Intro Fields
Intro Fields
1 Introduction
According to the catalogue, this course is “Overview of Quantum Field The-
ory”. Why “Overview”? When we teach Classical Mechanics or Quantum
Mechanics, we teach the groundwork of those subjects, at least up to some
point, and thus we call the courses “Introductions to” rather than “Overviews
of”. Giving an overview is an unusual way to start teaching one of the broad
conceptual frameworks to Physicists. We know that these theoretical frame-
works need to built on a firm conceptual basis. All physicists need that
understanding and so we don’t try to include a preview of time dependant
perturbation theory in freshman physics courses.
Quantum Field theory is also a broad conceptual framework underlying
much of how physicists understand their field. It is, except possibly for
the much more abstract string theory, the only framework for understanding
elementary particle physics at energies high enough for relativity to matter.
In ordinary Quantum Mechanics, the basic operators are the degrees of free-
dom, usually the positions and momenta of each particle. With sufficient
energy available, particles can be created, so the coordinates describing the
positions or spins of a list of individual particles cannot suffice to describe
the state of a system, and thus the quantum mechanical framework used in
nonrelativistic physics is inadequate.
Quantum field theory is also extensively used in condensed matter physics,
even though here there is not enough energy for electrons and positrons to
be created. In condensed matter physics one often describes the system
not in terms of fundamental particles or even nuclei or atoms, but in terms
of effective particles such as holes, phonons or excitons, and these are not
conserved.
Even in atomic physics, because intermediate states in perturbation the-
ory may not conserve energy, the possibility of particle creation has effects,
such as the Lamb shift in hydrogenic atoms, and larger effects in the inner
shells of heavy atoms, for which the potential energies may be comparable
to the rest masses of the electrons.
615: Introduction Last Latexed: September 12, 2013 at 10:18 2
We should note here that the space in which q lives is not necessarily
ordinary three dimensional Euclidean space, or even a Euclidean space of
any number of dimensions, though for single particle motion it is ordinary
space. For example, for a pendulum swinging about a fixed axis, the sole
615: Introduction Last Latexed: September 12, 2013 at 10:18 5
coordinate may be θ, the angle with respect to the downward direction, which
lives on a circle (with θ = 0 and θ = 2π the same point). If it is swinging
from a fixed point rather than a fixed axis, its coordinates are the polar angle
θ and the azimuthal angle φ, which together live on a two-sphere S 2 , that is,
the surface of a three dimensional ball.
Classical mechanics of a conservative system can also be described in
terms of a Hamiltonian, which depends on momenta pj as well as the coor-
dinates, but does not depend on the time derivatives of the coordinates:
where it is understood that on the right hand side the q̇j ’s are to be replaced
by their values in terms of q and p. When Eq. (1) is not solvable for q̇, we
have a complicated situation related to gauge invariance. We will see that
this situation does arise for the electromagnetic field.
The equations of motion in Hamiltonian form are
∂H ∂H
q̇j = , −ṗj = ,
∂pj ∂qj
which determine the classical path the system takes through phase space,
(q(t), p(t)).
We can also relate any explicit time dependences of L and H by
∂L ∂H
− = .
∂t ∂t
615: Introduction Last Latexed: September 12, 2013 at 10:18 6
In the continuum limit, it is more appropriate to use the linear mass density
µ = m/a, Young’s modulus Y and the Lagrangian density. Young’s modulus
is the force required to produce a unit extension per unit length, and as the
force required to stretch one spring, of length a, is F = k(ηj+1 − ηj ), we have
F = Y (ηj+1 − ηj )/a, so k = Y /a. Thus we may write the lagrangian as
1 2
1 ηj+1 − ηj
η̇j2 − Y
X X
L=a Lj = a µ .
2 2 a
j j
R
and L = L(x)dx.
