png2pdf
png2pdf
DigitalCommons@USU
6-6-2022
Part of the Applied Mathematics Commons, Cosmology, Relativity, and Gravity Commons, Elementary
Particles and Fields and String Theory Commons, and the Geometry and Topology Commons
Recommended Citation
Torre, Charles G., "Introduction to Classical Field Theory" (2022). All Complete Monographs. 3.
https://digitalcommons.usu.edu/lib_mono/3
C. G. Torre
Department of Physics
Utah State University
Version 1.4
June 2022
2
About this text
This is a quick and informal introduction to the basic ideas and mathematical
methods of classical relativistic field theory. Scalar fields, spinor fields, gauge
fields, and gravitational fields are treated. The material is based upon lecture
notes for a course I teach from time to time at Utah State University on
Classical Field Theory.
The following is version 1.4 of the text. It is roughly the same as version
1.3. The update to 1.4 includes:
3
4
Contents
Preface 9
2 Klein-Gordon field 17
2.1 The Klein-Gordon equation . . . . . . . . . . . . . . . . . . . 17
2.2 Solving the KG equation . . . . . . . . . . . . . . . . . . . . . 18
2.3 A small digression: one particle wave functions . . . . . . . . . 20
2.4 Variational principle . . . . . . . . . . . . . . . . . . . . . . . 21
2.5 Getting used to δ. Functional derivatives. . . . . . . . . . . . . 26
2.6 The Lagrangian . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.7 The Euler-Lagrange equations . . . . . . . . . . . . . . . . . . 28
2.8 Jet Space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.9 Miscellaneous generalizations . . . . . . . . . . . . . . . . . . 33
2.9.1 External “sources” . . . . . . . . . . . . . . . . . . . . 33
2.9.2 Self-interacting field . . . . . . . . . . . . . . . . . . . 34
2.9.3 KG in arbitrary coordinates . . . . . . . . . . . . . . . 35
2.9.4 KG on any spacetime . . . . . . . . . . . . . . . . . . . 40
2.10 PROBLEMS . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
5
6 CONTENTS
11 Goodbye 249
11.1 Suggestions for Further Reading . . . . . . . . . . . . . . . . . 249
Preface
This document was created to support a course in classical field theory which
gets taught from time to time here at Utah State University. In this course,
hopefully, you acquire information and skills that can be used in a variety
of places in theoretical physics, principally in quantum field theory, particle
physics, electromagnetic theory, fluid mechanics and general relativity. As
you may know, it is usually in courses on such subjects that the techniques,
tools, and results we shall develop here are introduced – if only in bits and
pieces as needed. As explained below, it seems better to give a unified,
systematic development of the tools of classical field theory in one place. If
you want to be a theoretical/mathematical physicist you must see this stuff
at least once. If you are just interested in getting a deeper look at some
fundamental/foundational physics and applied mathematics ideas, this is a
good place to do it.
The traditional physics curriculum supports a number of classical1 field
theories. In particular, there is (i) the “Newtonian theory of gravity”, based
upon the Poisson equation for the gravitational potential and Newton’s laws,
and (ii) electromagnetic theory, based upon Maxwell’s equations and the
Lorentz force law. Both of these field theories appear in introductory physics
courses as well as in upper level courses. Einstein provided us with another
important classical field theory – a relativistic gravitational theory – via
his general theory of relativity. This subject takes some investment in geo-
metrical technology to adequately explain. It does, however, also typically
get a course of its own. These courses (Newtonian gravity, electrodynam-
ics, general relativity) are traditionally used to cover a lot of the concepts
1
Here, and in all that follows, the term “classical” is to mean “not quantum”, e.g., as
in “the classical limit”. Sometimes people use “classical” to also mean non-relativistic; we
shall definitely not being doing that here. Indeed, every field theory we shall consider is
“relativistic”.
9
10 Preface
and methodology of classical field theory. The other field theories that are
important (e.g., Dirac, Yang-Mills, Klein-Gordon) typically arise, physically
speaking, not as classical field theories but as quantum field theories, and it
is usually in a course in quantum field theory that these other field theories
are described. So, in a typical physics curriculum, it is through such courses
that a student normally gets exposed to the tools and results of classical field
theory. This book reflects an alternative approach to learning classical field
theory, which I will now try to justify.
The traditional organization of material just described, while natural in
some ways, overlooks the fact that field theories have many fundamental fea-
tures in common – features which are most easily understood in the classical
limit – and one can get a really good feel for what is going on in “the big pic-
ture” by exploring these features in a general, systematic way. Indeed, many
of the basic structures appearing in classical field theory (Lagrangians, field
equations, symmetries, conservation laws, gauge transformations, etc.) are
of interest in their own right, and one should try to master them in general,
not just in the context of a particular example (e.g., electromagnetism), and
not just in passing as one studies the quantum field. Moreover, once one has
mastered a sufficiently large set of such field theoretic tools, one is in a good
position to discern how this or that field theory differs from its cousins in
some key structural way. This serves to highlight physical and mathematical
ingredients that make each theory special.
From a somewhat more pragmatic point of view, let me point out that
most quantum field theory texts rely quite heavily upon one’s facility with the
techniques of classical field theory. This is, in part, because many quantum
mechanical structures have analogs in a classical approximation to the theory.
By understanding the “lay of the land” in the classical theory through a
course such as this one, one gets a lot of insight into the associated quantum
field theories. It is hard enough to learn quantum field theory without having
to also assimilate at the same time concepts that are already present in the
much simpler setting of classical field theory. So, if you are hoping to learn
quantum field theory some day, this class should help out quite a bit.
A final motivation for the creation and teaching of this course is to support
the research activities of a number of faculty and students here at Utah State
University. The geometric underpinnings of classical field theory feature in a
wide variety of research projects here. If you want to find out what is going
on in this research – or even participate – you need to speak the language.
I have provided a number of problems you can use to facilitate your
11
learning the material. They are presented first within the text in order to
amplify the text and to give you a contextual hint about what is needed
to solve the problem. At the end of each chapter the problems which have
appeared are summarized for your convenience.
I would like to thank the numerous students who have endured the rough
set of notes from which this document originated and who contributed nu-
merous corrections. I also would like to thank Joseph Romano for his (unfor-
tunately rather lengthy) list of corrections. Finally, I would like to acknowl-
edge Ian Anderson for his influence on my geometric point of view regarding
Lagrangians and differential equations.
12 Preface
Chapter 1
13
14 CHAPTER 1. WHAT IS A CLASSICAL FIELD THEORY?
For example, if the medium is the air surrounding you, u could represent
its compression/rarefaction relative to some standard value. For small dis-
placements, the medium can often be well-modeled by supposing that the
displacement field u satisfies the wave equation
1
u,tt − ∇2 u = 0, (1.1)
c2
where the comma notation indicates partial derivatives, ∇2 is the Laplacian,
and c is a parameter representing the speed of sound in the medium. We say
that u is the field variable and that the wave equation is the field equation.
∇ · E = 4πρ, (1.5)
∇ · B = 0, (1.6)
1 ∂E 4π
∇×B− = j, (1.7)
c ∂t c
1 ∂B
∇×E+ = 0, (1.8)
c ∂t
where ρ is the electric charge density and j is the electric current density.
As we shall see later, the electromagnetic field is also fruitfully described
using potentials. They are defined via
1 ∂A
E = −∇φ − , B = ∇ × A. (1.9)
c ∂t
In each of these examples, the field variable(s) are u, φ, (E, B) – or
(φ, A), and they are determined by PDEs. The “infinite number of degrees
of freedom” idea is that, roughly speaking, the general solution to the field
equations (wave, Poisson, or Maxwell) involves arbitrary functions. Thus
the space of solutions to the PDEs – physically, the set of field configurations
permitted by the laws of physics – is infinite dimensional.
1
We are using Gaussian units for the electromagnetic field.
16 CHAPTER 1. WHAT IS A CLASSICAL FIELD THEORY?
Chapter 2
Klein-Gordon field
The simplest relativistic classical field is the Klein-Gordon field. It and its
various generalizations are used throughout theoretical physics.
17
18 CHAPTER 2. KLEIN-GORDON FIELD
you will see in this text.) You can see that the Klein-Gordon (KG) equation
is just a simple generalization of the wave equation, which it reduces to when
m = 0. In quantum field theory, quantum states of the Klein-Gordon field
can be characterized in terms of “particles” with rest mass m and no other
structure (e.g., no spin, no electric charge, etc.) So the Klein-Gordon field
is physically (and mathematically, too) the simplest of the relativistic fields
that one can study.
If you like, you can view the Klein-Gordon equation as a “toy model” for
the Maxwell equations which describe the electromagnetic field. The quan-
tum electromagnetic field is characterized by “photons” which have vanishing
rest mass, no electric charge, but they do carry intrinsic “spin”. Coherent
states of the quantum electromagnetic field contain many, many photons and
are well approximated using “classical” electromagnetic fields satisfying the
Maxwell equations. Likewise, you can imagine that coherent states involv-
ing many “Klein-Gordon particles” (sometimes called scalar mesons) are well
described by a classical scalar field satisfying the Klein-Gordon equation.
The KG equation originally arose in an attempt to give a relativistic
generalization of the Schrödinger equation. The idea was to let ϕ be the
complex-valued wave function describing a spinless particle of mass m. But
this idea didn’t quite work out as expected (see below for a hint as to what
goes wrong). Later, when it was realized that a more viable way to do
quantum theory in a relativistic setting was via quantum field theory, the KG
equation came back as a field equation for a quantum field whose classical
limit is the KG equation above. The role of the KG equation as a sort of
relativistic Schrödinger equation does survive the quantum field theoretic
picture, however. The story is too long to go into in this course, but we will
give a hint as to what this means in a moment.
3/2 Z
1
ϕ(t, r) = d3 k ϕ̂k (t) eik·r , (2.3)
2π R3
Problem:
1. Verify (2.4)–(2.9).
Let us pause and notice something familiar here. Granted a little Fourier
analysis, the KG equation is, via (2.5), really just an infinite collection of
uncoupled “harmonic oscillator equations” for the real and imaginary parts
of ϕ̂k (t) with “natural frequency” ωk . Thus we can see quite explicitly how
the KG field is akin to a dynamical system with an infinite number of degrees
of freedom. Indeed, if we label the degrees of freedom with the Fourier wave
vector k then each degree of freedom is a harmonic oscillator. It is this
interpretation which is used to make the (non-interacting) quantum Klein-
Gordon field: each harmonic oscillator is given the usual quantum mechanical
treatment.
20 CHAPTER 2. KLEIN-GORDON FIELD
and 3/2 Z
− 1
ϕ := d3 k a∗k e−ik·r+iωk t (2.11)
2π R3
∂ϕ+ √
i = −∇2 + m2 ϕ+ , (2.12)
∂t
where the square root operator is defined via Fourier analysis to be3
3/2 Z
√
1
−∇2 + m2 ϕ+ := d3 k ωk ak eik·r−iωk t . (2.13)
2π R3
the positive frequency solutions are sometimes called the “one-particle wave
functions”.
In a quantum field theoretic treatment, the (normalizable) positive fre-
quency solutions represent the wave functions of KG particles. What about
the negative frequency solutions? There are difficulties in using the negative
frequency solutions to describe KG particles. For example, you can easily
check that they satisfy a Schrödinger equation with a negative kinetic energy,
which is unphysical. Moreover, the relativistic inner product with respect to
which one can normalize the positive frequency solutions leads to a negative
norm for the negative frequency solutions. This means that the negative fre-
quency part of the solution to the KG equation cannot be used to describe the
quantum mechanics of a single particle. It turns out that when one couples
a single quantum mechanical KG particle to its environment one invariably
brings the negative frequency solutions into play, thus destroying the single
particle quantum description (see, for example, the “Klein paradox”). In
quantum field theory (as opposed to quantum mechanics), the negative fre-
quency solutions are interpreted in terms of the possibility for destruction
of particles or the creation of anti-particles. Quantum field theory, you see,
allows for creation and destruction of particles. But now we are going too
far afield. . .
The region of integration R can be anything you like at this point, but
typically we assume
We restrict attention to fields such that S[ϕ] exists. For example, we can
assume that ϕ is always a smooth function of compact spatial support. The
value of the integral, of course, depends upon which function ϕ we choose,
so the formula (2.15) assigns a real number to each function ϕ. We say that
S = S[ϕ] is a functional of ϕ. We will use this functional to obtain the KG
equation by a variational principle. When a functional can be used in this
manner it is called an action functional for the field theory.
The variational principle goes as follows. Consider any family of func-
tions, labeled by a parameter λ, which includes some given function ϕ0 at
λ = 0. We say we have a one-parameter family of fields, ϕλ . As a random
example, we might have
2 +y 2 +z 2 )
ϕλ = cos(λ(t + x)) e−(x . (2.17)
You can think of ϕλ as defining a curve in the “space of fields” which passes
through the point ϕ0 .4 We can evaluate the functional S along this curve;
the value of the functional defines an ordinary function of λ, again denoted
S:
S(λ) := S[ϕλ ]. (2.18)
Note that different choices of curve ϕλ will determine different functions S(λ).
We now define a critical point ϕ0 of the action functional S[ϕ] to be a “point
in the space of fields”, that is, a field ϕ = ϕ0 (x), which defines a critical
point of the function S(λ) for any curve passing through ϕ0 .
This way of defining a critical point is a natural generalization to the space
of fields of the usual notion of critical point from multi-variable calculus.
Recall that in ordinary calculus a critical point (x0 , y0 , z0 ) of a function f is
a point where all the first derivatives of f vanish,
This is equivalent to the vanishing of the rate of change of f along any curve
through (x0 , y0 , z0 ). To see this, define the parametric form of a curve ~x(λ)
passing through ~x(0) = (x0 , y0 , z0 ) via
The tangent vector T (λ) to the curve at the point ~x(λ) has Cartesian com-
ponents
T~ (λ) = (x0 (λ), y 0 (λ), z 0 (λ)), (2.22)
and the rate of change of a function f = f (x, y, z) along the curve at the
point ~x(λ) is given by the directional derivative along T~ :
It should be apparent from this equation that the vanishing of the rate of
change of f along any curve passing through the point ~x0 is equivalent to the
vanishing of the gradient of f at ~x0 , which is the same as the vanishing of all
the first derivatives of f at ~x0 . It is this interpretation of “critical point” in
terms of vanishing rate of change along any curve that we generalize to the
infinite-dimensional space of fields.
We shall show that the critical points of the functional S[ϕ] correspond
to functions on spacetime which solve the KG equation. To this end, let
us consider a curve that passes through a putative critical point ϕ at, say,
λ = 0.5 This is easy to arrange. For example, let ϕ̂λ be any curve in the
space of fields. Define ϕλ via
5
We now drop the distracting subscript 0.
24 CHAPTER 2. KLEIN-GORDON FIELD
We call δS the first variation of the action; its vanishing for all curves through
ϕ is the condition for a critical point. We can compute δS explicitly by
applying (2.25) to S[ϕλ ]; we find
Z
d4 x ϕ,t δϕ,t − ∇ϕ · ∇δϕ − m2 ϕδϕ ,
δS = (2.26)
R
The critical point condition means that δS = 0 for all variations of ϕ, and
we want to see what that implies about the critical point ϕ.
To this end we observe that δϕ is a completely arbitrary function (aside
from regularity and boundary conditions to be discussed below). To see this,
let ψ be any function you like and consider the curve
ϕλ = ϕ + λψ, (2.28)
so that
δϕ = ψ. (2.29)
To make use of the requirement that δS = 0 must hold for arbitrary δϕ, we
“integrate by parts” in δS via the divergence theorem:
Z
δS = d4 x − ϕ,tt + ∇2 ϕ − m2 ϕ δϕ
R
hZ it2 Z t2 Z
3
+ d x ϕ,t δϕ − dt d2 A n · ∇ϕδϕ
R3 t1 t1 r→∞
(2.30)
Here the last two terms are the boundary contributions from ∂R. For con-
creteness, I have assumed that
R = [t1 , t2 ] × R3 . (2.31)
The last integral is over the “sphere at infinity” in R3 , with n being the
outward unit normal to that sphere.
If you need a little help seeing where (2.30) came from, the key is to write
ϕ,t δϕ,t − ∇ϕ · ∇δϕ = ∂t (ϕ,t δϕ) − ∇ · (∇ϕδϕ) − ϕ,tt δϕ + (∇2 ϕ)δϕ. (2.32)
2.4. VARIATIONAL PRINCIPLE 25
The first term’s time integral is easy to perform, and the second term’s spatial
integral can be evaluated using the divergence theorem.
To continue with our analysis of (2.30), we make two assumptions re-
garding the boundary conditions to be placed on our various fields. First, we
note that ϕ, and ϕλ , and hence δϕ, must vanish at spatial infinity (r → ∞
at fixed t) in order for the action integral to converge. Further, we assume
that ϕ and δϕ vanish as r → ∞ fast enough so that in δS the boundary
integral over the sphere at infinity vanishes. One way to do this systemat-
ically is to assume that all fields have “compact support” in space, that is,
at each time t they all vanish outside of some bounded region in R3 . Other
asymptotic conditions are possible, but since the area element (dA) in the
integral over the sphere grows like r2 the integrand should fall off faster than
1/r2 as r → ∞ for this boundary term to vanish.
Secondly, we hold fixed the initial and final values of the fields – a step
which should be familiar to you from the variational formulation of classical
mechanics. To this end we fix two functions
φ1 , φ2 : R3 → R (2.33)
and we assume that at t1 and t2 , for any allowed ϕ (not just the critical
point),
ϕ|t1 = φ1 , ϕ|t2 = φ2 . (2.34)
The functions φ1 and φ2 are fixed but arbitrary, subject to the asymptotic
conditions as r → ∞. Now, for our one parameter family of fields we also
demand
ϕλ |t1 = φ1 , ϕλ |t2 = φ2 , (2.35)
which forces
δϕ|t1 = 0 = δϕ|t2 . (2.36)
This forces the vanishing of the first term in the boundary contribution to
δS in (2.30).
With these boundary conditions, we see that the assumption that ϕ is a
(smooth) critical point implies that
Z
d4 x −ϕ,tt + ∇2 ϕ − m2 ϕ δϕ
0= (2.37)
R
26 CHAPTER 2. KLEIN-GORDON FIELD
for any function δϕ subject to (2.36) and the asymptotic conditions just
described. Now, it is a standard theorem in calculus6 that this implies
− ϕ,tt + ∇2 ϕ − m2 ϕ = 0 (2.38)
everywhere in the region R.
This, then, is the variational principle for the KG equation. The critical
points of the KG action, subject to the two types of boundary conditions we
described (asymptotic conditions at spatial infinity and initial/final boundary
conditions), are the solutions of the KG equation.
I would understand if, at this point, you are thinking: “Why would I
want to replace a relatively simple PDE with all this complicated variational
stuff?” The payoff for this investment in variational technology turns out to
be quite large, as I hope you will see by the end of this course.
S[ϕ] = dt L. (2.48)
t1
In classical mechanics, the Lagrangian is a function of the independent vari-
ables, the dependent variables (the “degrees of freedom”), and the derivatives
of the dependent variables. In field theory we have, in effect, degrees of free-
dom labeled by spatial points. We then have the possibility of expressing the
Lagrangian as an integral over space (a sum over the degrees of freedom). In
the KG theory we have that
Z
L= d3 x L, (2.49)
R3
28 CHAPTER 2. KLEIN-GORDON FIELD
where
1
(ϕ,t )2 − (∇ϕ)2 − m2 ϕ2
L= (2.50)
2
is called the Lagrangian density. At a point (t, x, y, z), the Lagrangian density
for the KG field depends on the values of the field ϕ and its first derivatives
at (t, x, y, z). We say that L is a local function of the field.7 Theories like
the KG theory which admit an action which is a spacetime integral of a local
Lagrangian density are called local field theories.
Finally, notice that the Lagrangian for the KG theory can be viewed as
having the same structure as that for a finite dimensional dynamical system
in non-relativistic Newtonian mechanics, namely, L = T − U , where T can
be viewed as a kinetic energy for the field,
Z
1
T = d3 x ϕ2,t (2.51)
R3 2
and U plays the role of potential energy:
Z
1
d3 x (∇ϕ)2 + m2 ϕ2 .
U := (2.52)
R3 2
Evidently, we can view 21 ϕ2,t as the kinetic energy density, and view (∇ϕ)2 +
m2 ϕ2 as the potential energy density.
Here
V 0 = ϕ,t δϕ, V i = −(∇ϕ)i δϕ. (2.55)
The term involving V α is a four-dimensional divergence and leads to the
boundary contributions to the variation of the action via the divergence the-
orem. Assuming the boundary conditions are such that these terms vanish,
we see that the functional derivative of the action is computed by (1) varying
the Lagrangian density, and (2) rearranging terms to move all derivatives of
the field variations into divergences and (3) throwing away the divergences.
We now give a slightly more general way to think about this last com-
putation, which is very handy for certain purposes. This point of view is
developed more formally in the next section.
