QED ICTP Note0
QED ICTP Note0
analogously a field describing a plane wave propagating with momentum p should transform
into a field propagating a plane wave with momentum Λp under a Lorentz transformation, i.e.
Analogously for other representation of the Lorentz group (spinorial and tensorial representation
fields).
Λp
x0 x0 + ξ p
1
2 CHAPTER 1. CLASSICAL FIELD THEORY
where the index r can stand for any internal and/or space-time index.
It follows that the product of scalar fields computed at the same point is also a scalar field,
while a generic product of fields computed at the same point will be a generic tensor field
transforming as the product of the representation of the constituent fields. Notice that we are
only acting on the fields, without touching the coordinates. Therefore the product of a scalar
field with a function of the coordinates will not be a proper covariant field, e.g.
∂
V µ (x) ≡ g µν ∂ν φ(x) → g µν ∂ν φ(Λ−1 x) =g µν Λ−1 ν ρ φ(Λ−1 x)
∂(Λ−1 x)ρ
=g µν Λ−1 ν ρ gρσ V σ (Λ−1 x) = Λµ ν V µ (Λ−1 x) . (1.7)
Analogously when applied to fields in other representations or when multiple derivatives are
applied to the same field. Hence a product of fields and derivatives where all the space time
indices are saturated is a scalar field.
Note that the four-volume integral of a scalar field is Poincaré invariant
Z Z Z Z
−1
d x φ(x) → d x φ(Λ x − ξ) = d x| det Λ|φ(x) = d4 x φ(x)
4 4 4
(1.8)
where we assumed that integral extends to infinity (both in space and time) and is well defined.
∂ ∂
∂ν F 0µν = Λµ ρ Λν σ ∂ν F ρσ (Λ−1 x) = Λµ ρ F ρν (Λ−1 x) = Λµ ρ F ρν (x0 ) = 0 . (1.9)
∂(Λ−1 x)ν ∂x0ν
φ(x) = 0 , (1.10)
2
( + m )φ(x) = 0 , (1.11)
φ(x) = j(x) , (1.12)
φ(x) = V [φ(x)] , (1.13)
1.2. FIELD EQUATIONS AND LAGRANGIAN 3
describing respectively a free massless scalar field, a free massive scalar field, a free massless
scalar field coupled to a source, a self interacting scalar field with potential V (φ).
Note that when interactions are present (products of fields in the equations of motion) the
fields must be computed at the same point. If interactions between different space-time points are
present (e.g. φ(x)φ(y)) they would in general mediate instantaneous interactions with potential
catastrophic consequences for causality. Locality of interactions will thus always be assumed in
what follows.
For reasons that will become clear later we will be mainly interested to field equations
with up to 2 derivatives per fields. Such equations involve up to 2 time derivative. If boundary
conditions for the fields and their first time derivatives are given at some time t0 (or equivalently
on a space-like surface)
φr (t0 , ~x) = φ(0)
r (~x) , φ̇r (t0 , ~x) = φ̇(0)
r (~x) , (1.14)
the equations admit a unique solution.
The field on each point ~xi in space can be thought as an independent degree of freedom
qi (t) = φ(t, ~xi ), field equations thus correspond to the usual equations of motion for an infinite
system of degrees of freedom.
We can thus associate a Lagrangian and an action to the system:
Z Z X Z
S = dtL(qi (t), q̇i (t)) = dt ~
Li (qi (t), qi+1 (t), q̇i (t)) = dt d3 x L(φ(x), ∇φ(x), φ̇(x))
i
(1.15)
where the second step follows from locality1 and in the third we introduced the Lagrangian
density L defined as Z
L = d3 xL . (1.16)
Being the action S dimensionless in natural units, it follows that L has dimensions 4 in energy.
Field equations can now be derived by extremizing the action S[φ(x)]. Poincaré invariance
requires that if φ̄(x) is a solution of the equations of motion, the same should be true for the
transformed φ̄0 (x). It follows that if φ̄(x) extremizes the action, δS[φ̄(x)] = 0, so should the
transformed solution on the transformed action, δS 0 [φ̄0 (x)] = 0. A sufficient condition for this
to happen is that the action be invariant under Poincaré transformations S[φ(x)] = S[φ0 (x)].