This Lagrangian, however, will not be of much use until we figure out what
is meant by varying it with respect to each dynamical degree of freedom or its
corresponding velocity. In the discrete case we have the canonical momenta
pj = ∂L/∂ η̇j , where the derivative requires holding all η̇j fixed, for j 6= i, as
well as all ηk fixed. This extracts one term from the sum 21 µ aη̇j2 , and this
P
1 ∂ X
p(x = ja) = lim a L(η(x), η̇(x), x)|x=aj .
a→0 a ∂ η̇j
j
for any function f (x), provided x is contained in the interval (x1 , x2 ). Thus
δ ℓ 1 ℓ
Z Z
p(x) = dx′ µη̇ 2 (x′ , t) = dx′ µη̇(x′ , t)δ(x′ − x) = µη̇(x, t).
δ η̇(x) 0 2 0
δ ℓ ∂η ′ ′ ′ ∂2η
Z
L=− dx′ Y (x )δ (x − x) = Y 2 ,
δη(x) 0 ∂x ∂x
∂2η
µη̈(x, t) − Y 2 = 0.
∂x
1 0 0 0
0 −1 0 0
gµν = g µν = ,
with xµ = g µν xν ,
0 0 −1 0
0 0 0 −1
615: Introduction Last Latexed: September 12, 2013 at 10:18 9
Note
1 0 0 0
∂ ν 0 1 0 0
x = δµν =
.
∂xµ 0 0 1 0
0 0 0 1
The spatial part of a contravariant vector will be represented by ~x (boldface
in the book, whose notation we otherwise follow). The dot product of two D-
vectors is understood to use the Minkowski metric, p2 := g µν pµ pν = p20 −~p 2 =
E 2 − p~ 2 , which for a particle is m2 , classically at least.
Let us consider a scalar field φ(~x, t). As for the chain of masses on springs,
the Lagrangian will be a spatial integral of the Lagrangian density,
Z
L(t) = L dD−1x.
As for the chain, the Lagrangian density will depend on spatial as well as
temporal first derivatives of φ, and may also depend on φ itself:
L = L (φ, ∂µ φ, xµ ) ,
δL
π(~x, t) := ,
δ φ̇(~x, t)
1
Peskin defines π(~x, t) := ∂L/∂ φ̇(~x) which I agree with, but with some explanation.
615: Introduction Last Latexed: September 12, 2013 at 10:18 10
Before we give an example, let us note that we have not taken a very
relativistic attitude towards our Lagrangian density, for the momentum den-
sity π(~x, t) is its variation with respect to the time derivative of φ, but we
haven’t varied with respect to the spatial derivatives. We might consider the
canonical momentum π(~x) as one component of a four vector
δ δL(~x ′ , t)
Z
D−1 ′
L(t) = d x .
δ(∂µ φ(~x, t)) δ(∂µ φ(~x, t))
Our procedures do not fully treat δ/δ(∂µ φ(~x, t)) covariantly, however, for we
have been treating φ and φ̇ as completely independant variables, while φ and
∇φ are not. Thus the δ variation above is not well defined.
As for the mass chain, we have
δ ∂φ(~x ′ , t) ∂
′ j = ′ j δ D−1 (x′ − x),
δφ(~x, t) ∂x ∂x
where δ D−1 (~x ′ − ~x)Ris the D − 1 dimensional version of the Dirac delta, zero
unless ~x ′ = ~x with dD−1x′ f (~x ′ )δ D−1 (~x ′ − ~x) = f (~x). So we are not treating
∂j φ as independent of φ.
If we considered a different form of variation, δ̄, in which φ and ∇φ ~ are
considered independent and the Lagrangian density is L = L(φ, φ̇, ∇φ), then ~
our original δ variation can be expressed in terms of δ̄, with
δL δ
Z
= ~
dD−1 x′ L(φ, φ̇, ∇φ)(~
x ′)
δφ(~x) δφ(~x)
δ
Z
= dD−1 x′ ~ x ′)
L(φ, φ̇, ∇φ)(~
δφ(~x)
(
δ̄
Z
= dD−1 x′ ~
L(φ, φ̇, ∇φ)(~x ′) +
δ̄φ(~x)
" # )
δ̄ ~ ′ δ ′
L(φ, φ̇, ∇φ)(~
x) ∂j φ(~x )
δ̄∂j φ(~x ′ ) δφ(~x)
615: Introduction Last Latexed: September 12, 2013 at 10:18 11
~
(
∂L(φ, φ̇, ∇φ)
Z
D−1 ′
= d x (~x ′ )δ D−1 (~x ′ − ~x) +
∂φ
~
)
∂L(φ, φ̇, ∇φ) ′ ∂ D−1 ′
(~x ) ′ j δ (~x − ~x),
∂(∂j φ) ∂x
~
∂L(φ, φ̇, ∇φ) ~
∂ ∂L(φ, φ̇, ∇φ)
= (~x) − j (~x)
∂φ ∂x ∂(∂j φ)
Let us define
δ̄ Z
δ̄L(~x ′ , t) ∂L(φ, ∂µ φ)
π µ (~x, t) = L(t) = dD−1 x′ = (~x, t).