First, we view the formula giving the definition of the Lagrangian as a
function of 9 variables
L = L(x, ϕ, ϕα ), (2.56)
where now, formally, ϕ and ϕα are just a set of 5 variables upon which the
Lagrangian density depends.8 (The KG Lagrangian density does not actually
depend upon xα except through the field, so in this example L = L(ϕ, ϕα ),
but it is useful to allow for this possibility in the future.) This 9 dimensional
space is called the first jet space for the scalar field. From this point of view,
the field ϕ does not depend upon xα and neither does ϕα . The fields are
recovered as follows. For each function f (x) there is a field obtained as a
graph in the 5-dimensional space of (xα , ϕ), specified by ϕ = f (x). Similarly,
in this setting we do not view ϕα as the derivatives of ϕ; given a function
f (x) we can extend the graph into the 9-dimensional space (xα , ϕ, ϕα ) via
(ϕ = f (x), ϕα = ∂α f (x)). We can keep going like this. For example, we
could work on a space parametrized by (xα , ϕ, ϕα , ϕαβ ), where ϕαβ = ϕβα
parametrizes the values of the second derivatives. This space is the second
jet space; it has dimension 19 (exercise)! Given a field ϕ = f (x) we have a
graph in this 19 dimensional space given by (xα , f (x), ∂α f (x), ∂α ∂β f (x)).
Next, for any formula F (x, ϕ, ϕα ) built from the coordinates, the fields,
and the first derivatives of the fields, introduce the total derivative
∂F ∂F ∂F
Dα F (x, ϕ, ϕα ) = α
+ ϕα + ϕαβ . (2.57)
∂x ∂ϕ ∂ϕβ
8
Notice that we temporarily drop the comma in the notation for the derivative of ϕ.
This is just to visually enforce our new point of view. You can mentally replace the comma
if you like. We shall eventually put it back to conform to standard physics notation.
30 CHAPTER 2. KLEIN-GORDON FIELD
The total derivative just implements in this new setting the calculation of
spacetime derivatives of F via the chain rule. In particular, if we imagine
substituting a specific field, ϕ = f (x) into the formula F , then F becomes a
function F of x only:
F(x) = F (x, f (x), ∂α f (x)). (2.58)
The total derivative of F , when restricted to ϕ = f (x), is the same as the
derivative of F:
Dα F (x, ϕ, ϕα ) = ∂α F(x). (2.59)
ϕ=f (x)
δS
= E(L) = −ϕ,tt + ∇2 ϕ − m2 ϕ, (2.64)
δϕ
ϕ=ϕ(x)
The reason I introduce you to all this jet space formalism is that often
in field theory we want to manipulate a formula such as L(x, ϕ, ∂ϕ) using
∂L
the ordinary rules of multivariable calculus, e.g., calculate ∂x α , so that we
Problems:
L = Dα W α , (2.66)
where
W α = W α (ϕ). (2.67)
Show that
E(L) ≡ 0. (2.68)
so the EL equations are trivial (0 = 0).
a (local) function of the spacetime location, the fields, and the derivatives of
the fields to any finite order. We write
L = L(x, ϕ, ∂ϕ, ∂ 2 ϕ, . . . , ∂ k ϕ). (2.69)
Viewed this way, the Lagrangian is a function on a large but finite-dimensional
space called the k th jet space for the field theory. We denote this space by J k .
Remarkably, if we vary L we can always rearrange things so that all deriva-
tives of δϕ appear inside a total divergence. We have the Euler-Lagrange
identity:
δL = E(L)δϕ + Dα V α , (2.70)
where the general form for the Euler-Lagrange derivative is given by
∂L ∂L ∂L ∂L
E(L) := − Dα + Dα Dβ − · · · + (−1)k Dα1 · · · Dαk ,
∂ϕ ∂ϕ,α ∂ϕ,αβ ∂ϕ,α1 ···αk
(2.71)
and where the general form of the total derivative operator on a function
F = F (x, ϕ, ∂ϕ, ∂ 2 ϕ, . . . , ∂ k ϕ), (2.72)
is given by
∂F ∂F ∂F ∂F
Dα F = + ϕ ,α + ϕ,αβ + · · · + ϕ,αα1 ···αk (2.73)
∂xα ∂ϕ ∂ϕ,β ∂ϕ,α1 ···αk
Here we use the comma notation for (would-be) derivatives in conformation
with standard notation in physics. Notice that the total derivative of a
function on J k is a function on J k+1 .
From the total derivative formula, it follows that divergences have trivial
Euler-Lagrange derivatives
E(Dα V α ) = 0. (2.74)
This reflects the fact that the Euler-Lagrange derivative corresponds to the
functional derivative of the action integral in the case that the action func-
tional is differentiable. In particular, the Euler-Lagrange derivative ignores
all terms on the boundary of the domain of integration of the action integral.
In order to make contact between jet space and the usual calculus of
variations, one evaluates jet space formulas on a specific function ϕ = ϕ(x)
via
∂ϕ(x) ∂ 2 ϕ(x)
ϕ = ϕ(x), ϕ,α = , ϕ,αβ = , .... (2.75)
∂xα ∂xα ∂xβ
2.9. MISCELLANEOUS GENERALIZATIONS 33
In this way formulas defined as functions on jet space become formulas in-
volving only the spacetime. A good framework for doing all this is to view jet
space as a fiber bundle over spacetime. A particular KG field defines a cross
section of that fiber bundle which can be used to pull back various structures
to the base space.
Problems:
1
L = (ϕ2,t − (∇ϕ)2 − m2 ϕ2 ) − σϕ. (2.77)
2
34 CHAPTER 2. KLEIN-GORDON FIELD
1 1
V (ϕ) = − a2 ϕ2 + b2 ϕ4 . (2.81)
2 4
We shall explore the physical utility of this potential a bit later.
Problem:
g(V~ , W
~ ) = gαβ V α W β = −V t W t + V x W x + V y W y + V z W z . (2.82)
where the components of the inverse metric in the inertial Cartesian coordi-
nates happens to be the same as the components of the metric:
−1 0 0 0
0 1 0 0
g αβ =
0 0 1 0 .
(2.85)
0 0 0 1
The matrices gαβ and g αβ are symmetric and they are each other’s inverse:
Let me take a moment to spell out how the metric behaves under a
change of coordinates. Call the old coordinates xα = (t, x, y, z). Call the
2.9. MISCELLANEOUS GENERALIZATIONS 37
This compensates the change in the coordinate volume element in the action
integral:
4
p ∂ x̂ 4 ∂x p p
d x̃ − det(ĝ) = det( ) d x det( ) − det(g) = d4 x − det(g),
∂x ∂ x̂
(2.92)
so that the same formula (2.90) can be used in any coordinates provided you
use the metric appropriate to that coordinate system. Notice that while the
Lagrangian does not explicitly depend upon the coordinates when using an
inertial Cartesian coordinate system it may depend upon the coordinates
in general. Indeed, just switching (x, y, z) to spherical polar coordinates will
introduce explicit coordinate dependence in the Lagrangian, as you can easily
verify.
It is now a straightforward exercise to show that the EL equations of the
KG Lagrangian (2.90) take the form:
p p
∂α − det(g)g ∂β ϕ(x) − m2 − det(g)ϕ(x) = 0.
αβ
(2.93)
Problem:
At this point I want to try to nip some possible confusion in the bud.
While we have a geometric prescription for computing the KG Lagrangian in
any coordinates using the metric and so forth, it is not a good idea to think
that there is but one KG Lagrangian for all coordinate systems. Strictly
speaking, different coordinate systems will, in general, lead to different La-
grangians. This comment is supposed to be completely analogous to the
previously mentioned fact that, while we can compute “the KG equation”
in any coordinate system, each coordinate system leads, in general, to a
different PDE. Likewise, we have different functions L(x, ϕ, ∂ϕ) in different
2.9. MISCELLANEOUS GENERALIZATIONS 39
2.10 PROBLEMS
1. Verify (2.4)–(2.9).
Dα j α = 0, when ( − m2 )ϕ = 0. (3.2)
43
44 CHAPTER 3. SYMMETRIES AND CONSERVATION LAWS
∂ϕ(x) ∂ k ϕ(x)
j α (x) := j α (x, ϕ(x), ,..., ) (3.4)
∂x ∂xk
such that
∂ α
j (x) = 0. (3.5)
∂xα
You can easily see in inertial Cartesian coordinates that our definition of
a conserved current j α = (j 0 , j 1 , j 2 , j 3 ) simply says that the field equations
imply a continuity equation for the density ρ ≡ j 0 (a function on spacetime)
and the current density ~j = (j 1 , j 2 , j 3 ) (a time dependent vector field on
space) associated with any solution ϕ(x) of the field equations:
∂ρ
+ ∇ · ~j = 0, (3.6)
∂t
where
∂ϕ(x) ∂ k ϕ(x)
ρ(x) = j 0 (x, ϕ(x), ,..., ), (3.7)
∂x ∂xk
and
∂ϕ(x) ∂ k ϕ(x)
(~j(x))i = j i (x, ϕ(x), ,..., ), i = 1, 2, 3. (3.8)
∂x ∂xk
The utility of the continuity equation is as follows. Define the total charge
contained in the region V of space at a given time t to be
Z
QV (t) := d3 x ρ(t, ~x). (3.9)
V
3.2. CONSERVATION OF ENERGY 45
Note that the total charge is a functional of the field, that is, its value depends
upon which field you choose. The integral over V of the continuity equation
(3.6) implies that Z
d ~ .
~j · dS
QV (t) = − (3.10)
dt ∂V
Problem:
We call the right hand side of (3.10) the net flux into V . We say the charge QV
is conserved since we can account for its time rate of change purely in terms
of the flux into or out of the region V . In this sense there is no “creation”
or “destruction” of the charge, although the charge can move from place to
place.
With suitable boundary conditions, one can choose V such that charge
cannot enter or leave the region and so the total charge is constant in time.
In this case we speak of a constant of the motion. For example, we have
seen that a reasonable set of boundary conditions to put on the KG field
(motivated, say, by the variational principle) is to assume that the KG field
vanishes at spatial infinity. Let us then consider the region V to be all of
space, that is, V = R3 . If the fields vanish at spatial infinity fast enough,1
the flux will vanish asymptotically and we will have
dQV
= 0. (3.11)
dt
and
∂i j i = −(∇ϕ,t ) · (∇ϕ) − ϕ,t ∇2 ϕ. (3.15)
All together, we get
∂α j α = ϕ,t ϕ,tt − ∇2 ϕ + m2 ϕ
It follows from the identity (3.16) that if ϕ(x) is a solution to the Klein-
Gordon equation then the resulting vector field j α (x) defined in (3.12), (3.13)
will satisfy ∂α j α = 0.
The conserved charge QV associated with this conservation law is called
the energy of the KG field in the region V and is denoted by EV :
Z
1
d3 x ϕ2,t + (∇ϕ)2 + m2 ϕ2 .
EV = (3.17)
V 2
There are various reasons why we view this as an energy. First of all, if you
put in physically appropriate units, you will find that EV has the dimensions
of energy. The best reason comes from Noether’s theorem, which we shall
discuss later. For now, let us recall that the Lagrangian has the form
L = T − U, (3.18)
EV = T + U (3.21)
Problem:
2. What becomes of conservation of energy for the KG field when an
external source is present, as in (2.77)? How do you physically interpret
this state of affairs?
1
(~j(i) )l = −(∇ϕ)l ϕ,i + δil (∇ϕ)2 − (ϕ,t )2 + m2 ϕ2 .
(3.24)
2
Problem:
3. Verify that the currents (3.23), (3.24) are conserved. (If you like, you
can just fix a value for i, say, i = 1 and check that j1α is conserved.)
have the dimensions of momentum (if one takes account of the various di-
mensionful constants that we have set to unity). The name can also be
understood from the fact that the each of the three charge densities ρ(i) cor-
responds to a component of the current densities for the energy conservation
law. Roughly speaking, you can think of this quantity as getting the name
“momentum” since it defines the “flow of energy”. In a little while we will get
an even better explanation from Noether’s theorem. Finally, recall that the
total momentum of a system is represented as a vector in R3 . The Cartesian
components of this vector in the case of a KG field are the P(i) .
Our conservation laws were defined for the KG field on flat spacetime and
the formulas were given in inertial Cartesian coordinates xα = (t, xi ) such
that the metric takes the form
with
gαβ = diag(−1, 1, 1, 1). (3.29)
The formulas (3.26) or (3.27) are in fact correct on any spacetime provided
the metric in the chosen coordinates is specified. Note that the energy-
momentum tensor is symmetric:
If desired, one can view the formula for T as defining a collection of functions
on jet space representing a formula for a tensor field on spacetime. More
0
1
precisely, we can view T as a mapping from J into the 2 tensor fields on
spacetime.
Using inertial coordinates xα = (t, ~x) on flat spacetime you can check
that the conserved energy current has components given by
α
jenergy = −Ttα ≡ −g αβ Ttβ . (3.31)
ϕ = g αβ ϕ,αβ . (3.34)
Problem:
4. Show that the six currents (3.36) are conserved. (Hint: Don’t panic!
This is actually the easiest one to prove so far, since you can use
momentum form. The second charge density, ρ(0)(i) , when integrated over
a region V yields a conserved charge which can be interpreted, roughly, as
the “center of mass-energy at t = 0” in that region. Just as energy and
momentum are two facets of a single, relativistic energy-momentum, you can
think of these two conserved quantities as forming a single relativistic form
of angular momentum.
Let us note that while the energy-momentum conserved currents are (in
Cartesian coordinates) local functions of the fields and their first derivatives,
the angular momentum conserved currents are also explicit functions of the
spacetime coordinates. Thus we see that conservation laws may be, in gen-
eral, functions on the full jet space (x, ϕ, ∂ϕ, ∂ 2 ϕ, . . .).
You are familiar with such “curves in field space” from our discussion of
the variational calculus. There we considered a single (though arbitrary)
curve through a critical point. Here we make a stronger assumption and
assume that the transformation defines a curve through each point in the
space of fields via a formula which is a local function of the field and its
derivatives.3 We view this family of curves as defining a transformation of
any field ϕ, where the transformation varies continuously with the parameter
λ, and such that λ = 0 is the identity transformation.
A simple example of a one parameter family of transformations is the
scaling transformation:
1 αβ 1
g ∂α ϕλ (x)∂β ϕλ (x) + m2 ϕ2λ = − g αβ ∂α ϕ(x)∂β ϕ(x) + m2 ϕ2 .
−
2 2
(3.46)
An equivalent way to express this is that
∂ ∂ϕλ (x)
L(x, ϕλ (x), ) = 0. (3.47)
∂λ ∂x
I think you can see why this is called a “symmetry”. While the KG field is
certainly changed by the symmetry transformation, from the point of view
of the Lagrangian nothing is changed by the transformation.
3
The tangents to all these curves defines a vector field in jet space.
3.7. INFINITESIMAL SYMMETRIES 53
Our definition of variational symmetries did not rely in any essential way
upon the continuous nature of the transformation. For example, you can
easily see that the discrete transformation
ϕ → −ϕ (3.48)
Problems:
5. Consider the real scalar field with the double-well self-interaction po-
tential (2.81). Show that ϕ → ϕ̂ = −ϕ is a variational symmetry. Con-
sider the 3 constant solutions to the field equation (which you found
in the problem just after (2.81)) and check that this symmetry maps
these solutions to solutions.
ϕλ = eλ ϕ, (3.50)
we get
δϕ = ϕ, (3.51)
which shows quite clearly that δϕ is built from ϕ so that, while it is a function
on spacetime for a given field ϕ = ϕ(x), this spacetime function varies from
point to point in the space of fields. Likewise for time translations:
δϕ = ϕ,t . (3.53)
Of course, just as it is possible to have a constant vector field, it is possi-
ble to have a continuous transformation whose infinitesimal form happens
3.7. INFINITESIMAL SYMMETRIES 55
∂ ∂ϕλ (x)
0= L(x, ϕλ (x), )
∂λ ∂x λ=0
∂L ∂L
= δϕ + δϕα
∂ϕ ∂ϕ,α
= δL. (3.57)
4
Contrast this with idea of a critical point. An infinitesimal variational symmetry is a
particular family of field variations that does not change the Lagrangian for any choice of
the field. The critical point is a particular field such that the action does not change for
any field variation.
56 CHAPTER 3. SYMMETRIES AND CONSERVATION LAWS
That this condition is sufficient follows from the fact that it must hold at
all points in the space of fields, so that the derivative with respect to λ
vanishes everywhere on the space of fields. Thus one often checks whether
a continuous transformation is a variational symmetry by just checking its
infinitesimal condition (3.57).
L0 = L + Dα V α , (3.58)
E(Dα V α ) = 0 (3.59)
so that
E(L0 ) = E(L) + E(Dα V α ) = E(L). (3.60)
Therefore, it is reasonable – and we shall see quite useful – to consider gener-
alizing our notion of symmetry. We say that a transformation is a divergence
symmetry if the Lagrangian only changes by the addition of a divergence. In
infinitesimal form, a divergence symmetry satisfies
δL = Dα W α , (3.61)
for some spacetime vector field W α , built locally from the scalar field and its
derivatives, W α = W α (x, ϕ, ∂ϕ, . . . ). Of course, a variational symmetry is
just a special case of a divergence symmetry arising when W α = 0.
Problem:
Time translation,
δϕ = ϕ,t , (3.62)
defines a divergence symmetry of the KG Lagrangian. Let’s see how this
works. We begin by writing the Lagrangian as
1 αβ
g ϕ,α ϕ,β + m2 ϕ2 ,
L=− (3.63)
2
Physically, the presence of this symmetry reflects the fact that there is
no preferred instant of time in the KG theory. A shift in the origin of time
t → t + constant does not change the field equations.
everywhere in the space of fields. But, at any given point in field space, we
always have the identity, valid for any kind of variation, which defines the
Euler-Lagrange expression:
δL = E(L)δϕ + Dα V α , (3.68)
where
∂L ∂L
E(L) = − Dα , (3.69)
∂ϕ ∂ϕ,α
and5
∂L
Vα = δϕ. (3.70)
∂ϕ,α
This identity holds for any field variation. By hypothesis, our field varia-
tion δϕ is some expression built from x, ϕ, and its derivatives that has the
property that δL = 0. Consequently, for the infinitesimal symmetry trans-
formation δϕ we have the relation
Dα W α = E(L)δϕ + Dα V α , (3.73)
which implies
Dα (V α − W α ) = −E(L)δϕ, (3.74)
5
There is an ambiguity in the definition of V α here which we shall ignore for now to keep
things simple. We will confront it when we study conservation laws in electromagnetism.
3.10. TIME TRANSLATION SYMMETRY AND CONSERVATION OF ENERGY59
δL = Dα W α , (3.75)
∂L
jα = δϕ − W α . (3.76)
∂ϕ,α
With
∂L
δϕ = Dt ϕ = ϕ,t , W α = δtα L, = −g αβ ϕ,β , (3.78)
∂ϕ,α
we can apply the results of the previous section to obtain a conserved current:
which is our expression of the conserved energy current in terms of the energy-
momentum tensor.
It is worth pointing out that the existence of the time translation sym-
metry, and hence conservation of energy, is solely due to the fact that the
KG Lagrangian has no explicit t dependence; no other structural features of
the Lagrangian play a role. To see this, consider any Lagrangian whatsoever
satisfying
∂
L(x, ϕ, ∂ϕ, . . .) = 0. (3.81)
∂t
Here I should emphasize that this partial derivative is really a partial deriva-
tive – it only applies to the explicit coordinate dependence of the Lagrangian.
So, for example, in the jet space context of the present discussion we have
∂t (tϕ) = ϕ, Dt (tϕ) = ϕ + tϕ,t . (3.82)
From the identity
∂L ∂L ∂L
Dt L = + ϕ,t + ϕ,tα (3.83)
∂t ∂ϕ ∂ϕ,α
we have – provided ∂t L = 0:
∂L ∂L
ϕ,t + ϕ,tα = Dt L, (3.84)
∂ϕ ∂ϕ,α
which can be interpreted as saying the time translation δϕ = ϕ,t yields a
divergence symmetry (3.77), leading to conservation of energy. One says
that the conserved current for energy is the Noether current associated to
time translational symmetry.
where
W α = (0, ni L), (3.88)
∂
ni L(x, ϕ, ∂ϕ, . . .) = 0, (3.89)
∂xi
so that
∂L ∂L
ni ϕ,i + ni ϕ,ij = ni Di L (3.90)
∂ϕ ∂ϕ,j
j α = (ρ, j i ), (3.91)
with
ρ = ϕ,t n̂ · ∇ϕ, (3.92)
and
1
j i = −ϕ,i n̂ · ∇ϕ + ni (∇ϕ)2 − ϕ2,t + m2 ϕ2 .
(3.93)
2
Since the direction n̂ is arbitrary, it is easy to see that we really have three in-
dependent conservation laws corresponding to 3 linearly independent choices
for n̂. These three conservation laws correspond to the conservation laws for
momentum that we had before. The relation between ρ and j i here and ρ(i)
and ~j(i) there is given by
You can see that the translational symmetry in the spatial direction defined
by n̂ leads to a conservation law for the component of momentum along
n̂. Thus the three conserved momentum currents are the Noether currents
associated with spatial translation symmetry.
62 CHAPTER 3. SYMMETRIES AND CONSERVATION LAWS
Problem:
8. Verify that the Noether currents associated with a spacetime transla-
tion do yield the energy-momentum tensor.
3.13. ANGULAR MOMENTUM REVISITED 63
xα −→ Sβα xβ , (3.101)
We have then
Sγα Sδβ gαβ = gγδ . (3.103)
Consider a 1-parameter family of such transformations, S(λ), so that
α
α α
∂Sβ
Sβ (0) = δβ , =: ωβα (3.104)
∂λ λ=0
Differentiate both sides with respect to λ and set λ = 0 to find the infinites-
imal transformation:
δϕ = (ωβα xβ )ϕ,α , (3.109)
with an antisymmetric ωαβ as above. It is now a short computation to check
that, for the KG Lagrangian,
δL = Dα ωβα xβ L .
