m
2 (2π)3 4k02
† 2 2 2 † 2 2 2
+â(k)â (k) +k0 + k + m + â (k)â(k) +k0 + k + m
† † 2ik0 x0 2 2 2
+â (k)â (−k)e −k0 + k + m
d3 k
Z
1 2 † † ˆ
= 2k0 â(k)â (k) + â (k)(k)
2 (2π)3 4k02
d3 k
Z
1 † † ˆ
= k0 â(k)â (k) + â (k)(k) , (150)
2 (2π)3 2k0
where the δ functions in the first step emerge from the integral over x and
where we have eliminated the terms ââ and ↠↠by realising that due to
the relativistic energy-momentum relation k02 = k 2 + m2 . Therefore, the
Hamilton operator for the real scalar field is given by
d3 k
Z
1 † † ˆ
Ĥ = k0 â(k)â (k) + â (k)(k) (151)
2 (2π)3 2k0
It suggests that the Quantum Field Theory for a real scalar field can be in-
terpreted as a continuous sum of harmonic oscillators, permeating all space.
Simple States: Ground State and First Excited State Following the
same logic already present in the harmonic oscillator in Quantum Mechanics
we introduce a ground state – the “vacuum” – |0i which is annihilated by
any annihilation operator,
65
States containing fields (or particles) with momenta k i are generated by
repeated application of the corresponding creation operators
↠(k 1 ) |0i = |k 1 i
↠(k 1 )↠(k 2 ) |0i = |k 1 k 2 i . . . . (153)
suggests that the normalisation of the state equals the (infinite) spatial vol-
ume – a veritable divergence. This is actually not a surprising finding, after
all, the uncertainty principle tells you that a particle with completely fixed
momentum has no localisation. Our particle here, with its fixed momentum
represents a plane wave, filling all volume. If the volume is infinite – which it
is for us to have a continuous spectrum – such states have no normalisation.
The solution to this conundrum is to “smear” the state with a modulating
function f (k), and to define
|ki −→ |kif = f (k)↠(k) |0i (156)
which will lead to perfectly normalisable states, if
Z
d3 k |f (k)|2 < ∞ . (157)
66
Ground-State Energy The ground state |0i is an eigenstate of the Hamil-
tonian; calculating its energy E0 we arrive at
d3 k
Z i
1 h
† †
E0 = h0| Ĥ |0i = 0 k0 â(k)â (k) + â (k)â(k) 0
2 (2π)3 2k0
d3 k
Z
1 D
†
E
= k 0 â(k)â (k)0
0
2 (2π)3 2k0
Z
1
= d3 kk0 δ 3 (0) = ∞ , (158)
2
the product of the volume in both position and momentum space, infinity
for a Quantum Field Theory in an infinite volume. A simple solution is to
just subtract the ground state energy, by redefining the Hamiltonian as
d3 k
Z
1 h
† †
i
Ĥ −→ :Ĥ: = k0 : â(k)â (k) + â (k)â(k) : (161)
2 (2π)3 2k0
and, finally,
d3 k
Z
:Ĥ: = k0 ↠(k)â(k) . (162)
(2π)3 2k0
This obviously cures the divergence stemming from h0| â(k)↠(k) |0i in the
ground state energy and similar observables. In the remainder of this lecture
we will always assume implicit normal ordering, if not stared otherwise.
67
h i 0
o
+ ↠(k), â(k 0 ) eik·x−ik ·y
d3 k
Z n o
= e−ik·(x−y) − eik·(x−y)
(2π)3 2k0
= ∆+ (x − y) − ∆− (x − y) , (163)
where we have used the commutator relations for the creation and annihi-
lation operators to arrive at two δ-functions that allowed us to perform the
k 0 -integration, and where we have also introduced the two terms ∆± (x − y).
0 = ∂ µ ∂µ + m2 ∆(x − y)
d3 k h −ik·(x−y)
Z i
= ∂ µ ∂µ + m2 ik·(x−y)
e − e
(2π)3 2k0
Z 3
d k h −ik·(x−y) i
µ 2 ik·(x−y)
= ∂ ∂ µ + m e − e
(2π)3 2k0
d3 k
Z
2 2
h −ik·(x−y) ik·(x−y)
i
= −k + m e − e (164)
(2π)3 2k0
where the term k 2 − m2 in the last line guarantees that the overall
expression vanishes.
68
4.4 Second Quantisation of the Complex Scalar Field
Lagrangian and Hamilton and Field Operators Starting with the
Lagrangian of Eq. (104),
L = (∂µ φ∗ )(∂ µ φ) − m2 φ∗ φ ,
∂L ∂L
π = = φ̇∗ and π ∗ = = φ̇ (165)
∂ φ̇ ∂ φ̇∗
d3 k
Z h i
φ(x) = â(k) e−ik·x + b̂† (k) eik·x
(2π)3 2k0
d3 k
Z h i
φ∗ (x) = b̂(k) e−ik·x + ↠(k) eik·x . (168)
(2π)3 2k0
69
we arrive at commutator relations, for example,
h i Z
0 0
â(k), ↠(k 0 ) = d3 xd3 x0 e+ik·x−ik ·x
h i
× k0 φ̂(t, x) + iπ̂ ∗ (t, x), k00 φ̂∗ (t, x0 ) − iπ̂(t, x0 )
Z n h i
0 0
= d3 xd3 x0 e+ik·x−ik ·x −ik0 φ̂(t, x), π̂(t, x0 )
h io
+ ik00 π̂ ∗ (t, x), φ̂∗ (t, x0 )
Z
0 0
d3 xd3 x0 e+ik·x−ik ·x (k0 + k00 )δ 3 (x − x0 )
=
Z
0
= d3 xe+i(k−k )·x (k0 + k00 ) = 2k0 (2π)3 δ 3 (k − k 0 ) , (170)
d3 k
Z h i
† †
:Ĥ: = k 0 â (k)â(k) + b̂ (k)b̂(k) , (172)
(2π)3 2k0
and it looks like the Hamilton operator for the sum of two free real scalar
fields. This further fortifies the idea that we are presented by two kinds of
particles – those created and annihilated by ↠and â, and those created and
annihilated by b̂† and b̂, and that the vacuum is annihilated by both â and
b̂,
â(k) |0i = b̂(k) |0i = 0 . (173)
It is therefore natural to introduce two number operators for the two kinds
of particles,
d3 k
Z
N̂a = ↠(k)â(k)
(2π)3 2k0
d3 k
Z
N̂b = b̂† (k)b̂(k) . (174)
(2π)3 2k0
It is easy to check that they are indeed number operators counting the
number of a and b fields in a given state |ψi. Denoting
E Yna h nb h
iY i
(a) (a) (a) (b) (b) (b) † †
k
1 2k . . . k k
na 1 k 2 . . . k nb = â (k i ) b̂ (k i |0i ,
) (175)
i=1 i=1
70
it is easy to show that
D E
(a) (a) (a) (b) (b) (a) (a) (b) (b)
k 1 k 2 . . . k na k 1 k 2 . . . k nb N̂a k 1 k 2 . . . k (a)
(b) (b)
na k 1 k 2 . . . k nb = na .
(176)
Current and Charge As noted in Sec. 3.3, the Lagrangian for the com-
plex scalar field enjoys invariance under the “gauge transformation”
cf. Eq. (106). This leads to a conserved current given by Eq. (113)
←→
j µ = i φ∗ (∂ µ φ) − (∂ µ φ∗ )φ ≡ iφ∗ ∂ µ φ ,
where we added a factor i to ensure that the current is a real number. This
factor, obviously, does not change the fact that ∂µ j µ = 0. Of course the
current can be promoted to a current operator by replacing the fields with
field operators,
←→
ĵ µ = iφ̂∗ ∂ µ φ̂ . (177)
d3 k h †
Z i
†
= â (k)â(k) − b̂ (k) b̂(k) = N̂a − N̂b . (178)
(2π)3 2k0
This suggests that our two particle types a and b have opposite charged with
qa,b = ±1 . (179)
d3 k h †
Z i
†
:Q̂: = â (k)â(k) − b̂ (k)b̂(k) = N̂a − N̂b (180)
(2π)3 2k0
commutes with the Hamiltonian, i.e. [:Ĥ:, :Q̂:] = 0. This is indeed the case,
and we leave this proof for the problems below.
71
4.5 Problems & Solutions
1. States and Operators of the Real Scalar Field
E12 |k 1 k 2 i = Ĥ |k 1 k 2 i .
(c) show that the number operator N̂ counts the number of quanta:
N̂ |k 1 k 2 . . . k n i = n|k 1 k 2 . . . k n i (181)
Solution
|k 1 i = ↠(k 1 )|0i
|k 1 k 2 i = ↠(k 1 )|k 2 i = ↠(k 1 )↠(k 2 )|0i = ↠(k 2 )↠(k 1 )|0i
d3 k
Z
1 h
† †
i
:Ĥ: = k 0 : â (k)â(k) + ââ (k)(k) :
2 (2π)3 (2k0 )
d3 k
Z h i
†
= k 0 â (k)â(k) ,
(2π)3 (2k0 )
Ek1 k2 |k 1 k 2 i = Ĥ|k 1 k 2 i
d3 k †
Z
1
† †
= 3
â (k)â(k) â (k 1 )â (k 2 ) 0
2 (2π)
d3 k
Z
1 †
h
†
i
†
= â (k ) â(k), â (k 1 â (k 2 )
)
2 (2π)3
+ ↠(k)↠(k 1 )â(k)↠(k 2 ) 0
d3 k
Z
1
q
† † 3 2 2
= â (k)â (k 2 ) (2π) δ(k − k 2 ) 2 k + m
2 (2π)3
72
h i
+ ↠(k)↠(k 1 ) â(k), ↠(k 2 )
† † †
− â (k)â (k 1 )â (k 2 )â(k) 0
d3 k
Z
1
q
† † 3 2 2
= â (k)â (k 2 ) (2π) δ(k − k 1 ) 2 k + m
2 (2π)3
q
+ ↠(k)↠(k 1 ) (2π)3 δ(k − k 2 ) 2 k 2 + m2 − 0 0
q
1
q E
= 2 k 1 + m + 2 k 2 + m ↠(k 1 )↠(k 2 ) 0
2 2 2 2
2
= (E1 + E2 ) |k 1 k 2 i
as anticipated.
(c) With the number operator given by
d3 k
Z
N̂ = ↠(k)â(k)
(2π)3 (2k0 )
d3 k
Z h i
† † † † †
[N̂ , â (q)] = â (k)â(k)â (q) − â (q)â (k)â(k)
(2π)3 (2k0 )
d3 k
Z h
= ↠(k)(2π)3 (2q0 )δ 3 (k − q)
(2π)3 (2k0 )
i
+↠(k)↠(q)â(k) − ↠(q)↠(k)â(k)
= ↠(q) .
N̂ |k 1 k 2 . . . k n i = N̂ ↠(k 1 )|k 2 . . . k n i
= ↠(k 1 )(N̂ + 1)|k 2 k 3 . . . k n i = ↠(k 1 )↠(k 2 )(N̂ + 2)|k 3 . . . k n i = . . .
= ↠(k 1 )↠(k 2 ) . . . ↠(k n )n|0i = n|k 1 k 2 . . . k n i
as anticipated.
73
Solution
(a) write down the Lagrangian two real scalar fields φ1,2 of equal
mass and determine their canonical momenta π1,2
(b) construct the Hamiltonian from the fields and their momenta
(c) demand suitable commutators for fields and momenta
(d) express the fields in terms of creation and annihilation operators
and determine their commutation relations
(e) introduce the complex scalar fields φ and φ∗ as linear combi-
nations of φ1,2 , express them through creation and annihilation
operators. Fix the commutators of the creation and annihilation
operators that have not been explicitly calculated so far.
(f) calculate the commutator of the number operators with the cre-
ation and annihilation operators of the two fields, the commuta-
tor between the two number operators and with the charge and
Hamilton operator. Show that the charge is conserved by showing
that indeed [:Ĥ:, :Q̂:] = 0.
Solution
(a) Lagrangian for two fields as sum of two Lagrangians for single
fields
2
X 1
(∂µ φi )(∂ µ φi ) − m2i φ2i
L=
2
i=1
assume m1 = m2 .
∂L ∂L
πi = = = φ̇i
∂(∂t φi ) ∂ φ̇i
74
(b) As usual, Hamiltonian density given by
2 h 2
X i X 1h i
H= πi φ̇i − L = (φ̇i )(φ̇i ) + (∇φi )(∇φi ) + m2i φ2i
2
i=1 i=1
(c) Commutators in the usual way: field operators and and their
conjugate momentum operators do not commute, everything else
does:
h i
φ̂i (x, t), π̂j (y, t) = iδij δ 3 (x − y)
h i
φ̂i (x, t), φ̂j (y, t) = π̂i (x, t), π̂j (y, t) = 0
d3 k
Z h i
−ik·x † ik·x
φ̂i (x) = âi (k)e + â i (k)e
(2π)3 (2k0 )
d3 k
Z h i
−ik·x † ik·x
π̂i (x) = −ik â
0 i (k)e + ik â
0 i (k)e
(2π)3 (2k0 )
and therefore
Z h i
0
d3 xe−ik ·x k00 φ̂i (x) + iπ̂i (x)
d3 k
Z
0
= d3 xe−ik ·x
(2π)3 (2k0 )
h i
âi (k)e−ik·x (k00 + k0 ) + â†i (k)eik·x (k00 − k0 )
d3 k
Z h 0
= 3
d 3
x âi (k)e−i(k+k )·x−ik0 t (k00 + k0 )
(2π) (2k0 )
0
i
+â†i (k)ei(k−k )·x+ik0 t (k00 − k0 )
d3 k
Z h
= âi (k)(2π)3 δ 3 (k + k 0 )(k00 + k0 )e−ik0 t
(2π)3 (2k0 )
i
+â†i (k)(2π)3 δ 3 (k − k 0 )(k00 − k0 )eik0 t
1 h 0 0 −ik00 t
i 0
= 0 âi (k )(2k 0 )e + 0 = âi (k 0 )e−ik0 t ,
(2k0 )
p
where we have used that k0 = k 2 + m2 = (−k)2 + m2 = k00
p
75
0
Multiplying on both sides with eik0 t and replacing k 0 → k yields
Z h i
âi (k) = d3 xeik·x k0 φ̂i (x) + iπ̂i (x)
Taking the Hermitean conjugate and using that φ̂†i = φ̂i and
π̂i† = π̂i for real fields implies that
Z h i
† 3 −ik·x
âi (k) = d xe k0 φ̂i (x) − iπ̂i (x)
therefore8
π = φ̇∗ and π ∗ = φ̇ .
8
Expressing the fields φ1,2 through φ and φ∗ ,
1 i
φ1 = √ (φ + φ∗ ) and φ2 = − √ (φ − φ∗ ) and
2 2
yields the Lagrangian
L = (∂µ φ)(∂ µ φ∗ ) − m2 φ∗ φ
and the conjugate momenta are given, as before, by π = ∂L/∂ φ̇ and π ∗ = ∂L/∂ φ̇∗ .
76
Suitably combining the expansions for φ1 and φ2 yields
d3 k
Z
â1 (k) + iâ2 (k) −ik·x
φ̂(x) = √ e
(2π)3 (2k0 ) 2
#
↠(k) + i↠(k)
+ 1 √ 2 eik·x
2
d3 k
Z h i
−ik·x † ik·x
= â + (k)e + â − (k)e
(2π)3 (2k0 )
d3 k
â1 (k) − iâ2 (k) −ik·x
Z
φ̂∗ (x) = 3
√ e
(2π) (2k0 ) 2
#
â†1 (k) − iâ†2 (k) ik·x
+ √ e
2
d3 k
Z h i
−ik·x † ik·x
= â − (k)e + â + (k)e ,
(2π)3 (2k0 )
where
˙
Using that π̂i = φ̂i and that [φ1 , π2 ] = 0 and similar, the equal-
time commutators read
h i
φ̂(x, t), π̂(y, t)
1 h ˙ ˙
i
= φ̂1 (x, t) + iφ̂2 (x, t), φ̂1 (x, t) − iφ̂2 (x, t)
2
1 nh i h io
= φ̂1 (x, t), π̂1 (y, t) + φ̂2 (x, t), π̂2 (y, t) = iδ 3 (x − y)
h 2 i
φ̂ (x, t), π̂ ∗ (y, t)
∗
1 h ˙ ˙
i
= φ̂1 (x, t) − iφ̂2 (x, t), φ̂1 (x, t) + iφ̂2 (x, t)
2
1 nh i h io
= φ̂1 (x, t), π̂1 (y, t) + φ̂2 (x, t), π̂2 (y, t) = iδ 3 (x − y)
h 2 i
φ̂(x, t), π̂ ∗ (y, t)
1 h ˙ ˙
i
= φ̂1 (x, t) + iφ̂2 (x, t), φ̂1 (x, t) + iφ̂2 (x, t)
2
1 nh i h io
= φ̂1 (x, t), π̂1 (y, t) − φ̂2 (x, t), π̂2 (y, t) = 0
2
and similarly for all commutators of fields with fields and mo-
menta with momenta.