Given that the action is the space-time integral of the Lagrangian density, this corresponds to
require L to be a scalar. In particular L must be a scalar function of
~
L(φ(x), ∇φ(x), φ̇(x)) = L(φ(x), ∂µ φ(x)) . (1.17)
We are now ready to derive the equation of motion from our action
Z
S[φr (x)] = d4 x L(φr (x), ∂µ φr (x)) (1.18)
Ω
where the index r represents for short any field or space-time index. To make the integral finite
we took a finite 4d box Ω with −L < xi < L and t1 < t < t2 , and impose some boundary
conditions on the space boundaries:
φr (xi = −L) = φr (xi = L) or φr (xi = ±L) = const . (1.19)
1
More precisely locality requires that the Lagrangians Li depend only on the variables (qj (t), q̇j (t)) with j in
a finite neighborhood of i.
4 CHAPTER 1. CLASSICAL FIELD THEORY
Because of locality the boundary effects due to the choice of boundary conditions should vanish
in the limit L → ∞.
In analogy to classical mechanics with a finite number of degrees of freedom we require
that the action is extremized around the solution of the equation of motion φ(x) for arbitrary
fluctuations δφ(x) that vanish at t = t1 and t = t2 :
δS[φr (x)] = S[φr (x) + δφr (x)] − S[φr (x)] = 0 , δφr (x)|t=t1,2 = 0 , (1.20)
the boundary conditions (1.19) give respectively the following extra constraints
We thus get
Z
4 δL δL
0 = δS = d x δφr + δ∂µ φr
Ω δφr δ∂µ φr
Z Z
4 δL δL 4 δL
= d x − ∂µ δφr + d x ∂µ δφr (1.22)
Ω δφr δ∂µ φr Ω δ∂µ φr
that vanish because of the boundary conditions (we used Stokes’ theorem in the last term).
Exploiting the arbitrariness of δφr inside the 4-volume Ω eq. (1.22) implies the equation of
motion
δL δL
∂µ − = 0. (1.24)
δ∂µ φr δφr
1.2.1 Examples
Free real scalar field
Consider the most generic local Lagrangian density of a single real scalar field, quadratic in the
field and with up to one derivative per field
We can always rescale the action and thus the Lagrangian by a constant without changing
the equation of motion. We choose A = 21 which is called the canonical normalization. The
coefficient in front of φ(x)2 now becomes B/2A, which we rename −m2 /2, so that
1 1
L = ∂µ φ(x)∂ µ φ(x) − m2 φ(x)2 , (1.26)
2 2
the choice of the sign in front of the second term will be clear in a moment. From the kinetic
term (the first one) we can read the energy dimensions of the field: the Lagrangian density
has dimension 4, each derivative has dimension 1, so a scalar field canonically normalized has
dimension 1. The constant m2 has thus dimension 2.
1.2. FIELD EQUATIONS AND LAGRANGIAN 5
where the matrices K and M can be chosen symmetric without loss of generality and we further
require that all eigenvalues of K are strictly positive. If O1 and O2 are two orthogonal matrices
and D is a diagonal invertible matrix than it follows that
O1T D−1 O2T O2 DO1 = 1 . (1.30)
If we define
φ̃ = O2 DO1 φ (1.31)
we choose O1 to diagonalize K into KD = O1 KO1T , we choose D such that D−1 K D D−1 =1
and we choose O2 to diagonalize D−1 O1 M 2 O1T D−1 into MD
2 we get
1 1 1X
L = ∂µ φ̃T ∂ µ φ̃ − φ̃T MD
2
φ̃ = (∂µ φ̃i )2 − MD
2 2
ii φ̃i . (1.32)
2 2 2
i
So every quadratic Lagrangian can be brought into a factorized canonical form by means of a
field redefinition.
6 CHAPTER 1. CLASSICAL FIELD THEORY
1.3 Hamiltonian
Exploiting the analogy of a field φ(t, ~x) with an infinite continuum of degrees of freedom in
classical mechanics qi (t) = φ(t, ~x) we can introduce also an Hamiltonian. To proceed we need
to introduce the analogue of the conjugate momenta of the variables qi (t), i.e. the conjugate
momentum density Πr
∂L δL(φ, ∂µ φ)
pi = −→ Πr (t, ~x) = (1.33)
∂ q̇i δ φ̇r (t, ~x)
which we use to get rid of the variables φ̇r and write an Hamiltonian only functions of the
fundamental variable φr and Πr :
Z Z Z
X
3
3 δL
H= pi q̇i − L = d x Πr φ̇r − L = d x φ̇r − L = d3 x H (1.34)
i
δ φ̇ r
where the Hamiltonian should be taken as a function of the fundamental variables only H =
H(φr , Πr ) and we introduced the Hamiltonian density H. Unlike the Lagrangian the Hamiltonian
density is not a scalar under Lorentz transformation, afterall the energy is not a scalar quantity
as well. In the Hamiltonian formalism Lorentz invariance is not manifest anymore, but we will
see that at the quantum level it makes unitarity manifest.