(3.110)
(Here I am still using the jet space notation with the total derivative D.)
Thus the Lorentz transformations define a divergence symmetry for the KG
Lagrangian. The resulting Noether current is given by
where M α(γ)(δ) are the conserved currents associated with relativistic angular
momentum.
Problem:
9. Derive (3.110).
where
Sβα Sδγ ηαγ = ηβδ . (3.114)
These symmetries are, naturally enough, called spacetime symmetries since
they involve transformations in spacetime. These symmetry transformations
have a nice geometric interpretation which goes as follows.
Given a spacetime (M, g) we can consider the group of diffeomorphisms,
which are smooth mappings of M to itself with smooth inverses. Given a
diffeomorphism
f : M → M, (3.115)
there is associated to the metric g a new metric f ∗ g via the pull-back. In
coordinates xα on M the diffeomorphism f is given as
xα → f α (x), (3.116)
and the pullback metric has components related to the components of g via
∂f γ ∂f δ
(f ∗ g)αβ (x) = gγδ (f (x)). (3.117)
∂xα ∂xβ
We say that f is an isometry if
f ∗ g = g. (3.118)
LX g = 0. (3.120)
Problem:
ϕλ = ϕ + λ (3.122)
3.16. THE CHARGED KG FIELD AND ITS INTERNAL SYMMETRY67
Problem:
11. Show that this Lagrangian is the sum of the Lagrangians for two (real-
valued) KG fields ϕ1 and ϕ2 with m1 = m2 and with the identification
1
ϕ = √ (ϕ1 + iϕ2 ). (3.125)
2
∂L ∂L ∂L ∂L
− Dµ = 0, − Dµ = 0. (3.128)
∂ϕ1 ∂ϕ1,µ ∂ϕ2 ∂ϕ2,µ
In the latter case, the field equations — equivalent to those above — can be
computed via
∂L ∂L ∂L ∂L
− Dµ = 0, ∗
− Dµ ∗ = 0. (3.129)
∂ϕ ∂ϕ,µ ∂ϕ ∂ϕ,µ
Problem:
12. Using ϕ = ϕ1 +iϕ2 , show that the field equations (3.128) and (3.129) are
equivalent. You should be able to do this for an arbitrary Lagrangian
density, but you should at least do it for the one given in (3.124).
In any case, one has doubled the size of the field space. As we shall see,
the new “degrees of freedom” that have been introduced allow for a notion
of conserved electric charge.6
It is easy to see that the Lagrangian (3.124) for the complex KG field
admits the continuous symmetry
∂L ∂L ∂L ∗ ∂L
δL = δϕ + δϕ,α + δϕ + δϕ∗
∂ϕ ∂ϕ,α ∂ϕ ∗ ∂ϕ∗,α ,α
∗ ∂L ∂L ∗
= Eϕ (L)δϕ + Eϕ∗ (L)δϕ + Dα δϕ + δϕ . (3.132)
∂ϕ,α ∂ϕ∗,α
From the phase symmetry we know that when we set δϕ = iϕ it follows that
δL = 0, so we have
∗ ∂L ∂L ∗
0 = Eϕ (L)iϕ − Eϕ∗ (L)iϕ + Dα iϕ − iϕ . (3.133)
∂ϕ,α ∂ϕ∗,α
Using
∂L
= g αβ ϕ∗,β , (3.134)
∂ϕ,α
∂L
= g αβ ϕ,β , (3.135)
∂ϕ∗,α
we get a conserved current
Problem:
70 CHAPTER 3. SYMMETRIES AND CONSERVATION LAWS
Dα j α = 0, (3.137)
Note that the sign of this charge is indefinite: the charged KG field contains
both positive and negative charges. This charge can be used to model electric
charge in electrodynamics. It can also be used to model the charge which
interacts via neutral currents in electroweak theory.
Problem:
is the set of phases eiλ , labeled by λ, and with group multiplication being
ordinary multiplication of complex numbers:
ϕ : M → V. (3.141)
ϕ −→ r(g)ϕ. (3.142)
U † = U −1 , (3.143)
7
This way of defining SU(2) in terms of a representation provides the “defining repre-
sentation”.
72 CHAPTER 3. SYMMETRIES AND CONSERVATION LAWS
det U = 1, (3.144)
ϕ : M → C2 , (3.148)
so we now have two charged KG fields or, equivalently, four real KG fields.
You can think of ϕ as a 2-component column vector whose entries are complex
functions on spacetime. Let U (λ) be any one parameter family of SU(2)
transformations, as described above. We assume that
U (0) = I. (3.149)
We define
ϕλ = U (λ)ϕ. (3.150)
8
The elements of SU(2) can be parametrized by a unit vector and an angle, just as are
elements of the rotation group SO(3). This is related to the fact that SU(2) provides a
spinor representation of the group of rotations.
3.18. SU(2) SYMMETRY 73
δϕ = iτ ϕ, (3.151)
Note that
δϕ† = −iϕ† τ † = −iϕ† τ. (3.153)
By the way, you can see that τ is traceless and Hermitian by considering
our formula for U (θ, n) above, or by simply noting that U (λ) satisfies
τ = ai σ i , (3.155)
δL = Dα W α , (3.169)
for some W α locally constructed from x, ϕa , ϕa,α , etc. Then the following is
a conserved current:
j α = η α (F ) − W α . (3.170)
Noether’s theorem, as it is conventionally stated – more or less as above,
shows that symmetries of the Lagrangian beget conservation laws. But the
scope of this theorem is actually significantly larger. It is possible to prove
a converse to the result shown above, to the effect that to each conservation
law for a system of Euler-Lagrange equations there is a corresponding sym-
metry of the Lagrangian. Indeed, Noether’s theorem establishes a one-to-one
correspondence between conservation laws and symmetries of the Lagrangian
for a wide class of field theories (including the KG field and its variants that
have been discussed up until now). There is even more than this! But it is
time to move on. . . .
9
There is an ambiguity in the definition of η α here which we shall ignore for now to keep
things simple. We will confront it when we study conservation laws in electromagnetism.
76 CHAPTER 3. SYMMETRIES AND CONSERVATION LAWS
j α = ( − m2 )∂ α ϕ. (3.171)
S αβ = k α ϕ, β − k β ϕ, α , (3.173)
where k α = k α (x) is any given vector field on spacetime. Now make a current
via
j α = Dβ S αβ . (3.174)
It is easy to check that such currents are always conserved, irrespective of
field equations, because the order of differentiation is immaterial:
Dα j α = Dα Dβ S αβ = Dβ Dα S αβ = −Dβ Dα S βα = −Dα j α =⇒ Dα j α = 0.
(3.175)
These sorts of conservation laws are “trivial” because they do not really
reflect properties of the field equations but rather simple derivative identities
analogous to the fact that the divergence of the curl is zero, or that the curl
of the gradient is zero. Indeed, the current above is divergence free for any
function ϕ(x), whether or not it satisfies any field equations.
3.20. “TRIVIAL” CONSERVATION LAWS 77
Here dn−1 A is the area element of the boundary, n is the covariant unit
normal to the boundary, and i = 1, 2, . . . , n. From the continuity equation,
the time rate of change of QV arises from the flux through B:
Z
d
QV = − dn−1 A ni j i . (3.178)
dt B
Problem:
16. Let S be a two dimensional surface in Euclidean space with unit normal
~n and boundary curve C with tangent d~l. Show that
Z Z
1
2 ij
d S ni S ,j = V~ · d~l, (3.181)
S 2 C
where
1
V i = ijk Sjk . (3.182)
2
We have seen there are two kinds of conservation laws that are in some
sense trivial. We can combine these two kinds of triviality. So, for example,
the current
j α = Dβ (k [β (x)Dα] ϕ) + (ϕ − m2 ϕ)Dα ϕ (3.183)
is trivial.
We can summarize our discussion with a formal definition. We say that
a conservation law j α is trivial if there exists a skew-symmetric tensor field
S αβ – locally constructed from the fields and their derivatives – such that
Keep in mind that ∗ω is really a 3-form locally constructed from the field
and its derivatives, that is, it is a 3-form-valued function on the jet space
for the theory. The exterior derivative in (3.188) is a total derivative. As
you know, an exact 3-form is of the form dβ for some 2-form β. If there is a
2-form β locally constructed from the fields such that
then clearly ∗ω is closed modulo the field equations. This is just the differ-
ential form version of a trivial conservation law. Indeed, the anti-symmetric
tensor field that is the “potential” for the conserved current is given by
1
S αβ = αβγδ βγδ . (3.190)
2
Let me mention and dispose of a common point of confusion concerning
trivial conservation laws. This point of confusion is why I felt compelled to
80 CHAPTER 3. SYMMETRIES AND CONSERVATION LAWS
occasionally stick in the phrase “locally constructed from the field” in the
discussion above. For simplicity, I will use the flat metric and Cartesian
coordinates on the spacetime manifold M = R4 in what follows. To expose
the potential point of confusion, let me remind you of the following standard
result from multivariable calculus. Let V α be a vector field on Minkowski
space, expressed in the usual inertial Cartesian coordinates. V α is not to be
viewed as locally constructed from the field, except in the trivial sense that it
does not depend upon the fields at all, only the spacetime point, V α = V α (x).
If V α is divergence free,
∂α V α = 0, (3.191)
then there exists an antisymmetric tensor field S αβ such that
V α = ∂β S αβ . (3.192)
This is just the dual statement to the well-known fact that all closed 3-forms
(indeed, all closed forms of degree higher than 0) on R4 are exact i.e., the De
Rham cohomology of R4 is trivial. This result might tempt you to conclude
that all conservation laws are trivial! Unlike the case in real life, you should
not give in to temptation here. There are two reasons. First, a conservation
law should not be viewed as just a single divergence-free vector field on the
spacetime manifold M . A conservation law is a formula which assigns a
divergence-free vector field to each solution of the field equations. Each field
configuration will, in principle, define a different conserved current. Second,
as we have been saying, each of the (infinite number of) divergence-free vector
fields is to be locally constructed from the fields, i.e., are functions on jet space
(rather than just x space). Put differently, the conserved current at a point
x is required to depend upon the values of the fields and their derivatives at
the point x. The correct notion of triviality is that a conserved current j α is
trivial if for each field configuration it is (modulo the field equations) always a
divergence of a skew tensor field S αβ that is itself locally constructed from the
fields. If we take a conservation law and evaluate it on a particular solution
to the field equations, then we end up with a divergence-free vector field on
M (or a closed 3-form on M , if you prefer). If M = R4 we can certainly write
this vector field as the divergence of an antisymmetric tensor on M (or as the
exterior derivative of a 2-form on M ). But the point is that for non-trivial
conservation laws there is no way to construct all the antisymmetric tensors
(2-forms) for all possible field configurations using a local formula in terms of
the fields and their derivatives. So, while conservation laws are in many ways
3.22. PROBLEMS 81
Problem:
17. Consider two possible choices, η1 and η2 , for η defined in (3.167). They
are related by
η2α = η1α + Dβ S αβ , (3.193)
for some skew tensor S locally constructed from the fields and field
variations. Suppose there is a divergence symmetry of the Lagrangian.
Show that the conserved currents j1 and j2 , constructed via (3.170)
using η1 and η2 , respectively, differ by a trivial conservation law.
3.22 PROBLEMS
3. Verify that the currents (3.23), (3.24) are conserved. (If you like, you
can just fix a value for i, say, i = 1 and check that j1α is conserved.)
10
One says the conservation laws are “horizontally” closed forms on the jet space.
82 CHAPTER 3. SYMMETRIES AND CONSERVATION LAWS
4. Show that the six currents (3.36) are conserved. (Hint: Don’t panic!
This is actually the easiest one to prove so far, since you can use
5. Consider the real scalar field with the double well self-interaction poten-
tial (2.81). Show that ϕ̂ = −ϕ is a symmetry. Consider the 3 constant
solutions to the field equations (which you found in the problem just
after (2.81)) and check that this symmetry maps these solutions to
solutions.
9. Derive (3.110).
ϕλ = ϕ + λ
11. Show that the Lagrangian (3.124) is the sum of the Lagrangians for
two (real-valued) KG fields ϕ1 and ϕ2 with m1 = m2 and with the
identification
1
ϕ = √ (ϕ1 + iϕ2 ).
2
3.22. PROBLEMS 83
12. Using ϕ = ϕ1 +iϕ2 , show that the field equations (3.128) and (3.129) are
equivalent. You should be able to do this for an arbitrary Lagrangian
density, but you should at least do it for the one given in (3.124).
16. Let S be a two dimensional surface in Euclidean space with unit normal
~n and boundary curve C with tangent d~l. Show that
Z Z
2 ij
d S ni S ,j = V~ · d~l, (3.194)
S C
where
1
V i = ijk Sjk . (3.195)
2
17. Consider two possible choices, η1 and η2 , for η defined in (3.167). They
are related by
η2α = η1α + Dβ S αβ , (3.196)
for some skew tensor S locally constructed from the fields and field
variations. Suppose there is a divergence symmetry of the Lagrangian.
Show that the conserved currents j1 and j2 , constructed via (3.170)
using η1 and η2 , respectively, differ by a trivial conservation law.
84 CHAPTER 3. SYMMETRIES AND CONSERVATION LAWS
Chapter 4
85
86 CHAPTER 4. THE HAMILTONIAN FORMULATION
The 2-form is called the symplectic form, and the pair (Γ, Ω) is known mathe-
matically as a symplectic manifold. Physicists call it the phase space or state
4.1. REVIEW OF THE HAMILTONIAN FORMULATION OF MECHANICS87
space for the mechanical system. The dimensionality of Γ is 2n, where the
integer n is the number of degrees of freedom of the system.
The second ingredient in a Hamiltonian system is a function, H : Γ → R,
called the Hamiltonian. The Hamiltonian defines how states of the system
evolve in time. Given coordinates z i for Γ, the dynamical evolution of the
mechanical system is a curve z i = z i (t) in the phase space defined by the
ordinary differential equations:
∂H
ż i (t) = Ωij , (4.7)
∂z j
z=z(t)
where Ωij are the components of the inverse symplectic form. The ODEs
(4.7) are known as Hamilton’s equations of motion. More generally, given
any function G : Γ → R, there is an associated foliation of Γ by curves
z i = z i (s) defined by
∂G
ż i (s) = Ωij j . (4.8)
∂z
z=z(s)
You can check that this form of Ω is closed and non-degenerate. The co-
ordinates (q, p) are called canonical coordinates (or “canonical coordinates
and momenta”). The existence of such coordinates is a direct consequence of
the non-degeneracy of Ω and the condition (4.6). There are infinitely many
canonical coordinate systems.1 As a nice exercise, you should determine the
components Ωij of the symplectic form in canonical coordinates. In canonical
coordinates the Hamilton equations take the familiar textbook form
∂H ∂H
q̇ a = , ṗa = − . (4.10)
∂pa ∂q a
Problem:
1
They are all related by “canonical transformations”.
88 CHAPTER 4. THE HAMILTONIAN FORMULATION
Problem:
4.1. REVIEW OF THE HAMILTONIAN FORMULATION OF MECHANICS89
2. Given L = L(q, q̇), give a formula for the EL equations linearized about
a solution q a = q a (t).
Ω = dΘ. (4.16)
In formulas, this prescription says that the value of the symplectic 2-form on
a pair of tangent vectors δ1 q a (t) and δ2 q a (t) at the point of Γ specified by
q a (t) is given by
∂ 2L ∂ 2L
Ω(δ1 q, δ2 q) = a b (δ1 q δ2 q − δ1 q δ2 q ) + a b (δ1 q̇ b δ2 q a − δ1 q a δ2 q̇ b ),
b a a b
∂ q̇ ∂q ∂ q̇ ∂ q̇
(4.17)
where it is understood that all expressions involving q(t) and q̇(t) are eval-
uated at the given solution, and that all instances of δq(t) are evaluated at
solutions of the equations of motion linearized about q(t).
The formula (4.17) appears to depend upon the time at which the solu-
tions and the linearized solutions are evaluated. But it can be shown that Ω
does not depend upon t by virtue of the EL equations satisfied by q a (t) and
their linearization satisfied by δq a (t).
Problem:
3. Show that the symplectic form given in (4.17) does not depend upon
time. (Hint: Differentiate (4.17) with respect to time and then use the
EL equations and their linearization.)
Since we are currently viewing the phase space as the set of solutions to
the EL equations, the Hamiltonian should define one parameter families of
solutions corresponding to translations in time of those solutions. For sim-
plicity in what follows, I will suppose we are working with a system whose
Lagrangian is not explicitly dependent upon time, L = L(q, q̇).2 I will now
show how the canonical energy function,
∂L
E = q̇ i − L, (4.19)
∂ q̇ i
will implement (4.18) for time translations.
To begin, it is a little easier to rewrite (4.18) as
where δq is any solution to the linearized equations and X defines the lin-
earized solution defined by infinitesimal time translations3 :
X = q̇ a . (4.21)
This matches the second and fourth terms in (4.22), and (4.20) is verified.
An equivalent way to think about time evolution makes use of an iden-
tification of the phase space with the space of initial data for the solutions.
This is the most common approach found in textbooks. Since we have chosen
L = L(q, q̇) the EL equations are second-order ODEs whose solution space is
parametrized by initial data at some time t0 which we denote (q a (t0 ), q̇ a (t0 )).
We assume that there is a bijective correspondence between solutions and
initial conditions; we can therefore identify Γ with the initial conditions. A
tangent vector at a point of Γ is then a pair (δq(t0 ), δ q̇(t0 )). It is then easy
to restrict Θ and Ω to t = t0 and interpret them as forms on the space of
initial data. Henceforth we drop the notation pertaining to t0 . If we define
∂L
pa = , (4.25)
∂ q̇ a
then
Θ = pa dq a , Ω = dpa ∧ dq a . (4.26)
So, if we make a change of variables (q, q̇) −→ (q, p) on Γ via (4.25) then
(q a , pa ) are canonical coordinates and momenta for Γ. The Hamiltonian
which generates time evolution is then given by the canonical energy for-
mula:4
H = pa q̇ a − L, (4.27)
where it is understood that H = H(q, p) – all velocities q̇ a being eliminated
in terms of momenta by the inverse formula q̇ a = q̇ a (q, p) to (4.25).5 It is
easy to check that the relation (4.18) now yields (4.10), and with the choice
(4.27) the Hamilton equations are equivalent to the original EL equations.
Problem:
4. A harmonic oscillator is defined by the Lagrangian and EL equations
1 1
L = mq̇ − mω 2 q 2 , q̈(t) = −ω 2 q(t). (4.28)
2 2
What does the symplectic structure look like when the space of solu-
tions to the EL equations is parametrized according to the following
formulas?
4
We denote the Hamiltonian by H to distinguish it from E since the former is a
function of (q, p) while the latter is a different function of (q, q̇).
5
The non-degeneracy condition (4.14) ensures the local existence of an inverse.
92 CHAPTER 4. THE HAMILTONIAN FORMULATION
δL = −ϕ,tt + ∇2 ϕ − m2 ϕ − V 0 (ϕ) δϕ + Dα W α ,
(4.31)
We can now mimic the construction of the Hamiltonian formalism from the
previous section.
We let the phase space Γ consist of a suitable6 function space of solutions
ϕ to the EL equations,
ϕ − m2 ϕ − V 0 (ϕ) = 0. (4.33)
Assuming the solutions vanish at spatial infinity, the variational identity for
the Lagrangian evaluated on Γ reads:
Z
d
δL = d3 x W 0 . (4.34)
dt R3
We define a 1-form Θ by its linear action on δϕ, a tangent vector to Γ:
Z Z
3 0
Θ(δϕ) = d xW = d3 x ϕ,t δϕ. (4.35)
R3 R3
6
“Suitable” could be, for example, smooth solutions to the field equations with com-
pactly supported initial data.
4.2. HAMILTONIAN FORMULATION OF THE SCALAR FIELD 93
To get this result I used the divergence theorem and the requirement that
the solutions and their linearization vanish asymptotically.
Up to this point the phase space has been defined implicitly inasmuch as
we have not given an explicit parametrization of the set of solutions to the
94 CHAPTER 4. THE HAMILTONIAN FORMULATION
field equations. There are various ways to parametrize the space of solutions
depending upon how much analytic control you have over this space. Let
us use the most traditional parametrization, which relies upon the existence
of a well-posed initial value problem: for every pair of functions on R3 ,
(φ(~x), π(~x)) there is uniquely determined a solution ϕ(x) of the field equations
such that at a given initial time t = t0 :
ϕ(t0 , ~x) = φ(~x), ∂t ϕ(t0 , ~x) = π(~x). (4.41)
We can thus view a point in Γ as just a pair of functions (φ, π) on R3 . Using
this parametrization of solutions, from (4.35) and (4.36):
Z
Θ(δφ, δπ) = d3 x π δφ. (4.42)
R3
Z
Ω({δφ1 , δπ1 }, {δφ2 , δπ2 }) = d3 x (δ2 π δ1 φ − δ1 π δ2 φ) . (4.43)
R3
Hopefully you recognize the pattern familiar from particle mechanics where
Θ = pi dq i and Ω = dpi ∧dq i . Indeed, if one views the field as just a mechanical
system with an infinite number of degrees of freedom labeled by the spatial
point ~x, then one can view the integrations over ~x as the generalizations of the
various sums over degrees of freedom which occur in particle mechanics. For
this reason people often call φ(~x) the “coordinate” and π(~x) the “momentum”
for the scalar field. Although we shall not worry too much about precisely
what function spaces φ and π live in, it is useful to note that both of these
functions must vanish at spatial infinity if the symplectic structure is to be
defined.