77
The non-vanishing commutators of the creation and annihilation
operators read
h i 1h i
â+ (k), â†+ (q) = â1 (k) + iâ2 (k), â†1 (k) − iâ†2 (k)
2
1 nh i h io
= â1 (k), â†1 (q) + â2 (k), â†2 (q) = (2k0 )(2π)3 δ 3 (k − q)
2
h i 1h i
â− (k), â†− (q) = â1 (k) − iâ2 (k), â†1 (k) + iâ†2 (k)
2
1 nh i h io
= â1 (k), â†1 (q) + â2 (k), â†2 (q) = (2k0 )(2π)3 δ 3 (k − q)
2
(†) (†)
As the a± commute with the a± we also have
h i h i
N̂± , â†∓ (q) = N̂± , â∓ (q) = 0 .
78
d3 k d3 q
Z
=
(2π)3 (2k0 ) (2π)3 (2q0 )
h i
â†+ (k)â+ (k)â†− (q)â− (q) − â†− (q)â− (q)â†+ (k)â+ (k)
= 0,
because the positive charge â+ and â†+ commute with their neg-
ative charge counterparts, as seen above.
This implies that the commutator of Hamilton or Charge operator
with the number operators also vanish. For example, with the
Hamilton operator from Eq. (172):
h i Z d3 k d3 q
Ĥ, N̂+ = k0
(2π)3 2k0 (2π)3 2q0
† † †
â+ (k)â+ (k) + â− (k)â− (k), â+ (q)â+ (q)
d3 k d3 q
Z
= k0
(2π)3 2k0 (2π)3 2q0
â†+ (k)â+ (k)â†+ (q)â+ (q) − â†+ (q)â+ (q)â†+ (k)â+ (k)
† † † †
+ â− (k)â− (k)â+ (q)â+ (q) − â+ (q)â+ (q)â− (k)â− (k)
d3 k d3 q
Z
= k0
(2π)3 2k0 (2π)3 2q0
(2π)3 2q0 δ 3 (k − q)â†+ (k)â+ (q) + â†+ (k)â†+ (q)â+ (k)â+ (q)
− (2π)3 2q0 δ 3 (k − q)â†+ (q)â+ (k) − â†+ (q)â†+ (k)â+ (q)â+ (k)
+0 = 0,
4. Momentum Operator
79
The total four-momentum operator of a real scalar field is given by
d3 k
Z
:P̂ µ:= k µ ↠(k)â(k) .
(2π)3 2k0
(a) Show that P̂ µ can be expressed in terms of the field operator φ̂(x)
and the conjugate momentum operator π̂(x) as
Z
:P̂ := d3 x :π̂(x)∂ µ φ̂(x):
µ
Solution
(a) Inserting the expansion of the field operator and its conjugate
momentum through plane waves and creation and annihilation
operators we have
d3 k d3 q
Z h
µ −ik·x †
:P̂ : = d3 x : −ik 0 â(k)e − â (k)eik·x
(2π 3 )2k0 (2π 3 )2q0
i
iq µ −â(q)e−iq·x + ↠(q)eiq·x :
d3 k d3 q q µ
Z
= d3 x
(2π 3 ) (2π 3 ) 4q0
h
: −â(k)â(q)e−i(k+q)·x + ↠(k)↠(q)ei(k+q)·x
i
+â(k)↠(q)e−i(k−q)·x + ↠(k)â(q)ei(k−q)·x :
d3 q q µ h
Z
= : −â(−q)â(q)e−2iq0 t + ↠(−q)↠(q)e2iq0 t
(2π 3 ) 4q0
i
+â(q)↠(q) + ↠(q)â(q) :
d3 q q µ †
Z
= â (q)â(q)
(2π 3 ) 2q0
The first two terms in the second-to last line vanishes because â(q)
commutes with â(−q), and similarly for the “daggered” operators.
We can therefore replace
q µ e−2iq0 t
Z
d3 q â(−q)â(q)
q0
q µ e−2iq0 t 1
Z
d3 q
= â(−q)â(q) + â(q)â(−q) ,
q0 2
80
showing for each component of q µ that this is an integration of an
odd function over an even integration space. The same reasoning
holds also true for the daggered operators. Therefore only the
two last terms survive and we have shown that indeed
d3 k
Z Z
µ µ †
:P̂ : = k â (k)â(k) = d3 x :π̂(x)∂ µ φ̂(x):
(2π)3 2k0
[P̂ µ , φ̂(x)]
d3 k d3 q
Z h i
µ † −iq·x † iq·x
= k â (k)â(k), â(q)e + â (q)e
(2π)3 2k0 (2π)3 2q0
d3 k d3 q
Z n h i
µ −iq·x †
= k e â (k), â(q) â(k)
(2π)3 2k0 (2π)3 2q0
h io
+ eiq·x ↠(k) â(k), ↠(q)
d3 k
Z n o
µ −ik·x ik·x †
= k −e â(k) + e â (k)
(2π)3 2k0
d3 k
Z n o
µ −ik·x ik·x †
= (−i∂ ) e â(k) + e â (k) = −i∂ µ φ̂(x) ,
(2π)3 2k0
as demanded.
∆1 (x − y) = ∆+ (x − y) + ∆− (x − y)
(b) show that ∆1 (x − y) does not vanish outside the light-cone, i.e.
that ∆1 (x − y) 6= 0 for (x − y)2 < 0.
Solution
∆+ (x − y) = h0|φ̂(x)φ̂(y)|0i
d3 k d3 k 0
Z D E
−ikx+ik0 y † 0
= e 0 â(k)â (k ) 0
(2π)3 (2k0 ) (2π)3 (2k00 )
81
d3 k d3 k 0
Z D h i E
−ikx+ik0 y † 0
= e 0 â(k), â (k ) 0
(2π)3 (2k0 ) (2π)3 (2k00 )
d3 k d3 k 0
Z
0
e−ikx+ik y 0 (2π)3 (2k0 )δ 3 (k − k 0 ) 0
= 3 3 0
(2π) (2k0 ) (2π) (2k0 )
d3 k
Z
= e−ik(x−y)
(2π)3 (2k0 )
d3 k
Z
∆− (x − y) = eik(x−y)
(2π)3 (2k0 )
d3 k d3 k 0
Z
−ik(y−x) (2π)3 (2k0 )δ 3 (k − k 0 ) 0
= e 0
(2π)3 (2k0 ) (2π)3 (2k00 )
d3 k d3 k 0
Z D h i E
−ik0 y+ikx 0 †
= e 0 â(k ), â (k) 0
(2π)3 (2k0 ) (2π)3 (2k00 )
= h0|φ̂(y)φ̂(x)|0i
Therefore, as demanded
D h i E
∆(x−y) = ∆+ (x−y)+∆− (x−y) = 0 φ̂(x)φ̂(y) + φ̂(y)φ̂(x) 0
(b)
d3 k
Z
−ik(x−y) ik(x−y)
∆1 (x − y) = e + e
(2π)3 (2k0 )
d3 k
Z
x0 →y0
−ik(x−y) ik(x−y)
−→ e + e
(2π)3 (2k0 )
d3 k
Z
x→y
−→ 2 →∞
(2π)3 (2k0 )
d3 k
Z Z
P̂ µ = k µ ↠(k)â(k) = d3 xπ̂(x)∂ µ φ̂(x) ;
(2π 3 )2k0
h i
(b) Ĥ, ↠(k)â(q) ;
82
Solution
83
7. ∗ Properties of the Charge Operator In the following, consider a
free complex scalar field φ.
(a) Show that the (normal-ordered) charge operator is given by
Z h i
:Q̂: = d3 x :φ̂∗ (x)π̂ ∗ (x) − φ̂(x)π̂(x):
d3 k h †
Z i
†
= â (k)â(k) − b̂ (k)b̂(k)
(2π)3 2k0
Solution
d3 k d3 q
Z
1
= d3 x q0
2 (2π)3 2k0 (2π)3 2q0
h i
: e+i(k−q)·x ↠(k)â(q) − b̂† (k)b̂(q)
h i
− e−i(k−q)·x b̂(k)b̂† (q) − â(k)↠(q)
h i
− e+i(k+q)·x ↠(k)b̂† (q) − b̂† (k)↠(q)
h i
−i(k+q)·x
+e b̂(k)â(q) − â(k)b̂(q) :
d3 k d3 q
Z
1
= q0
2 (2π) 2k0 (2π)3 2q0
3
h i
: e+i(k0 −q0 )x0 δ 3 (k − q) ↠(k)â(q) − b̂† (k)b̂(q)
h i
− e−i(k0 −q0 )x0 δ 3 (k − q) b̂(k)b̂† (q) − â(k)↠(q)
h i
− e+i(k0 +q0 )x0 δ 3 (k + q) ↠(k)b̂† (q) − b̂† (k)↠(q)
h i
+ e−i(k0 +q0 )x0 δ 3 (k + q) b̂(k)â(q) − â(k)b̂(q) :
84
d3 k h †
Z
1 †
i
= â (k)â(k) − b̂ (k)b̂(k) .
2 (2π)3 2k0
(b) To calculate the commutator let us first take a look at one typical
term, namely
[P̂ µ , π̂(x)] = [P̂ µ , ∂t φ̂∗ (x)] = −i∂t ∂ µ φ̂∗ (x) = −i∂ µ π̂(x)
we see that
h i Z h i h i
µ
:Q:, :P̂ : = − d x 3
P̂ µ , φ̂∗ (x) π̂ ∗ (x) + φ̂∗ (x) P̂ µ , π̂ ∗ (x)
h i h i
µ µ
− P̂ , φ̂(x) π̂(x) + φ̂(x) P̂ , π̂(x)
Z
= i d x [∂ µ φ̂∗ (x)]π̂ ∗ (x) + φ̂∗ (x)[∂ µ π̂ ∗ (x)]
3
µ µ
− [∂ φ̂(x)]π̂(x) − φ̂(x)[∂ π̂(x)]
Z h i
3 µ ∗ ∗
= i d x ∂ φ̂ (x)π̂ (x) − φ̂(x)π̂(x) = 0
d3 k h †
Z i
iπ †
P̂ = exp − â (k)â(k) − ηP â (k)â(−k) ,
2 (2π)3 2k0
where the phase ηP = ±1 is the intrinsic parity of the field. Fields
with ηP = 1 are scalars and those with ηP = −1 are pseudoscalars.
Prove that [P̂, Ĥ] = 0.
Solution
85
[f (Ô), X̂] = 0 if [Ô, X̂] = 0, because functions of operators are defined
through their series expansion. This means that we have to show that
d3 k
Z
† †
0 = â (k)â(k) − ηP â (k)â(−k) ,
(2π)3 2k0
d3 q
Z
†
q0 â (q)â(q)
(2π)3 2q0
d3 k d3 q
Z Z h i
† †
= q 0 â (k)â(k), â (q)â(q)
(2π)3 2k0 (2π)3 2q0
h i
† †
− ηP â (k)â(−k), â (q)â(q)
d3 k d3 q
Z Z
= q0
(2π)3 2k0 (2π)3 2q0
(2π 3 )2q0 δ 3 (k − q)↠(k)â(q) − ↠(k)↠(q)â(k)â(q)
−(2π 3 )2q0 δ 3 (k − q)↠(q)â(k) + ↠(q)↠(k)â(q)â(k)
h
− ηP (2π 3 )2q0 δ 3 (k + q)↠(k)â(q) − ↠(k)↠(q)â(−k)â(q)
− (2π 3 )2q0 δ 3 (k − q)↠(q)â(−k) + ↠(q)↠(k)â(q)â(−k)
d3 k d3 q
Z Z
† †
= −ηP q0 â (k)â(−k) − â (k)â(−k) = 0
(2π)3 2k0 (2π)3 2q0
86
5 Fermions
In this section we will get acquainted with the Dirac equation, which intro-
duces not only fermions, but also provides insight into anti-particles.
The Dirac equation emerges through linearisation of the Klein-Gordon equa-
tion, after realising that its quadratic form yields negative energy solutions.
Such a linearised form, however, only satisfies the original energy-momentum
relation – essentially the kernel of the free Klein-Gordon equation – if the
fields have an even number of components, at least two. This proves to be a
blessing, as it allows us to describe spin-1/2 particles, and the corresponding
fields are dubbed “spinors”. Insisting on maintaining that the spinors sat-
isfy the Klein-Gordon equation for massive particles leads to spinors with
four components - two more than necessary for spin-1/2 particles. These
additional degrees of freedom are identified with negative energy solutions
and interpreted as anti-particles. As before, in the case of the scalar fields,
this implies that the energy spectrum of the theory is unbounded from be-
low. Consequently the vacuum is not empty, and i fact it contains short-
lived quantum fluctuations of particle+anti-particle with opposite energy,
momentum and spin.
The Dirac equation has been covered ubiquitously in the literature. Keeping
in mind that we use a somewhat different (and in my opinion, more mod-
ern) normalisation, it would maybe be a good idea to also take a look at
Sections 4.1 and 4.2 of Hatfield [3] or Sections 3.1-3.4 and 3.6 in Peskin &
Schroeder [1], the latter section more of some extended reading. It is also
worthwhile to check out Chapter 2 of Itzykson & Zuber [?], if you can find
it somewhere.
Fourier-transforming it into
(E 2 − p2 − m2 )φ = 0 −→ E 2 = p2 + m2 (182)
we realise that, due to its quadratic form, nothing prevents us from con-
structing solutions with negative energies. Assuming plane wave solutions
for the fields, φ(x) ∼ exp(ikx) the charge or probability density for the
complex scalar field is given by
87
which can be translated into a real, unit-free quantity, that is more ap-
propriate for a probability density. This is achieved through a suitable
normalisation, for example
i µ
j µ → j̃ µ = j , (184)
2m
such that ρ = k0 /m. For negative-energy solutions, though, this would
result in negative probability densities, which are extremely hard to inter-
pret. Ultimately, the appearance of these solutions mean that the energy
spectrum of the theory is not bound from below and there is no lowest en-
ergy ground-state. In other words, there is nothing that prevents us from
producing more and more particles, by pairing positive and negative energy
solutions – clearly an unacceptable problem for the interpretation of the the-
ory. Ultimately this shows that it is impossible to produce a single-particle
theory when imposing Lorentz-invariance as a construction paradigm.
Of course, we know by now that this issue can be completely circumnav-
igated by identifying the negative energy-solutions as anti-particles, parti-
cles with positive energy but opposite charge that propagate backwards in
time. However, when Dirac introduced his famous equation in 1928 this
anti-particles were not discovered yet, and it was in fact his work that in-
troduced anti-particles as a meaningful theoretical concept that emerges
naturally when combining Quantum Mechanics and Special Relativity into
Quantum Field Theory.
88
αi and β must satisfy the following relations
{αi , αj } = αi αj + αj αi = 2δij
{αi , β} = 0
β 2 = αi2 = 1
Tr(αi ) = Tr(β) = 0 . (187)
This implies that the eigenvalues of αi and β are ±1, and the combination of
them being traceless and having these eigenvalues suggests that they must
be of an even dimension, i.e. dim(αi , β) = 2, 4, . . . . Focusing on the case
of lowest dimension, 2×2 matrices, we can see straightaway that the αi can
be identified with the Pauli matrices, αi = σi , where
0 1 0 −i 1 0
σ1 = , σ2 = , and σ3 = . (188)
1 0 i 0 0 −1
This however won’t work for massive theories where m 6= 0: There is just no
fourth candidate matrix for β that satisfies all the properties of Eq. (187).
This has two implications: First of all, for massless theories, we could stick
with two-component fields ψ, also known as Weyl spinors. And secondly,
for massive theories like the ones we’re going to pursue, we must use four-
component fields – the Dirac spinors – and have four-dimensional αi and β
matrices:
0 σi 1 0
αi = and β = . (189)
σi 0 0 −1
Direct calculation shows that they enjoy the following anti-commutator re-
lation
{γ µ , γ ν } = γ µ γ ν + γ ν γ µ = 2g µν . (191)
2
In addition, γ 0 = γ 0† is Hermitean with γ 0 = 1, while the γ i = −γ i† are
anti-Hermitean, with (γ i )2 = −1.
89
Dirac Equation Multiplying the Dirac equation, expressed through the
α and β matrices, Eq. (185), from the left with γ 0 we arrive at
(iγ µ ∂µ − m1)ηξ ψξ = (i∂/ − m)ψ = 0 . (192)
In the equation above, Eq. (192) the components of the Dirac equation in
“spinor space” have been made explicit, indicated by the indices η and ξ. It
is important to stress that this exhibits the fact that there are two spaces
in the Dirac equation, namely the “normal” Minkowski space with index µ,
incorporating the external Lorentz symmetry of space-time, and this spinor
space. The γ matrices and the spinor ψ have multiple components in this
space, and the mass term is diagonal in this space, indicated by the 1-
symbol. As before, the Lorentz indices µ etc. run from 0 to 3, while the
Dirac or spinor indices run from 1 to 4.