δ̄(∂µ φ(~x, t)) δ̄(∂µ φ(~x, t)) ∂(∂µ φ)
Thus we have
δL ~
∂L(φ, φ̇, ∇φ) ~
∂ ∂L(φ, φ̇, ∇φ)
= (~x) − j (~x)
δφ(~x) ∂φ ∂x ∂(∂j φ)
∂L ∂
= − j π j (~x)
∂φ(~x) ∂x
∂ ∂L(φ, ∂ν φ) ∂L
− = 0.
∂xµ ∂(∂µ φ) ∂φ
Our first important example will be the Klein-Gordon field, in three spa-
tial dimensions, with a lagrangian density
1 1 1 1 1
L = φ̇2 − (∇φ)2 − m2 φ2 = (∂µ φ)(∂ µ φ) − m2 φ2
2 2 2 2 2
615: Introduction Last Latexed: September 12, 2013 at 10:18 12
∂µ ∂ µ φ + m2 φ = 0.
This will be postponed until the end of this chapter. First let’s take a step
into Quantum Mechanics.
3 Quantum Mechanics
In quantum mechanics the question of what path a system takes through
coordinate space is no longer meaningful, nor can a system be at a point in
phase space, for the Heisenburg uncertainty principle says the most localized
a state can be still requires a given volume in phase space. The possible
states of a system can be described in terms of wave functions on coordinate
space, and quantum mechanics (Schrödinger’s equation, for example) tells
us how those functions evolve with time. Thus we have a unitary operator,
615: Introduction Last Latexed: September 12, 2013 at 10:18 13
the time-evolution operator, which gives the amplitude for a state |ψ1 i at
one time to be in the state |ψ2 i a time t later, hψ2 | e−iHt/h̄ |ψ1 i, where we
have assumed the Hamiltonian is time-independent. In fact, this transition
amplitude can be understood in a functional integral formulation of quantum
mechanics as a sum over all possible paths Γ of eiSΓ /h̄ , for the system to go
from one configuration to another at a later time, rather than, as in classical
mechanics, choosing the one path that extremizes the action.
Actually, as we are going to be dealing only with quantum systems, it is
convenient to use units with h̄ = 1, thereby measuring energies and (ordi-
nary) momenta in units of inverse meters, as we have already set c = 1.
Let us look at a simple transition amplitude for a free nonrelativistic
particle, to move from a definite position ~x0 at time 0 to a point ~x at time t.
Here H = ~p 2 /2m, so the amplitude is
2
h~x| e−iHt |~x0 i = h~x| e−ip t/2m |~x0 i (2)
d3 p
Z
2
= 3
h~x| e−ip t/2m |~p i h~p |~x0 i (3)
(2π)
d3 p −i~p 2 t/2m+i~p·(~x−~x0 )
Z
= e (4)
(2π)3
m 3/2 im(~x−~x0 )2 /2t
= e . (5)
2πit
Here2 we are using states normalized so that h~x ′ | ~xi = δ 3 (~x ′ − ~x), with
momentum eigenstates normalized so that h~p ′ | p~i = (2π)3 δ 3 (~p ′ − ~p), and
completeness is
d3 p
Z
1= |~p i h~p | ,
(2π)3
and h~x| p~ i = eipx . This explains the second and third lines above. To get to
the last line, it is useful to know the Gaussian integral3
s
1 ~ ~
πD B · A−1 · B
Z
~ x
dD xe −~
x·A·~
x+B·~
= e4 ,
det A
2
Later we will change that normalization to be consistent with Peskin and Schroeder,
q p
p iPS = 2 p~ 2 + m2 |~p ihere .
|~
3
See http://en.wikipedia.org/wiki/Gaussian integral.