The formula (4.43) makes it easy to check that the symplectic form is
non-degenerate as it should be. Can you construct a proof?
Problems:
5. Show that the symplectic structure (4.43) is non-degenerate. (Hint:
A 2-form Ω is non-degenerate if and only if Ω(u, v) = 0, ∀v implies
u = 0.)
I will now show that these equations can be put into Hamiltonian form,
∂φ δH
=
∂t δπ
∂π δH
=− , (4.46)
∂t δφ
where the Hamiltonian H generating time evolution along the foliation of
Minkowski spacetime by hypersurfaces t = const. is given by the energy as
defined in this inertial reference frame. In terms of our parametrization of Γ
using initial data:
Z
3 1 2 1 1 2 2
H[φ, π] = dx π + ∇φ · ∇φ + m φ + V (φ) , (4.47)
R3 2 2 2
where ∇ denotes the usual spatial gradient. In preparation for calculating
the Hamilton equations (4.46), let us consider the functional derivatives of
H. To do this we vary φ and π:
Z
d3 x π δπ + ∇φ · ∇δφ + m2 φ δφ + V 0 δφ .
δH = (4.48)
R3
Integrating by parts in the term with the gradients, and using the fact that
φ vanishes at infinity to eliminate the boundary terms, we get
Z
d3 x π δπ + −∇2 φ + m2 φ + V 0 δφ .
δH = (4.49)
R3
This means
δH
= π(~x), (4.50)
δπ(~x)
96 CHAPTER 4. THE HAMILTONIAN FORMULATION
and
δH
= −∇2 φ(~x) + m2 φ(~x) + V 0 (φ(~x)). (4.51)
δφ(~x)
You can now see that the Hamilton equations (4.46) are equivalent to the
field equations (4.33).
Finally, using standard techniques of classical mechanics it is possible to
deduce the Hamiltonian from the Lagrangian. Recall that the KG Lagrangian
is (in a particular inertial reference frame (t, ~x))
Z Z
3 3 1 2 2 2 2
L= d xL = dx ϕ − (∇ϕ) − m ϕ − V (ϕ) . (4.52)
R3 R3 2 ,t
At any fixed time t, view φ ≡ ϕ(t, ~x) and φ̇ ≡ ϕ,t (t, ~x) as independent fields
on R3 . Define the canonical momentum as the functional derivative of the
Lagrangian with respect to the velocity:
δL
π= = φ̇. (4.53)
δ φ̇
The Lagrangian now assumes the canonical form:
Z Z
3 3 1 2 1 1 2 2
L= d x π φ̇ − dx π + ∇φ · ∇φ + m φ + V (φ) , (4.54)
R3 R3 2 2 2
so that the Hamiltonian is given by the familiar Legendre transformation:
Z Z
3 3 1 2 1 1 2 2
H= d x π φ̇ − L = dx π + ∇φ · ∇φ + m φ + V (φ) .
R3 R3 2 2 2
(4.55)
Two important applications of this formula are the canonical Poisson bracket
relations,
[φ(~x), π(~y )] = δ(~x − ~y ), (4.67)
and the Hamilton equations:
∂φ(~x, t) ∂π(~x, t)
= [φ(~x), H], = [π(~x), H]. (4.68)
∂t ∂t
Problem:
7. Verify (4.67) follows from (4.66) and that (4.68) agrees with (4.44),
(4.45).
δH = [H, G] = 0. (4.72)
The field momentum is a constant of the motion for solutions of the scalar
field equation (4.33). We showed this earlier in the special case of the Klein-
Gordon field (V (ϕ) = 0), but it is not hard to see that this result generalizes
to the case of equation (4.33), as it must because the Lagrangian density still
has the spatial translation symmetry.
Problem:
8. Show that P (~v ) is a constant of the motion for solutions of the scalar
field equation (4.33).
100 CHAPTER 4. THE HAMILTONIAN FORMULATION
In terms of our parametrization of the phase space using initial data, the
field momentum along ~v is given by
Z
P (~v ) = d3 x πφ,i v i , (4.75)
R3
Let us compute the Poisson bracket of the field momentum with the Hamilto-
nian. There are a few ways to organize this computation, but let us emphasize
the role of P as an infinitesimal generator of translations. Our strategy is
to use (4.48) in conjunction with the infinitesimal change in (φ, π) under the
canonical transformation generated by P (~v ). We have (try it!)
and
δπ(~x) ≡ [π(~x), P (~v )] = v i π,i ≡ ~v · ∇π. (4.77)
This is indeed what one expects to be the change in a function under an
infinitesimal translation along ~v . The change in the KG Hamiltonian (4.47)
for any changes in the fields is given by (4.48). Using (4.76) and (4.77) we
get
Z
d3 x π ~v · ∇π + ∇φ · ∇(~v · ∇φ) + m2 φ ~v · ∇φ + V 0 ~v · ∇φ
δH =
3
ZR
3 1 2 1 1 2 2
= d x ∇ · ~v π + ∇φ · ∇φ + m φ + V (φ) . (4.78)
R3 2 2 2
Next, we use the divergence theorem to convert this integral to a surface
integral “at infinity”. The asymptotic decay of the fields needed to make the
energy finite in the first place then implies that the fields decay fast enough
such that this integral vanishes. Thus the Hamiltonian is invariant under
spatial translations. As already noted, this is equivalent to the fact that the
momentum P (~v ) is conserved. Of course, we already knew this from our
discussion of Noether’s theorem.
Here is an example for you to try.
Problem:
9. Consider the massless, free scalar field. Show that
Z
Q= d3 x π(~x) (4.79)
R3
4.5. PROBLEMS 101
4.5 PROBLEMS
2. Given L = L(q, q̇), give a formula for the EL equations linearized about
a solution q a = q a (t).
3. Show that the symplectic form given in (4.17) does not depend upon
time. (Hint: Differentiate (4.17) with respect to time and then use the
EL equations and their linearization.)
1 1
L = mq̇ 2 − mω 2 q 2 , q̈(t) = −ω 2 q(t). (4.80)
2 2
What does the symplectic structure look like when the space of solu-
tions to the EL equations is parametrized according to the following
formulas?
7. Verify (4.67) follows from (4.66) and that (4.68) agrees with (4.44),
(4.45).
Problems:
103
104 CHAPTER 5. ELECTROMAGNETIC FIELD THEORY
~ x, t) the electric
Show that for any function φ(~x, t) and vector field A(~
and magnetic fields defined by
~
~ = −∇φ − ∂ A ,
E ~ =∇×A
B ~ (5.5)
∂t
satisfy (5.2) and (5.4).
1
t0 = γ(t − vx), x0 = γ(x − vt), y 0 = y, γ=√ z 0 = z, ,
1 − v2
(5.7)
~ ~ ~ 0 ~0
the electric and magnetic fields change (E, B) → (E , B ), where
j α = (ρ, j i ), i = 1, 2, 3. (5.10)
where indices are raised and lowered with the usual Minkowski metric.
5.2. ELECTROMAGNETIC LAGRANGIAN 105
4. Show that the scalar and vector potentials, when assembled into the
4-potential
Aµ = (−φ, Ai ), i = 1, 2, 3, (5.12)
are related to the electromagnetic tensor Fµν by
Fµν = ∂µ Aν − ∂ν Aµ . (5.13)
Show that this formula for Fµν solves the homogeneous Maxwell equa-
tions Fαβ,γ + Fβγ,α + Fγα,β = 0.
∂xβ
A0α (x0 ) = Aβ (x(x0 )). (5.15)
∂xα0
In any case, A is called the “Maxwell field”, the “electromagnetic field”,
the “electromagnetic potential”, the “gauge field”, the “4-vector potential”,
the “U(1) connection”, and some other names as well, along with various
mixtures of these.
As always, having specified the geometric nature of the field, the field
theory can be defined by choosing a Lagrangian. To define the Lagrangian
1
It can be shown using techniques from the inverse problem of the calculus of variations
that there is no variational principle for Maxwell’s equations built solely from (E, ~ B)
~
(equivalently from Fαβ ) and their derivatives.
106 CHAPTER 5. ELECTROMAGNETIC FIELD THEORY
we introduce the field strength tensor F , also known as the “Faraday tensor”,
or as the “curvature” of the gauge field A. We write
F = Fαβ (x)dxα ⊗ dxβ , Fαβ = ∂α Aβ − ∂β Aα . (5.16)
The field strength is in fact a two-form (an anti-symmetric (02 ) tensor field):
Fαβ = −Fβα , (5.17)
and we can write
1 1
F = Fαβ (dxα ⊗ dxβ − dxβ ⊗ dxα ) = Fαβ dxα ∧ dxβ , (5.18)
2 2
where the anti-symmetric tensor product is known as the “wedge product”,
denoted with ∧. In terms of differential forms, the field strength is the
exterior derivative of the Maxwell field:
F = dA. (5.19)
This guarantees that dF = 0, which is equivalent to the homogeneous
Maxwell equations Fαβ,γ + Fβγ,α + Fγα,β = 0. So, the use of potentials
solves half the Maxwell equations; the only remaining Maxwell equations to
be considered are F αβ ,β = 4πj α .
The 6 independent components of F in an inertial Cartesian coordinate
chart (t, x, y, z) define the electric and magnetic fields as perceived in that
reference frame. Note, however, that all of the definitions given above are
in fact valid on an arbitrary spacetime manifold in an arbitrary system of
coordinates.
The Lagrangian for electromagnetic theory – on an arbitrary spacetime
(M, g) – can be defined by the n-form (where n = dim(M )),
1
L = − F ∧ ∗F = L dx1 ∧ dx2 ∧ · · · ∧ dxn , (5.20)
4
where ∗F is the Hodge dual defined by the spacetime metric g. In terms of
components in a coordinate chart we have
1
(∗F )αβ = αβ γδ Fγδ , (5.21)
2
and the Lagrangian density given by
1√
L=− −gF αβ Fαβ , (5.22)
4
5.2. ELECTROMAGNETIC LAGRANGIAN 107
where
F αβ = g αγ g βδ Fγδ . (5.23)
Of course, we can – and usually will – restrict attention to the flat spacetime
in the standard Cartesian coordinates for explicit computations. It is always
understood that F is built from A in what follows.
Let us compute the Euler-Lagrange derivative of L. For simplicity we
will work on flat spacetime in inertial Cartesian coordinates so that
We have
1
δL = − F αβ δFαβ
2
1
= − F αβ (δAβ,α − δAα,β )
2
= −F αβ δAβ,α
= F αβ ,α δAβ + Dα −F αβ δAβ .
(5.25)
E β (L) = F αβ ,α , (5.26)
F αβ ,α = 0. (5.27)
There are some equivalent expressions of the field equations that are worth
knowing about. First of all, we have that
We write this using the wave operator (which acts component-wise on the
1-form A) and the operator
via
Aβ − (div A),β = 0. (5.31)
You can see that this is a modified wave equation.
A more sophisticated expression of the field equations, which is manifestly
valid on any spacetime, uses the technology of differential forms. Recall that
on a spacetime one has the Hodge dual, which identifies the space of p-forms
with the space of n − p forms. This mapping is denoted by
α → ∗α. (5.32)
So that
∗d∗F = 0 ⇐⇒ Fαβ ,α = 0. (5.37)
5.2. ELECTROMAGNETIC LAGRANGIAN 109
Problems:
F αβ ,α = −4πj β . (5.40)
(2) Use the results from the preceding problem to show that the Maxwell
equations with sources have no solution unless the vector field repre-
senting the sources is divergence-free:
∂α j α = 0. (5.42)
(3) Show that this condition is in fact the usual continuity equation
representing conservation of electric charge.
8. Show that E ~ ·B
~ is relativistically invariant (unchanged by a Lorentz
transformation). Express it in terms of potentials and show that it is
just a divergence, with vanishing Euler-Lagrange expression.
110 CHAPTER 5. ELECTROMAGNETIC FIELD THEORY
F = dA, (5.43)
A → A0 = A + dΛ, (5.44)
where Λ : M → R, then
Problem:
9. For any function Λ = Λ(x), define
A0α = Aα + ∂α Λ. (5.46)
Show that
A0α,β − A0β,α = Aα,β − Aβ,α . (5.47)
Show that, in terms of the scalar and vector potentials, this gauge
transformation is equivalent to
φ → φ0 = φ − ∂t Λ, ~→A
A ~0 = A
~ + ∇Λ. (5.48)
A → A0 (5.49)
we have
F → F. (5.50)
5.3. GAUGE SYMMETRY 111
Since Λ is an arbitrary smooth function, we can choose the first two deriva-
tives of Λ to vanish on the initial hypersurface so that A0 and A are distinct
solutions with the same initial data.
To uniquely determine the potential A from Cauchy data and the Maxwell
equations one has to add additional conditions on the potential beyond the
field equations. This is possible since one can adjust the form of A via gauge
transformations. It is not too hard to show that one can gauge transform
any given potential into one which satisfies the Lorenz gauge condition:
∂ α Aα = 0. (5.54)
To see this, take any potential, say, Ã and gauge transform it to a potential
A = Ã + dΛ such that the Lorenz gauge holds; this means
∂ α ∂α Λ = −∂ α Ãα . (5.55)
Viewing the right-hand side of this equation as given, we see that to find
such a gauge transformation amounts to solving the wave equation with a
given source, which can always be done.
In the Lorenz gauge the Maxwell equations are just the usual, hyper-
bolic wave equation for each inertial-Cartesian component of the 4-vector
potential,
Aα = −4πjα . (5.56)
It is therefore tempting to suppose that one can thus identify the EM field
theory with 4 copies of a massless KG field theory. Things are a little more
interesting than that: the 4-potential still must satisfy (5.54), and even in the
Lorenz gauge the potentials are not uniquely determined! Pick any solution
Λ of the wave equation Λ = 0 and use Λ to make a gauge transformation.
You can easily check that the transformed potentials still satisfy the condition
(5.54). It is possible to show that the Lorenz condition along with this
residual gauge freedom ultimately permits the elimination of two functions
worth of information from A.2 One says that the electromagnetic field has
“two degrees of freedom per spatial point”; this is intimately related to the
two helicity states of photons in the associated quantum theory.
2 ~ = 0.
In detail, one can choose Λ such that A0 = 0 and, therefore, so that ∇ · A
5.4. NOETHER’S SECOND THEOREM IN ELECTROMAGNETIC THEORY113
so that δL = 0.
For any variation, the first variational identity is
δL = E β δAβ + Dα −F αβ δAβ ,
(5.63)
where
E β = F αβ ,α . (5.64)
For a variation induced by an infinitesimal gauge transformation we know
that δL = 0, so the variational identity tells us that
0 = E β ∂β σ + Dα −F αβ ∂β σ ,
(5.65)
which is valid for any function σ. Now we take account of the fact that the
function σ is arbitrary. We rearrange the derivatives of σ to get them inside
a divergence:
0 = −Dβ E β σ + Dα −F αβ ∂β σ + F αβ ,β σ
(5.66)
Restrict this equation to a potential A = A(x), then integrate the result over
a spacetime region R:
Z Z
αβ
−F αβ ∂β σ + F αβ ,β σ dΣα .
0=− F ,αβ σ + (5.67)
R ∂R
This must hold for any function σ; we can use the fundamental theorem
of variational calculus to conclude that the Euler-Lagrange equations must
satisfy the differential identity
Dβ E β = 0, (5.68)
which you proved directly in a previous homework problem. Note that this
says the Euler-Lagrange expression is divergence-free, and that this holds
whether or not the field equations are satisfied – it is an identity arising due
to the gauge symmetry of the Lagrangian.
Compare our results above to Noether’s first theorem. We have seen that
the gauge symmetry – being a continuous variational symmetry – leads to a
divergence-free vector field, as it must by Noether’s first theorem. But we
now have a new ingredient: the gauge symmetry is built from an arbitrary
function of all the independent variables xα so that the gauge transformation
can be localized to an arbitrary location in spacetime. This leads to the vector
5.5. NOETHER’S SECOND THEOREM 115
Problem:
10. Consider the electromagnetic field coupled to sources with the La-
grangian density
1
Lj = − F αβ Fαβ + 4πj α Aα (5.69)
4
Show that this Lagrangian is gauge invariant (up to a divergence) if
and only if the spacetime vector field j α is chosen to be divergence-free.
What is the Noether identity in this case?
δL = Ea (L)δϕa + Dα η α , (5.70)
δL = Dα W α (Λ), (5.73)
116 CHAPTER 5. ELECTROMAGNETIC FIELD THEORY
0 = δL − Dα W α = Ea (L)[D(Λ)]a + Dα (η α − W α ), (5.74)
Now imagine integrating by parts each term in the linear operator [D(Λ)]a so
that all derivatives of Λ are removed. The boundary terms that arise vanish
with our boundary conditions on ΛA . This process defines the formal adjoint
D∗ of the linear differential operator D:
Z Z
a
Ea [D(Λ)] = ΛA [D∗ (E)]A . (5.76)
R R
You can easily check out this argument via our Maxwell example. The
gauge transformation is defined by the exterior derivative on functions:
[D(Λ)]α = ∂α Λ. (5.79)
5.6. THE CANONICAL ENERGY-MOMENTUM TENSOR 117
W α = 0. (5.81)
Dα E α = 0 (5.84)
We have then4
δAα = bβ Aα,β . (5.87)
This implies that
δFµν = bα Fµν,α (5.88)
and hence the translations define a divergence symmetry:
1 1
δL = − bγ F αβ Fαβ,γ = Dγ (− bγ F αβ Fαβ ). (5.89)
2 4
Recalling the variational identity:
δL = F αβ ,α δAβ + Dα −F αβ δAβ ,
(5.90)
Dα j α = 0, (5.92)
Problem:
11. Verify equations (5.87)–(5.92).
Since this conservation law exists for each constant vector bα , we can sum-
marize these conservation laws using the canonical energy-momentum tensor
α αβ 1 α µν
Tγ = F Aβ,γ − δγ F Fµν , (5.93)
4
which satisfies
Dα Tβα = 0, (5.94)
modulo the field equations. We view the energy-momentum tensor as a
collection of conserved currents labeled by the index γ.
4
It is worth noting that this formula is not gauge-invariant; it really only defines the
change in the fields due to a translation modulo a gauge transformation. We will address
this issue soon.
5.7. IMPROVED MAXWELL ENERGY-MOMENTUM TENSOR 119
A −→ A + dΛ (5.95)
we have
Tβα −→ Tβα + F αµ ∂µ ∂β Λ. (5.96)
In order to see what to do about this, we need to use some of the flexibility
we have in defining conserved currents. This is our next task.
It is possible to show that all local and gauge invariant expressions must
depend on the vector potential only through the field strength. Consequently,
the currents are not gauge invariant because of the explicit presence of the
potentials A. With that in mind we write
1 α µν
Tγ = F Fγβ − δγ F Fµν − F αβ Aγ ,β
α αβ
4
1 α µν
= F Fγβ − δγ F Fµν − Dβ (F αβ Aγ ) + Aγ Dβ F αβ . (5.97)
αβ
4
According to §3.20, the last two terms are trivial conservation laws. So,
modulo a set of trivial conservation laws, the canonical energy-momentum
tensor takes the gauge-invariant form
1
Tγα = F αβ Fγβ − δγα F µν Fµν . (5.98)
4
This tensor is called the “gauge-invariant energy-momentum tensor” or the
“improved energy-momentum tensor” or the “general relativistic energy-
momentum tensor”. The latter term arises since this energy-momentum
tensor serves as the source of the gravitational field in general relativity and
can be derived using the variational principle of that theory.
The improved energy-momentum tensor has another valuable feature rel-
ative to the canonical energy-momentum tensor (besides gauge invariance).
The canonical energy-momentum tensor,
α αβ 1 α µν
Tγ = F Aβ,γ − δγ F Fµν , (5.99)
4
120 CHAPTER 5. ELECTROMAGNETIC FIELD THEORY
1
Tγα = F αβ Fβγ − δγα F µν Fµν . (5.103)
4
is symmetric:
Tαβ = Tβα . (5.104)
Why is all this important? Well, think back to the KG equation. There,
you will recall, the conservation of angular momentum, which stems from
the symmetry of the Lagrangian with respect to the Lorentz group, comes
from the currents
δL = E(L) + Dα η α (5.106)
5.8. THE HAMILTONIAN FORMULATION OF ELECTROMAGNETISM.121
1
W α = − bα Fµν F µν , (5.107)
4
along with
The lack of gauge invariance snuck into the calculation via the formula for
the change in A under an infinitesimal translation, δAβ = bγ Aβ,γ . We de-
stroyed gauge invariance with this formula since its right hand side is not
gauge invariant. A gauge invariant formula for the change of A under an
infinitesimal translation can be gotten by accompanying the translation with
a gauge transformation:
where the integral takes place at some chosen value for x0 ≡ t and we used
the spatial divergence theorem along with boundary conditions at spatial
infinity to get the second equality. The 1-form Θ which is normally used to
construct the symplectic 2-form is then defined by
Z Z
3 β0
Θ(δA) = d x F δAβ = d3 x F i0 δAi , (5.114)
R3 R3
Problem:
12. Show that the 2-form (5.115) does not depend upon the time at which
it is evaluated.