Dirac Equation for ψ † The nature of the equation above suggest that
the Hermitean conjugate spinor ψ † represents a second, independent field,
similar to φ∗ and φ. Straightforward Hermitean conjugation of Eq. (185)
results in
∂ψ † (x, t)
−i = i∇ψ † (x, t) · α† + mψ † (x, t)β † , (193)
∂t
and multiplying from the right with β † = β = γ 0 yields
←−
−iψ † (x, t)∂µ γ †µ = mψ † (x, t) . (194)
2
Using γ 0 = 1 and γ † = (βα)† = αβ = β(βα)β = γ 0 γγ 0 while defining the
“barred” spinor ψ̄ = ψ † γ 0 allows to find the E.o.M. for the barred spinor as
←−
ψ̄(i ∂/ + m) = 0 . (195)
Lagrangian It is easy to check that the two E.o.M. for the spinors ψ and
ψ † can be obtained from the free Dirac Lagrangian
←→
L = ψ̄(x) i ∂/ − m ψ(x) , (196)
where
←→ 1
a ∂ b = [a(∂b) − (∂a)b] . (197)
2
The E.o.M. for ψ (ψ̄) are obtained,as usual, by varying the Lagrangian with
respect to ψ̄ (ψ):
∂L ∂L 1 →−
− ∂µ = − mψ + [i∂/ψ − ∂µ (−iγ µ ψ)] = i ∂/ − m ψ = 0
∂ ψ̄ ∂(∂µ ψ̄) 2
∂L ∂L 1h ← − i ←−
− ∂µ = − mψ̄ − ψ̄(i ∂/ ) + ∂µ (iψ̄γ µ ) = −ψ̄ i ∂/ + m = 0 .
∂ψ ∂(∂µ ψ) 2
(198)
90
Conserved Current It is relatively straightforward to construct a con-
served current from the two E.o.M. Eqs (192) and (195): multiply the former
from the left with ψ̄ and the latter from the right with ψ and add. This
results in
→
− ←− →
− ← −
0 = ψ̄ · (i ∂/ − m)ψ + ψ̄ · (i ∂/ + m) · ψ = iψ̄( ∂/ + ∂/ )ψ (199)
∂µ j µ = ∂µ iψ̄γ µ ψ .
(200)
d3 p −ip·x
Z
ip·x
ψη (x) = e uη (p) + e v η (p) , (201)
(2π)3
where we have made explicit the spinor index η. This expansion moves the
spinor index to the u and v spinors, i.e. they are objects with four entries,
and the Dirac matrices act on these indices10 . To construct them, it is
sufficient to realise that the E.o.M. become a system of linear equations for
the eigenstates u(p) and v(p). Let us first solve this equation for a particle
at rest, p = 0, p0 = E = m, leading to
implies that the third and fourth component of u and the first and second
component of v must be zero. Both u and v therefore have two independent
9
But, although they look like vectors because of the four-components, they differ from
four-vectors in how they behave under Lorentz transformations. Simply put: spinor index
6= Lorentz index)
10
Positive and negative energy solutions ψ± are of course related to the wave factors
such that
and we will recycle them later when quantising the Dirac fields.
91
solutions each, and the corresponding eigenstates can be readily identified
with the two spin states: u(1/2) describe positive-energy particles with spin
up/down, and v (1/2) decibel negative-energy particles with spin up/down.
Choosing normalised “eigenvectors” then results in
1 0
0 , u(2) (0) = 1 ,
u(1) (0) =
0 0
0 0
0 0
0 0
v (1) (0) = (2)
1 , v (0) = 0 . (204)
0 1
92
q
where the energy E = p2 + m2 > 0 and the normalisation is given by
√
η= E + m. (209)
Note that we have normalised the spinors such that, apart from the norm η
the first component of the spinors equals 1.
What is left to do now is to explictly check that the spinors indeed satisfy
their equations of motion, i.e. that (p/ − m)u(1,2) (p) and (p/ + m)v (1,2) (p)
vanish. For example, for u(1) and v (1) we find
/ − m)u(1) (p)
(p
E−m −pz −p− 1
0
0 E−m −p+ pz 0
=
pz
−(E + m)
pz p− 0 E+m
p+
p+ −pz 0 −(E + m) E+m
p2
E − m − E+m
−p+ pz +p+ pz
E+m
= = 0;
−pz + (E+m)pz
E+m
(E+m)p+
−p+ + E+m
Similar calculations for u(2) and v (2) prove that the spinors indeed satisfy
the equations of motion.
ū(i) (p)u(j) (p) = 2mδij = −v̄ (i) (p)v (j) (p) . (211)
93
A simple calculation exemplifies how to calculate such spinor products. For
example for i = j = 1 we find
p2z
(i) (j) (i)† 0 (j) 2 p+ p−
ū u = u γ u = η 1 + 0 − −
(E + m)2 (E + m)2
E 2 + 2Em + m2 − p2 2m(E + m) 2m
= η2 2
= η2 2
= η2 , (212)
(E + m) (E + m) E+m
† E 2 + 2Em + m2 + p2 2E(E + m) 2E
u(i) u(j) = η 2 2
= η2 2
= η2 = 2E .
(E + m) (E + m) E+m
(213)
Therefore,
u(i)† (p)u(j) (p) = v (i)† (p)v (j) (p) = 2p0 δij
2 2
(i) (i)
X X
u(i)
α ūβ = (p
/ + m)αβ , vα(i) v̄β = (p/ − m)αβ . (215)
i=1 i=1
Using Eq. (207) and directly calculate the spinor products, i.e. the terms
uū we see that this holds in fact true. It is important to stress that the
product of “column vector” and “row vector” is not a scalar product but
generates a matrix.
94
priori no well-defined ground state. Dirac circumnavigated this problem
by demanding that the negative energy solutions are all fill, and that the
v-states are “holes” in this otherwise full “sea” of negative energy solutions.
This obviously abandons any notion of the resulting Quantum Field The-
ory describing just one particle – which is possible in Quantum Mechanics.
Adding Special Relativity to the mix implies that the resulting Quantum
Field Theory indeed can only be realised as a multi-particle theory. It is
then not surprising that the vacuum is not “empty”; instead it can have
short-time quantum fluctuations of particle+anti-particle (hole), with op-
posite energy, momentum, and spin such that the overall quantum numbers
(all 0) are conserved. We will now move on to quantise this theory.
It is worth noting here that our conjugate momenta differ from the usual
form in textbooks by a factor of 1/2, stemming from our vey literal inter-
pretation of the derivative of Eq. (197) in the Lagrangian, Eq. (196).
95
and where we used that π̂β = ψ̂β† .
2
d3 p X h −ip·x
Z i
(i) ip·x ˆ† (i)
ψ(t, x) = e b̂i (p)u (p) + e d i (p)v (p)
(2π)3 2p0
i=1
2 h
d3 p
Z i
e−ip·x dˆi (p)v̄ (i) (p) + eip·x b̂†i (p)ū(i) (p) γ 0
X
ψ † (t, x) =
(2π)3 2p0
i=1
(220)
and all others vanishing, the anti-commutators of Eq. (218) are fulfilled.
For example:
n o
ψ̂α (t, x), ψ̂β† (t, y)
d3 p d3 q
Z n o
−ip·x−iq·y (β) 0 (α) ˆβ (q)
= e v̄ (q)γ u (p) b̂α (p), d
(2π)3 2p0 (2π)3 2q0
n o
+e−ip·x+iq·y ū(β) (q)γ 0 u(α) (p) b̂α (p), b̂†β (q)
n o
+e+ip·x−iq·y v̄ (β) (q)γ 0 v (α) (p) dˆ†α (p), dˆβ (q)
n o
+e +ip·x+iq·y (β)
ū 0 (α)
(q)γ v (p) dˆ†α (p), b̂†β (q)
d3 p d3 q
Z
3 3
= 2p0 δαβ (2π) δ (p − q)
(2π)3 2p0 (2π)3 2q0
−ip·x+iq·y (β)† (α) +ip·x−iq·y (β)† (α)
e u (q)u (p) + e v (q)v (p)
96
d3 p
Z
−ip·(x−y) (β)† (α) +ip·(x−y) (β)† (α)
= δαβ e u (p)u (p) + e v (p)v (p)
(2π)3 2p0
d3 p
Z
−ip0 ·(t−t)+ip·(x−y) +ip0 ·(t−t)−ip·(x−y)
= δαβ 2p0 δαβ e +e
(2π)3 2p0
d3 p
Z h i
−ip·(x−y) +ip·(x−y)
= δαβ e + e = 2δαβ δ 3 (x − y) , (222)
(2π)3
in agreement with Eq. (218). We realise that due to the equal times, the
exponentials of the time differences vanish; in addition, because α and β
are external parameters, we cannot use Einstein’s convention of summing
over repeated indices, since this would eliminate these parameters and the
right-hand side of the anti-commutator would not depend on them. Simply
put, the α and β are not indices in some space but label the spin-states of
the fermions and cannot be summed over. Finally, we used that δ 3 (x − y) =
δ 3 (y − x).
States To construct states with one and more particle states, we first
realise that
b̂†1,2 (p)/b̂1,2 (p) creates/annihilates positive-energy electrons with spin
up/down and momentum p;
+, p, ↑ = b† (p)|0i
1
−, p, ↓ = d† (p)|0i .
2 (223)
While this looks straightforward, things become more interesting when con-
sidering two-electron states, both with positive energy, momentum p, and
one spin up and one spin down:
+, p, ↑; +, p, ↓ = b† (p)b† (p) |0i = −b† (p)b† (p) |0i ,
1 2 2 1 (224)
97
Hamilton Operator To promote the Hamilton density of Eq. (217) to an
operator it is sufficient to replace the fields with field operators. Plugging in
the expansion in terms of creation and annihilation operators, using ūγ 0 =
u† and v̄γ 0 = v † , and integrating over space, we find
2
d3 p d3 q
Z X
3
Ĥ = d x
(2π)3 2p0 (2π)3 2q0
i,j=1
←
→
e−ip·x dˆi (p) v̄ (i) (p) + eip·x b̂†i (p) ū(i) (p) iγ · ∇ + m
× −iq·x
e (j)
b̂j (q)u (q) + eiq·x dˆ†j (q)v (j) (q)
2
d3 p d3 q
Z X
3
= d x
(2π)3 2p0 (2π)3 2q0
i,j=1
1
−i(p+q)·x ˆ (i) (j)
e di (p)b̂j (q) v̄ (p) γ · (−p + q) + m u (q)
2
1
−i(p−q)·x ˆ ˆ† (i) (j)
+ e di (p)dj (q) v̄ (p) γ · (−p − q) + m v (q)
2
1
+ e+i(p−q)·x b̂†i (p)b̂j (q) ū(i) (p) γ · (+p + q) + m u(j) (q)
2
1
† †
+ e +i(p+q)·x
b̂i (p)dˆj (q) ū (p)
(i) (j)
γ · (+p − q) + m v (q)
2
2
d3 q
Z X
=
(2π)3 4q02 i,j=1
h i
dˆi (−q)b̂j (q) v̄ (i) (−q) +γ · q + m u(j) (q) e−2iq0 x0
h i
dˆi (q)dˆ†j (q) v̄ (i) (q) −γ · q + m v (j) (q)
+
h i
b̂†i (q)b̂j (q) ū(i) (q) +γ · q + m u(j) (q)
+
h i
b̂†i (−q)dˆ†j (q) ū(i) (−q) −γ · q + m v (j) (q) e+2iq0 x0 ,
+
(226)
98
and ūγ0 = u† and v̄γ0 = v † ,
2 n
d3 q
Z
1 X h i
ˆi (−q)b̂j (q) v (i)† (−q)u(j) (q) e−2iq0 x0
Ĥ = d
2 (2π)3 2q0
i,j=1
h i h i
− dˆi (q)dˆ†j (q) v (i)† (q)v (j) (q) + b̂†i (q)b̂j (q) u(i)† (q)u(j) (q)
h i
† ˆ† (i)† (j) +2iq0 x0
+ b̂i (−q)dj (q) u (q)v (−q) e
With the orthogonality relations of Eq. (214) and their counterparts for
terms u(i)† (−q)v (j) (q) and v (i)† (−q)u(j) (q), the first and the last term in the
bracket above vanish. We finally arrive at the Hamiltonian
2 h
d3 q
Z i
† †
X
Ĥ = q 0 b̂ (q) b̂i (q) − ˆ
di (q)dˆ (q) , (228)
i i
(2π)3 2q0
i=1
which exhibits the same problems with infinite ground state energy as its
counterpart of the Klein-Gordon field, cf.. Sec. 4.2. We cure this, again,
by applying normal-ordering, Eq. (160) for bosons, which for fermion fields,
however, comes with an extra minus sign to encode the Pauli exclusion
principle,
:dˆ†i (q)dˆi (q):= − :dˆi (q)dˆ†i (q):= dˆ†i (q)dˆi (q) . (229)
Therefore, the normal-ordered Hamiltonian is given by
2 h
d3 q
Z i
† ˆ† (q)dˆi (q) .
X
Ĥ = q 0 b̂i (q) b̂i (q) + di (230)
(2π)3 2q0
i=1
Introducing number operators N̂± for particles with positive and negative
energy, electrons and positrons,
2 2
b̂†i (q)b̂i (q) and N̂− (q) = dˆ†i (q)dˆi (q) ,
X X
N̂+ (q) = (231)
i=1 i=1
we see that the Hamiltonian merely sums the energies of these particles
d3 q
Z h i
:Ĥ: = q 0 N̂+ (q) − N̂− (q) . (232)
(2π)3 2q0
99
we arrive at
2 h
d3 q
Z i
† ˆ† (q)dˆi (q) .
X
:Q̂:= b̂i (q)b̂i (q) − di (234)
(2π)3 2q0
i=1
d3 q
Z h i
:Q̂: = N̂+ (q) − N̂− (q) , (235)
(2π)3 2q0
and the overall charge of the system is given by the difference of the total
numbers of positively and negatively charged particles. It is a straightfor-
ward exercise to show that the charge is conserved, by asserting that the
commutator of the charge and Hamilton operator vanishes; we leave this as
an exercise.
100
5.3 Problems & Solutions
1. Dirac equation and Anti-Commutators
Show how the anti-commutation relations for the α and β matrices
follow from the requirement that the solutions to the Dirac E.o.M.
∂
i ψ(x, t) = [−iα · ∇ + βm] ψ(x, t)
∂t
also satisfy the KG equation.
Solution
∂2
− ψ(x, t) = [−iα · ∇ + βm]2 ψ(x, t)
∂t2
= −(α · ∇)2 + β 2 m2 − im(α · ∇ · β + βα · ∇) ψ(x, t)
X 3 3
X
= − αi ∂i αj ∂j + β 2 m2 − im (αi β + βαi ) ∂i ψ(x, t)
i,j=1 i=1
3
" #
!
X
= − ∂i2 + m2 ψ(x, t) = [−∇2 + m2 ]ψ(x, t)
i=1
where we have imposed equality with the relevant part of the Klein-
Gordon equation in the final line. Direct comparison with individual
terms shows that:
3
X 3
X
− αi αj ∂i ∂j = − ∂i2
i,j=1 i=1
2 2 2
+β m = m
3
X
−im (αi β + βαi ) ∂i = 0
i=1
and therefore
101
2. Commutators with the Dirac Hamilton Operator
Calculate the following commutators:
Solution
Ĥ = α · p̂ + βm ,
i iijk
[Ĥ, Ŝ ] = − [α · p̂ + βm, αj αk ]
4
102
iijk l iijk l
= − p̂ [αl , αj αk ] = − p̂ [(2δlj − αj αl )αk − αj (2δlk − αl αk )]
4 4
iijk j i
p̂ αk − p̂ αj = −iijk p̂j αk = − ip̂ × α
k
= −
2
and therefore
(e) [Ĥ, Ĵ] = −p̂ × α:
This proves that neither orbital nor spin angular momentum are
conserved quantities for the free fermions described by the Dirac
equation, and only their total angular momentum is conserved.
Solution
Using the Dirac γ matrices of Eq. (??) we then obtain an equation for
the two components as
E − m −σ · p ψ+
= 0,
σ · p −E − m ψ−
where pµ = (E, p). To solve this system, its determinant must vanish
and we arrive at
E − m −σ · p
0 = = −(E 2 − m2 ) + (σ · p)2 = −E 2 + p2 + m2 ,
σ · p −E − m
(E − m)ψ+ − (σ · p)ψ− = 0
(σ · p)ψ+ − (E + m)ψ+ = 0
103
implying that
σ·p
ψ− = ψ+
E+m
and therefore he positive energy solutions are given by
ψ+
u(p) = u+ (E, p) = σ · p
ψ+
E+m
with two basic spinors ψ+ for spinup and spin-down solutions given by
(±) 1 0
ψ+ = and .
0 1
with two basic spinors ψ+ for spinup and spin-down solutions given by
(±) 1 0
ψ− = and .
0 1
The last thing to note is that for the negative-energy solutions there
emerged a relative sign between energy and momentum, which makes
the assignment of a plane-wave factor tricky. Therefore the v-spinors
where introduced such that
σ·p
ψ−
v(p) = u− (−E, −p) = E + m
ψ−
104
(b) prove, by explicit calculation, that
2 2
(i) (i)
X X
u(i)
α ūβ = (p/ + m)αβ , vα(i) v̄β = (p/ − m)αβ
i=1 i=1
Solution
√
(a) Scalar products ūu and v̄v, for η = E + m:
T
1 0 0 1
0 1
0 1 0 0
0 0
ū(1) (p)u(1) (p) = η 2
pz pz
E+m 0 −1 0
0
E+m
p− p+
E+m 0 0 −1 0 E+m
!