As an example consider a massive real scalar field, its conjugate momentum
δL
Π(x) = = φ̇(x)
δ φ̇
the Hamiltonian density will read
1 1 ~ 1
H = Π(x)2 − L = Π(x)2 + (∇φ(x))2
+ m2 φ(x)2 (1.35)
2 2 2
Note that in this case the Hamiltonian is the sum of three squares so it is positive definite and in
particular is bounded from below. The Hamiltonian is minimized by the classical configuration
φ(x) = 0 which has minimum energy and it’s called the classical vacuum. Notice that a wrong
sign in front of the mass term would result in a Hamiltonian unbounded from below, with obvious
implications regarding the stability of φ(x) = 0 configuration.
For a generic Lagrangian
1
L = (∂µ φ)2 − V (φ) (1.36)
2
where V (φ) is a generic function of φ, the Hamiltonian
1 1 ~
H = Π(x)2 − L = Π(x)2 + (∇φ(x))2
+ V (φ(x)) (1.37)
2 2
will still be bounded from below if the potential V (φ) is. The field configurations that minimize
the potential are called the vacuum configurations. They have Π(x) = φ̇(x) = 0 and ∇φ(x) = 0,
i.e. φ(x) = φ0 =const, and minimize the potential, i.e. V (φ) ≥ V (φ0 ).
Now consider again eq. 1.35 in the discretized limit φ(xi , t) → qi (t)
X1 1
2 2 2 2
H= p + (qi − qi−1 ) + m qi , (1.38)
2 i ∆x2
i
1.4. EXPLICIT SOLUTIONS OF FREE FIELD THEORY 7
where the second terms arises from the discretization of the spatial derivative. We can recognize
the Hamiltonian of an infinite number of coupled harmonic oscillators. The coupling is controlled
by the derivative term that couples the nearest neighbors oscillators. The system can easily be
solved exactly by diagonalizing the system of oscillators, we will see that this simply corresponds
to take the Fourier transform.
( + m2 )φ(x) = 0 .
where the Dirac-δ imposes the constraint p2 = m2 and we used the step functions θ to separate
the two branches with positive and negative frequencies with coefficients ãp~ and b̃p~ respectively,
which are arbitrary functions of the spacial momentum p~ only.
Using the fact that
δ(p0 − ωp ) δ(p0 + ωp )
δ(p2 − m2 ) = δ((p0 )2 − ωp2 ) = + ,
2ωp 2ωp
the solution can be rewritten as
d3 p 1
Z h i
φ(x) = ãp~ e−i(ωp t−~p~x) + b̃p~ ei(ωp t+~p~x)
(2π)3 2ωp
d3 p 1
Z h i
= ãp~ e−i(ωp t−~p~x) + b̃−~p ei(ωp t−~p~x) (1.39)
(2π)3 2ωp
where we integrated over p0 and in the last step we changed p~ to −~ p in the integral.
For a real scalar field φ(x) = φ? (x) which implies that b̃−~p = ap?~ so we finally have
d3 p 1 h
Z i
−i(ωp t−~
p~
x) ? i(ωp t−~
p~
x)
φ(x) = ãp~ e + ãp~ e
(2π)3 2ωp
d3 p 1
Z
= Re(ã p
~ ) cos(ω p t − p
~ ~
x ) + Im(ãp~ ) sin(ωp t − p
~~
x ) (1.40)
(2π)3 ωp
8 CHAPTER 1. CLASSICAL FIELD THEORY
Notice that the final form of the solution is just the linear combination of general solutions of
simple harmonic oscillators, one for each momentum p~ with frequency ωp . The Fourier transform
has thus diagonalized the Hamiltonian, so that in momentum space each mode p~ is associated
to an independent harmonic oscillator.
We now consider the same equation in the presence of a source j(x):
( + m2 )φ(x) = j(x) ,
1 1
L = [∂µ φ(x)]2 − m2 φ(x)2 + j(x) φ(x) .