The new feature that appears here is that the 2-form Ω is in fact degen-
erate. This means that there exists a vector ~v such that Ω(~u, ~v ) = 0 for all
~u. Let us see how this happens in detail.
5.8. THE HAMILTONIAN FORMULATION OF ELECTROMAGNETISM.123
=− d3 x Λ ∂i (∂t δ1 Ai − ∂ i δ1 At ), (5.117)
R3
where integration by parts and the divergence theorem were used to get the
last equality. The boundary term “at infinity” vanishes since Λ has compact
support. Next, recall that the field variations δA represent tangent vectors to
the space of solutions of the Maxwell equations (5.111) and so are solutions
to the linearized equations. Because the Maxwell equations are linear, their
linearization is mathematically the same:
∂ β (∂α δAβ − ∂β δAα ) = 0. (5.118)
Setting α = 0 we get
∂ i (∂t δAi − ∂i δAt ) = 0. (5.119)
This means that (5.117) vanishes and we conclude that the “pure gauge”
tangent vectors (5.116) are degeneracy directions for the 2-form Ω.
Degenerate 2-forms are often called “pre-symplectic” because there is
a canonical procedure for extracting a unique symplectic form on a smaller
space from a pre-symplectic form. We will not develop this elegant geometric
result here. Instead, we will proceed in a useful if more roundabout route
by examining the Hamiltonian formulation of the theory that arises when we
parametrize the space of solutions to the field equations (5.111) with initial
data.
potential and the conjugate momentum is then given by minus the electric
field. We define
Qi (~x) = Ai (t = 0, ~x), P i (~x) = F i0 (t = 0, ~x). (5.120)
Because of the definition of the field strength tensor in terms of the vector
potential, we have
Pi = ∂t Ai − ∂i At , (5.121)
so we can write
Q̇i = Pi + ∂i At . (5.122)
Next, consider the equations of motion with time and space (as given in an
inertial reference frame) explicitly separated:
∂t F αt + ∂i F αi = 0, (5.123)
which yields four equations:
∂t F jt + ∂i F ji = 0 (5.124)
∂i F ti = 0, (5.125)
These equations allow us to view the Maxwell equations as determining a
curve (Qi = Qi (t, ~x), P i = P i (t, ~x)) in the space of initial data along with
a constraint on the canonical variables. Equations (5.122) and (5.124) are
equivalent to the evolution equations:
Q̇i = δij P j + ∂i At , (5.126)
Ṗ j = ∂j F ij , (5.127)
Problem:
5.8. THE HAMILTONIAN FORMULATION OF ELECTROMAGNETISM.125
13. Show that if the constraint (5.128) holds at one time and the canonical
variables evolve in time according to (5.126), (5.127) then (5.128) will
hold at any other time. (Hint: Consider the time derivative of (5.128).)
Notice that the equations (5.126), (5.127) are evolution equations for
(Qi , P i ) only. The time component At is not determined by these equations,
it simply defines a gauge transformation of the Qi as time evolves.
Equations (5.126), (5.127) and (5.128) determine solutions to the Maxwell
equations as follows. Specify 6 functions on R3 , namely (qi (~x), pi (~x)), where
pi is divergence-free, ∂i pi = 0. Pick a function At (t, ~x) any way you like.
Solve the evolution equations (5.126), (5.127) subject to the initial conditions
(Qi (0) = q i , Pi (0) = pi ) to get (Qi (t), P i (t)). In the given inertial reference
frame define
Fti = P i , Fij = ∂i Qj − ∂j Qi . (5.129)
The resulting field strength tensor Fαβ satisfies the (source-free) Maxwell
equations, as you can easily verify.
Viewing the spacetime fields as Qi (t, ~x), P i (t, ~x), φ(t, ~x), the Lagrangian can
be viewed as a functional of (Qi , P j , φ) and can be written in the Hamiltonian
“pq̇ − H” form (exercise):
Z
3 i 1 i 1 ij i
L[Q, P, φ] = d x P Q̇i − (Pi P + Fij F ) + P ∂i φ
R3 2 2
Z
3 i 1 i 1 ij i
= d x P Q̇i − (Pi P + Fij F ) + φ∂i P (5.134)
R3 2 2
To get the second equality I integrated by parts and used the divergence
theorem on the last term. At each time t, we will assume that (Qi , P i )
vanish sufficiently rapidly at infinity so that boundary term vanishes. As
usual, the EL equations (using functional derivatives) for P i reproduce the
definition (5.132) so that the EL equations for (At , Qi ) then yield the Maxwell
equations. Notice that φ enters as a Lagrange multiplier enforcing the (Gauss
law) constraint on the canonical variables,
∂i P i = 0, (5.135)
Problem:
δL d δL δL d δL δL
i
− = 0, − = 0, = 0, (5.136)
δP dt δ Ṗ i δQi dt δ Q̇i δφ
The first two terms are what you might expect: they define the energy of the
electromagnetic field (once you recognize that 14 Fij F ij = 12 B 2 is the magnetic
energy density). You can see that the first term (the electric energy) is akin to
the kinetic energy of a particle, while the second term (the magnetic energy)
is akin to the potential energy of a particle. The term we want to focus on
is the third term – what’s that doing there? Well, first of all note that this
term does not affect the value of the Hamiltonian provided the canonical
variables satisfy the constraint (5.135). Secondly, let us consider Hamilton’s
equations:
δH
Q̇i = = Pi − ∂i φ, (5.138)
δP i
δH
Ṗ i = − = ∂i F ij . (5.139)
δQi
From (5.138), which is secretly the relation between the electric field and
the potentials, you can see that the last term in the Hamiltonian is precisely
what is needed to generate the gauge transformation term we already found
in (5.126).
Problem:
15. Using the Poisson brackets
Z
3 δM δN δM δN
[M, N ] = dx − , (5.140)
R3 δQi (~x) δP i (~x) δP i (~x) δQi (~x)
show that Z
G=− d3 x Λ(~x) ∂i P i (5.141)
R3
is the generating function for gauge transformations
δQi = [Qi , G] = ∂i Λ, δP i = [P i , G] = 0. (5.142)
5.9 PROBLEMS
1. Maxwell’s equations for the electric and magnetic field (E, ~ B)
~ associ-
ated to charge density and current density (ρ, ~j) are given by
~ = 4πρ,
∇·E
~ = 0,
∇·B
~
~ − 1 ∂ E = 4πσ,
∇×B
c ∂t
~
∇×E~ + ∂ B = 0.
∂t
~ x, t) the electric
Show that for any function φ(~x, t) and vector field A(~
and magnetic fields defined by
~
~ = −∇φ − 1 ∂ A ,
E ~ =∇×A
B ~
c ∂t
satisfy (5.2) and (5.4).
E x0 = E x , E y0 = γ(E y − vB z ), E z0 = γ(E z + vB y )
B x0 = B x , B y0 = γ(B y + vE z ), B z0 = γ(B z − vE y ).
√
(Here γ = 1/ 1 − v 2 .) Show that this is equivalent to saying that Fµν
are the components of a spacetime tensor of type 02 . Show that the
two quantities E~ ·B~ and E 2 − B 2 do not change under the boost.
5.9. PROBLEMS 129
j α = (ρ, j i ), i = 1, 2, 3.
where indices are raised and lowered with the usual Minkowski metric.
4. Show that the scalar and vector potentials, when assembled into the
4-potential
Aµ = (−φ, Ai ), i = 1, 2, 3,
are related to the electromagnetic tensor Fµν by
Fµν = ∂µ Aν − ∂ν Aµ .
Show that this formula for Fµν solves the homogeneous Maxwell equa-
tions Fαβ,γ + Fβγ,α + Fγα,β = 0.
F αβ ,α = −4πj β .
These are the Maxwell equations with prescribed electric sources having
a charge density ρ and current density ~j, where
j α = (ρ, ~j).
130 CHAPTER 5. ELECTROMAGNETIC FIELD THEORY
Use the results from the preceding problem to show that the Maxwell
equations with sources have no solution unless the vector field repre-
senting the sources is divergence-free:
∂α j α = 0.
Show that this condition is in fact the usual continuity equations rep-
resenting conservation of electric charge.
8. Show that E ~ ·B
~ is relativistically invariant (unchanged by a Lorentz
transformation). Express it in terms of potentials and show that it is
just a divergence, with vanishing Euler-Lagrange expression.
A0α = Aα + ∂α Λ.
Show that
A0α,β − A0β,α = Aα,β − Aβ,α .
Show that, in terms of the scalar and vector potentials, this gauge
transformation is equivalent to
φ → φ0 = φ − ∂t Λ, ~→A
A ~0 = A
~ + ∇Λ.
10. Consider the electromagnetic field coupled to sources with the La-
grangian density
1
Lj = − F αβ Fαβ + 4πj α Aα
4
Show that this Lagrangian is gauge invariant (up to a divergence) if
and only if the spacetime vector field j α is chosen to be divergence-free.
What is the Noether identity in this case?
5.9. PROBLEMS 131
12. Show that the 2-form (5.115) does not depend upon the time at which
it is evaluated.
13. Show that if the constraint (5.128) holds at one time and the canonical
variables evolve in time according to (5.126), (5.127) then (5.128) will
hold at any other time. (Hint: Consider the time derivative of (5.128).)
δL d δL δL d δL δL
i
− = 0, − = 0, = 0,
δP dt δ Ṗ i δQi dt δ Q̇i δφ
show that Z
G=− d3 x Λ(~x) ∂i P i
R3
is the generating function for gauge transformations
δQi = [Qi , G] = ∂i Λ, δP i = [P i , G] = 0.
132 CHAPTER 5. ELECTROMAGNETIC FIELD THEORY
Chapter 6
Scalar Electrodynamics
Let us have an introductory look at the field theory called scalar electrody-
namics, in which one considers a coupled system of Maxwell and charged KG
fields. There are an infinite number of ways one could try to couple these
fields. There is one particularly interesting way, physically speaking, and
this is the one we shall be exploring. Mathematically, too, this particular
coupling has many interesting features which we shall explore.
To understand the motivation for the postulated form of scalar electro-
dynamics, it is easiest to proceed via Lagrangians. For simplicity we will
restrict attention to flat spacetime in inertial Cartesian coordinates, but our
treatment is easily generalized to an arbitrary spacetime.
1
Lj = − F αβ Fαβ + 4πj α Aα . (6.1)
4
Incidentally, given the explicit appearance of Aα , one might worry about
the gauge symmetry of this Lagrangian. But it is easily seen that the gauge
transformation is a divergence symmetry of this Lagrangian. Indeed, under
133
134 CHAPTER 6. SCALAR ELECTRODYNAMICS
idea is that the EL equations for A will give the Maxwell equations with the
KG current as the source. The EL equations for the scalar field will now
involve A, but that is reasonable since we expect the presence of the electro-
magnetic field to affect the sources. But there is one big problem with this
Lagrangian: it is no longer gauge invariant! Recall that the gauge invariance
of the Maxwell Lagrangian with prescribed sources made use of the fact that
the current was divergence-free. But now the current is only divergence-free
when the field equations hold. The key to escaping this difficulty is to let
the KG field participate in the gauge symmetry. This forces us to modify
the Lagrangian as we shall now discuss.
Problem:
The Lagrangian (6.8) still admits the U (1) phase symmetry of the charged
KG theory, but because this Lagrangian depends explicitly upon A it will
not be gauge invariant unless we include a corresponding transformation of
ϕ. We therefore extend the gauge transformation to be:
Aα −→ Aα + ∂α Λ, (6.9)
Dα ϕ −→ e−iqΛ Dα ϕ, (6.11)
Dα ϕ∗ −→ eiqΛ Dα ϕ∗ . (6.12)
For this reason Dα is sometimes called the gauge covariant derivative. There
is a nice geometric interpretation of this covariant derivative, which we shall
discuss later. For now, because of this “covariance” property of D, it follows
that the Lagrangian (6.8) is gauge invariant with respect to the transforma-
tions (6.9), (6.10).
The Lagrangian for scalar electrodynamics is taken to be
1
LSED = − F αβ Fαβ − Dα ϕ∗ Dα ϕ − m2 |ϕ|2 . (6.13)
4
We now discuss some important structural features of this Lagrangian.
If we expand the gauge covariant derivatives we see that
1
LSED = − F αβ Fαβ −∂ α ϕ∗ ∂α ϕ−m2 |ϕ|2 +iqAα ϕ∗ ∂ α ϕ − ϕ∂ α ϕ∗ + iqAα |ϕ|2 .
4
(6.14)
This Lagrangian is the sum of the electromagnetic Lagrangian, the free
charged KG Lagrangian, and a j · A “interaction term”. The vector field
that is contracted with A is almost the conserved current (6.4), but is modi-
fied by the last term involving the square of the gauge field, which is needed
6.2. MINIMAL COUPLING: THE GAUGE COVARIANT DERIVATIVE137
for invariance under the gauge transformation (6.9), (6.10) and for the cur-
rent to be conserved when the new form of the field equations are satisfied.
The EL equations for the Maxwell field are of the desired form:
∂β F αβ = −4πJ α , (6.15)
where the current is defined using the covariant derivative instead of the
ordinary derivative:
iq
J α = − (ϕ∗ Dα ϕ − ϕDα ϕ∗ ) . (6.16)
4π
As you will verify in the problem below, this current can be derived from
Noether’s first theorem applied to the U (1) phase symmetry of the La-
grangian (6.13). Thus we have solved the gauge invariance problem and ob-
tained a consistent version of the Maxwell equations with conserved sources
using the minimal coupling prescription.
Problems:
2. Derive (6.15), (6.16) from (6.13).
3. Verify that (6.16) is the Noether current coming from the U (1) sym-
metry of the Lagrangian and that it is indeed conserved when the field
equations for ϕ hold.
One more interesting feature to ponder: the charged current (6.16) serving
as the source for the Maxwell equations is built from the KG field and the
Maxwell field. Physically this means that one cannot say the charge “exists”
only in the KG field. In an interacting system the division between source
fields and fields mediating interactions is somewhat artificial and arbitrary.
This is physically reasonable, if perhaps a little unsettling. Mathematically,
this feature stems from the demand of gauge invariance. Just like the vector
potential, the KG field is no longer uniquely defined - it is subject to a gauge
transformation as well! In the presence of interaction, the computation of
the electric charge involves a gauge invariant combination of the KG and
electromagnetic field. To compute, say, the electric charge contained in a
volume V one should take a solution (A, ϕ) of the coupled Maxwell-KG
equations and substitute it into
Z
1
QV = d3 x iq (ϕ∗ D0 ϕ − ϕD0 ϕ∗ ) . (6.17)
4π V
This charge is conserved and gauge invariant.
138 CHAPTER 6. SCALAR ELECTRODYNAMICS
Aµ −→ Aµ + ∂µ α, (6.23)
and
Dα ϕ∗ := (∂α − iqAα )ϕ∗ , (6.25)
which satisfies
Dµ (e−iqα(x) ϕ) = e−iqα(x) Dµ ϕ. (6.26)
Then with the Lagrangian modified via
∂µ ϕ → Dµ ϕ, (6.27)
so that
LKG = −Dα ϕ∗ Dα ϕ − m2 |ϕ|2 , (6.28)
we get the local U (1) symmetry, as shown previously. Thus the minimal
coupling rule that we invented earlier can be seen as a way of turning the
global U (1) symmetry into a local U (1) gauge symmetry. One also obtains
the satisfying mental picture that the electromagnetic interaction of charges
is the principal manifestation of this local phase symmetry in nature. Thus
140 CHAPTER 6. SCALAR ELECTRODYNAMICS
Problem:
In this way we have an interacting theory designed by local U (1) gauge sym-
metry. The parameter q, which appears via the gauge covariant derivative,
is a “coupling constant” and characterizes the strength with which the elec-
tromagnetic field couples to the charged aspect of the KG field. In the limit
in which q → 0 the theory becomes a decoupled juxtaposition of the non-
interacting (or “free”) charged KG field theory and the non-interacting (free)
6.3. GLOBAL AND LOCAL SYMMETRIES 141
where
iq ∗ α
Jα = − (ϕ D ϕ − ϕDα ϕ∗ ) . (6.37)
4π
Evidently, J is divergence-free when the scalar field equations of motion hold.
J is the conserved Noether current J corresponding to the electric charge
carried by the scalar field. This is the current that serves as source for the
Maxwell field. The presence of the gauge field renders the Noether current
suitably “gauge invariant”, that is, insensitive to the local U (1) transforma-
tion. It also reflects the fact that the equations of motion for ϕ, which must
be satisfied in order for the current to be conserved, depend upon the Maxwell
field as is appropriate since the electromagnetic field affects the motion of its
charged sources.
By construction, the theory of scalar electrodynamics admits the local
U (1) gauge symmetry. With α(x) being any function, the symmetry is
Aµ → Aµ + ∂µ α(x). (6.39)
There is a corresponding Noether identity (Noether’s second theorem, re-
member?). To compute it we consider an infinitesimal gauge transformation:
where Eϕ and Eϕ∗ are the scalar field EL expressions and E µ is the gauge
field EL expression. If we integrate this relation over a compact region and
choose α(x) to vanish at the boundary of this region, then the divergence
term vanishes. We can integrate by parts in the third term to get the Noether
identity
Dµ E µ + iq(ϕEϕ − ϕ∗ Eϕ∗ ) = 0. (6.43)
(Notice that this identity does not follow if the gauge transformation includes
a global part. ) The terms involving the EL expressions for the KG field are
the same as arise in the identity (6.36). Thus the Noether identity (6.43) can
also be written as
Dα Dβ F αβ = 0. (6.44)
Thanks to this Noether identity we can obtain – again! – the conservation
law of electric charge. We have the electromagnetic field equation
E β ≡ Dα F αβ − 4πJ β = 0. (6.45)
iq ∗ α
Jα = − (ϕ D ϕ − ϕDα ϕ∗ ) , (6.46)
4π
features in the Maxwell equations via the EL equation E α = 0 for the gauge
field Aα , where
E β = F,α
αβ
− 4πJ β . (6.47)
6.4. A LOWER-DEGREE CONSERVATION LAW 143
The relation between these 2 quantities can be obtained using Stokes theo-
rem: Z Z Z Z
0
χ −χ= ω− ω= ω= dω, (6.61)
S0 S S 0 −S V
0
where ∂V = S − S. In particular, if ω is a closed p-form, that is, dω = 0,
then χ = χ0 and the integral χ is independent of the choice of the space S in
the sense that χ is unchanged by any continuous deformation of S. A simple
illustration of this result is the following.
Problem:
5. Consider two concentric circles of radii a and b in the “x-y plane”. Let
V be the area between the circles so that its boundary S consists of the
concentric circles. In Cartesian coordinates with origin at the center of
~ be a vector field defined in V by
the circles, let A
~ = xŷ − yx̂ .
A (6.62)
x2 + y 2
ϕ : M → C. (6.63)
4
Notice that this is the “global” version of the U (1) symmetry transformation.
5
To read more about such things, have a look at arxiv.org/abs/hep-th/9706092 .
6.5. SCALAR ELECTRODYNAMICS AND FIBER BUNDLES 147
π: E → M (6.64)
where
π −1 (x) = C, x ∈ M. (6.65)
M is called the base space. The space π −1 (x) ≈ C is the fiber over x. Since
C is a vector space, this type of fiber bundle is called a vector bundle. For
us, M = R4 and it can be shown that for a contractible base space such as
Rn there is always a (non-unique) diffeomorphism that makes possible the
global identification:
E ≈ M × C. (6.66)
For a general base space M , such an identification will only be valid locally.
Next, recall that a cross section of E (often just called a “section”) is a
map (or graph)
σ: M → E (6.67)
satisfying
π ◦ σ = idM . (6.68)
Using coordinates (x, z) adapted to (6.66), we can identify a KG field ϕ(x)
with the cross section
σ(x) = (x, ϕ(x)). (6.69)
Thus, given the identification E ≈ M × C, we see that the bundle point
of view just describes the geometric setting of our theory: complex valued
functions on R4 . For the purposes of this discussion the most interesting issue
is that this identification is far from unique. Let us use coordinates (xα , z)
for E, where xα ∈ R4 and z ∈ C. Each set of such coordinates provides an
identification of E with M × C. Since we use a fixed (flat) metric on M ,
one can restrict attention to inertial Cartesian coordinates on M , in which
case one can only redefine xα by a Poincaré transformation. What is more
interesting for us in this discussion is the freedom to redefine the way that
the complex numbers are “glued” to each spacetime event.
Recall that to build the Lagrangian for the charged KG field we also had
to pick an inner product on the vector space C; of course we just used the
standard one
hz, wi = z ∗ w. (6.70)
148 CHAPTER 6. SCALAR ELECTRODYNAMICS
z → e−iα z, α ∈ R. (6.71)
We can make this change of coordinates on C for each fiber so that on π −1 (x)
we make the transformation
z → e−iα(x) z. (6.72)
There is no intrinsic way to compare points on different fibers, and this fact
reflects itself in the freedom to redefine our labeling of those points in a way
that can vary from fiber to fiber. We have seen this already; the change of
fiber coordinates z → e−iα(x) z corresponds to the gauge transformation of
the charged KG field:
ϕ(x) → e−iα(x) ϕ(x). (6.73)
When building a field theory of the charged KG field we need to take
derivatives. Now, to take a derivative means to compare the value of ϕ at
two neighboring points on M . From our fiber bundle point of view, this
means comparing points on two different fibers. Because this comparison is
not defined a priori, there is no natural way to take derivatives of a section of
a fiber bundle. This is closely related to the fact that the ordinary derivative
of the KG field does not transform homogeneously under a gauge transfor-
mation. Thus, for example, to say that a KG field is a constant, ∂α ϕ = 0,
is not an intrinsic statement since a change in the bundle coordinates will
negate it.