2 p2 + p2
p z x y
= η2 1 + 0 − −
(E + m)2 (E + m)2
E 2 + m2 + 2Em − p2 2m
= η2 2
= η2 = 2m
(E + m) E+m
T
1 0 0 1
0 0
0 1 0 0
0 1
ū(1) (p)u(2) (p) = η 2
pz
p−
E+m 0 −1 0
0
E+m
p− −pz
E+m 0 0 −1 0 E+m
2 pz p− pz p−
= η 0+0− + =0
(E + m)2 (E + m)2
T
0 0 0 10 1
1 1 0 00 0
ū(2) (p)u(1) (p) = η 2
p +
pz
E+m 0 −1 0
0
E+m
−pz p+
E+m 0 0 −1 0 E+m
2 pz p+ pz p+
= η 0+0− + =0
(E + m)2 (E + m)2
105
T
0 1 0 0 0 0
1 0 1 0 0 1
ū(2) (p)u(2) (p) = η 2
p+ p−
E+m
0 0 −1 0
E+m
−pz −pz
E+m 0 0 0 −1 E+m
E2 + 2
m + 2Em − p2 2m
= η2 = η2 = 2m
(E + m)2 E+m
pz T pz
E+m 1 0 0 0 E+m
p− 0 p+
1 0 0
v̄ (1) (p)v (1) (p) = η 2
E+m E+m
1 0 0 −1 0 1
0 0 0 0 −1 0
!
p2z p2x + p2y
= η2 + −1−0
(E + m)2 (E + m)2
p2 − E 2 − m2 − 2Em 2m
= η2 2
= −η 2 = −2m
(E + m) E+m
pz T p−
E+m 1 0 0 0 E+m
p− 0 −pz
1 0 0
v̄ (1) (p)v (2) (p) = η 2
E+m E+m
1 0 0 −1 0 0
0 0 0 0 −1 1
2 pz p− pz p−
= η − −0−0 =0
(E + m)2 (E + m) 2
p+ T pz
E+m 1 0 0 0 E+m
−pz 0 p+
1 0 0
v̄ (2) (p)v (1) (p) = η 2
E+m E+m
1 0 0 −1 0 1
0 0 0 0 −1 0
2 pz p+ pz p+
= η 2
− 2
−0−0 =0
(E + m) (E + m)
p+ T p−
E+m 1 0 0 0 E+m
−pz 0 −pz
1 0 0
v̄ (2) (p)v (2) (p) = η 2
E+m E+m
1 0 0 −1 0 0
0 0 0 0 −1 1
!
p2x + p2y p2z
= η2 + −1−0
(E + m)2 (E + m)2
p2 − E 2 − m2 − 2Em 2m
= η2 2
= −η 2 = −2m
(E + m) E+m
106
Scalar products of v̄u and ūv:
pz T
E+m 10 0 0 1
p− 01 0 0 0
v̄ (1) (p)u(1) (p) = η 2
E+m
pz
1 00 −1 0
E+m
p+
0 00 0 −1 E+m
pz pz
= η2 − =0
E+m E+m
pz T
E+m 1 0 0 0 0
p− 0 1 0 0 1
v̄ (1) (p)u(2) (p) = η 2
E+m p−
1 0 0 −1 0 E+m
−pz
0 0 0 0 −1 E+m
2 p− p−
= η − =0
E+m E+m
p+ T
E+m 10 0 0 1
−pz 01 0 0 0
v̄ (2) (p)u(1) (p) = η 2
E+m
pz
1 00 −1 0
E+m
p+
0 00 0 −1 E+m
p+ p+
= η2 − =0
E+m E+m
p+ T
E+m 1 0 0 0 0
−pz 0 1 0 0 1
v̄ (2) (p)u(2) (p) = η 2
E+m p−
1 0 0 −1 0 E+m
−pz
0 0 0 0 −1 E+m
2 −pz −pz
= η − =0
E+m E+m
T pz
1 1 0 0 0 E+m
p+
0 0 1 0 0
ū(1) (p)v (1) (p) = η 2
E+m
pz
E+m
0 0 −1 0 1
p−
E+m 0 0 0 −1 0
−pz −pz
= η2 − =0
E+m E+m
107
T p−
1 0 0 1
0 E+m
−pz
0 0
1 0 0
ū(1) (p)v (2) (p) = η 2
E+m
pz
E+m 0 −1 0
0 0
p−
E+m 0 0 −10 1
2 p− p−
= η − =0
E+m E+m
T pz
0 0 0 01 E+m
p+
1 1 0 0
0
ū(2) (p)v (1) (p) = η 2
E+m
p+
E+m 0 −1 0
0 1
−pz
E+m 0 0 −1 0 0
2 p+ p+
= η − =0
E+m E+m
T p−
0 0 0 10 E+m
−pz
1 0
1 0 0
ū(2) (p)v (2) (p) = η 2
E+m
p+
E+m 0 −1 0
0 0
−pz
E+m 0 0 −1
0 1
2 −pz −pz
= η − =0
E+m E+m
−pz −p−
E+m 0
0 E+m −p+ pz
=
−E + m
pz p− 0
p+ −pz 0 −E + m αβ
T
1 1 1 0 0 0
2
0 0 0 1 0 0
= η
pz
pz
0 0 −1 0
E+m E+m
p+ p−
E+m E+m 0 0 0 −1
108
T
0 0 1 0 0 0
1 1 0 1 0 0
+
p−
p+
0 0 −1 0
E+m E+m
−pz −pz
E+m E+m 0 0 0 −1 αβ
T T
1 1 0 0
0 0 1 1
2
= η +
pz −pz p− −p+
E+m E+m E+m E+m
p+ −p− −pz pz
E+m E+m E+m E+m
αβ
pz p−
− E+m − E+m
1 0
0 0 0 0
= (E + m)
pz 2 pz p−
pz
E+m 0 − (E+m)2 − (E+m) 2
p+ pz p+ p+ p−
E+m 0 − (E+m) 2 − (E+m) 2
0 0 0 0
p+ pz
− E+m
0 1 E+m
+
p− p+ p− pz p−
0 E+m − (E+m)
2 (E+m)2
pz p+ pz 2
pz
0 − E+m (E+m)2
− (E+m)2
E+m 0 −pz −p−
0 E+m −p+ pz
=
pz p− −E + m 0
p+ −pz 0 −E + m
5. γ Algebra
109
Prove the following identities
γ µ γµ = 4
γ µ γ ρ γµ = −2γ ρ
Tr(γ µ γ ν ) = 4g µν
Tr(γ µ γ ν γ ρ γ σ ) = 4 (g µν g ρσ + g µσ g νρ − g µρ g νσ )
Solution
110
6. Dirac Hamiltonian from Creation and Annihilation Opera-
tors
Show that for the Dirac field indeed the Hamiltonian (not! normal-
ordered) is given by
2 Z
d3 p
Z h i
† ˆi (p)dˆ† (p)
X
3
Ĥ = d xĤ = p 0 b̂i (p)b̂i (p) − d i
(2π)3 2p0
i=1
Solution
where multiplying from the right with γ 0 for the latter gives (remember
(γ 0 )2 = 1)
2
d3 p X h −ipx (i)
Z i
ψ̄ = e v̄ (p) ˆi (p) + eipx ū(i) (p)b̂† (p) ,
d i
(2π)3 2p0
i=1
and therefore
←
→
Z
Ĥ = d3 xψ̄(−iγ · ∇ + m)ψ
We use a lot of spinor identies, such as ū = u† γ 0 etc. in the following and try to make
11
/ − m)u(p) = 0
(p ←→ (p · γ + m)u(p) = Eγ 0 u(p)
(p
/ + m)v(p) = 0 ←→ (p · γ − m)u(p) = Eγ 0 u(p)
111
2
d3 p d3 q
Z X
3
= d x
(2π)3 2p0 (2π)3 2q0
i,j=1
h i ←
→
e−ipx v̄ (i) (p)dˆi (p) + eipx ū(i) (p)b̂†i (p) (−iγ · ∇ + m)
h i
−iqx (j) iqx (j) ˆ†
e u (q)b̂j (q) + e v (q)dj (q)
2
d3 p d3 q
Z
1 3
X
= d x
2 (2π)3 2p0 (2π)3 2q0
i,j=1
h i
e−ipx v̄ (i) (p)dˆi (p) + eipx ū(i) (p)b̂†i (p)
h i
× (γ · q + m)e−iqx u(j) (q)b̂j (q) + (−γ · q + m)eiqx v (j) (q)dˆ†j (q)
h i
− e−ipx v̄ (i) (p)dˆi (p)(−γ · p + m) + eipx ū(i) (p)b̂†i (p)(γ · p + m)
h i
× e−ipx u(j) (p)b̂j (p) + eipx v (j) (p)dˆ†j (p)
2
d3 p d3 q
Z
1 3
X
= d x
2 (2π)3 2p0 (2π)3 2q0
i,j=1
h i
e−ipx v̄ (i) (p)dˆi (p) + eipx ū(i) (p)b̂†i (p) Eq γ 0
h i
e−iqx u(j) (q)b̂j (q) − eiqx v (j) (q)dˆ†j (q)
h i
− e−ipx v̄ (i) (p)dˆi (p) − eipx ū(i) (p)b̂†i (p) Ep γ 0
h i
−iqx (j) iqx (j) ˆ†
e u (q)b̂j (q) + e v (q)dj (q)
2
d3 p d3 q
Z
1 X
= d3 x
2 (2π)3 2p0 (2π)3 2q0
i,j=1
h
−i(p+q)x †(i)
Eq e v (p)u (q)dˆi (p)b̂j (q)
(j)
112
+ e−i(p−q)x v †(i) (p)v (j) (q)dˆi (p)dˆ†j (q)
io
−ei(p+q)x u†(i)(p)v (j) (q)b̂†i (q)dˆ†j (q)
2
d3 p d3 q
Z
1 X
=
2 (2π)3 2p0 (2π)3 2q0
i,j=1
h
Eq (2π)3 δ 3 (p + q)v †(i) (p)u(j) (q)dˆi (p)b̂j (q)
i
+u†(i) (p)u(j) (p)b̂†i (p)b̂j (p) − v †(i) (p)v (j) (p)dˆi (p)dˆ†j (p)
h
−p0 v †(i) (p)u(j) (−p)dˆi (p)b̂j (−p) − u†(i)(p)v (j) (−p)b̂†i (q)dˆ†j (−p)
i
†(i) † †(i) †
−u (p)u (p)b̂i (p)b̂j (p) + v (p)v (p)dˆi (p)dˆj (p)
(j) (j)
2
d3 p
Z
1
u†(i) (p)u(j) (p)b̂†i (p)b̂j (p)
X
=
2 (2π)3 2p0
i,j=1
†(i) (j) ˆ ˆ †
− v (p)v (p)di (p)dj (p)
2 h
d3 p
Z i
† †
X
= p0 b̂ (q)b̂i (p) − ˆ
di (p) ˆ
d (q)
i i
(2π)3 2p0
i=1
7. Gordon identities
113
(a) Prove the Gordon identities
ν
2mū(p1 )γµ u(p2 ) = ū(p1 ) (p1 + p2 )µ + iσµν (p1 − p2 ) u(p2 )
ν
2mv̄(p1 )γµ v(p2 ) = −v̄(p1 ) (p1 + p2 )µ + iσµν (p1 − p2 ) u(p2 ),
where
i
σµν = [γµ , γν ]
2
(b) Prove that
Solution
114
which proves the Gordon identity for the u-spinors. For the ex-
pressions v-spinors the proof is completely analogous, the only
difference is the sign in front of the terms proportional to mass.
(b)
ν
ū(p1 ) σµν (p1 + p2 ) u(p2 )
i
= ū(p1 ) γµ (p/1 + p/2 ) − (p/1 + p/2 )γµ u(p2 )
2
i
= imū(p1 )u(p2 ) + ū(p1 ) γµ p/1 − p/2 γµ u(p2 )
2
= imū(p1 )u(p2 )
i ν ν
+ ū(p1 ) 2gµν p1 − p/1 γµ − 2gµν p2 + γµ p/2 u(p2 )
2
= iū(p1 )(p1 − p2 )ν u(p2 ) .
(c) To evaluate the current let us take a look at the argument first
(and keep in mind that we can replace ū(p2 )p/2 → ū(p2 )m and
/1 u(p1 ) → mu(p1 ).
p
p /2 = 2pν1 gµν − γµ p/1 p/2
/ 1 γµ p
= 2p1µ p/2 − γµ 2p1 · p2 − p/2 p/1
= 2 p1µ p/2 + p2µ p/1 − p1 · p2 γµ − p/2 γµ p/1
Jµ = ū(p2 )p
/1 γµ p
/2 u(p1 )
= ū(p2 ) 2 p1µ p/2 + p2µ p/1 − p1 · p2 γµ − p/2 γµ p/1 u(p1 )
2
= ū(p2 ) 2m(p1µ + p2µ ) − (2p1 · p2 + m )γµ u(p1 )
ν 2 2
= ū(p2 ) − 2miσµν q + (q + m )γµ u(p1 )
2 2 ν
= ū(p2 ) F1 (m, q )γµ + F1 (m, q )σµν q u(p1 )
where we have used the Gordon identity from part (a) in the last
step. Comparing coefficients we arrive at
F1 = (q 2 + m2 ) and F2 = −2im .
115
8. ∗ Dealing with γ5
We introduce γ5 , given by
i 0 1
5 0 1 2 3
γ5 = γ = iγ γ γ γ = − µνρσ γ µ γ ν γ ρ γ σ =
4! 1 0
γ µ† = γ 0 γ µ γ 0 .
γ5† = γ5
and that
{γ µ , γ5 } = 0 .
(b) Show that
exp[−iθγ5 ] = cos θ + iγ5 sin θ .
(c) Analyse the behaviour of the free Dirac field Lagrangian under
chiral phase transformations given by
ψ → ψ 0 = exp [iθγ5 ] ψ
h i
ψ † → ψ 0† = ψ † exp −iθγ5† ,
Solution
γ 0† = γ 0 γ 0 γ 0 = γ0
1 0 0 +σi 1 0
γ i† = 0 i 0
γ γγ =
0 −1 −σi 0 0 −1
†
0 −σi 0 σi
= = ,
+σi 0 −σi 0
because σi† = σi .
†
γ5† = iγ 0 γ 1 γ 2 γ 3 = −iγ 3† γ 2† γ 1† γ 0†
116
= −iγ 0 γ 3 γ 0 γ 0 γ 2 γ 0 γ 0 γ 1 γ 0 γ 0 γ 0 γ 0 = −iγ 0 γ 3 γ 2 γ 1
= iγ 0 γ 3 γ 1 γ 2 = −iγ 0 γ 1 γ 3 γ 2 = iγ 0 γ 1 γ 2 γ 3 = γ5 ,
{γ µ , γ5 } = i γ µ γ 0 γ 1 γ 2 γ 3 + γ 0 γ 1 γ 2 γ 3 γ µ
0γ0γ1γ2γ3 + γ0γ1γ2γ3γ0
µ = 0 : i γ
1γ2γ3 − γ0γ0γ1γ2γ3 = 0
= γ
1 0 1 2 3 0 1 2 3 1
µ = 1 : i γ γ γ γ γ + γ γ γ γ γ
0 2 3 0 2 3
=γ γ γ −γ γ γ =0
= =0
µ = 2 : i γ2γ0γ1γ2γ3 + γ0γ1γ2γ3γ2
= −γ 0 γ 1 γ 3 + γ 0 γ 1 γ 3 = 0
3 0 1
µ=3 : i γ γ γ γ γ +γ γ γ γ γ 2 3 0 1 2 3 3
= γ0γ2γ3 − γ0γ1γ2 = 0
(b) To see how this works, remember that fnctions with matrices as
arguments can be defined by their Taylor series, and therefore
∞
X
exp[iθγ5 ] = (iθγ5 )k .
k=0
and we see that even powers of γ5 are the unit matrix, while odd
powers yield the γ5 . Thus
ψ̄ → ψ̄ 0 = ψ † exp[iγ5† θ]† γ0 .