615: Introduction Last Latexed: September 12, 2013 at 10:18 14
m3 ∞ √
Z
−iτ 1+v2
h~x| e−iHt |~x0 i = dv v sin(vr)e (12)
2π 2 r 0
m3 d Z ∞ √
2
= − 2 dv cos(vr)e−iτ 1+v (13)
2π r dr 0
Gradshtein and Ryzhik 3.914 tells us that the last integral is on the border
iτ √
of the valid region and should be √ 2 2
K 1 ( r 2 − τ 2 ).
r −τ
It is pleasing to find that, except for the non-covariant normalization fac-
tor, the dependence is on the invariant m2 ((~x − ~x0 )2 − t2 ) but it is not so
615: Introduction Last Latexed: September 12, 2013 at 10:18 15
pleasant to see that the function does not vanish for positive, spacelike, argu-
ments. Evaluation
√ by steepest descent4 shows that for r ≫ τ the propagator
2 2
goes like e− r −τ , so it falls exponentially but is not identically zero, as rela-
tivity should have it. Thus we have a finite probability that the particle has
moved to a position it only could have gotten to by moving faster than the
speed of light!
We will see that the resolution of this contains
the creation of particle — antiparticle pairs, and ∆ x = ct
the probability of finding a particle at a distance t x
x > ct from where one particle was is due to cre-
ation of another particle from the vacuum, fol- x0
lowed by the antiparticle member of the pair later
annihilating our original particle.
So we see that the quantum mechanics of a single particle gives unphysical
results in a relativistic theory, and we need to consider rather quantizing the
field.
4 Field Quantization
Now consider again the Klein Gordon field described by the “coordinates”
φ(~x, t) and the canonical momenta π(~x, t), with the Hamiltonian density
1 1 1
H = π 2 + (∇φ)2 + m2 φ2 .
2 2 2
How do we quantize this? If we go back to the lattice, we would say
[Pi , qj ] = −iδij .
Our momenta π, however, are momentum densities, so the normalization we
should expect is d3 x[π(~x), φ(~x ′ )] = −i provided ~x ′ is in the integration
R
region, and 0 otherwise. This is what a Dirac delta was invented for, so
[π(~x), φ(~x ′ )] = −iδ 3 (~x − ~x ′ ).
Of course the coordinates commute with each other, as do the momenta:
[φ(~x), φ(~x ′ )] = 0, [π(~x), π(~x ′ )] = 0.
4
See, for example, Arfken, “Mathematical Methods for Physicists”, 2nd Ed., pp 373-
376.
615: Introduction Last Latexed: September 12, 2013 at 10:18 16
The classical mechanics of the wave equation is very simple, with solutions
corresponding to all 3-momenta ~k, with
q
φ(~x, t) ∝ ei~p·~x−iωt , with ω = ± p2 + m2 ,
so more generally the field can be expanded in terms of coefficients for each
momentum,
d3 p i~p·~x−iωt
Z
φ(~x, t) = e φ̃(~p).