A definition of derivatives of sections of the fiber bundle requires the in-
troduction of additional structure beyond the bundle and the metric. (One
often introduces this structure implicitly!) This additional structure is called
a connection and the resulting notion of derivative is called the covariant
derivative defined by the connection. A connection can be viewed as a def-
inition of how to compare points on neighboring fibers. If you are differ-
entiating in a given direction, the derivative will need to associate to that
direction a linear transformation (actually, a phase transformation) which
“aligns” the vector spaces/fibers and allows us to compare them. Since the
derivative involves an infinitesimal motion in M , it turns out that this fiber
transformation is an infinitesimal phase transformation, which involves mul-
tiplication by a pure imaginary number (think: eiα = 1 + iα + . . . ). So,
6.5. SCALAR ELECTRODYNAMICS AND FIBER BUNDLES 149
To continue the analogy with differential geometry a bit further, you see that
the field ϕ is playing the role of a vector, with its vector aspect being the
fact that it takes values in the vector space C and transforms homogeneously
under the change of fiber coordinates, that is, the gauge transformation. The
complex conjugate can be viewed as living in the dual space to C, so that it is
150 CHAPTER 6. SCALAR ELECTRODYNAMICS
iq ∗ µ
Jµ = − (ϕ D ϕ − ϕDµ ϕ∗ ) (6.77)
4π
is divergence free with respect to the ordinary derivative, which is the correct
covariant derivative on “scalars”.
Do we really need all this fancy mathematics? Perhaps not. But, since all
the apparatus of gauge symmetry, covariant derivatives, etc., which show up
repeatedly in field theory, arises so naturally from this geometric structure,
it is clear that this is the right way to be thinking about gauge theories.
Moreover, there are certain results that would, I think, be very hard to come
by without using the fiber bundle point of view. I have in mind certain
important topological structures that can arise via global effects in classical
and quantum field theory. These topological structures are, via the physics
literature, appearing in the guise of “monopoles” and “instantons”. Such
structures would play a very nice role in a second semester for this course, if
there were one.
6.6 PROBLEMS
3. Verify that (6.16) is the Noether current coming from the U (1) sym-
metry of the Lagrangian and that it is indeed conserved when the field
equations for ϕ hold.
5. Consider two concentric circles of radii a and b in the “x-y plane”. Let
V be the area between the circles so that its boundary S consists of the
concentric circles. In Cartesian coordinates with origin at the center of
~ be a vector field defined in V by
the circles, let A
~ = xŷ − yx̂ .
A (6.78)
x2 + y 2
~ vanishes in V . Show that the line integral of
Show that the curl of A
~ around either of the circles gives the same result, independent of a
A
and b.
152 CHAPTER 6. SCALAR ELECTRODYNAMICS
Chapter 7
Spontaneous symmetry
breaking
We now will take a quick look at some of the classical field theoretic underpin-
nings of “spontaneous symmetry breaking” (SSB) in quantum field theory.
Quite generally, SSB can be a very useful way of thinking about phase tran-
sitions in physics. In particle physics, SSB is used, in collaboration with the
“Higgs mechanism”, to give masses to gauge bosons (and other elementary
particles) without destroying gauge invariance.
153
154 CHAPTER 7. SPONTANEOUS SYMMETRY BREAKING
ϕ̃ = ϕ + const. (7.2)
∂ α ∂α ϕ = 0 (7.3)
Problem:
1. While every symmetry of a Lagrangian is a symmetry of its EL equa-
tions, it is not true that every symmetry of the field equations is a
symmetry of the Lagrangian. Consider the massless KG field. Show
that the scaling transformation ϕ̃ = (const.)ϕ is a symmetry of the
field equations but is not a symmetry of the Lagrangian.
If the Lagrangian and its field equations represent the “laws”, then the
solutions of the field equations are the “states” of the field that are allowed by
the laws. The function ϕ(x) is an allowed state of the field when it solves the
field equations. A symmetry of a given “state”, ϕ0 (x) say, is then defined
to be a transformation of the fields, ϕ → ϕ̃[ϕ], which preserves the given
solution
ϕ̃[ϕ0 (x)] = ϕ0 (x). (7.4)
Since symmetry transformations form a group, such solutions to the field
equations are sometimes called “group-invariant solutions”.
Let us consider an elementary example of group-invariant solutions. Con-
sider the KG field with mass m. Use inertial Cartesian coordinates. We have
7.1. SYMMETRY OF LAWS VERSUS SYMMETRY OF STATES 155
Problem:
2. Derive the result (7.5).
1 d 2 df
(r ) = 0. (7.8)
r2 dr dr
This is the principal reason one usually makes a “symmetry ansatz” for
solutions to field equations which involves fields invariant under a subgroup
K of the symmetry group G of the equations. It is not illegal to make other
kinds of ansatzes, of course, but most will lead to inconsistent equations or
equations with trivial solutions.
Having said all this, I should point out that just because you ask for
group invariant solutions according to the above scheme it doesn’t mean you
will find any! There are two reasons for this. First of all, it may be that
there are no (non-trivial) fields invariant with respect to the symmetry group
you are trying to impose on the state. For example, consider the symmetry
group ϕ → ϕ + const. we mentioned earlier for the massless KG equation.
You can easily see that there are no functions which are invariant under
that transformation group. Secondly, the reduced differential equation may
have no (or only trivial) solutions, indicating that no (interesting) solutions
exist with that symmetry. Finally, I should mention that not all states have
symmetry - indeed the generic states are completely asymmetric. States
with symmetry are special, physically simpler states than what you expect
generically.
To summarize, field theories may have two types of symmetry. There
may be a group G of symmetries of its laws – the symmetry group of the
Lagrangian (and field equations). There can be symmetries of states, that
is, there may be a transformation group (usually a subgroup of G) which
preserves certain solutions to the field equations.
1 1 1
L0 = − ∂α ϕ∂ α ϕ − (− a2 ϕ2 + b2 ϕ4 ). (7.9)
2 2 4
7.2. THE “MEXICAN HAT” POTENTIAL 157
Problem:
3. Compute the first variation δE of the functional (7.11). Show that it
vanishes when evaluated on fields ϕ = 0, ± ab .
You can easily check that the solutions given by |ϕ| = 0, a/b are critical
points of this energy functional. As before, the maximally symmetric state
ϕ = 0 is unstable. The circle’s worth of states (7.18) are quasi-stable in the
following sense. Any displacement in field space yields a non-negative change
in energy. To see this, write
ϕ = ρeiΘ , (7.21)
where ρ and Θ are spacetime functions. The energy takes the form
Z
3 1 2 1 i 1 2 2 i 1 2 2 1 2 4
E= dx ρ + ρ,i ρ, + ρ (Θ,t + Θ,i Θ, ) − a ρ + b ρ . (7.22)
2 ,t 2 2 2 4
1 a4
Z
3 1 2 1 i 1 a 2 2 i 2 2
E=− 2+ dx δρ + δρ,i δρ, + (δΘ,t + δΘ,i δΘ, ) + a δρ
4b 2 ,t 2 2 b
(7.23)
Evidently, all displacements except δρ = 0, δΘ = const. increase the energy.
The displacements δρ = 0, δΘ = const. do not change the energy, as you
might have guessed, since they correspond to displacements along the circular
locus of minima of the potential energy function. The states (7.18) are the
lowest energy states – the ground states. Thus the lowest energy is infinitely
degenerate – the set of ground states (7.18) is topologically a circle. That
these stable states form a continuous family and have less symmetry than the
unstable state will have some physical ramifications which we will unravel
after we take a little detour.
7.3. DYNAMICS NEAR EQUILIBRIUM AND GOLDSTONE’S THEOREM161
ϕ + a2 ϕ − b2 ϕ3 = 0. (7.25)
Problem:
4. Using (7.24) expand the action functional for (7.25) (see (7.9)) to
quadratic order in δϕ. Show that this approximate action, viewed
as an action functional for the displacement field δϕ, has (7.26) as its
Euler-Lagrange field equation.
162 CHAPTER 7. SPONTANEOUS SYMMETRY BREAKING
Evidently, the linearized equation (7.26) is a linear PDE for the displace-
ment field δϕ. In general, this linear PDE has variable coefficients due to the
presence of ϕ0 . But if the given solution ϕ0 is a constant in spacetime, the
linearized PDE is mathematically identical to a Klein-Gordon equation for
δϕ with mass given by (−a2 +3b2 ϕ20 ). The mass at the minima, ϕ0 = ±a/b, is
2a2 . The mass at the maximum, ϕ0 = 0, is −a2 . The negative mass-squared
is a symptom of the instability of this state of the field.
From the way the linearized equation is derived, you can easily see that
any displacement field δϕ constructed as an infinitesimal symmetry of the
field equations will automatically satisfy the linearized equations when the
fields being used to build δϕ satisfy the field equations. Indeed, this fact
is the defining property of an infinitesimal symmetry of the field equations.
Here is a simple example.
Problem:
5. Consider time translations ϕ(t, x, y, z) → ϕ̃ = ϕ(t+λ, x, y, z). Compute
the infinitesimal form δϕ of this transformation as a function of ϕ and
its derivatives. Show that if ϕ solves the field equation coming from
(7.9) then δϕ solves the linearized equation (7.26).
ρ − ρ∂α Θ ∂ α Θ + a2 ρ − b2 ρ3 = 0, (7.28)
∂α (ρ2 ∂ α Θ) = 0. (7.29)
There are two things to notice here. First, the symmetry under Θ → Θ +
const. is manifest – only derivatives of Θ appear. Second, the associated
conservation law is the content of (7.29).
7.3. DYNAMICS NEAR EQUILIBRIUM AND GOLDSTONE’S THEOREM163
Let us consider the linearization of these field equations about the circle’s
worth of equilibria ρ = a/b, Θ = const. We could proceed precisely as
before, of course. But it will be instructive to perform the linearization in
the Lagrangian. To this end we write
a
ρ= + δρ, Θ = θ + δΘ, (7.30)
b
where θ = const., and we expand the Lagrangian to quadratic order in δρ,
δΘ:
1 1 a 2 1 a4
L = − ∂α δρ ∂ α δρ − ∂α δΘ ∂ α δΘ − 2 − a2 δρ2 + O(δϕ3 ). (7.31)
2 2 b 4b
Evidently, in a suitably small neighborhood of equilibrium the complex scalar
field can be viewed as 2 real scalar fields (δρ, δΘ); one of the fields (δρ) has
a mass m = a and the other field (δΘ) is massless.
To get a feel for what just happened, let us consider a very similar U (1)
symmetric theory, just differing in the sign of the quadratic potential term
in the Lagrangian. The Lagrangian density is
1 1
L0 = − ∂α ϕ ∂ α ϕ∗ − µ2 |ϕ|2 − b2 |ϕ|4 . (7.32)
2 4
There is only a single Poincaré invariant critical point, ϕ = 0, which is a
global minimum of the energy and which is also U (1) invariant, so the U (1)
symmetry is not spontaneously broken in the ground state. In the vicinity
of the ground state the linearized Lagrangian takes the simple form3
1
L0 = − ∂α δϕ ∂ α δϕ∗ − µ2 |δϕ|2 + O(δϕ3 ). (7.33)
2
Here of course we have the Lagrangian of a complex-valued KG field δϕ with
mass µ; equivalently, we have two real scalar fields with mass µ.
To summarize thus far: With a complex KG field described by a potential
such that the ground state shares all the symmetries of the Lagrangian, the
physics of the theory near the ground state is that of a pair of real, massive
KG fields. Using instead the Mexican hat potential, the ground state of the
complex scalar field does not share all the symmetries of the Lagrangian –
there is spontaneous symmetry breaking – and the physics of the field theory
3
We do not use polar coordinates which are ill-defined at the origin.
164 CHAPTER 7. SPONTANEOUS SYMMETRY BREAKING
near equilibrium is that of a pair of scalar fields, one with mass and one
which is massless.
To some extent, it is not too hard to understand a priori how these re-
sults occur. In particular, we can see why a massless field emerged from the
spontaneous symmetry breaking. For Poincaré invariant solutions – which
are constant functions in spacetime – the linearization of the field equations
about a Poincaré invariant solution involves: (1) the derivative terms in the
Lagrangian, which are quadratic in the fields and, since the Poincaré in-
variant state is constant, are the same in the linearization as for the full
Lagrangian; (2) the Taylor expansion of the potential V (ϕ) to second order
about the constant equilibrium solution ϕ0 . Because of (1), the mass terms
comes solely from the expansion of the potential in (2). Because of the U (1)
symmetry of the potential, through each point in the set of field values there
will be a curve (with tangent vector given by the infinitesimal symmetry)
along which the potential will not change. Because the symmetry is broken,
this curve connects all the ground states of the theory. Taylor expansion
about the ground state in this symmetry direction can yield only vanishing
contributions because the potential has vanishing derivatives in that direc-
tion. Thus the broken symmetry direction(s) defines direction(s) in field
space which correspond to massless fields in an expansion about equilibrium.
This is the essence of the (classical limit of the) Goldstone theorem: to each
broken continuous symmetry generator there is a massless field.
1 1 1 1 1
L = − F αβ Fαβ − ∂α ρ∂ α ρ − ρ2 (∂α Θ + qAα )(∂ α Θ + qAα ) + a2 ρ2 − b2 ρ4 .
4 2 2 2 4
(7.36)
The Poincaré invariant ground state(s) can be determined as follows. As
we have observed, a Poincaré invariant function ϕ is necessarily a constant.
Likewise, it is too not hard to see that the only Poincaré invariant (co)vector
is the the zero (co)vector Aα = 0. Consequently, the Poincaré invariant
ground state, as before, is specified by
a
ρ= , Θ = const. Aα = 0. (7.37)
b
As before, the U (1) symmetry of the theory is not a symmetry of this state,
but instead maps these states among themselves. As before, we want to
expanding to quadratic order about the ground state. To this end we write
a
Aα = 0 + δAα , ρ= + δρ, Θ = Θ0 + δΘ, (7.38)
b
where Θ0 is a constant. We also define
1
Bα = δAα + ∂α δΘ. (7.39)
q
Ignoring terms of cubic and higher order in the displacements (δA, δρ, δΘ)
we then get
1 1 aq 2
L ≈ − (∂α Bβ − ∂β Bα )(∂ α B β − ∂ β B α ) − Bα B α
4 2 b
1 1
− ∂α δρ ∂ α δρ − a2 δρ2
2 2
(7.40)
Problem:
As you can see, excitations of ρ around the ground state are those of a
scalar field with mass a, as before. To understand the rest of the Lagrangian
(7.40) we need to understand the Proca Lagrangian:
1 1
LP roca = − (∂α Bβ − ∂β Bα )(∂ α B β − ∂ β B α ) − κ2 Bα B α . (7.41)
4 2
For κ = 0 this is just the usual electromagnetic Lagrangian. Otherwise...
Problem:
( − κ2 )Bα = 0, ∂ α Bα = 0. (7.42)
the U (1) gauge symmetry. The change of variables (7.39) which we used to
put the (linearized) Lagrangian into the Proca form amounts to modifying
A by a gauge transformation. The effect of this is that the field Θ provides
a “longitudinal” mode to the field B, corresponding to its acquisition of a
mass.
Finally, we point out that the Higgs phenomenon can be generalized con-
siderably. We do not have time to go into it here, but the idea is as follows.
Consider a system of matter fields with symmetry group of its Lagrangian
being a continuous group G and with a ground state which breaks that sym-
metry to some subgroup of G. Couple these matter fields to gauge fields, the
latter with gauge group which includes G. Excitations of the theory near the
ground state will have the gauge fields corresponding to G acquiring a mass.
This is precisely how the W and Z bosons of the weak interaction acquire
their effective masses at (relatively) low energies.
7.5 PROBLEMS
( − κ2 )Bα = 0, ∂ α Bα = 0. (7.43)
Chapter 8
Recall that spin is an “internal degree of freedom”, that is, an internal bit
of structure for the particle excitations of a quantum field whose classical
approximation we have been studying. The spin gets its name since it con-
tributes to the angular momentum conservation law. In quantum field the-
ory, the KG equation is an equation useful for describing relativistic particles
with spin 0; the Maxwell and Proca equations are equations used to describe
particles with spin 1. One of the many great achievements of Dirac was to
devise a system of differential equations that is perfectly suited for studying
the quantum theory of relativistic particles with “spin 12 ”. This is the Dirac
equation and it used in quantum field theory to describe electrons, neutri-
nos, quarks, etc. Originally, this equation was obtained from trying to find
a relativistic analog of the Schrödinger equation. The idea is that one wants
a PDE that is first order in time, unlike the KG or the 4-potential form of
Maxwell equations. Here we shall simply define the Dirac equation without
trying to give details regarding its historical derivation. Then we will explore
some of the simple field theoretic properties associated with the equation.
169
170 CHAPTER 8. THE DIRAC FIELD
γµ : C4 → C4 , µ = 0, 1, 2, 3 (8.3)
γµ γν + γν γµ = 2ηµν I. (8.4)
We can define
iI 0
γ0 = , (8.6)
0 −iI
and
0 −iσj
γj = , j = 1, 2, 3. (8.7)
iσj 0
Here we are using a 2 × 2 block matrix notation.
The γ-matrices are often called, well, just the “gamma matrices”. They
are also called the “Dirac matrices”. The Dirac matrices satisfy
and
γ0† = −γ0−1 , γi† = γi−1 (8.9)
The Dirac matrices are not uniquely determined by their anti-commutation
relations. It is possible to find other, equivalent representations. How-
ever, Pauli showed that if two sets of matrices γµ and γµ0 satisfy the anti-
commutation relations and the Hermiticity relations of the gamma matrices
as shown above, then there is a unitary transformation U on C4 such that
γµ0 = U −1 γµ U. (8.10)
8.1. THE DIRAC EQUATION 171
γ µ ∂µ ψ + mψ = 0. (8.11)
Here we raise the “spacetime index” on γµ with the flat spacetime metric:
γ µ = η µν γν . (8.12)
γ µ ∂µ ψ = −mψ, (8.17)
172 CHAPTER 8. THE DIRAC FIELD
0 = (γ ν ∂ν + mI)(γ µ ∂µ ψ + mψ)
= ( − m2 )ψ, (8.18)
≡ I, m2 ≡ m2 I. (8.19)
Note: That each component of the Dirac field satisfies the KG equation is
necessary for ψ to satisfy the Dirac equation, but it is not sufficient.
The preceding computation allows us to view the Dirac operator as a
sort of “square root” of the KG operator. This is one way to understand the
utility of the Clifford algebra.
I have no choice but to assign you the following.
Problem:
Elements of the Poincaré group are thus labeled by pairs (L, a).
Problem:
(g2 g1 · x)α = Lα2β (Lβ1γ xγ + aβ1 ) + aβ2 = (Lα2β Lβ1γ )xγ + Lα2β aβ1 + aα2 , (8.24)
so that
g2 g1 = (L2 L1 , L2 · a1 + a2 ). (8.25)
174 CHAPTER 8. THE DIRAC FIELD
where cg,h are constants which depend upon a pair of group elements g and
h. A representation is now the special case cg,h = 1, ∀ g, h. The projective
representations are of interest because of quantum theory. Recall that the set
of states of a physical system are identified with elements of a vector space
(in fact a Hilbert space) with the proviso that any two linearly dependent
vectors – two vectors which differ by a scalar multiple – define the same
state. When considering the transformation of states by symmetries, one
thus considers linear transformations, but the group homomorphism property
is only required up to a constant rescaling, as in (8.27).
It can be shown that all irreducible unitary, projective representations of
the Poincaré group are characterized by two parameters, physically identified
with the spin and the mass. Nature only seems to take advantage of the
representations in which the mass is real and positive and the spin is a non-
negative integer or half integer.1 In quantum field theory, the spin of the
representation is principally controlled by the geometrical type of the field.
The principal role of the field equations is then to pick out the mass of the
representation and to ensure its irreducibility. So far we have considered
a scalar field and a 1-form (or 4-vector field), corresponding to spin-0 and
spin-1. The Dirac field is a new type of geometric object: a spinor field,
corresponding to spin- 21 .
To better understand this last statement, we now want to see how the
Poincaré group is represented as a transformation group of the classical fields.
A field representation of the Poincaré group G of transformations assigns to
1
The other allowed representations have imaginary mass, or vanishing mass and con-
tinuous spin. Both mathematical possibilities appear to have unphysical properties.
8.2. REPRESENTATIONS OF THE POINCARÉ GROUP 175
ϕ → ϕ0 = Ψg (ϕ), (8.28)
To understand this point of view, consider our previously studied fields. The
Klein-Gordon field was defined as a function or scalar field, ϕ : R4 → R, so
that we have the representation of the Poincaré group on the space of KG
fields given by:
ϕ(x) → ϕ0 (x) = ϕ(L−1 · (x − a)). (8.30)
It is easy to check that the space of KG fields, equipped with this trans-
formation rule provides a field representation of the Poincaré group. This
representation is called the spin-0 representation. The electromagnetic field
was defined in terms of a 1-form A,
A : R4 → T ∗ M. (8.31)
A = Aα dxα (8.32)
via
Aα (x) → A0α (x) = Lβα Aβ (L−1 · (x − a)). (8.33)
This transformation law leads to the spin-1 representation of the Poincaré
group.