117
←→
Ignoring for a moment the ∂ notation, the Lagrangian trans-
forms as
L → L0 = ψ̄ 0 i∂/ − m ψ 0 = ψ † exp[iγ5† θ]† γ0 i∂/ − m exp[iθγ5 ]ψ
h i h i
= ψ † cos θ − iγ5† sin θ γ0 i∂/ − m cos θ + iγ5 sin θγ5 ψ
= ψ † (cos2 θ)γ0 (i∂/) + (cos θ sin θ) γ5† γ 0 ∂/ − γ 0 ∂/γ5
i
+(sin2 θ)γ5† γ 0 (i∂/)γ5 ψ
† 2 0 † 0
−mψ γ0 cos θ + i cos θ sin θ γ γ5 − γ5 γ
+γ5† γ0 γ5 sin2 θ ψ
† 2 0 0
= ψ (cos θ)γ0 (i∂/) + cos θ sin θ γ5 γ ∂/ − γ ∂/γ5
2
+ sin θγ5 γ0 (i∂/)γ5 ψ
† 2 0 0
−mψ γ0 cos θ + i cos θ sin θ γ γ5 − γ5 γ
2
+ γ5 γ0 γ5 sin θ ψ
† 2 2
= ψ γ0 cos θ + sin θ (i∂/) ψ
† 2 2
−ψ γ0 cos θ − sin θ + γ0 iγ5 cos θ sin θ mψ
= iψ̄∂/ψ − mψ̄ cos(2θ) + iγ5 sin(2θ) ψ
118
(b) Prove that the spin operator indeed satisfies the spin/angular
momentum algebra (a SU (2) algebra), given by
[Ŝ i , Ŝ j ] = iijk Ŝ k .
Solution
119
1 ikl jnl
= − γk γn − γn γk
4
1 ij kn in jk
= − g g −g g γk γn − γn γk
4
1 ij j i ij i j 1 i j j i
= − −3g − γ γ + 3g + γ γ = − γ γ −γ γ
4 4
2 1 ijk ilm
Ŝ = γj γk γl γm
16
1 jl km 1
= g g − g jm g kl γj γk γl γm = γj γk γ j γ k − γj γk γ k γ j
16 16
1 h j k j
i 1 h i
= (2gjk − γk γj )γ γ + 3γj γ = −2γk γ k + 3γk γ k + 3γj γ j
16 16
12 3
= − =− .
16 4
120
6 Electrodynamics
In this section we will quantise electrodynamics, by quantising the free vec-
tor potential that gives rise to the (free) electromagnetic fields. It turns
out that this results in a somewhat more involved procedure; while the vec-
tor potential has four components, which we would naively treat as four
independent quantities – scalar fields – and quantise them accordingly, the
gauge invariance of the fields implies that in fact there are only two physically
meaningful degrees of freedom. This means that, naively exercised, our algo-
rithm of second quantisation would lead to a degree of “over-quantisation”,
i.e. trying to quantise objects that cannot and should not be quantised in a
consistent and physically meaningful way. The solution to this is to fix the
gauge before quantising the fields, which is nothing but the imposition of
additional external conditions.
The quantisation of the four potential is, as indicaed, a somewhat tricky
business. In my opinion, the best explannations of the procedure can be
found in Chapter 5 of Hatfield’s book [3], and in Section 3.2 of Itzykson &
Zuber [12].
E2 − B2 1
L = = − F µν Fµν ,
2 4
where we have set the current to zero, j µ = 0 and moved a factor of 4π into
the vanishing j µ Aµ term in the first expression. It is simple to show that
under the gauge transformations of Eq. (120) ,
Aµ → A0µ = Aµ − ∂ µ Λ ,
Reminding ourselves of the connection of the field strength tensor with the
electric and magnetic fields E and B, Eq. (122), invariance of the fields
under gauge transformations is manifest.
This has two implications, which are worth making explicit: First of all, al-
though we will explicitly quantise the vector potential Aµ and only indirectly,
through it, the fields, the latter are the physical quantities, measurable in
121
every day life 12 . Secondly, and in the context of what follows more impor-
tantly, we may use special forms of the gauge transformation, Eq. (120), to
eliminate some components of Aµ without impacting on the physics. But
this also implies that there are less than four physically meaningful degrees
of freedom encoded in the vector potential, and we will have to deal with
the problem of how to quantise a system that has less physical degrees of
freedom than the field that is used for its description.
Fixing the Gauge Let us discuss now some of the conditions that can
be imposed on Aµ , which effectively fix the gauge. Looking at the form
of the field strength tensor it is worth noting that F 00 = 0, which implies
that there is no conjugate momentum for the temporal component of Aµ .
Defining them, as before, through
∂L
πµ = (237)
∂ Ȧµ
and specialising on µ = 0 yields
∂L
π0 = = 0. (238)
∂ Ȧ0
This motivates us to use a temporal gauge defined by
Zt
Λ(t, x) = dt0 A0 (t0 , x) (239)
−∞
Coulomb vs. Lorentz vs. Axial Gauge It turns out, however, that
this does not yet entirely fix the gauge and an additional condition has to
be applied. Three types of gauge, with different calculational advantages
and disadvantages in different situations are frequently found:
Coulomb gauge, defined through
∇ · A = 0. (240)
∂µ Aµ = 0 . (241)
Az = 0 . (242)
12
The impact of a finite vector potential in regions where the fields vanish is subject of
the Aharonov-Bohm effect, which is discussed, for example, in Chapter 2.6 of Sakurai’s
book [11].
122
Polarisation Vectors and Degrees of Freedom To build on this idea
of gauge fixing, let us analyse in some more detail what this actually implies.
Most transparently this can be done in Coulomb gauge. For free fields, i.e.
with j = 0 and, in particular the charge density ρ = j 0 = 0, A0 is not
a dynamical degree of freedom: its derivative w.r.t. time is not present
in the free field Lagrangian and hence its conjugate momentum vanishes.
The temporal part of the gauge in Eq. (239) fixes this constant then to
A0 = 0, making the lack of dynamical relevance explicit. This leaves only
the spatial components of A, A, and the field strength tensor is composed
of the components of ∇ × A, as F ij = ∂ i Aj − ∂ j Ai .
But imposing the Coulomb gauge condition by demanding that the diver-
gence of A vanishes, ∇ · A = 0 we exposed that there is a longitudinal com-
ponent of A, AL . By definition of it being longitudinal, ∇ × AL = 013 . This
implies that yet another component of F vanishes, or, differently put, we
see that also AL is not dynamically relevant. This shows that the Coulomb
gauge is the one where the longitudinal degree of freedom vanishes, AL = 0.
Not surprisingly, imposing two conditions on the four-vector Aµ eliminates
two of its components, and we are left with two degrees of freedom. The
logic above, eliminating the temporal and longitudinal parts of A from the
dynamical degrees of freedom means that we are left with two transverse
degrees of freedom AT .
To make the physics of this more explicit, let us see how this works out in
practice. Assume we want to describe a quantum of electromagnetism, a
photon, with momentum k. It’s four-momentum of course is given by
q
k µ = (ω, k) with ω = k0 = k 2 . (243)
The relevant degrees of freedom for the photon are its two remaining po-
larisations. They are usually denoted by λ = {1, 2} and represented by
polarisation vectors µ(λ) (k). Fourier transformation of the gauge conditions
above then become conditions on products of the three-momentum and the
polarisation three-vector; while the temporal gauge condition implies 0 = 0
we have:
Demanding additionally that the polarisation vectors are real and ortho-
normal we have
123
A simple way to guarantee this is to orient k along the z-axis. Then
∂L
0
π = = 0
ν ∂L ∂ Ȧ0
π = =⇒ (247)
∂ Ȧν πi
∂L
= = −Ei .
∂ Ȧi
This makes the anticipated problem of vanishing conjugate momen-
tum for A0 manifest. Further down the line it will prevent us from
quantising it, because we will not be able to produce a non-vanishing
commutator between this field component and its conjugate momen-
tum: for our choice of Lagrangian, it is guaranteed that [A0 , π 0 ] = 0
irrespective of what we try to do and therefore quantisation of A0 is
bound to fail.
E2 + B2
H = Ȧµ πµ − L = + E · ∇A0 , (248)
2
where the last term obviously vanishes if we set A0 = 0.
124
Dealing with A0 : Gauss’ law To re-iterate: the fact that π 0 = 0 means
that also the field operator vanishes and hence commutes with every field
operator. Therefore A0 is not a dynamical variable, and Ȧ0 = 0. This
means that A0 is not an operator but an ultimately inconsequential number
in our construction of a quantum field theory. However, there is a direct
consequence of it being not dynamical:
∂L
= 0 −→ ∇ · E = 0, (250)
∂A0
Gauss’ law in the absence of sources14
We would of course be tempted to implement this as a wonderfully physical
constraint on the field operators. But this would lead to yet another way to
see the problem with the procedure. Going back to the commutation rela-
tions, and forming a divergence we would arrive at, somewhat schematically,
X ∂ h i
 i (t, x), Êj (t, y)
∂y j
j
h i X ∂
= Âi (t, x), ∇ · Ê j (t, y) = −i δij j δ 3 (x − y) . (251)
∂y
j
This is difficult, because while the left hand side of the second line vanishes,
due to Gauss’ law, the right hand side doesn’t. This implies that we cannot
implement Gauss’ law as an operator equation.
∇ · Ê |ψi = 0 (252)
and would classify all states that do not fulfil this criterion as unphysical
and ignore them. It is a bit cumbersome to show that this doesn’t work
either and in fact would also violate the commutation relations.
The next weaker constraint, however, works. Demanding that for physical
states Gauss’ law is satisfied as expectation value,
125
encapsulates this part of Maxwell’s equation as an average. We will come
back to its implications at a somewhat later state.
d3 k ik·(x−y)
Z
3 tr ki kj
δij δ (x − y) −→ δij (x − y) = e δij − 2 .
(2π)3 k
(254)
It is easy to show that the gradient of the transverse δ-function with respect
to x or y vanishes, because derivatives will produce a term ±ki multiplying
the rounded bracket, and
X kj k 2
ki kj
ki δij − 2 = kj − 2 = 0. (256)
i
k k
This means that, with the modified commutator relation, ∇ · E now com-
mutes with every meaningful operator, and in particular
h i
Âi (t, x), ∇ · Ê(t, y) = 0 . (257)
126
1. Â is a gauge-dependent quantity and therefore essentially unphysical.
It cannot directly be measured, and therefore, any potentially harmful
a-causal behaviour may not have physical implications.
2. careful calculations reveals that while [Âi , Êj ] may not vanish for
space-like distances, the commutators of the physical E and B fields
and their components do vanish, irrespective of the use of the trans-
verse δ function.
0 0
while for circular polarisations we could write
0
1 1
µ(λ=1,2) (k) = √
. (262)
2 ±i
0
Using four-vectors instead of three vectors means that we replace the transver-
sality condition with k µ · µ = 0, keeping the ortho-normality condition of
Eq. (245).
127
6.3 Lorentz Gauge
Modifying the Lagrangian One of the issues with the Coulomb gauge
is that it is not Lorentz-invariant. To achieve this invariance, though, we
will need to demand commutator relations that fully reflect this symmetry,
h i
µ (t, x), π̂ν (t, y) = igµν δ 3 (x − y) . (263)
This implies, obviously, that all four components of the vector potential have
a conjugate momentum, and, in particular, that π 0 doesn’t vanish. Since
π 0 emerges by differentiation of the Lagrangian w.r.t Ȧ0 , we must modify
the Lagrangian such that this derivative does not vanish any more. This is
achieved by modifying the free-field Lagrange density,
1 1 α
L = − F µν Fµν −→ L = − F µν Fµν − (∂µ Aµ )2 . (264)
4 4 2
128
commutation relations we would like to use, namely the ones in Eq. (263).
We also cannot impose the constraint ∂ · A = 0 as an operator equation,
because this would imply that π 0 = 0, and we would not be able to recover
our postulated commutator relations.
This means that we are forced down an avenue that we briefly considered in
the case of the Coulomb gauge, by demanding that we implement the gauge
condition as a condition on physical states. We realise very quickly that it
cannot be realised as a condition on physical states |ψi,
∂ · Â |ψi = 0 , (267)
for the following reason. Consider the expectation value of the commutator
relation Eq. (263), specified for µ = ν = 0:
h i
hψ| Â0 (t, x), π̂ 0 (t, y) |ψi = iδ 3 (x − y) hψ|ψi . (268)
enforces that the l.h.s. of Eq. (268) must vanish, while the r.h.s. does not.
This rules out the weaker constraint as a viable, consistent option.
This leaves us the only option to encode the gauge condition by demand-
ing that it holds only true for expectation values of physical states, i.e.,
demanding that
We will use this after we defined polarisation vectors and expanded the field
operators in plane waves and creation and annihilation operators.
129
where we have chosen four linearly independent polarisation vectors as
1 0 0 0
0 1 (2) 0 (3) 0
(0) (k) =
(1)
, (k) = , (k) = , (k) = .
0 0 1 0
0 0 0 1
(274)
For simplicity we assumed that the photon momentum is oriented along the
positive z-axis, k k ez It is easy to check that the polarisation vectors satisfy
where the differnence between labels (λ) for the polarisation vectors and
their components - the Lorentz index µ has been made explicit. A simple
calculation shows that the commutators of Eq. (263) are satisfied, if the
only non-vanishing commutator of the creation and annihilation operators
is given by
h i
âλ (k), â†κ (q) = −(2π 3 )(2k0 )gλκ δ 3 (k − q) . (276)
Physical States So, let us now take a closer look at some of the states
and their energies. Start with the by now familiar assertion that the vacuum
reduces to zero when the one of the annihilation operators is applied,
130
Now, let us analyse one of the more tricky states: a scalar photon, modulated
by some well-behaved function f (k),
d3 k
Z
|1S i = f (k)â†0 (k) |0i . (279)
(2π)3 2k0
d3 k d3 k 0
Z Z
h1S |1S i = f (k)f ∗ (k 0 )h0|â(k 0 , 0)↠(k, 0)|0i
(2π)3 2k0 (2π)3 2k0
d3 k
Z
= − h0|0i |f (k)|2 < 0 . (280)
(2π)3 2k0
The minus sign of course stems from the relative sign in the metric, or,
when followed through, from the “-”-sign in front of the right-hand side
of the commutator in Eq. (276). Phrased differently, the combinations of
positive and negative energy solutions that are still allowed destroys the
positive definiteness of the norm.
So let us impose the gauge constraint ∂ · Â+ |ψi = 0. Evaluating ∂ · Â+ we
of course only take into account the positive energy solutions and arrive at
3 h
d3 k
Z X i
µ (λ) −ik·x
∂ · Â+ = −i k µ (k)â λ (k) e . (281)
(2π)3 2k0
λ=0
We can simplify this further by realising that for λ = {1, 2} the polarisations
are orthogonal to the momentum, ⊥ k and therefore k · = 0. This leaves
us with two surviving polarisations, scalar (λ = 0) and longitudinal (λ = 3).
Our constraint on physical states |ψi therefore becomes
0 = ∂ · Â+ |ψi
d3 k
Z h i
= −i k · (0)
(k)â0 (k) − k · (3)
(k)â3 (k) e−ik·x . (282)
(2π)3 2k0
131
6.4 Problems & Solutions
1. Polarisation vectors in Coulomb gauge
Assume a momentum parallel k to the z-axis √ and introduce left- and
right-circular polarisations (L),(R) (k) = 1/ 2(0, 1, ±i, 0). Show by
explicit calculation that for the spatial components of the polarisation
vectors
X (λ) ∗(λ) kj ki
i (k)j (k) = δij − 2
λ=1,2
k
Solution
0 0
132
Expert level: To write this in more convenient form, introduce a time-
like unit-length four-vector η (for simplicity we can assume η = (1, 0, 0, 0)),
and
k µ − η µ (k · η)
µ 1 0
k̃ = p −→ .
(k · η)2 − k 2 |k| k
Thus, for our choice of η, k̃ just becomes the direction of the three-
momentum of k, with no temporal component.
Expressed through these vectors,
2
X
(λ) (λ)
µ (k)ν (k) = −gµν + ηµ ην − k̃µ k̃ν .
λ=1
and
tr
[Ai (x, t), Ej (y, t)] = −iδij (x − y)
and show with explicit calculation that the modification does not affect
the physically observable fields.
Solution
(a) The commutator of the electric fields is trivial; because Êi = −π̂i ,
h i
Êi (x, t), Êj (y, t) = π̂i (x, t), π̂j (y, t) = 0
133
d3 k ik(x−y)
Z
(y) (y)
= iijk ∂k δ 3 (x − y) = iijk ∂k e
(2π)3
d3 k
Z
= −ijk kk eik(x−y)
(2π)3
while for the “trans.” case
h i h i
Êi (x, t), B̂j (y, t) = − π̂i (x, t), (∇(y) × Â(y, t))j
h i
(y) (y) (tr)
= −jkl ∂k π̂i (x, t), Âl (y, t) = ijkl ∂k δil (x − y)
d3 k
Z
ki kl (y)
= ijkl 3
δil − 2 ∂k eik(x−y)
(2π) k
3
Z
d k ki kl kk
= −jkl δ k
il k − eik(x−y)
(2π)3 k2
d3 k
Z
= −ijk kk eik(x−y) ,
(2π)3
identical to the “reg.” case, as advertised. In going from the
second-to-last to the last line we have used that the product of
a symmetric combination (kj kk ) with the anti-symmetric Levi-
Civita tensor (the jkl ) vanishes.