(2π)3
Note that as φ(~x) is real (or, in quantum mechanics, hermitean), φ̃ is not
real but rather satisfies the condition φ† (~p) = φ(−~p). Thus we would do
better to write
Z
d3 p 1 i~
p·~
x−iωp~ t † −i~p·~
x+iωp~ t
φ(~x, t) = q ap~ e + ap~ e ,
(2π)3 2ωp~
√
which is automatically Hermitean. The factor 1/ 2ωp~ is introduced to set
the scale of a and a† conveniently, as we will explain in a minute. From π = φ̇
we gather that
d3 p ωp~
Z r
π(~x, t) = −i ap~ ei~p·~x−iωp~ t − ap†~ e−i~p·~x+iωp~ t ,
(2π)3 2
At time t = 0 it may be more convenient, by changing the dummy variable
~p into −~p in the a† terms, to write these as
d3 p 1
Z
†
φ(~x, 0) = 3
q ap~ + a−~p ei~p·~x ,
(2π) 2ωp~
d3 p ωp~
Z r
π(~x, 0) = −i ap~ − a†−~p ei~p·~x ,
(2π)3 2
Then
[φ(~x), φ(~x ′ )]
0
d3 p d 3 p′ 1
Z Z
′ ′
′
ei~p·~x+i~p ·~x
3 ′
[φ(~x), π(~x )] = iδ (~x − ~x ) =
q
(2π) 3 3
(2π) 2 ωp~ ω~p ′
[π(~x), π(~x ′ )] 0
[ap~ + a†−~p , a~p ′ + a†−~p ′ ]
or
[ap~ , a~p ′ ] = 0, [ap~ , a~p† ′ ] = (2π)3 δ 3 (~p − ~p ′ ), [ap†~ , a~p† ′ ] = 0.
Thus we have commuting operators for different ~p, and for each p~ we have a
set of ordinary harmonic oscillator raising and lowering operators, or at least
we would have if we were dealing with momenta discretized by working in a
finite box.
The Hamiltonian (at t = 0) can be written
1 1 1
Z
H = d x π 2 (~x) + (∇φ)2 (~x) + m2 φ2 (~x)
3
2 2 2
( q
1
Z Z 3
dp
Z 3 ′
d p ωp~ ω~p ′ †
†
= d3 x − ap~ − a−~p a~
p ′ − a
−~p ′
2 (2π)3 (2π)3 2
2 ′
)
m − ~p · ~p †
†
′
+ q ap~ + a−~p a~p ′ + a−~p ′ ei(~p +~p)·~x
2 ωp~ ω~p ′
d3 p
(
1 ωp~
Z
† †
= − a p
~ − a −~p a −~
p − a p
~
2 (2π)3 2
2 2
)
m + p~ †
†
+ ap~ + a−~p a−~p + ap~
2ωp~
d3 p
( )
1
Z
† †
= ωp~ ap~ ap~ + ap~ ap~
2 (2π)3
d3 p 1
Z
† †
= ωp~ ap~ ap~ + [ap~ , ap~ ]
(2π)3 2
3
dp 1 3
Z
†
= ω p
~ ap a
~ p~ + δ (0)
(2π)3 2
and the coupling of this energy to gravitation, we can ignore this constant,
though infinite, contribution to the energy of every state in the system —
only energy differences have any effect. So we will drop this constant and
write Z
d3 p
H= ωp~ ap†~ ap~ .
(2π)3
We can find the possible states of the field theory by examining the com-
mution relations of H with ap~ and ap†~ . As
and
[a~p† ′ a~p ′ , ap†~] = a~p† ′ [a~p ′ , ap†~ ] = (2π)3 δ 3 (~p ′ − p~ ) ap†~,
we have
d 3 p′
Z
[H, ap~ ] = 3
ω~p ′ [a~p† ′ a~p ′ , ap~ ] = −ωp~ ap~ ,
(2π)
Z
d 3 p′
[H, ap†~ ] = 3
ω~p ′ [a~p† ′ a~p ′ , a† ] = ωp~ ap†~ .
(2π)
If |ψi is a state with energy E, (H |ψi = E |ψi), then ap~ |ψi is a state
with energy E − ωp~ , unless ap~ |ψi = 0. Thus the lowest energy state of the
system must be the |ψi for which ap~ |ψi = 0 for all p~. This is called the
vacuum state, and has energy zero (after having dropped the constant term
from H). Thus we will describe this state as |0i, with H |0i = 0, ap~ |0i = 0
for all p~.
We can create other states from the vacuum state by applying raising
operators ap†~ , repeatedly, to the vacuum state. From their commutator with
H we see that the energy is
Hap†~1 ap†~2 |0i = [H, ap†~1 ]ap†~2 |0i + ap†~1 [H, ap†~2 ] |0i + ap†~1 ap†~2 H |0i
= ωp~1 ap†~1 ap†~2 |0i + ap†~1 ωp~2 ap†~2 ] |0i + 0
= (ωp~1 + ωp~2 )ap†~1 ap†~2 |0i
so the state ap†~1 ap†~2 |0i has energy (is an eigenstate of H with eigenvalue)
ωp~1 + ωp~2 . This is what a state of two noninteracting particles, each of mass
m, with momenta p~1 and p~2 respectively, should have.