2 2
Again using Fourier decomposition
d4 p −ipx d4 p −ipx
Z Z
φ(x) = e φ̃p j(x) = e j̃p
(2π)4 (2π)4
j̃p
[p2 − m2 ]φ̃p + j̃p = 0 , or φ̃p = −
p2 − m2
d4 p
Z
i
φ(x) = ij̃p e−ipx + free solution
(2π) p − m2
4 2
Focusing on the first term the solution can conveniently be written as the convolution
Z
φ(x) = i d4 y G(x − y)j(y) , (1.41)
which is the equation for a massive scalar field with a localized source. The solution for G(x) is
then formally
d4 p
Z
i
G(x) = e−ipx .
(2π) p − m2
4 2
Notice that the integral over p0 has singularities over the path of integration so that the integral
is not well defined. The singularities are simple poles located at p0 = ±ωp and can be avoided
by deforming the path of integration onto the complex p0 plane. The ambiguity of the choice
of path is related to the freedom in the boundary conditions of the equation of motion, i.e. to
different choices of the coefficients ãp~ for the free part of the solution.
We discuss next some important cases.
1.4. EXPLICIT SOLUTIONS OF FREE FIELD THEORY 9
vanishes for t < 0. This is simple to check: for t < 0 we can close the contour of integration
on the complex p0 plane with a path extending at |p0 | → ∞ with Im(p0 )>0 so to close a loop
on the upper half plane. The loop does not contain any poles so the integral vanishes. The
contribution from the arch at |p0 | → ∞ also vanishes because the argument of the exponential
in (1.42) goes to −∞ for t < 0 and Im(p0 )>0. Therefore eq. (1.42) vanishes as well.
For t > 0 we can close the contour in the lower half plane. The contribution from the arch at
|p0 | → ∞ and Im(p0 )<0 still vanishes for the same reason as before, but now the loop encloses
the two singularities. Using Cauchy theorem the integral is thus
d3 p 1 −i(ωp t−~p~x)
Z
GR (x) = e − e i(ωp t−~
p~
x)
θ(t) = θ(t)[∆+ (x) − ∆− (x)] .
(2π)3 2ωp
where we defined the functions
d3 p 1 ∓i(ωp t−~p~x)
Z
±
∆ (x) ≡ e
(2π)3 2ωp
which satisfy the relation ∆− (x) = ∆+ (−x) = ∆+ (x)? and correspond to the Green functions
computed on a loop around each single pole at p0 = ±ωp respectively. Such Green functions are
indeed the coefficients of the ãp~ and ãp?~ parameters of the free equation, so that the ambiguity
in the path of integration is a particular choice for ãp~ in the free part of the solution.
The retarded Green function corresponds to the field response to a point-like pulse source
localized at the origin with vanishing boundary condition for t < 0. From eq. (1.42) is clear that
GR (x) = GR (Λx) so that GR can only be a function of the Lorentz invariant combination x2 .
Since the response vanishes for t < 0 and at t = 0 for ~x 6= 0, using Lorentz invariance it follows
that the response vanishes for any xµ except inside the future light-cone, i.e. x2 ≥ 0 and x0 ≥ 0.
The convoluted solution in eq. (1.41) is also clear. The response to a generic source will
just be the linear superposition of the responses from each space-time point of the source, each
producing a signal inside their own future light cone.
The physical meaning of this solution is the time reversal of the previous one: instead as a
source j(x) should be interpreted as a sink and the solution correspond to what field configuration
in the past could end up producing a sink j(x) localized in the origin.
The solution is non-vanishing everywhere and is complex. We will compute the explicit form in
the following chapter where we will explain the physical meaning of the correspoding boundary
conditions.
Before concluding notice once again that the difference between the different choices, i.e. the
ambiguity in the choice of path of integration, correspond to a different choice of the coefficients
for the free part of the equation. For example
where α is the parameter of the transformation and δφr (x) can be a generic local functions of
all fields.
By definition the action is invariant under such transformation, i.e. δS/δα = 0 identically
for any field configuration φr (x) even not satisfying the equation of motion. This implies that
the Lagrangian density must be invariant too, up to a total derivative, i.e.
δL
= ∂µ K µ . (1.44)
δα
2
This definition must be slightly modified at the quantum level...
1.5. SYMMETRIES AND CONSERVED CURRENTS 11
where in the second line we used the equations of motion. If we equate eqs. (1.44) and (1.45)
we get the identity
δL
∂µ δφr = ∂µ K µ , (1.46)
δ(∂µ φr )
which means that on the equation of motion there exist a conserved current j µ defined as
δL
jµ ≡ δφr − K µ ,
δ∂µ φr
∂µ j µ = 0 . (1.47)
For trasformations that are also symmetries of the Lagrangian the last term vanishes.