Problem:
3. Verify that the transformation laws (8.30) and (8.33) define represen-
tations of the Poincaré group.
2
Recall that a Lie algebra is a vector space V equipped with a skew-symmetric bilinear
mapping [·, ·] : V × V → V – the Lie bracket – which satisfies the Jacobi identity.
8.3. THE SPINOR REPRESENTATION 177
Problem:
3
The transformation of the value of a scalar field is just the identity transformation.
4
We shall see that this representation of the Lie algebra will lead to a projective rep-
resentation of the corresponding group.
178 CHAPTER 8. THE DIRAC FIELD
4. Using
γµ γν + γν γµ = 2ηµν I, (8.46)
show that
1 1
[S(ω), S(χ)] = ω µν χαβ [Sµν , Sαβ ] = [ω, χ]αβ Sαβ = S([ω, χ]). (8.47)
4 2
This shows that the linear transformations S(ω) on the space C4 represent
the Lie algebra of the Lorentz group. This representation is called the (in-
finitesimal) spinor representation of the (infinitesimal) Lorentz group. The
Lorentz transformation L(ω) defined by the matrix ω α β is represented on
vectors in C4 – “spinors” – by the matrix exponential R(ω):
The fact that S(ω) represents the Lie algebra ensures that the exponential
will (projectively) represent the Lorentz group. As with scalar and vector
fields, the translational part of the Poincaré group is represented trivially on
the values of the spinor.
We have considered how C4 – the values of taken by a Dirac field – can
provide a representation of the Lie algebra of the Poincaré group. We now
extend this representation to a field representation (of the connected com-
ponent of the identity of the Poincaré group) by adding in the Lorentz and
translational transformations to the argument of the field and exponentiat-
ing. We specify a Poincaré transformation5 by using a skew tensor ω, and a
constant vector b. We define the transformation ψ → ψ 0 by
δ1 δ2 ψ − δ2 δ1 ψ = δ3 ψ, (8.51)
5
We are restricting to the component connected to the identity transformation this way.
For simplicity I will suppress the interesting discussion of how to represent the remaining
transformations (time reversal and spatial reflection).
8.3. THE SPINOR REPRESENTATION 179
where
ω3β α = [ω1 , ω2 ]β α , α β
bα3 = ω1β α β
b2 − ω2β b1 + bα1 − bα2 . (8.52)
Comparing with (8.41) and (8.42) we see that the infinitesimal transforma-
tions (8.50) do indeed represent the Lie algebra of the Poincaré group.
As I have mentioned in the preceding discussion, the “spinor represen-
tation” of the Poincaré group we have been describing is not quite a true
representation of the group, but rather a projective representation. (As we
just saw, we do have a true representation of the Lie algebra.) To see why
I say this, consider a Poincaré transformation consisting of a rotation by 2π
about, say, the z-axis. You can easily check that this is generated by an
infinitesimal transformation with bα = 0 and
which is just showing that a rotation by 2π is the same as the identity trans-
formation. (Check this calculation as a nice exercise.) The corresponding
transformation ψ → ψ 0 of the spinor field is determined by
1
S(ω) = ω αβ [γα , γβ ] = π[γ1 , γ2 ] = iπ diag(1, 1, −1, −1). (8.56)
4
So that
eS(ω) = −I, (8.57)
and
ψ 0 (x) = −ψ(x). (8.58)
Problem:
5. Verify (8.56) through (8.58).
180 CHAPTER 8. THE DIRAC FIELD
Problem:
6. Verify that L in (8.60) is real. (Note that we have the identity (γ 0 γ µ )† =
γ 0 γ µ .)
To compute the field equations we vary the fields and put the result in
the standard Euler-Lagrange form:
1
δL = −δ ψ̄(γ µ ∂µ ψ+mψ)−(−∂µ ψ̄γ µ +mψ̄)δψ− ∂µ (ψ̄γ µ δψ−δ ψ̄γ µ ψ). (8.61)
2
Evidently, the EL expression for ψ̄ is
so the EL equations coming from varying ψ̄, namely Eψ̄ = 0, yields the Dirac
equation for ψ The EL equations coming from varying ψ are determined by
Problem:
7. Show that the equations Eψ = 0 are equivalent to the Dirac equation
for ψ.
More interesting perhaps is the way in which the spinor representation makes
the Lagrangian Lorentz-invariant. We need to consider how the Lagrangian
changes when we make a transformation
We have
1
δL = − [δ ψ̄γ µ ψ,µ + ψ̄γ µ (Dµ δψ) − (Dµ δ ψ̄)γ µ ψ − ψ̄,µ γ µ δψ] − m(δ ψ̄ψ + ψ̄δψ).
2
(8.67)
To see how to proceed, we need a few small results. First, we have that
Next, if δψ is as given in (8.66), then using (8.4) and (8.8) it follows that
Using these facts we can compute the change in the Lagrangian under an
infinitesimal Lorentz transformation to be
Problem:
8. Prove (8.70).
8.6. ENERGY 183
8.6 Energy
It is instructive to have a look at the conserved energy for the Dirac field,
both to illustrate previous technology and to motivate the use of Grassmann-
valued fields.
From (8.65) the time translation symmetry is a divergence symmetry with
Problems:
9. Derive (8.72). Show that L = 0 when the field equations hold. There-
fore, modulo a trivial conservation law and a scaling by (−1), we can
define the conservation of energy via
1
j µ = − [ψ̄γ µ ψ,0 − ψ̄,0 γ µ ψ]. (8.73)
2
10. Check that the current (8.73) is conserved when the Dirac equation
holds.
∂i ψ = 0. (8.75)
184 CHAPTER 8. THE DIRAC FIELD
γ 0 ∂t ψ + mψ = 0. (8.76)
Let us write
a(t)
b(t)
ψ=
c(t) .
(8.77)
d(t)
Since γ0 = diag(i, i, −i, −i) we have that γ 0 = diag(−i, −i, i, i) so that the
Dirac equation reduces to the decoupled system
−iȧ + ma = 0
−iḃ + mb = 0
iċ + mc = 0
id˙ + md = 0. (8.78)
This energy is clearly not bounded from below. By choosing initial conditions
with c0 and d0 sufficiently large in magnitude we can make the energy as
negative as we wish. Physically, this is a disaster since it means that one can
extract an infinite amount of energy from the Dirac field by coupling it to
other dynamical systems. Note that, while we are free to consider redefining
the energy density by a change of sign, this won’t help because ρ is not
bounded from above either. One could try to avoid this problem by simply
decreeing that one must only consider solutions in which c = d = 0. This
would work if the only thing in the universe were the Dirac field. But when
8.7. ANTI-COMMUTING FIELDS 185
Up until now, we have always assumed that fields could be built out of
functions which take values in a commutative algebra (e.g., real numbers).
Insofar as the classical field theory is the“classical limit” of a quantum field
theory, it turns out that this commutative algebra assumption is reasonable
for bosonic fields (integer spin). But for fermionic fields, e.g., spin 1/2 fields
like the Dirac field, one can show that the “classical limit” of the operator
algebra leads to fields which anti-commute. This leads one to formulate a
classical Dirac field theory as a theory of Grassmann-valued spinors. In the
following I will briefly outline how this goes.
A finite-dimensional Grassmann algebra A is a (real or complex) vector
space V with a basis χα , α = 1, 2, . . . , n, and equipped with a product such
that
χα χβ = −χβ χα . (8.82)
7
More precisely: fields are viewed as “unbounded, self-adjoint operator-valued distri-
butions on spacetime”.
186 CHAPTER 8. THE DIRAC FIELD
ψ → eiα ψ, α ∈ R, (8.89)
or, infinitesimally,
δψ = iψ. (8.90)
You can easily check that this transformation is a symmetry of the Dirac
Lagrangian density (8.60). Noether’s first theorem then provides a conserved
current.
Problem:
11. Show that the conserved current associated to the variational symmetry
(8.89) by Noether’s theorem is given by
j α = iψγ α ψ. (8.91)
Verify that this vector field is divergence-free when the Dirac equation
is satisfied.
1 µ 1
L=− ψ̄γ Dµ ψ − (Dµ ψ̄)γ µ ψ − mψ̄ψ − Fαβ F αβ . (8.95)
2 4
Here we have defined
Dµ ψ̄ = ∂µ ψ̄ + iqAµ ψ̄. (8.96)
Problem:
12. Calculate the field equations for ψ and A from this Lagrangian density.
Notice that, unlike the case with scalar electrodynamics, the electric
4-current that acts as a source for the electromagnetic field does not
depend upon the electromagnetic field.
8.9 PROBLEMS
8.9. PROBLEMS 189
{γµ , γν } := γµ γν + γν γµ = 2ηµν I.
3. Verify that the transformation laws (8.30) and (8.33) define represen-
tations of the Poincaré group.
4. Using
γµ γν + γν γµ = 2ηµν I,
show that
1 µν αβ 1
ω χ [Sµν , Sαβ ] = [ω, χ]αβ Sαβ = S([ω, χ]).
4 2
7. Show that the equations Eψ = 0 (see (8.63)) are equivalent to the Dirac
equation for ψ.
8. Prove (8.70).
9. Derive (8.72). Show that L = 0 when the field equations hold. There-
fore, modulo a trivial conservation law, we can define the conservation
of energy via
1
j µ = − [ψ̄γ µ ψ,0 − ψ̄,0 γ µ ψ].
2
10. Check that the current (8.73) is conserved when the Dirac equation
holds.
190 CHAPTER 8. THE DIRAC FIELD
11. Show that the conserved current associated to the variational symmetry
(8.89) by Noether’s theorem is given by
j α = iψγ α ψ.
Verify that this vector field is divergence-free when the Dirac equation
is satisfied.
12. Calculate the field equations for ψ and A from the Lagrangian density
(8.95). Notice that, unlike the case with scalar electrodynamics, the
electric 4-current that acts as a source for the electromagnetic field does
not depend upon the electromagnetic field.
Chapter 9
191
192 CHAPTER 9. NON-ABELIAN GAUGE THEORY
where
n = (n1 , n2 , n3 ), (n1 )2 + (n2 )2 + (n3 )2 = 1, (9.6)
and
0 1 0 −i 1 0
σ1 = σ2 = , σ3 = . (9.7)
1 0 i 0 0 −1
are the Pauli matrices. Note that there are three free parameters in this
group, corresponding to θ and the two free parameters defining ni .
The group SU(2), as represented on C2 , acts on the fields in the obvious
way:
ϕ(x) → U ϕ(x). (9.8)
One parameter subgroups U (λ) are defined by any curve in the 3-d parameter
space associated with θ and n. As an example, set n = (1, 0, 0) and θ = λ
to get
cos λ i sin λ
U (λ) = . (9.9)
i sin λ cos λ
The infinitesimal form of the SU(2) transformation is
δϕ = τ ϕ, (9.10)
Note that
δϕ† = −ϕ† τ. (9.12)
We can write
τ = −iaj σj , (9.13)
where ai ∈ R3 . Thus infinitesimal transformations can be identified with a
three-dimensional vector space with a basis
δj ϕ = iσj ϕ. (9.14)
Problem:
1. Verify that the ek in (9.15) satisfy the su(2) Lie algebra (9.16) as ad-
vertised.
These currents “carry” the SU(2) charge possessed by the scalar fields.
Problems:
2. (a) show that the Lagrangian (9.17) is invariant with respect to the
transformation (9.8); (b) compute its Euler-Lagrange equations; (c)
derive the conserved currents (9.19) from Noether’s theorem; (d) verify
that the currents (9.19) are divergence-free when the Euler-Lagrange
equations are satisfied.
3. Show that if any one of the currents (9.19) is conserved, then the other
2 are automatically conserved. (Hint: Consider the behavior of the
currents under SU(2) transformations of the fields.)
U : C2 → C2 , U † = U −1 , det U = 1, (9.21)
becomes a “local” symmetry:
Aµ = Aiµ ei , (9.25)
where ei are the basis for su(2) defined in (9.15). Thus, if you like, you can
think of the gauge field as 4 × 3 = 12 real fields Akµ (x) labeled according to
their spacetime (index µ = 0, 1, 2, 3) and “internal” su(2) structure (index
k = 1, 2, 3).
The gauge covariant derivative of ϕ is now defined as2
Dµ ϕ = ∂µ ϕ + Aµ ϕ. (9.26)
A fixed, given gauge field Aµ just defines another way to compare phases of
fields at different points. Nothing is gained symmetry-wise by its introduc-
tion. Indeed, under a gauge transformation
= U (x) (∂µ + Aµ ) ϕ
= U (x)Dµ ϕ. (9.29)
Granted this set-up, we can then define the Lagrangian for ϕ using the “min-
imal coupling prescription
∂µ → Dµ , (9.30)
so that we have
L = −η αβ hDα ϕ, Dβ ϕi − m2 hϕ, ϕi. (9.31)
Since, at the moment, we view the gauge field as fixed/prescribed, strictly
speaking, this Lagrangian is no more gauge invariant than our original La-
grangian (9.17),3 which it includes as the special case A = 0, but it has
2
For simplicity we absorb a coupling constant into the definition of A.
3
This is completely analogous to our discussion of the behavior of the KG Lagrangian
under a coordinate transformation, back in §2.9.3.
198 CHAPTER 9. NON-ABELIAN GAUGE THEORY
We can interpret this last result as follows. The Lie algebra su(2), like
any Lie algebra, is a vector space. You can easily check that the set of 2 × 2
9.3. INFINITESIMAL GAUGE TRANSFORMATIONS 199
ξ → U ξU −1 . (9.40)
δξ = [τ, ξ] (9.42)
Problem:
4. Show that the adjoint representation (9.40) has the infinitesimal form
(9.42).
The point of these observations is to show that we can now extend our
definition of the covariant derivative to ξ via the adjoint representation. We
200 CHAPTER 9. NON-ABELIAN GAUGE THEORY
δϕ = τ ϕ, (9.43)
We can associate a group element g[γ] ∈ SU (2) to each point on this curve
by, in effect, exponentiating this infinitesimal transformation. More precisely,
we define a group transformation at each point along the curve by solving
the differential equation
d
g(s) = −γ̇ µ (s)Aµ (γ(s))g(s), (9.49)
ds
subject to the initial condition
g(0) = I. (9.50)
We say that ϕ has been parallelly propagated from x1 to x2 along the curve
γ if4
ϕ(x2 ) = g(1)ϕ(x1 ). (9.51)
The idea is that, given the curve, the group transformation g(1) is used to
define the relationship between the spaces C2 at x1 and x2 . Equivalently, we
can define the parallel propagation of ϕ along the curve γ to be defined by
solving the equation
d
ϕ(γ(s)) = −γ̇ µ (s)Aµ (γ(s))ϕ(γ(s)). (9.52)
ds
Thus the gauge field A defines what it means to have elements of C2 to be
“parallel”, that is, to stay “unchanged” as we move from point to point in
M . In particular, we say that the field ϕ is not changing (relative to A) or
is parallelly propagated along the curve if
d
0= ϕ(γ(s)) + γ̇ µ (s)Aµ (γ(s))ϕ(γ(s))
ds
= γ̇ µ (ϕ,µ (γ) + Aµ (γ)ϕ(γ))
= γ̇ µ Dµ ϕ(γ). (9.53)
You can see how the covariant derivative determines the rate of change of ϕ
along the curve. So how a field is changing from point to point is determined
by the choice of gauge field or “connection”.
Let me emphasize that the parallel propagation of a field (relative to a
given connection) from one point to another is, in general, path-dependent,
4
“Parallel propagation” is also called ”parallel transport”.
202 CHAPTER 9. NON-ABELIAN GAUGE THEORY
i.e., depends upon the choice of curve connecting the points. This is because
the form of the ODE shown above will depend upon γ. We will say a little
more about this below.
It is possible to give a formal series solution to the differential equation
of parallel propagation. Indeed, if you have studied quantum mechanics, you
will perhaps note the similarity of the equation
d
g(s) = −γ̇ µ (s)Aµ (γ(s))g(s). (9.54)
ds
to the Schrödinger equation for the time evolution operator U (t) of a time-
dependent Hamiltonian H(t):
∂U (t)
i~ = H(t)U (t). (9.55)
∂t
In quantum mechanics one solves the Schrödinger equation using the “time-
ordered exponential”; here we can use the analogous quantity, the “path-
ordered exponential”:
∞
X Z 1 Z sn Z s2
n
g(s) = (−1) dsn dsn−1 · · · ds1 γ̇ µ (sn )Aµ (γ(sn )) · · · γ̇ µ (s1 )Aµ (γ(s1 ))
n=0 0 0 0
Z 1
≡ P exp − ds γ̇ µ Aµ .
0
(9.56)
Dµ ϕ = 0, (9.57)
such fields are called covariantly constant, or just “constant”, or just “par-
allel”. Covariantly constant fields have the property that their value at any
point is parallel with its value at any other point – relative to the definition of
“parallel” provided by the connection A and independently of any choice of
curve connecting the points. As we shall see below, the existence of parallel
fields requires a special choice of connection.
It is instructive to point out that everything we have done for SU(2) gauge
theory could be also done with the group U (1) in SED. Now the connections
9.5. GEOMETRICAL INTERPRETATION: CURVATURE 203
where Fµν defines a Lie algebra-valued 2-form, the curvature of the gauge
field,
i
Fµν = Fµν ei , (9.60)
given by
Fµν = ∂µ Aν − ∂ν Aµ + [Aµ , Aν ], (9.61)
and
i
Fµν = ∂µ Aiν − ∂ν Aiµ + i jk Ajµ Akν . (9.62)
Problem:
5. Verify the results (9.59) – (9.62) on the YM curvature.
Dµ ϕ = 0, (9.63)
204 CHAPTER 9. NON-ABELIAN GAUGE THEORY
Therefore, assuming ϕ 6= 0,
i
Fµν ei ϕ = 0 =⇒ Fµν = 0. (9.67)
∂µ Aν − ∂ν Aµ + [Aµ , Aν ] = 0, (9.68)
This condition, which says that the connection is “flat”, is also the integra-
bility condition for the existence of SU(2)-valued functions U satisfying
∂µ U + Aµ U = 0. (9.69)
Thus, if the connection is flat then there exists U (x) : M → SU (2) such that
This means that the gauge field is just a gauge transformation of the trivial
connection A = 0. Thus we see that, at least locally,5 all flat connections are
gauge-equivalent to the zero connection.
Finally, let us consider the behavior of the curvature under a gauge trans-
formation.
Problem:
6. Let U (x) : M → SU (2) define a gauge transformation, for which
Show that
Fµν → U (x)Fµν U −1 (x). (9.72)
5
I say “locally” because arguments based upon integrability conditions only guarantee
local solutions to the differential equations.
9.6. LAGRANGIAN FOR THE YM FIELD 205
Thus while the gauge field transforms inhomogeneously under a gauge trans-
formation, the curvature transforms homogeneously. The curvature is trans-
forming according to the adjoint representation of SU(2) on su(2) in which,
with
τ ∈ su(2), U ∈ SU (2), (9.73)
we have
τ → U τ U −1 . (9.74)
Infinitesimally, if we have a gauge transformation defined by
Problem:
7. Verify (9.76).
The curvature of the YM field differs from the curvature of the Maxwell
field in a few key ways. First, the YM curvature is really a trio of 2-forms,
while the Maxwell curvature is a single 2-form. Second, the YM curvature is
a non-linear function of the the gauge field in contrast to the linear relation
F = dA arising in Maxwell theory. Finally, the YM curvature transforms
homogeneously under a gauge transformation, while the Maxwell curvature
is gauge invariant. You can now interpret this last result as coming from the
fact that the adjoint representation of an Abelian group like U (1) is trivial
(exercise).
If we view A as fixed, i.e., given, then it is not a field variable but instead
provides explicit functions of xα which appear in Lϕ = Lϕ (x, ϕ, ∂ϕ). Gauge-
related connections A will give the same Lagrangian if we transform ϕ as
206 CHAPTER 9. NON-ABELIAN GAUGE THEORY
well, but this has no immediate field theoretic consequence, e.g., it doesn’t
provide a symmetry for Noether’s theorem. The Lagrangian in this case is
viewed as describing the dynamics of KG fields interacting with a prescribed
YM field. The full, gauge invariant theory demands that A also be one of
the dynamical fields. Following the examples of the electrically charged KG
field and Dirac field, we need to adjoin to this Lagrangian a term providing
dynamics for the YM field.6 This is our next consideration.
We can build a Lagrangian for the YM field A by a simple generaliza-
tion of the Maxwell Lagrangian. Our guiding principle is gauge invariance.
Now, unlike the Maxwell curvature, the YM curvature is not gauge invariant,
rather, it is gauge “covariant”, transforming homogeneously under a gauge
transformation via the adjoint representation of SU(2) on su(2). However, it
is easy to see that the trace of a product of su(2) elements is invariant under
this action of SU(2) on su(2):
tr (U τ1 U −1 )(U τ2 U −1 ) = tr {τ1 τ2 } .
(9.78)
This implies that the following Lagrangian density is gauge invariant:
1
LY M = tr (Fµν F µν ) . (9.79)
2
Writing
i
Fµν = Fµν ei , (9.80)
and using
1
tr(ei ej ) = − δij , (9.81)
2
we get
1
LY M = − δij F µν i Fµνj
. (9.82)
4
So, the YM Lagrangian is arising really as a sum of 3 Maxwell-type La-
grangians. However there is a very significant difference between the Maxwell
and YM Lagrangians: the Maxwell Lagrangian is quadratic in the gauge field
while the YM Lagrangian includes terms cubic and quartic in the gauge field.