615: Introduction Last Latexed: September 12, 2013 at 10:18 19
is a state of N particles of momenta p~1 , . . . , p~N , and we can get further ev-
idence by asking for the total momentum P~ of this state. The momentum
operator may be derived from translation invariance, as any continuous sym-
metry of a theory corresponds to a conserved quantity, which for translations
is the total momentum. We will discuss this relationship, called Noether’s
theorem, after we make a few more observations about the quantum mechan-
ics of a free scalar field. For the moment, let’s just accept that
~ R 3
~ x, t). Expanding that in terms of a and a† ,
P (t) = − d x φ̇(~x, t) ∇φ(~
d3 p′ ω~p ′
Z Z r
P~ (0) = − d3 x(−i) a ~
p ′ − a †
−~p ′ ei~p ′ ·~
x
(2π)3 2
d3 p 1
Z
†
3
q ap~ + a−~p (i~p) ei~p·~x
(2π) 2ωp~
d3 p 1
Z
† †
= − a −~p − ap~ ap~ + a−~p ~p
(2π)3 2
Z
d3 p
= p~ a† ap~ ,
(2π)3 p~
where the integral over x has given us a (2π)3 δ 3 (~p +~p ′ ) which we have used to
do the d3 p′ integral, and in the last line we have dropped from the integrand
terms which are antisymmetric under p~ ↔ −~p (including d3 p ~pδ 3 (0).)
R
Notice that as ap~ |0i = 0 for all p~, P~ |0i = 0. Also, because
d 3 p′ ′ †
Z
[P~ , ap†~ ] = ~p a~p ′ (2π)3 δ 3 (~p − ~p ′ ) = p~ ap†~ ,
(2π)3
P~ ap†~1 ap†~2 |0i = [P~ , ap†~1 ]ap†~2 |0i + ap†~1 [P~ , ap†~2 ] |0i + ap†~1 ap†~2 P~ |0i
= p~1 ap†~1 ap†~2 |0i + ap†~1 p~2 ap†~2 ] |0i + 0
= (~p1 + ~p2 )ap†~1 ap†~2 |0i
615: Introduction Last Latexed: September 12, 2013 at 10:18 20
so ap†~1 ap†~2 |0i is an eigenstate of P~ with total momentum p~1 + p~2 , and of
H with total energy ωp~1 + ωp~2 , just as one would expect for a state of two
noninteracting particles with momenta p~1 and p~2 respectively. Thus it is clear
that we can construct multiparticle states by applying a† ’s to the vacuum
state.
The normalization, however, is not what Peskin and Schroeder use, which
is instead
q q q
|~p i = 2ωp~ ap†~ |0i , |~p1 , ~p2 i = 2ωp~1 2ωp~2 ap†~1 ap†~2 |0i , . . .
for reasons they explain, which I will not repeat. (I will also switch from ωp~
to Ep~ to be consistent with them, though I don’t see why.) For 1-particle
states, this means
h~p |~q i = 2Ep~(2π)3 δ 3 (~p − ~q )
and
d3 p 1Z
1I1-part = |~p i h~p |
(2π)3 2Ep~
is the projection operator onto single particle states. This normalization is
Lorentz invariant, for the obviously (orthochronous5 ) Lorentz invariant
d4 p d3 p ∞
Z Z Z
δ(p2 − m2 )Θ(p0 ) = dp 0
δ (p 0 2
) − Ep
2
~
(2π)3 (2π)3 0
d3 p d(p0 )2 0 2
Z Z ∞
2
= δ (p ) − E p
~
(2π)3 0 2p0
d3 p 1
Z
= .
(2π)3 2Ep~
Now that we understand ap~ and ap†~ as operators which annihilate or create
one particle of momentum ~p, we see that the field operator φ(~x) can
either create or annihilate a particle at ~x.