(c) Let us finally calculate the commutator of the magnetic fields.
h i
B̂i (x, t), B̂j (y, t)
h i
= − (∇(y) × Â(x, t))i (∇(y) × Â(y, t))j
h i
(x) (y)
= ikl imn ∂k ∂m Âl (x, t), Ân (y, t) = 0
3. Momentum Operator P̂ µ
The four-momentum operator is given by
P̂ µ = (Ĥ, P̂ ),
where
Z
1 h i
Ĥ = − d3 x π̂ µ π̂µ + ∇µ ∇µ
2
Z h i
P̂ i = − d3 x ∂t µ ∂ i µ
134
To show that this represents indeed the right structure, show that
[P̂ µ , Aν ] = −i∂ µ Aν .
Solution
Z
1 h i
Ĥ = − d3 x π̂ µ π̂µ + ∇µ ∇µ
2
Z h i
P̂ i = − d3 x ∂t µ ∂ i µ
Let us start with the commutator of the Hamiltonian and the vector
potential,
Z
ν 1 h i
[Ĥ,  ] = − d3 y π̂ µ (y)π̂µ (y) + ∇µ (y)∇µ (y), Âν (x)
2
Z
1 n h i h i o
= − d3 y π̂ µ (y) π̂µ (y), Âν (x) + π̂ µ (y), Âν (x) π̂µ (y)
2
Z
i
d3 yδ(x − y) π̂ µ (y)gµν + g µν π̂µ (y) = iπ̂ ν (x) = −i∂ 0 Âν (x) ,
=
2
and turn now to the ith component of the momentum operator. To
use this let us remind ourselves that
π µ = F µ0 − αg µ0 (∂ · A) = ∂ µ A0 − ∂ 0 Aµ − αg µ0 (∂ · A)
This suggests that, in the commutator, we can replace ∂0 Aν with −π ν ,
and we find
Z h i
[P̂ i , Âν ] = − d3 y (∂y0 µ (y))(∂ i µ (y)), Âν (x)
Z h i
= + d3 y π̂ µ (y), Âν (x) ∂ i µ (y)
Z
µν
= −ig d3 yδ 3 (x − y)∂ i µ (y) = −i∂ i Aν (x)
135
(a) Find a general solution for the vector potential inside the capac-
itor made by the two plates, ignore the effect of the limited size
L.
(b) Quantise the electromagentic field.
(c) Find the Hamiltonian and show that the vacuum energy is given
by
" ∞ r #
L2
Z
d2 k X nπ 2 r
E= 2 2 2
k1 + k2 + 2 2
+ k1 + k2 .
2 (2π)2 a
n=1
f (k) = Θ(Λ − k)
Solution
Keeping in mind that the vector potential satisfies the wave equa-
tion
∂t2 − ∇2 A = 0 ,
136
we assume a solution factorising into the form
with unit polarisation vectors e1,2,3 , where 1,2 live in the xy-
plane and 3 = (0, 0, 1). pointing into the x, y, and z-direction.
We therefore arrive at
Zi = ai sin(k3 z) + bi cos(k3 z)
137
aλ (k1 , k2 , n)e−iωk,n t+ik1 x+ik2 y
(λ) nπz (λ) nπz
× k (k, n) sin + 3 (k, n) cos
a a
a†λ (k1 , k2 , n)eiωk,n t−ik1 x−ik2 y
∗(λ) nπz (λ) nπz
× k (k, n) sin + 3 (k, n) cos
a a
d2 k
Z
+ a3 (k1 , k2 , 0)e−iωk,n t+ik1 x+ik2 y
(2π)2 2ωk,0
† iωk,n t−ik1 x−ik2 y
+ a3 (k1 , k2 , 0)e 3 .
and this difference stems from the fact that the momentum in
z-direction only takes discrete values, encoded in the n and n0 .
(c) As before the (not! normal-ordered, therefore a factor of 1/2
in front) Hamiltonian is given by a sum over all modes, and we
arrive at
(∞ Z 2
1 X dk1 dk2 X †
Ĥ = âλ (k, k2 , n)âλ (k, k2 , n)
2 (2π)2 2ωk, n
n=1 λ=1
†
+ âλ (k, k2 , n)âλ (k, k2 , n)
Z
dk1 dk2
+ ↠(k, k2 , 0)â3 (k, k2 , 0)
(2π)2 2ωk, 0 3
†
+ â3 (k, k2 , 0)â3 (k, k2 , 0)
138
For the calculation of the ground-state energy we realise that â |0i
vanishes and that we use the commutator for terms like h0| â↠|0i.
We arrive at
E = h0| Ĥ |0i
* (∞ Z 2
1 X dk1 dk2 †
X
= 0 ωk, n âλ (k, k2 , n), âλ (k, k2 , n)
2 (2π)2 2ωk, n
n=1 λ=1
Z
dk1 dk2 †
+ ωk, 0 â3 (k , k2 , 0)â3 (k, k 2 , 0) 0
(2π)2 2ωk, 0
* (∞ Z 2
1 X dk1 dk2 X
= 0 ω (2π)2 2ωk, n δ 2 (0)
k, n
2 (2π)2 2ωk, n
n=1 λ=1
Z
dk1 dk2 2 2
+ 2
ωk, 0 (2π) 2ωk, 0 δ (0) 0
(2π) 2ωk, 0
Z " ∞ #
1 2 2
X
= d kδ (0) 2 ωk, n + ωk,0
2
n=1
L2
Z
dxdy ik1 x+ik2 y
δ 2 (0) = e = ,
(2π)2 (2π)2
kk =0
139
The normalised difference reads
E − E0
∆ε =
L2
∞ r
" r
d2 k 1
Z
2 + k 2 + nπ
X 2
= k 2 + k 2 + k
1 2 1 2
(2π)2 2 a
n=1
Z∞ r
nπ
− dn k12 + k22 +
a
0
∞ Z∞ p
2 ∞ p
1√
Z
π X
= 3
du u+ u + n2 − dn u + n2 ,
4a 2
0 n=1 0
Z∞ p √
π u
− dn u + n2 f
a
0
Z∞
∞
π2 1 X
= F (0) + F (n) − 2 F (n) ,
4a3 2
n=1 0
where we have introduced
Z∞ p √
2
π u
F (n) = du u + n f .
a
0
140
d3 F (n)
000
F (0) = = −4
dn3 n=0
π2 π2
4
∆ε ≈ − 3 =−
4a 30 ∗ 4! 720a3
∂∆ε π3
f =− = ,
∂a 240a4
which for a = 1µm and L = 1 cm is approximately f ≈ 10−8 N.
This is the Casimir effect, measured for the first time in 1958.
17
17
It states that the difference of a sum and an integral of the same function can be
expressed by the Bernoulli numbers bk and (multiple) derivatives of the function as
m Zm
X X Bk h (k−1) i
f (n) − dxf (x) = ∞ f (m) − f (k−1) (n) .
n
k!
n k=1
141
7 Time-Ordered Products
Until now we have quantised various free elementary fields: real and complex
scalars, Dirac-spinors, and the vector fields of electrodynamics. The result-
ing structure in each case can be condensed into a sequence of algorithmic
steps, which, starting from a Lagrange density, resulted in the expansion
of field operators as products of plane waves and creation and annihilation
operators, and we succeeded in expressing “static” global quantities such as
the Hamilton or charge operators through the latter.
This implies an underlying causal structure if the theory: in non-relativistic
field theories, which we do not discuss here, evolution is forward in time,
and for the analysis of causality it is usually sufficient to concentrate on
the positive–energy solutions only. This changes in relativistic field the-
ories, where both forward and backward evolution, and therefore positive
and negative energy solutions, are included. It is important to realise the
interplay with causality requirements of the theory - the simplest one is that
the commutator of two fields must vanish for space-like distances. It turns
out that in non-relativistic theory this cannot be achieved, which actually
should not come as a surprise. If you do not embed relativity in your formal-
ism you cannot expect to obtain relativisticly sensible results from it. In the
relativistic field theories we have discussed here, the positive and negative
energy solutions could be arranged such that the commutator of two fields
vanishes outside the light-cone, i.e. for space-like distances, but remains
finite inside the light-cone.
In this chapter we will build further on the logic and discussion started in
Sec. 4.3, and we will analyse the propagation of particles. This first step
towards a dynamic picture of quantised field theories is deeply connected
to the notion of Green’s functions, which will return to us in this chapter,
and called propagators. They will fortify the notion of the negative-energy
solutions as anti-particles, which travel backwards in time, with opposite
charge. We will also see how the wrong commutator (or anti-commutator)
for a given statistics (Bose–Einstein vs. Fermi–Dirac) destroys the causality
structure of the theory.
In this chapter of the lecture notes I have amalgamated time-ordered prod-
ucts for the three Quantum Field Theories we discussed so far – for some
additional reading, I’d like to refer you to Sections 3.5, 4.3, and Chapter 5
of Hatfield’s book [3], or maybe take a look at Sections 2.4, 3.5, and possibly
4.8 of Peskin & Schroeder [1].
142
7.1 Greens Functions: Non-Relativistic Quantum Mechanics
What is the Green’s Function? Consider the time-dependent Schrödinger
equation for a point particle,
∇2
i∂ i∂
− Ĥ |ψ(t)i = + − V̂ |ψ(t)i = 0 . (284)
∂t ∂t 2m
It is formally solved through the introduction of Green’s function, G(t, x; t0 , x0 ):
Z
ψ(t, x) = hx|ψ(t)i = d3 x0 G(t, x; t0 , x0 )ψ(t0 , x0 ) . (285)
The interpretation is clear: the wave function ψ(t, x) at time t and position
x in position space is constructed as the superposition of all wave functions
at all positions x0 at an earlier time t0 , and the Green’s function parame-
terises the “strength” of the connection. Because it this has been couched in
the framework of non-relativistic Quantum Mechanics the maximal velocity
of causation (speed of light in relativistic physics) is infinite, and the connec-
tion is instantaneous. The Green’s function G is also called the (retarded)
propagator.
with the boundary condition that it vanishes for t0 > t. This allows to
rewrite it as
and
Zt
h i
Û (t, t0 ) = exp −i dτ Ĥ(τ ) −→ exp −iĤ(t − t0 ) (289)
t0
143
Free-Particle Propagator: Direct Solution in Momentum Space
As a simple example consider a free point particle in Quantum Mechanics.
Its propagator (Green’s function) G0 is a solution to
i∂
− Ĥ0 G0 (t, x; t0 , x0 ) =
∂t
∇2
i∂
+ G0 (t, x; t0 , x0 ) = δ(t − t0 )δ 3 (x − x0 ) . (290)
∂t 2m
A simple way to solve this equation is by Fourier-transforming on, resulting
in
!
p2
ω− G0 (ω, p) = 1 (291)
2m
and therefore
1
G0 (ω, p) = . (292)
ω − p2 /2m
d3 p dω exp[ip · (x − x0 ) − iω(t − t0 )]
Z
0 0
G0 (t, x; t , x ) = . (293)
(2π)3 2π ω − p2 /(2m)
d3 p dω exp[ip · (x − x0 ) − iω(t − t0 )]
Z
(R) 0 0
G0 (t, x; t , x ) = . (294)
(2π)3 2π ω − p2 /(2m) − i+
144
Therefore the overall result is given by
" #
d3 p ip2 (t − t0 )
Z
(R) 0 0 0 0
G0 (t, x; t , x ) = Θ(t − t ) exp − + ip · (x − x )
(2π)3 2m
3 Z
ip2i (t − t0 )
Y dpi 0 0
= Θ(t − t ) exp − + ipi (xi − xi )
(2π)3 2m
i=1
3 Z
"
i(t − t0 ) m(xi − x0i ) 2
0
Y dp i
= Θ(t − t ) exp − pi −
2π 2m t − t0
i=1
im(xi − x0i )2
+
2(t − t0 )
3
s
im(xi − x0i )2
0
Y 1 2mπ
= Θ(t − t ) exp
2π i(t − t0 ) 2(t − t0 )
i=1
s 3 " 3
#
0 −im im X
0 2
= Θ(t − t ) exp (xi − xi )
2π(t − t0 ) 2(t − t0 )
i=1
3
im(x − x0 )2
0 −im 2
= Θ(t − t ) exp . (296)
2π(t − t0 ) 2(t − t0 )
We have made use of the fact that we can write this integral as a product
of three integrals, one for each spatial component of p, then completed the
squares in each component of p, rendering this a product of three Gaussian
integrals.
145
In going from the first to the second line we have use the fact that momentum
and position space kets are connected through a Fourier transform,
d3 p −ip·x
Z
|xi = e p , (298)
(2π)3
2
and in going to the third line we have replaced the operator p̂ with the
eigenvalues p corresponding to its eigenkets p . Of course, they form an
ortho-normal base, such that their scalar product is a δ function,
0
p p i = (2π)3 δ 3 (p − p0 ) . (299)
Including the Θ-function which connects the transition amplitude with the
restarted propagator, this is exactly the same result we already obtained
with the more direct method in momentum space.
+ m2 G0 (x, x0 ) = iδ 4 (x − x0 ) .
(300)
−i
−p2 + m2 G0 (p) = i −→
G0 (p) = . (301)
p2 − m2
(302)
When expanding these products in terms of the creation and annihilation
operators, it is worth noting that â |0i = h0| ↠= 0 and that we therefore
146
can replace products â↠with the commutators of the two operators, when
they are sandwiched between vacuum states. This leads to
cf. Eq. (163), where in both cases the sign indicates whether the energies,
i.e. the k0 come with the correct or wrong sign for a wave the evolves
in positive or negative time direction. This means that the propagator is
composed of two components: a +-component of forward propagation, and
a − component of backward propagation of a particle with four-momentum
k.
147
0
dωk d3 k −ie−ik·(x−x )
Z
1 1
= lim +
→0+ (2π)4 2k0 k0 − ωk − i k0 + ωk − i
0
dωk d3 k −ie−ik·(x−x )
Z
= lim
→0+ (2π)4 k02 − ωk2 + i
0
dk0 d3 k −ie−ik·(x−x )
Z
= lim
→0+ (2π)4 k02 − (k 2 + m2 ) + i
d4 k −ik·(x−x0 ) −i
Z
= 4
e
(2π) k − m2 + i+
2
d4 k −ik·(x−x0 )
Z
= e ∆F (k) = ∆F (x − x0 ) . (306)
(2π)4
This is obviously the Fourier transform of our propagator from Eq. (301),
and it confirms that indeed propagators are time-ordered products. It is
also called the Feynman propagator of the theory.
In the last step we have used that p/p/ = p2 . There are a couple of things
worth noting of this propagator. As before, it exhibits a pole for on-shell
particles, where p2 = m2 , and, as before, we will cure this by shifting the
pole in the complex plane by i+ . This is in complete analogy to the case
of scalar particles. In addition we realise that the numerator, (p/ + m),
represents a matrix in Dirac space. This is not a surprise, as the propagator
connects two Dirac spinors and their components. What is structurally more
interesting is that this matrix is the completeness relation for the u-spinors
from Eq. (215), and we will see the emergence of analogous terms later
when we discuss the propagator of the photon field. However, to build more
confidence into our interpretation of the propagator we will now check if we
can recover it as a time-ordered product of two spinor fields.
148
But we must also include the opposite case of a negative-energy fermion to
go from y to x, taking into account Fermi statistics. Making time-ordering
explicit through Θ-functions we therefore arrive at
iSF (y, x)γ 0 = h0| ψ̂(y)ψ̂ † (x) |0i Θ(y0 − x0 ) − h0| ψ̂ † (x)ψ̂(y) |0i Θ(x0 − y0 ) ,
(310)
or, in a more compact form
Since ψ̂ and ψ̄ˆ are spinors, SF is a matrix in spinor space, as already antici-
pated. Let us now include the expansion of the fermion fields in place waves
and creation and annihilation operators. Making spinor indices explicit, us-
ing the fact that b̂ |0i = dˆ|0i = 0 and 0̄b̂† = h0| dˆ† , and taking into account
ˆ = [b̂† , dˆ† ] = 0, which makes their products vanish, we arrive at
that [b̂, d]
+
i
− e+ik·y−iq·x dˆj (q)dˆ†i (k)vβ (k)v̄α(j) (q)Θ(x0 − y0 ) 0
(i)
* Z 2 h
d3 k X (i)
e−ik·(y−x) uβ (k)ū(j)
= 0 3 2k α (k)Θ(y0 − x0 )
(2π) 0 i,j=1
+
i
(i)
− e+ik·(y−x) vβ (k)v̄α(j) (k)Θ(x0 − y0 ) 0
d3 k h −ik·(y−x)
Z
= 0 e (k/ + m)βα Θ(y0 − x0 )
(2π)3 2k0
+
i
− e+ik·(y−x) (k/ − m)βα Θ(x0 − y0 ) 0 . (312)
Repeating the same steps of replacing the Θ functions with integrals, suit-
ably closing the contours and keeping track of the relative signs in the re-
placement of the “dummy” energy with the real energy, we arrive at
d4 k −ik·(y−x)
Z
k/ + m
iSF (y, x)βα = e i 2 , (313)
(2π)4 k − m2 + i+ βα
149
the Fourier transform of Eq. (311), modified by the now familiar i prescrip-
tion.