As usual we can take a four-volume and integrate the relation above to get
Z Z t2 Z Z t2 Z
0= d4 x ∂µ j µ (x) = dt ∂t d3 x j 0 (x) + ~ · ~j(x)
dt d3 x∇
Ω t1 t1 Ω
=⇒
Z t2 Z
Q(t2 ) − Q(t1 ) = dt ~ · ~j(x)
dΣ
t1 ∂Ω
Z
Q(t) ≡ d3 x j 0 (x) (1.48)
the variation of the total charge Q(t) (the integral of the charge density j 0 (x)) is given by the
integrated charge flux exchanged along the spatial boundaries, if we take the spatial volume to
be infinite, or make sure the flux on the boundary vanishes, the charge is conserved.
Examples
• Consider a complex scalar field with Lagrangian
j µ = i(φ∂ µ φ? − φ? ∂ µ φ) , (1.51)
12 CHAPTER 1. CLASSICAL FIELD THEORY
• Consider now a theory with N identical real scalar fields φi=1...N with Lagrangian density
1 1
L = (∂ µ φi )(∂µ φi ) − m2 φ2i , (1.53)
2 2
where repeated indices are summed over. The Lagrangian is invariant under a O(N )
transformation which rotates the scalar fields among themselves
where ta with a = 1 . . . N (N − 1)/2 are the antisymmetric generators of the SO(N ) group
in the fundamental representation, the field transformation can be written as
φi → φi + αa taij φj . (1.56)
δφr = −∂ν φr ,
K µ = −δνµ L , (1.61)
1.5. SYMMETRIES AND CONSERVED CURRENTS 13
Lorentz transformations
The transformation of a field in a generic representation of the Lorentz group can be represented
as
φr (x) → φ0r (x) = Srs (Λ)φs (Λ−1 x) , (1.65)
where S(Λ) is the representation of the action of the Lorentz group on the field φr , e.g. the
identity matrix for scalar fields, Λ itself for vector fields or a more complicated expression for
spinors and tensors. Under infinitesimal transformations the following expressions hold
Λµ ν = δνµ + µ ν
Srs (Λ) = Srs (1 + ) = δrs + µν Σµν
rs
φr (x) → φr (x) + µν Σµν µ ν
rs φs (x) − ν x ∂µ φr (x)
µν 1 µ ν ν µ
= φr (x) + µν Σrs φs (x) + x ∂ φr (x) − x ∂ φr (x) (1.66)
2
where we used the antisymmetric property of µν in the last expression.
As in the previous case we can use the fact that the Lagrangian density transforms as a
scalar field
δL µν δL 1 µ ν 1
j ρ,µν = Σ φs (x) + [x ∂ φr (x) − xν ∂ µ φr (x)] − (xµ g νρ L − xν g µρ L)
δ∂ρ φr rs δ∂ρ φr 2 2
δL µν 1 µ ρν
= Σ φs (x) + (x T − xν T ρµ ) . (1.70)
δ∂ρ φr rs 2
δL µν
Mρµν ≡ 2j ρ,µν = 2 Σ φs (x) + xµ T ρν − xν T ρµ
δ∂ρ φr rs
∂ρ Mρµν = 0 . (1.71)
Z Z
δL µν
J µν = d3 x M0µν = d3 x 2 Σrs φs (x) + xµ P ν − xν P µ (1.72)
δ φ̇r
which are the sum of the intrinsic spin associated to the Lorentz representation of the field and
the orbital momentum ~x × p~.
The other three charges associated to the boosts are
Z
i 0i 3 δL 0i i i
K =J = d x 2 Σrs φs (x) + tP − x H (1.74)
δ φ̇r
P~
Z
1
d3 x H ~x = const + t (1.75)
H H
Notice that these latter charges have an explicit time dependence, therefore
d i ∂
K ={K i , H} + K i = 0 ,
dt ∂t
∂
{K i , H} = − K i 6= 0 (1.76)
∂t
when we will quantize the theory this will imply that boost charges do not commute with the
Hamiltonian—eigenstates of the boost operator will not have definite energy. Therefore they
will not be useful to classify Hamiltonian eigenstates, this is not surprising as energy is not
invariant under Lorentz boosts afterall.