So far we have only considered these operators at a given time, as we
generally do in Hamiltonian mechanics, or the Schrödinger picture, where
the operators are fixed but the states transform with H ≡ i∂/∂t. In the
Heisenberg picture, the operators are time dependent, with
φ(xµ ) = φ(~x, t) = eiHt φ(~x, 0)e−iHt ,
5
Proper orthochronous Lorentz transformations are those which preserve the future
direction of time (orthochronous) and which do not reverse right and left hands (proper).
That is, Λ00 > 0 and det Λµν > 0.
615: Introduction Last Latexed: September 12, 2013 at 10:18 21
and ν ρ
φ(xµ ) = eiHt−iP ·~x φ(~0, 0)e−iHt+iP ·~x = eiPν x φ(~0, 0)e−iPρ x .
~ ~
6
If ai and a†i satisfy [ai , a†j ] = δij , Mij is any c-number matrix, and O any function of
a and a† ,
† †
ea Ma O(a, a† )e−a Ma = O(e−M a, a† eM ).
615: Introduction Last Latexed: September 12, 2013 at 10:18 22
1 ∞ p2 π 2π
Z Z Z
= dp sin θdθe−ipr cos θ dφ
(2π)3 0 2Ep 0 0 0110
1 Z∞ p2 eipr − e−ipr 1010
= dp 1010
(2π)2 0 2Ep ipr 10
−i Z ∞
p 1010
= 2
dp √ 2 eipr 10
8π r −∞ p + m2 p
To evaluate this integral9 we note that for r > 0 we may add a semicircle in
the upper half complex p plane at |p| = ∞ to close the contour. Then we
may deform the contour to surround the cut along p = iy for
1
0
y ≥ m. Thus
0
1
0
1
−i
Z ∞ iy 0
1
0
1
µ
D(x′ − xµ ) = −i dy e−yr
0
1
q
8π 2 r m −i y2 − m2
0
1 p
−i ∞ iy 0
1
Z
+ i dy q e−yr 0
1
8π 2 r 2
i y −m
m 2
0
1
0
1
−i iy 0 m
1
Z ∞
= i dy q e−yr .
4π 2 r m 2
i y −m 2
While this does decrease exponentially (roughly as e−mr for large r), it is
certainly not zero (the integrand is positive everywhere), so we do not have
a zero probability of finding an “b” particle at a space-like separated point.
9
Again we see that the integral doesn’t quite converge. The oscillations which cause
this are thrown away in this argument at the ends of the arc at infinity. The justification
for treating these propagators as if they were well defined is to expect them to be convolved
with some wave packet which will smear out the oscillations. Another way to say this is
that the oscillations are an indication of delta functions which are irrelevant for nonzero
distance separations.
615: Introduction Last Latexed: September 12, 2013 at 10:18 24
[φ† (x), φ(y)] = h0| φ† (x)φ(y) |0i − h0| φ(y)φ†(x) |0i = D(x − y) − D(y − x),
were I have used the fact that the calculation of h0| φ(y)φ†(x) |0i is identical
to that for h0| φ† (y)φ(x) |0i, only replacing a and a† with b and b† . For
spacelike separations, however, x − y and y − x can be rotated into each
other, and as D is invariant under orthochronous Lorentz transformations,
which include rotations, the difference must vanish. Thus [φ† (x), φ(y)] = 0
for spacelike x − y. Of course for the complex field [φ(x), φ(y)] = 0 for all
x and y, as the a’s commute with both b and b† . We do, however, need to
be careful about [φ† (x), φ(y)] = 0 near xµ = y µ√ . For from the lagrangian
derived in your homework (with φ := (φ1 + iφ2 )/ 2)
L = ∂µ φ† ∂ µ φ − φ† φ,
we see that
δL
π(x) := = φ̇† (x),
δ φ̇(x)
and as we know that at equal times
we see that
∂ †
[φ (~x, t), φ(~y , t′ )] = −iδ 3 (~y − ~x).
∂t ′
t =t
So we see that we have singular behavior for xµ → y µ .
615: Introduction Last Latexed: September 12, 2013 at 10:18 25
d3 p 1
!