150
Time-Ordered Product To arrive at the same expression using time-
ordered products we employ the Feynman gauge from the beginning, where
the completeness relation for the polarisation vectors is given by
3
X
∗(λ)
(λ)
µ (k)ν (k) = gµν (321)
λ=0
Using the by now usual â |0i = 0 relation allows to simplify the time-ordered
product and the propagator reads
d3 k h
Z i
−ik·(x−y) +ik·(x−y)
= − igµν Θ(x0 − y0 )e + Θ(y0 − x0 )e
(2π)3 2k0
= − igµν [Θ(x0 − y0 )∆+ (x − y) + Θ(y0 − x0 )∆− (x − y)]
d4 k e−ik·(x−y)
Z
= − igµν . (322)
(2π)4 k 2 + i+
This is, of course, the Fourier transform of the propagator of Eq. (320)
151
7.5 Problems & Solutions
1. Green’s function for a free particle in Quantum Mechanics
Consider, once again, the free Hamiltonian, Ĥ0 = p̂2 /(2m) with eigen-
states labelled by their momentum, |pi.
ψp (x, t) = hx|pi
d3 p
Z
G(x, t; , x0 , t0 ) = Θ(t − t0 ) ψp (x, t)ψp∗ (x0 , t0 )
(2π)3
Solution
(a) To construct the wave function add the time evolution to a set
R 3 basis kets |pi = |p(t = 0)i and insert a 1 in
of momentum-space
the form of 1 = d p|pihp|
*
Zt
+
0 0
ψp (x, t) = hx|p(t)i = x exp −i dt Ĥ(t ) p(0)
0
* " # + * " # +
p̂2
= x exp −iEp t p(0) = x exp −i t p(0)
2m
152
* +* " # +
2
d3 p0 p̂
Z
0
= x p p0 exp −i t p(0)
(2π)3 2m
d3 p0 ix·p0 −ip2 t/(2m)
Z
2
= e e (2π)3 δ 3 (p0 − p) = eix·p−ip t/(2m)
(2π)3
d3 p
Z
0 0 0
G(x, t; , x , t ) = Θ(t − t ) ψ(x, t)ψ ∗ (x0 , t0 )
(2π)3
d3 p ip·(x−x0 )−ip2 (t−t0 )/(2m)
Z
0
= Θ(t − t ) e ,
(2π)3
exactly the form found in the lecture.
(c) For discrete energy eigenstates, the wave functions can be labelled
as ψn (x, t), and the integral over momentum eigenstates collapses
to a sum over energy eigenstates – in the end this is only a change
of basis from one orthonormal set to another. In this case,
X
G(x, t; , x0 , t0 ) = Θ(t − t0 ) ψn (x, t)ψn∗ (x0 , t0 )
En
and therefore
0
Θ(t − t0 )ψn (x, t)ψn∗ (x0 , t0 ) = Θ(t − t0 )e−iEn (t−t ) ψn (x)ψn∗ (x0 )
153
where
X ψ ∗ (x)ψn (x0 )
G(ret) (x, x0 , ω) = n
n
ω − En − i+
Using the prinicpal value decomposition above, we see that
h i X ψ ∗ (x)ψn (x0 )
(ret) 0 n
Re G (x, x , ω) = P
n
ω − En
h i X
Im G(ret) (x, x0 , ω) = π ψn∗ (x)ψn (x0 )δ(ω − En )
n
Solution
ˆ † ˆ†
Puse b̂|0i = d|0i = 0 = h0|
In the following we will Pb̂ = h0|d and the
completeness relations u(p)ū(p) = (p/+m) and v(p)v̄(p) = (p/−m)
iSF (y, x)βα = h0|T ψβ (y)ψ̄α (x) |0i
= h0|ψ(y)ψ̄(x)|0iΘ(y0 − x0 ) − h0|ψ̄(x)ψ(y)|0iΘ(x0 − y0 )
d3 p d3 q
Z
= 0
(2π)3 2p0 (2π)3 2q0
2 h i
e−ip·y b̂i (p)u(i) (p) + eip·y dˆ†i (p)v (i) (p)
X
β
i,j=1
h i
× eiq·x b̂†j (q)ū(j) (q)
+e −iq·x
dˆj (q)v̄ (j)
(q) 0 Θ(y0 − x0 )
α
d3 p d3 q
Z
− 0
(2π)3 2p0 (2π)3 2q0
2 h i
eiq·x b̂†j (q)ū(j) (q) + e−iq·x dˆj (q)v̄ (j) (q)
X
×
α
i,j=1
h i
−ip·y (i) ip·y ˆ† (i)
× e b̂i (p)u (p) + e di (p)v (p) 0 Θ(x0 − y0 )
β
154
d3 p d3 q
Z
= 0
(2π)3 2p0 (2π)3 2q0
2
e−i(p·y−q·x) b̂i (p)b̂†j (q)uβ (p)ū(j)
(i)
X
α (q)Θ(y0 − x0 )
i,j=1
i(p·y−q·x) ˆ ˆ† (i) (j)
−e dj (q)di (p)vβ (p)v̄α (q)Θ(x0 − y0 ) 0
d3 p d3 q
Z
= 0
(2π)3 2p0 (2π)3 2q0
2 n o
e−i(p·y−q·x) b̂i (p), b̂†j (q) uβ (p)ū(j)
(i)
X
α (q)Θ(y0 − x0 )
i,j=1
n o
i(p·y−q·x) ˆ ˆ† (i) (j)
−e dj (q), di (p) vβ (p)v̄α (q)Θ(x0 − y0 ) 0
d3 p d3 q
Z
2q0 (2π)3 δ 3 (p − q)δij
= 0
(2π)3 2p0 (2π)3 2q0
2
(i)
X
e−i(p·y−q·x) uβ (p)ū(j)
α (q)Θ(y0 − x0 )
i,j=1
(i)
ei(p·y−q·x) vβ (p)v̄α(j) (q)Θ(x0
− − y0 ) 0
d3 p
Z
e−ip·(y−x) (p/ + m)βα Θ(y0 − x0 )
= 0
(2π)3 2p0
i E
−eip·(y−x) (p/ − m)βα Θ(x0 − y0 ) 0
155
d4 p e−ip0 (y0 −x0 )+ip(y−x) (p/ + m) d4 p e−ip(y−x) (p/ + m)
Z Z
= =
(2π)4 Ep2 − p2 − m2 + i+ (2π)4 p2 − m2 + i+ )
In going from the secon to the third line we have put both terms onto
one denominator, replaced the integral over ω with an integral over
p0 , and we shifted the sign of the spatial integration, going from −p
to +p in the second term. In the second to last step we have first
replaced p20 = p2 + m2 and then identified Ep with p0 . Therefore, the
propagator reads
d4 p −ip·(y−x)
Z
p/ + m
iSF (y − x) = 4
e
(2π) p − m2 + i+
2
Solution
−iqy+ikx †
+ Θ(y0 − x0 )e â(q, κ)â (k, λ) 0
2
d3 k d3 q
Z Z X
= µ (k, λ)ν (q, κ)
(2π)3 (2k0 ) (2π)3 (2q0 )
λ, κ=1
3 3
× −(2π) δ (k − q)(2k0 )δκλ
−ikx+iqy −iqy+ikx
× 0 Θ(x0 − y0 )e
+ Θ(y0 − x0 )e 0
156
2
d3 k
Z X
= − 3
µ (k, λ)ν (k, λ)
(2π) (2k0 )
λ=1
ik(y−x) ik(x−y)
× Θ(x0 − y0 )e + Θ(y0 − x0 )e
(∂µ ∂ µ + m2 )G(x − y) = iδ 4 (x − y) .
157
A solution is readily obtained by Fourier transforming and inverting
this equation:
−i
(−pµ pµ + m2 )G̃(p) = i −→ G̃(p) =
p2 − m2 + i+
where the + takes care of the causality structure of the theory. Equat-
ing the propagator with the classical Green’s function yields the de-
sired result from above. The same also works for the Dirac equation,
where the Lagrangian
L = ψ̄(i∂/ − m)ψ
gives rise to the E.o.M.
(i∂/ − m)ψ = 0 ,
(i∂/ − m)G(x − y) = iδ 4 (x − y) .
158
−α∂ρ g ρσ ∂ν Aν
159
(a) derive the Euler-Lagrange equations of motion for ψ and its con-
jugate ψ † ;
(b) find the canonical momenta π and π † ;
(c) promote the fields and momenta to operators and demand equal
time commutation relations;
(d) expand the fields in plane waves and creation and annihilation
operators, taking into account that this is a non-relativistic field
theory in which negative-energy solutions are absent;
(e) calculate the commutators for the annihilation and creation op-
erators;
(f) express the Hamilton operator through the creation and annihi-
lation operators;
(g) calculate the free-field propagator
For this proof you will have to use that the δ-function can be
represented by r
a −ax2
δ(x) = lim e .
|a|→∞ π
Solution
∂L ∂L ∂L
0 = ∂t †
+∇ −
∂(∂t ψ ) ∂(∇ψ ) ∂ψ †
†
i 1 2 i
= − ∂t ψ − ∇ ψ − ∂t ψ + V (r)ψ
2 2m 2
1 2
= − i∂t + ∇ − V (r) ψ
2m
∂L ∂L ∂L
0 = ∂t +∇ −
∂(∂t ψ) ∂(∇ψ) ∂ψ
i 1 i
= ∂t ψ † − ∇2 ψ † + ∂t ψ † + V (r)ψ †
2 2m 2
1 2
= + i∂t − ∇ + V (r) ψ † .
2m
160
(b) conjugate momenta:
∂L i
π = = ψ†
∂ ψ̇ 2
∂L i
π† = = − ψ.
∂ ψ̇ † 2
(c) we will demand that the non-vanishing commutator is
h i
ψ̂(t, x), iψ̂ † (t, y) = iδ 3 (x − y)
161
(f) The Hamiltonian is given by
H = π ψ̇ + π † ψ̇ † − L
←
i † † †→ 1 † †
= ψ ψ̇ − ψ ψ̇ − iψ ∂ t ψ − (∇ψ ) · (∇ψ) − V (r)ψ ψ
2 2m
1
= (∇ψ † ) · (∇ψ) + V (r)ψ † ψ ,
2m
the sum of kinetic and potential energies, as expected.
(g) To evaluate the free propagator we will use the non-relativistic
energy-momentum relation Ek = k0 = k 2 /(2m) and the fact that
â |0i = 0. This allows to calculate free propagator as
G0 (x0 , x; y0 , y) = −iΘ(x0 − y0 ) h0| ψ(x0 , x)ψ † (y0 , y) |0i
d3 k d3 q −i(k0 x0 −q0 y0 ) i(k·x−iq·y)
Z
= −iΘ(x0 − y0 ) e e
(2π)3 (2π)3
D E
× 0 â(k)↠(q) 0
3 Z∞
"
m(xj − yj ) 2
Y dkj i(x0 − y0 )
= exp − kj −
2π 2m x0 − y0
i=1−∞
im(xj − yj )2
+
2(x0 − y0 )
3
im(xj − yj )2
r
Y m
= exp
2πi(x0 − y0 ) 2(x0 − y0 )
i=1
and therefore
3/2 " #
im(x − y)2
m
G0 (x0 , x; y0 , y) = −iΘ(x0 − y0 ) exp
2πi(x0 − y0 ) 2(x0 − y0 )
162
To show that this satisfies the definition of a Green’s function we
apply the differential kernel of the free E.o.M., and using ∇·x = 3
we find:
∇2
i∂
+ G0 (t, x; 0, 0)
∂t 2m
∇2 imx2
i∂ h m i3/2
= −i + Θ(t) exp
∂t 2m 2πit 2t
2
3i mx
= δ(t) − Θ(t) − 2
2t 2t
3im m2 x2 imx2
h
1 m i3/2
− − 2 exp
2m t t 2πit 2t
2
h m i3/2
imx
= δ(t) exp .
2πit 2t
with
t→0
a = −im/(2t) = m/(2it) −→ ∞
due to the δ(t) we see that indeed
∇2
i∂
+ G0 (t, x; 0, 0) = δ(t)δ 3 (x) .
∂t 2m
163
8 Interacting Fields
8.1 Perturbative Expansion: Born Series
Ĥ = Ĥ0 + V̂ , (323)
in the simplest case a potential. Going back to Eq. (286), where we have
defined the Green’s function this means that we now have
i∂t − Ĥ G(t, x; t0 , x0 )
= i∂t − Ĥ0 − V̂ G(t, x; t0 , x0 ) = δ(t − t0 )δ 3 (x − x0 ) . (324)
where we have for the moment suppressed the i prescription. The Fourier
transform of Eq. (324) can therefore be rewritten as
ˆ
1 ˆ
ω − H̃ G̃ = − Ṽ G̃ = 1 . (326)
G̃0
This can be formally solved, and
1 1
= − Ṽˆ (327)
G̃ G̃0
or
1
G̃ = . (328)
1
G̃0
− Ṽˆ
G(t, x; t0 , x0 ) = G0 (t, x; t0 , x0 )
Z
+ dτ d3 ξG0 (t, x; τ, ξ)V̂ (τ, ξ)G(τ, ξ; t0 , x0 ) , (329)
164
which can now be expanded in powers of interactions with the potential.
This is called the Born series or the perturbative expansion of the Green’s
function. For it to converge we implicitly assume that interactions with the
potential are sufficiently small. Replacing explicit time and space coordi-
nates with four-positions t, x → xi , the Born series therefore reads
G(xN ; x0 ) = G0 (xN ; x0 )
Z
+ dx1 G0 (xN ; x1 ) V̂ (x1 ) G(x1 ; x0 )
Z
+ dx1 dx2 G0 (xN ; x2 ) V̂ (x2 ) G(x2 ; x1 ) V̂ (x1 ) G(x1 ; x0 ) . . . , (330)
tN ≥ tN −1 ≥ tN −2 · · · ≥ t2 ≥ t1 ≥ t0 . (331)
Truncating this series after the first non-trivial term, i.e. after one interac-
tion with the potential is called the Born approximation.
Mf ←i = hf |ii . (332)
The definition of these states is subject on how they are being prepared (for
this initial state) or measure (the final state). For perturbation theory to
165
work, this means that they must be prepared or measured infinitely far away,
both in space and time, from the point where they collider – this assumes,
of course, that the interaction between the states vanishes with increasing
distance. This assumption of asymptotic states is crucial for us to be able
to calculate in a quantum field theory19 .
The S matrix There is yet another problem, while the states |ISi span-
ning the possible initial states of our collision are eigenstates in the initial-
state Fock space of the theory, the corresponding final states |FSi live in
the final-state Fock space. These two sets of states are related to each other
through the Ŝ-matrix such that |FSi = Ŝ |ISi. Therefore the transition
amplitude within the same Fock space is given by
Mf ←i = hf | Ŝ |ii . (333)
In this chapter we will discuss first steps on how to calculate the S-matrix
elements, i.e.
Ŝf i = Mf ←i = hf | Ŝ |ii . (334)
and the time evolution is distributed over both operators and states.
19
This is because the interacting fields are not identical to the free fields: the vacuum
of interacting and free theories is potentially different, and we only know how to quantise
the latter. This implies immediately that the states |ii and |f i are eigenstates of the free
field theory but usually not eigenstates of the interacting field theory. Their interactions
with a cloud of virtual particles around them, from the surrounding interacting vacuum,
will ultimately force us to renormalise the external field, a topic well beyond this lecture
course.
166
The exponential terms exp[−iĤ(t − t0 )] are known as the time evolution
operator,
Similar equations naturally also hold true for other operators in the inter-
action picture.
We now redefine the time-evolution operator in the interaction picture such
that the “free” time-evolution is factored out:
∂ Û (I) (t, t0 )
i = eiĤ0 (t−t0 ) Ĥ0 (t0 )e−iĤ(t−t0 ) − eiĤ0 (t−t0 ) Ĥ(t0 )e−iĤ(t−t0 )
∂t h i
= −eiĤ0 (t−t0 ) Ĥ(t0 ) − Ĥ0 (t0 ) e−iĤ(t−t0 ) = −eiĤ0 (t−t0 ) Ĥint (t0 )e−iĤ(t−t0 )
Û (I) (t, t0 )
Zt
Zt Zt1 tZn−1
∞
(I) (I) (I)
X
= dt1 (−i)n dt1 dt2 . . . dtn Ĥint (t1 )Ĥint (t2 ) . . . Ĥint (tn )
t0 n=0 t0 t0 t0
167
Zt
(I)
= T exp −i dt0 Ĥint (t0 ) , (342)
t0
1 m2 2 λ 4
L = (∂µ φ)(∂ µ φ) − φ − φ , (344)
2 2 4!
is probably the most used example on how to construct and evaluate inter-
acting field theories. The Taylor expansion of its S-matrix elements is thus
given by
Z
−iλ D h i E
hf |Ŝ|ii = hf |1̂|ii + d4 x f T φ̂4 (x) i
4!
−iλ 2
Z D h i E
+ d4 xd4 y f T φ̂4 (x)φ̂4 (y) i + . . . .
4!
(345)
168
S-Matrix vs. Creation and Annihilation Operators Let us now see,
how we can evaluate this expression. We will discuss a 2 → 2 scattering
process, where two φ-particles with momenta p1 and p2 scatter to become
two φ-particles with momenta q1 and q2 , p1 + p2 → q1 + q2 . This means
we will have to manipulate expressions like hq 1 q 2 ; out|p1 p2 ; ini between the
in-space and the out-space. For the sake of clarity we will keep a notation,
where we make it explicit to which space the states and operators belong.
Let us start by using creation and annihilation operators to move one of the
in-particles, p1 , from the state-ket into operators:
D E D E
q 1 q 2 ; out p1 p2 ; in = q 1 q 2 ; out ↠(p1 ; in) p2 ; in
D E
= q 1 q 2 ; out ↠(p1 ; out) p2 ; in
D E
+ q 1 q 2 ; out ↠(p1 ; in) − ↠(p1 ; out) p2 ; in .
(346)
The first term vanishes, unless one of the two momenta q 1,2 = p1 . But this
would mean that one particle would not really participate in the scattering,
something that is usually called a “disconnected diagram”. In such cases
we wouldn’t calculate an amplitude that contributes to a scattering cross
section, and we ignore contributions like this. This leaves us with the second
term. Here, it is important to realise that the in-operator lives at times
t = −∞, while the out-operator is positioned at time t = +∞. This will
help us when we re-express the creation and annihilation operators ↠and
â with the field operators φ̂.
169
allows us to replace the in-space and out-space field operators φ̂in and φ̂rmout
with the field operators in the limits t → −∞ and t → +∞ resulting ulti-
mately in
D E D E
q 1 q 2 ; out p1 p2 ; in = q 1 q 2 ; out ↠(p1 ; in) − ↠(p1 ; out) p2 ; in
−ip1 ·x ←
→
Z
3
= −i lim d x e ∂ q 1 q 2 ; out φ̂(tf , x; out)
ti → −∞
tf → +∞
−φ̂(ti , x; in) p2 ; in
Ztf
−ip1 ·x ← →
Z
3 ∂
= −i lim dt d x e ∂ q 1 q 2 ; out φ̂(t, x) p2 ; in
ti → −∞ ∂t
tf → +∞ t
i
Z
4 2 −ip1 ·x
= −i d x E1 e q 1 q 2 ; out φ̂(t, x) p2 ; in
−ip1 ·x 2
+e ∂t q 1 q 2 ; out φ̂(t, x) p2 ; in
Z
4 2 2
−ip1 ·x
= −i d x ∇ + m e q 1 q 2 ; out φ̂(t, x) p2 ; in
−ip1 ·x 2
+e ∂t q 1 q 2 ; out φ̂(t, x) p2 ; in
Z
4 −ip1 ·x 2
= −i d x e x + m q 1 q 2 ; out φ̂(x) p2 ; in .
(350)
In going from the third to the fourth line we have used that e−ip1 ·x is a
solution for the Klein-Gordon equation, which allowed us to replace the
energy square E12 with (p21 + m2 ), and in going from the fourth to the fifth
line we have integrated by parts, which shifts the ∇2 from the plane wave
to the field operator. This step is possibly only because the interaction is
localised, φ4 (x) and we assume that the fields vanish fast enough for x → ∞
such that the surface terms equal zero.
In a similar way, we can “pull” a state from the final-sate bra through
annihilation operators into a field operator, and we arrive at
D E
q 1 q 2 ; out φ̂(x) p2 ; in
−iq1 ·y ←
→
Z
3
= −i lim d y e ∂ q 2 ; out φ̂out (tf , y; out)φ̂(x)
ti → −∞
tf → +∞
−φ̂(x)φ̂(ti , y; in) p2 ; in
Z h i
4 −iq1 ·y 2
= −i d ye y + m q 2 ; out T φ̂(y)φ̂(x) p2 ; in , (351)
170
where the time-ordering results from the limits for the temporal integration.
h i
× 0 T φ̂(y1 )φ̂(y2 )φ̂(x1 )φ̂(x2 ) 0
(352)
171
By using that the vacuum expectation value of any normal-ordered product
vanishes when sandwiched between vacua,
D E
0 :φ̂(x1 )φ̂(x2 ) . . . φ̂(xn ): 0 = 0 , (355)
and by realising that the Feynman propagators are just numbers, for ex-
ample Eq. (301), and that therefore the vacuum expectation number of any
product of them just equals their product,
h0 |∆F (x1 − x2 )∆F (x3 − x4 ) . . . | 0i = ∆F (x1 − x2 )∆F (x3 − x4 ) . . . (356)
we see that the vacuum expectation value of the time-ordered product of
fields reduces to a product of Feynman propagators and, possibly, “vertex
factors” related to interaction points, where three or more of these fields
interact.
0th -Order Let us now see how this plays out for the 0th -order term, where
we merely have the four field operators. This is equivalent to the term
hf |1̂|ii, the first term in the perturbative expansion of Eq. (345). Going
back to Eq. (352) we therefore end up with
q 1 q 2 1̂ p1 p2
Z
= d x1 d x2 d y1 d y2 e−i(p1 ·x1 +p2 ·x2 −q1 ·y1 −q2 ·y2 )
4 4 4 4
× ∆F (x1 − x2 )∆F (y1 − y2 )
"Z
d4 k1 d4 k2 e−ik1 (x1 −x2 ) e−ik2 (y1 −y2 )
×
(2π)4 (2π)4 k12 − m2 − i+ k22 − m2 − i+
d k1 d4 k2 e−ik1 (x1 −y1 ) e−ik2 (x2 −y2 )
Z 4
+
(2π)4 (2π)4 k12 − m2 − i+ k22 − m2 − i+
#)
d k1 d4 k2 e−ik1 (x1 −y2 ) e−ik2 (x2 −y1 )
Z 4
+
(2π)4 (2π)4 k12 − m2 − i+ k22 − m2 − i+
172
Z
= d x1 d x2 d y1 d y2 e−i(p1 ·x1 +p2 ·x2 −q1 ·y1 −q2 ·y2 )
4 4 4 4
Z 4
d k1 d4 k2 2 2 2
× 4 4
k12 − m2 k2 − m2
(2π) (2π)
"
e−ik1 (x1 −x2 ) e−ik2 (y1 −y2 )
×
k12 − m2 − i+ k22 − m2 − i+
e−ik1 (y1 −x1) e−ik2 (y2 −x2)
+
k12 − m2 − i+ k22 − m2 − i+
#)
e−ik1 (y2 −x1 ) e−ik2 (y1 −x2 )
+ 2
k1 − m2 − i+ k22 − m2 − i+
d4 k1 d4 k2
Z
k12 − m2 k22 − m2
= 4 4
(2π) (2π)
× δ 4 (k1 + p1 )δ 4 (k1 − p2 )δ 4 (k2 + q1 )δ 4 (k2 − q2 )
4 4
p21 2
p22 2
+ δ (p1 − q1 )δ (p2 − q2 ) −m −m . (357)
Looking at this, we realise that the first term will vanish if all external
particles have positive energies - which they should as we want to calculate
a physical cross section. This leaves us the last two terms where particle p1,2
transit directly, without interaction, into particles q1,2 or vice versa. The
absence of an interaction should not come as a surprise: as a starting point
we have only sandwiched the free theory between initial and final state.
1st -Order This however changes, when we go to the first order of pertur-
bation theory, or the first term with an interaction Hamiltonian sandwiched
between hf | and |ii, i.e. the second term on the right-hand side of Eq. (345).
In this case, and in order to arrive at connected diagram, i.e. those where
all external lines are connected through propagators, we will have to con-
nect the four outgoing particles with the interaction vertex. Integrating over
all possible permutations of possible connections and over all space for the
173
vertex position we arrive at
Z
iλ
d z q 1 q 2 : − φ̂4 (z) :
4
p p
4! 1 2
Z
iλ X
=− d x1 d x2 d y1 d y2 d z e−i(p1 ·x1 +p2 ·x2 −q1 ·y1 −q2 ·y2 )
4 4 4 4 4
4!
{x1 ,x2 ,y1 ,y2 }
× ∆F (x1 − z)∆F (x2 − z)∆F (z − y1 )∆F (z − y2 )
Z
= − iλ d x1 d x2 d y1 d y2 d z e−i(p1 ·x1 +p2 ·x2 −q1 ·y1 −q2 ·y2 )
4 4 4 4 4
Z 4
d k1 d4 k3 d4 k3 d4 k4
×
(2π)4 (2π)4 (2π)4 (2π)4
× x1 + m2 x2 + m2 y1 + m2 y2 + m2
"
e−ik1 ·(z−x1) e−ik2 ·(z−x2)
×
k12 − m2 + i+ k22 − m2 + i+
#)
e−ik3 ·(y1−z) e−ik4 ·(y2−z)
× 2
k3 − m2 + i+ k42 − m2 + i+
Z
= − iλ d x1 d x2 d y1 d y2 d z e−i(p1 ·x1 +p2 ·x2 −q1 ·y1 −q2 ·y2 )
4 4 4 4 4
Z 4
d k1 d4 k3 d4 k3 d4 k4
×
(2π)4 (2π)4 (2π)4 (2π)4
× k12 − m2 k22 − m2 k32 − m2 k42 − m2
"
e−ik1 ·(z−x1) e−ik2 ·(z−x2)
× 2 2
k1 − m2 + i+ k2 − m2 + i+
#)
e−ik3 ·(y1−z) e−ik4 ·(y2−z)
× 2
k3 − m2 + i+ k42 − m2 + i+
Z 4
d k1 d4 k3 d4 k3 d4 k4
= − iλ (2π)4 δ 4 (k1 + k2 − k3 − k4 )
(2π)4 (2π)4 (2π)4 (2π)4
× (2π)4 δ 4 (k1 − p1 )(2π)4 δ 4 (k2 − p2 )
4 4 4 4
× (2π) δ (k3 − q1 )(2π) δ (k4 − q2 )
174
interaction vertex, when taking into account the 4! combinations of combin-
ing the four external legs with the vertex.
Feynman Rules This finding allows us to formulate simpler rules for the
construction of amplitudes. The LSZ formula above guarantees that we only
have to take into account interaction vertices connecting the internal lines
for particles, and we know that they are given by the time-ordered products
– or commutators – of the fields. This gives rise to the Feynman rules for
the λφ4 theory, namely
−i
p =
p2 − m2 + i+
@
@t −iλ
@ = (359)
@ 4!
2nd -Order Amplitude Let us now construct a second order amplitude for
the 2 → 2-scattering, using the Feynman rules from Eq. (359). Labelling in-
coming particles as 1, 2 and outgoing particles as 3, 4, we find three different
diagrams, namely
Let us focus now on diagram (a) and translate it into an expression for the
amplitude. We have
2 Z
d4 k d4 q
(a) iλ −i −i
Ŝ = −
4! (2π)4 (2π)4 k2 − m + i q − m2 + i+
2 + 2
(4!)2
× (2π)4 δ(p1 + p2 − q − k)(2π)4 δ(q + k − p3 − p4 )
2
λ2
= (2π)4 δ(p1 + p2 − p3 − p4 )
2!
d4 k
Z
1
× , (361)
(2π) [k − m + i ][(P − k)2 − m2 + i+ ]
4 2 2 +
175
fact that there are two internal lines connecting the two vertices at positions
y1 and y2 , taking out a combinatorial factor of 2!. For diagrams (b) and (c)
we arrive at similar expressions, where P is modified to become P = p1 − p3
and P = p1 − p4 , respectively.
Closer inspection of the k-integration reveals that this diagram gives rise to
a logarithmic divergence. To see this, consider a limit where k becomes in-
finitely large, k → ∞. In this limit the integral assumes the asymptotic form
of d4 k/k 4 , and using polar coordinates in four dimensions, we can write this
as k 3 d3 Ωdk/k 4 , where d3 Ω takes care of the finite angular integrals. This
leaves us with a final integral dk/k which diverges for k → ∞. This consti-
tutes yet another divergence in Quantum Field Theory, and, similar to the
treatment before, it is cured by subtracting suitable terms, this time directly
in the Lagrangian. These terms are constructed after “regularising” the inte-
grals, i.e., after quantifying the degree of their divergence and its prefactors.
The overall procedure of dealing with these ultraviolet divergences is known
as “renormalisation”.
and multiply the result with the Lorentz-invariant flux that describes
√
the phase space density of the incoming particle beam (the term 1/(4 . . .)
in front of the overall expression).
Expressed as an equation and using that p21,2 = m21,2 , and making four-
momentum conservation explicit this therefore reads
1
σi→f = p
4 (p1 · p2 )2 − p21 p22
n n
!
d3 qi
Z Y
¯
X X
× 3
|Ŝf i |2 (2π)4 δ 4 p1 + p2 − qi (362)
(2π) 2Ei
i=1 d.o.f. i=1
For the case at hand, we have the first-order amplitude from Eq. (358).
Stripping out the overall four-momentum conservation it is given by Ŝf i =
iλ. Assuming incident momenta
p
p1,2 = (E, 0, 0, ± E 2 − m2 ) , (363)
176
we arrive at
λ2
σi→f = p
4 (2E 2 − m2 )2 − m4
d3 q1 d3 q2
Z
× (2π)4 δ 4 (p1 + p2 − q1 − q2 )
(2π)3 2E1 (2π)3 2E2
λ2 d3 q1
Z
= √ 3
(2π)δ(2E − E1 − E2 ) √
8E E − m2 2 (2π) 4E1 E2 E2 = q 21 +m2
Z 2
λ2 q 1 d|q 1 |d2 Ω1 q
= √ δ(2E − 2 q 21 + m2 )
32π 2 E E 2 − m2 4(q 21 + m2 )
λ2 (E12 − m2 )dE1
Z
= √ δ(2E − 2E1 )
32πE E 2 − m2 E12
r
λ2 m2
= 1 − (364)
32πE 2 E2
for the cross section at the lowest order in the couplnig constant, O(λ2 )
where we have used polar coordinates for the q1 -integration and realised
that d|q 1 | = dE1 . The cross section has units of inverse energy squared or
area and is usually given in units of “barn”, where
177
9 List of Problems
22 ∗ Poincare transformations
57 Energy-Momentum Tensor
4 72 States and Operators of the Real Scalar Field
73 Wave Functional from State Vectors
74 Two Real Scalar Fields Equal One Complex Scalar Field
79 Momentum Operator
81 Causality and Anti-Commutators (Real Scalars)
82 Commutators for Free Real Scalar Fields
84 ∗ Properties of the Charge Operator
178
Section Page Title
6 132 Polarisation Vectors in Coulomb Gauge
133 Equal-Time Commutators of E and B
134 Momentum Operator P̂ µ
135 ∗∗ Casimir Effect
179
References
[1] M.E. Peskin and D.V. Schroeder. ”An Introduction to Quantum Field
Theory”. http://www.physicsbook.ir/book/An%20Introduction%
20To%20Quantum%20Field%20Theory%20-%20M.%20Peskin,%20D.
%20Schroeder%20(Perseus,%201995).pdf. (probably the standard
reference for quantum field theory. The book is full of insight and
definitely worth a read; be aware that they use a somewhat different
normalisation convention for their fields.).
[2] D.J. Griffiths. ”Introduction to Elementary Particles”. http://
nuclphys.sinp.msu.ru/books/b/Griffiths.pdf. (a light-touch in-
troduction, especially he first few chapters. Recommended for addi-
tional reading to develop intuition and to contextualise some of the
material of the course.).
[3] B. Hatfield. ”Quantum Field Theory of Point Particles and Strings”.
http://www.fulviofrisone.com/attachments/article/483/
Hatfield%20QFT%20of%20particles%20and%20strings(T).pdf.
(this book is eminently readable. What I like best about it is that it
shows field quantisation in three parallel formalisms: Heisenberg oper-
ators, Schr”odinger wave functionals, and Feynman’s path integrals.
This is where – for me – the penny finally dropped. It took a quite
some time . . . .).
[4] M. Srednicki. ”Quantum Field Theory”. http://web.physics.ucsb.
edu/~mark/ms-qft-DRAFT.pdf. (a great resource and an excellent
book – this is the pre-publication version of one of the best recent
books on QFT, and I recommend it full-heartedly!).
[5] D. Tong. ”Quantum Field Theory”. https://www.damtp.cam.ac.uk/
user/tong/qft/qft.pdf. (great lecture notes on Quantum field The-
ory with some overlap with our course).
[6] S Coleman. ”Notes from Sidney Coleman’s Physics 253a: Quantum
Field Theory”. https://arxiv.org/pdf/1110.5013v4.pdf. (arguably
one of the best lectures of all time on Quantum Field Theory, by one
of the great minds in the field.).
[7] J.A. Dror. ”Useful Relations in Quantum Field Theory”. http://
pages.physics.cornell.edu/~ajd268/Notes/UsefulFormulas.pdf.
(quite a summary of things that we will work out during this lecture
course).
[8] K. Dullemond and K. Peeters. ”Introduction to Tensor Calcu-
lus”. http://www.ita.uni-heidelberg.de/~dullemond/lectures/
tensor/tensor.pdf. (exhaustive lecture notes on tensor calculus).
180
[9] H. Goldstein. ”Classical Mechanics”. https://detritus.
fundacioace.com/pub/books/Classical_Mechanics_Goldstein_
3ed.pdf. (standard book on classical mechanics).
181