−ipµ (x − y)µ −ipµ (x − y)µ
Z
= e − e (16)
(2π)3 2Ep~
p0 =Ep
p0 =−Ep
d4 p −ipµ (x − y)µ 2
Z
= 3
e δ(p − m2 )ǫ(p0 ) (17)
(2π)
I used the fact that the integration is symmetric under p~ to change the sign
of ~x − ~y in writing (16), and the property of the delta function that
X δ(x − xi )
δ(f (x)) =
xi ∋f (xi )=0
|f ′ (xi )|
√ 1 f
(with E = p~ 2 + m2 ) as the integral around the poles of 2
dp0 .
2πi p0 − E 2
∞
If x0 > y 0 , the integral −∞
R
dp0 , distorted
as in contour Γ0 , may be closed in the lower p0
half plane with an infinite semicircle, and the −E p Γ0 +E p
delta function contributions can be considered
as residues of poles of (p2 − m2 )−1 . That is,
because
Γ+
1 1 1 1
Res 2 2
= , Res 2 2
=− ,
p0 =E p − m 2E p0 =−E p − m 2E
we can write the commutator as
d3 p dp0 −1 −ipµ (x − y)µ
Z Z
†
h0| [φ (x), φ(y)] |0i = e for x0 > y 0
(2π)3 Γ0 2πi p2 − m2
615: Introduction Last Latexed: September 12, 2013 at 10:18 26
The minus sign is because our contour runs clockwise rather than counter-
clockwise.
For x0 < y 0, we can close the contour Γ0 with a semicircle Γ− in the upper
half plane. This closed contour contains no poles or other singularities, so the
integral over Γ0 Γ− vanishes. Our function, integrated over Γ0 , is known as
S
some infinitesimal but positive values of ǫ and ǫ′ . This places the poles just
above the negative real axis and just below the positive one, so dp0 can run
exactly on the real axis and pass to the sides of the poles Feynman requested.
Closing the contour with Γ+ in the lower half plane for x0 > y 0 or with Γ−
in the upper half plane for x0 < y 0 , we see that only the pole at p0 = Ep~
contributes for x0 > y 0, and that at p0 = −Ep~ for x0 < y 0 . The residues of
(p2 −m2 )−1 are +1/2Ep and −1/2Ep respectively, but the clockwise direction
of the contour Γ Γ+ gives an extra − sign in the x0 > y 0 case. Then we
S
have
DF (x − y) = θ(x0 − y 0)D(x − y) + θ(y 0 − x0 )D(y − x)
= θ(x0 − y 0) h0| φ† (x)φ(y) |0i + θ(y 0 − x0 ) h0| φ(y)φ†(x) |0i
=: h0| T φ† (x)φ(y) |0i . (19)
In the last expression, T is a kind of meta “time-ordering” operator, which
tells us to rewrite the operators following it in chronological sequence (right
to left increasing).
Thus far we have discussed only a free field — that is, as we have seen,
the solution to the equations of motion are all of the form that we have some
bunch of particles of various momenta, but each one has constant momentum,
and the number of particles is unchanged in time. This is elegant but not
very interesting — we need to include interactions to have interesting physics.
Unfortunately, interactions make the theory much more difficult to solve. The
simplest interaction our Klein-Gordon particles can have is with an external
classical source. In electromagnetism, this classical source is a 4-current j µ
as you considered for homework, which enters as an additional term −Aµ j µ
in the lagrangian. There are complications here, but for a real scalar field
the source is simpler, so that
1 1
L = (∂µ φ)2 − m2 φ2 + j(x)φ(x) =⇒ (∂ 2 + m2 )φ = j(x).
2 2
Thus the field φ(x) can be found using the retarded Green function
Z
φ(x) = φ0 + i d4 y DR (x − y)j(y).
The last section of Chapter 2 shows that the source creates particles of mo-
mentum p~ with a number expectation value given by the square of the fourier
transform q Z
j̃( p~ 2 + m2 , ~p) = d4 y eip·y j(y) √
.
p0 = ~ 2 +m2
p
615: Introduction Last Latexed: September 12, 2013 at 10:18 28
You should read through this section, but I will not lecture on it.