This means that, unlike the source-free Maxwell equations, the source-free
YM equations will be non-linear. Physically this means that the YM fields
are “self-interacting”.
6
Simply using (9.77) to describe the dynamics of the scalars and the YM field is unsat-
isfactory since the Euler-Lagrange equations for the gauge field imply that ϕ is covariantly
constant and hence that A is flat.
9.7. THE SOURCE-FREE YANG-MILLS EQUATIONS 207
where
δFµν = Dµ δAν − Dν δAµ , (9.85)
and
Dµ δAν = ∂µ δAν + [Aµ , δAν ]. (9.86)
Problem:
Aµ → U Aµ U −1 − (∂µ U )U −1 , (9.87)
L = LY M + Lϕ , (9.95)
where
Lϕ = − hDµ ϕ, Dµ ϕi + m2 hϕ, ϕi .
(9.96)
It is not hard to check that the EL equations for ϕ are given by
Dµ Dµ ϕ − m2 ϕ = 0, (9.97)
Dµ Fiµν + ϕ† ei Dν ϕ − Dν ϕ† ei ϕ = 0.
(9.98)
Dµ F µν − j ν = 0. (9.99)
You can see that the fields ϕ are coupled to the gauge field through a sort
of covariant version of the KG equation. The scalar field acts as a source of
the YM field via the current
jiµ = Dµ ϕ† ei ϕ − ϕ† ei Dµ ϕ .
(9.100)
Note that, just as in SED, the definition of the current involves the gauge
field itself.
Problem:
7
While calling ϕ the “source” is quite all right, it should be kept in mind that, owing
to the non-linearity of the source-free YM field equations, one can consider the YM field
as its own source!
210 CHAPTER 9. NON-ABELIAN GAUGE THEORY
Problem:
11. Verify by direct computation
Dν Dµ F µν = 0. (9.104)
(Note that this is not quite as trivial as in the U (1) case since covariant
derivatives appear.)
For the YM field coupled to its source described by Lϕ , the differential iden-
tity is
{Dν [Dµ Fiµν − jiν ]} + 2Re [Eϕ ei ϕ] = 0. (9.105)
We conclude then that the scalar current (9.100) satisfies a covariant conser-
vation equation when the field equations hold:
Dµ j µ ≡ ∂µ j µ + [Aµ , j µ ] = 0, modulo field equations. (9.106)
Problem:
9.9. NOETHER THEOREMS 211
Problems:
A = A1 e1 , A1 = αµ dxµ . (9.107)
Dµ φ = 0. (9.108)
14. Suppose that the gauge field A is such that there exists a covariantly
constant field φ = φi ei (e.g., as in the previous problem),
Dµ φ = 0. (9.109)
Show that J µ := φi jiµ (see (9.100)) defines a bona fide conserved cur-
rent.
212 CHAPTER 9. NON-ABELIAN GAUGE THEORY
ϕ : M → V, (9.110)
Dµ ϕ = ∂µ ϕ + Aµ ϕ. (9.111)
so that
Fµν = ∂µ Aν − ∂ν Aµ + [Aµ , Aν ]. (9.113)
The curvature is a g-valued 2-form. The behavior of the gauge field and
curvature under a gauge transformations is the same as when the gauge
group was SU(2):
α: V × V → R (9.116)
α(v, w) = 0 ∀ w =⇒ v = 0. (9.118)
9.11. CHERN-SIMONS THEORY 213
Next, let
β: g × g → R (9.119)
be a non-degenerate bi-linear form on the Lie algebra, invariant with respect
to the adjoint representation of G on g. If U ∈ G the adjoint representation,
AdU : g → g, is given by
AdU · τ = U τ U −1 , τ ∈ g. (9.120)
L = LA + Lϕ , (9.122)
where
LA = β(Fµν , F µν ), (9.123)
and
Lϕ = −α(Dµ ϕ, Dµ ϕ) − m2 α(ϕ, ϕ). (9.124)
I think you can see that, given the the transformation properties of the
covariant derivative and the curvature, and given the invariance properties
of the bi-linear forms α and β, the Lagrangian L is gauge invariant.
β = βij ω i ⊗ ω j , (9.126)
where ω i is the basis of 1-forms dual to the basis ei . This means that, with
v = v i ei , w = w i ei , (9.127)
we have
β(v, w) = βij v i wj . (9.128)
The adjoint-invariance of β implies that
c c
Cka βcb + Ckb βac = 0. (9.129)
c c
Along with the usual anti-symmetry of the structure constants, Cab = −Cba ,
this implies
c c
βca Cbd = βc[a Cbd] . (9.130)
The next ingredient we will need to build the Lagrangian is the 3-dimensional
permutation symbol.8 This object is denoted η αβγ , α, β, γ = 1, 2, 3, and is de-
fined to have the following properties: (i) it is totally anti-symmetric (changes
sign under any permutation of its components), and (ii) it has components
consisting of 0, ±1 where, in any coordinate system (with the same orienta-
tion),
η 123 = 1. (9.131)
The permutation symbol plays a fundamental role in defining the deter-
minant of a matrix. From the Leibniz formula for the determinant of a 3 × 3
matrix Bβα we have
Thus if (9.131) holds in any one coordinate system it will hold in all coordi-
nate systems, consistent with the definition of the permutation symbol. An
important point to make in this context is that nowhere did we need to use
a metric tensor to define the permutation symbol; such an object is always
available as a rank-n tensor density on any n-dimensional manifold.
Problem:
15. Show that one can likewise define a permutation symbol ηαβγ , which
is a covariant tensor density of weight minus one. (Do not introduce a
metric to lower the indices!)
To get the last equality I used (9.130). The curvature formula (9.113) ex-
pressed in the basis (9.125) is
i
Fµν = ∂µ Aiν − ∂ν Aiµ + Cjk
i
Ajµ Akν . (9.138)
We thus get
j
Erµ (L) = 2η µβγ βjr Fβγ . (9.139)
Because of the identities
β β
η µβγ ηµρσ = 2δ[ρ δσ] , βjr β rs = δjs , (9.140)
we can conclude
Erµ (L) = 0 ⇐⇒ i
Fµν = 0. (9.141)
The EL equations for the Chern-Simons Lagrangian say that the gauge field
(or connection) is flat!
I will conclude this brief introduction to Chern-Simons theory by exposing
the somewhat some remarkable symmetries of the Lagrangian. To begin, it
is reasonable to inquire whether there is a gauge symmetry. At a first glance,
existence of such a symmetry would appear doubtful since the Lagrangian is
not built in an invariant way from the curvature. Still, it is instructive to
look more closely. Use (9.138) to write the Lagrangian as
αβγ i j 1 j i k l
L = η βij Aα Fβγ − Ckl Aα Aβ Aγ
3
αβγ 1
=η β(Aα , Fβγ ) − β(Aα , [Aβ , Aγ ])
3
αβγ 2
=η β(Aα , Fβγ ) − β(Aα , Aβ Aγ ) . (9.142)
3
9.11. CHERN-SIMONS THEORY 217
Define
Ãα = U Aα U −1 − ∂α U U −1 , F̃βγ = U Fβγ U −1 . (9.143)
Use the gauge transformation formulas (9.114) and (9.115) and the invariance
conditions (9.121) and (9.130) to obtain
(9.144)
and
+ 3β(U −1 ∂α U U −1 ∂β U, Aγ ) − 3β(U −1 ∂α U, Aβ Aγ )
(9.145)
We now have
2
L̃ ≡ η αβγ β(Ãα , F̃βγ ) − β(Ãα , Ãβ Ãγ )
3
2 αβγ
= L + η β(∂α U U −1 , ∂β U U −1 ∂γ U U −1 )
3
− 2η αβγ β(U −1 ∂α U, ∂β Aγ ) + β(U −1 ∂α U U −1 ∂β U, Aγ ) . (9.146)
U −1 ∂α U U −1 = −∂α U −1 , (9.147)
so that
2
L̃ = L + η αβγ β(∂α U U −1 , ∂β U U −1 ∂γ U U −1 ) + ∂α η αβγ β(2U −1 ∂βU, Aγ ) .
3
(9.148)
The last term is a divergence; what about the second term? This term is
interesting. If the gauge transformation can be obtained by iterating in-
finitesimal transformations, that is, if it can be written as
You can check that in this case the term of interest in (9.148) can also be
written as a divergence. Thus infinitesimal gauge transformations will be
divergence symmetries of the Lagrangian. Noether’s second theorem then
applies implying the existence of differential identities satisfied by the Euler-
Lagrange equations. In this case these are equivalent to the “Bianchi identi-
ties”:
D[α Fβγ] = 0. (9.151)
Problem:
16. Show by direct computation just using the definition (9.113) and the
definition of covariant derivative that
It can happen that there are gauge transformations not obtained by it-
erating infinitesimal transformations as in (9.149). To explain this in detail
will take us too far afield, so let me just provide you with an example.
Problem:
17. Consider a Chern-Simons theory where M = S 3 and G = SU (2),
and βij = −2δij . Using the Euler angle coordinates xµ = (γ, β, α),
(0 < α < 2π, 0 < β < 2π, 0 < γ < π) for S 3 , define U : M → G by
Problem:
18. A “covariant” tensor like Aα transforms under (9.155) by
∂φν
Aµ (x) → Aν (φ(x)). (9.158)
∂xµ
Show that (9.157) corresponds to the infinitesimal form of this transfor-
mation law (using (9.156)) up to an infinitesimal gauge transformation
generated by V α Aα (see (9.46)).
10
For those who know about these things, this is the “gauge covariant Lie derivative”
of A.
220 CHAPTER 9. NON-ABELIAN GAUGE THEORY
Problem:
19. Apply Noether’s second theorem to the infinitesimal gauge and dif-
feomorphism symmetries of the Chern-Simons Lagrangian. What are
the corresponding differential identities satisfied by the Euler-Lagrange
expression (9.139)?
9.12 PROBLEMS
1. Verify that the ek in (9.15) satisfy the su(2) Lie algebra (9.16) as ad-
vertised.
2. (a) Show that the Lagrangian (9.17) is invariant with respect to the
transformation (9.8); (b) compute its Euler-Lagrange equations; (c)
derive the conserved currents (9.19) from Noether’s theorem; (d) verify
that the currents (9.19) are divergence-free when the Euler-Lagrange
equations are satisfied.
3. Show that if any one of the currents (9.19) are conserved, then the
other 2 are automatically conserved. (Hint: Consider the behavior of
the currents under SU(2) transformations of the fields.)
4. Show that the adjoint representation (9.40) has the infinitesimal form
(9.42).
14. Suppose that the gauge field A is such that there exists a covariantly
constant field φ = φi ei ,
Dµ φ = 0. (9.167)
Show that J µ := φi jiµ (see (9.100)) defines a bona fide conserved cur-
rent.
222 CHAPTER 9. NON-ABELIAN GAUGE THEORY
15. Show that one can define a permutation symbol ηαβγ , which is a co-
variant tensor density of weight minus one. (Do not introduce a metric
to lower the indices!)
16. Show by direct computation just using the definition (9.113) and the
definition of covariant derivative that
D[α Fβγ] = 0. (9.168)
11
Interestingly and importantly, with a proper normalization the action integral changes
by an additive integer under such gauge transformations. This allows gauge invariance in
the associated quantum field theory.
Chapter 10
223
224 CHAPTER 10. GRAVITATIONAL FIELD THEORY
It is a basic result from linear algebra that any quadratic form Q can be
put into canonical form by a change of basis (try googling “Sylvester’s law of
inertia”). What this means is that there always exists a basis for the vector
space upon which Q is defined such that the matrix of components of the
quadratic form takes the form
Q = diag(1 , 2 , . . . , n ), (10.4)
∇α v β = ∂α v β + Γβγα v γ , (10.6)
∇α wβ = ∂α wβ − Γγβα wγ . (10.8)
Problem:
d2 q µ α
µ dq dq
β
+ Γ αβ = 0, (10.10)
ds2 ds ds
dq µ dq ν
gµν
= κ, (10.11)
ds ds
where κ is a given constant; κ > 0 for spacelike curves, κ < 0 for timelike
curves and κ = 0 for lightlike curves.
Problem:
µ dq ν
2. Show that gµν dqds ds
is a constant of motion for (10.10).
2 2
Cαβγδ = Rαβγδ − (gα[γ Rδ]β − gβ[γ Rδ]α ) + Rgα[γ gδ]β .
n−2 (n − 1)(n − 2)
(10.18)
The curvature tensor is completely determined by the Weyl tensor, Ricci
tensor, and Ricci scalar through the following formula:
2 2
Rαβγδ = Cαβγδ + gα[γ Rδ]β − gβ[γ Rδ]α − Rgα[γ gδ]β .
n−2 (n − 1)(n − 2)
(10.19)
dq µ (0)
q µ (0) = xµ0 , = vµ. (10.21)
ds
Here is a useful mathematical result with an important physical interpre-
tation. Pick a spacetime event, i.e., a point p ∈ M . Events q sufficiently close
to p can be labeled using geodesics as follows. Find a geodesic starting at p
which passes through q. This geodesic will be unique if q is in a sufficiently
small neighborhood of p. Let uµ be the components of the tangent vector at
p which, along with p, provides the initial data for the geodesic which passes
228 CHAPTER 10. GRAVITATIONAL FIELD THEORY
∂rµ ∂q σ
v µ (s, λ) = + Γµνσ rν . (10.26)
∂s ∂s
The relative acceleration aµ (s, λ) of the neighboring geodesic labeled by λ+dλ
relative to the geodesic labeled by λ is the directional derivative of the relative
velocity in the direction of the geodesic labeled by λ:
∂v µ ∂q σ
aµ (s, λ) = + Γµνσ v ν . (10.27)
∂s ∂s
A nice exercise is to show that:
This is the precise sense in which spacetime curvature controls the relative
acceleration of geodesics. Note that while one can erect a freely falling refer-
ence frame such that (10.22) holds at some event, it is not possible to make
the curvature vanish at a point by a choice of reference frame. Thus the
geodesic deviation (10.28) is an immutable feature of the gravitational field.
One can say that gravity is geodesic deviation.
Problem:
3. Perhaps the simplest example of a “generally covariant” differential
equation is the general form of the geodesic equation in 2-d Euclidean
space. It arises as the Euler-Lagrange equations for a curve (x(λ), y(λ))
of the Lagrangian: p
L = ẋ2 + ẏ 2 . (10.32)
Show that the form of the EL equations do not depend upon the choice
of parameter λ. (Hint: consider a new parameter τ = f (λ); show that
the equations in terms of τ are of the same form as those in terms of
λ, irrespective of the choice of f .)
10.3. THE PRINCIPLE OF GENERAL COVARIANCE 231
∗ ∂φµ ∂φν
(φ g)αβ (x) = g̃µν (φ(x)). (10.33)
∂xα ∂xβ
More generally, given any tensor field we can relate its components φ∗ Tβ...
α...
(x)
µ α... µ
in terms of x and its components T̃β... (y) in terms of y :
∂φν ∂φ−1α
∗
(φ T )α...
β... (x) = ... µ...
. . . T̃ν... (φ(x)). (10.34)
∂xβ ∂y µ y=φ(x)
What this formula says in English is: If you apply the curvature formula to
the transformed metric, the result is the same as applying the formula to
the untransformed metric and then transforming the result. In this sense the
curvature formula is “the same” in all coordinate systems or, more elegantly,
is defined from the metric only using the underlying manifold structure. We
4
One way to see this is to take the line element defined with y α and express it in terms
α
of x using the mapping φ.
5
We could also phrase this in terms of a connection.
232 CHAPTER 10. GRAVITATIONAL FIELD THEORY
say that the curvature tensor is “naturally” defined in terms of the met-
ric. More generally, any tensor field obtained from the curvature using the
metric and covariant derivatives with the metric-compatible connection will
be “naturally” constructed from the metric. We will call such tensor fields
natural.6
The principle of general covariance is the requirement that the field equa-
tions for spacetime are naturally constructed from the metric (and any other
matter fields which may be present). Just considering spacetime (with no
matter), let the field equations be of the form G = 0, where G = G(g) is
a tensor constructed naturally from the metric g. Supposed g0 is solution
to the field equations, G(g0 ) = 0. Then, thanks to the property (10.35), we
have
G(φ∗ g0 ) = φ∗ G(g0 ) = 0. (10.36)
6
Some old-fashioned expositions of this subject simply call such an object a “tensor”.
This term is clearly ambiguous since any field of multi-linear maps on the tangent spaces,
cotangent spaces, and products thereof is, by definition, a tensor field, irrespective of its
naturality. We will use the modern terminology: “natural tensor” to denote things like
the curvature which obey (10.35).
10.4. THE EINSTEIN-HILBERT ACTION 233
The principal properties of the volume form (10.38) are contained in the
following problems:
Problems:
5. Show that the integral in (10.40) does not depend upon the choice of
coordinates.
so that
δg αβ = −g αγ g βδ δgγδ . (10.44)
Henceforth we must remember to make this exception in the usual tensor
notation for raising and lowering indices with the metric. Next, we recall a
couple of results from linear algebra. Let A be a non-singular square matrix.
We have the identity
ln(det(A)) = tr(ln(A)). (10.45)
Now let A(λ) be a non-singular square matrix depending upon a parameter
λ. It follows from (10.45) that
d −1 d
det(A(λ)) = det(A(λ))tr A (λ) A(λ) . (10.46)
dλ dλ
Problem:
6. Prove (10.45) and (10.46).
δg = gg αβ δgαβ . (10.47)
After the volume form, the next simplest natural Lagrangian density is
the Einstein-Hilbert Lagrangian:
1
L= (R − 2Λ), (10.50)
2κ
where κ is given in terms of Newton’s constant G and the speed of light c as
8πG
κ=
c4
Problem:
δSgrav 1
E αβ ≡ = − (Gαγ + Λg αγ ). (10.59)
δgαβ 2κ
Gαβ + Λg αβ = 0. (10.60)
These are usually called the “vacuum Einstein equations” (with a cosmolog-
ical constant). Using the definition of the Einstein tensor, and contracting
this equation with gαβ (“taking the trace”), we get (in 4-dimensions)
R = 4Λ. (10.61)
10.6. DIFFEOMORPHISM SYMMETRY AND THE CONTRACTED BIANCHI IDENTITY237
The vacuum equations are 10 coupled non-linear PDEs for the 10 compo-
nents of the metric, in any given coordinate system. Despite their complexity,
many solutions are known. The majority of the known solutions have Λ = 0,
but here is a pretty famous solution which includes Λ. It is called the “Kottler
metric”. It is also called the “Schwarzschild-de Sitter metric”. In coordinates
(t, r, θ, φ) it is corresponds to the line element given by
2m Λ 2 2 1
ds2 = −(1 − − r )dt + 2m Λ 2
dr2 + r2 (dθ2 + sin2 θdφ2 ). (10.63)
r 3 1− r
− 3
r
Problem:
This is the contracted Bianchi identity; it can be derived from the Bianchi
identity (10.15).
Problem:
∇α E αβ = 0. (10.78)
Finally, you might want to re-read §5.5. Then you will see that the discussion
above is just an instance of Noether’s second theorem.
equation of state” – and also provides a nice simple warm-up for incorpora-
tion of electromagnetic sources. Of course, we are very familiar with scalar
fields by now and, since this is a course in field theory, let us begin with this
matter field.11
We will use a Klein-Gordon field ϕ. We begin by recalling the definition
of its Lagrangian as a top-degree differential form on any curved spacetime
(M, g):
1
LKG = − g αβ ∇α ϕ∇β ϕ + m2 ϕ2 .
(10.79)
2
The Euler-Lagrange equation for ϕ is a curved spacetime generalization of
the original Klein-Gordon equation:
∇α ∇α ϕ − m2 ϕ = 0 (10.80)
You can think of the appearance of the metric in this equation (in the co-
variant derivative term) as bringing into play the effect of the gravitational
field on the “motion” of the scalar field.
Problem:
10. The following line element represents a class of “big bang” cosmological
models, characterized by a scale factor a(t).
There are now EL equations for the metric and for the scalar field. The EL
equation for the scalar field is (10.80), given above. The EL equations for
the metric take the form
1 1
(−Gαγ + Λg αγ ) + T αγ = 0, (10.83)
2κ 2
or
Gαβ + Λg αβ = κT αβ , (10.84)
where
1 1
T αγ = ∇α ϕ∇γ ϕ − g αγ ∇β ϕ∇β ϕ − g αγ m2 ϕ2 (10.85)
2 2
is the energy-momentum tensor of the scalar field. The equations (10.84) are
the celebrated Einstein field equations.
Problem:
11. With
Z
1 αβ
g ∇α ϕ∇β ϕ + m2 ϕ2
SKG [g, ϕ] = − (10.86)
M 2
2 δSKG
T αβ = p (10.87)
|g| δgαβ
Note that here the energy-momentum tensor is not defined via Noether’s
theorem but instead via the coupling of the scalar field to gravity. States of
the combined system of matter (modeled as a scalar field) interacting with
gravity satisfy the coupled system of 11 non-linear PDEs given by (10.80)
and (10.84), known as the Einstein-Klein-Gordon (or Einstein-scalar) system
of equations.
Problem:
10.8. THE CONTRACTED BIANCHI IDENTITY REVISITED 243
12. Show that the following metric and scalar field, given in the coordinate
chart (t, r, θ, φ) define a solution to the Einstein-scalar field equations
for m = 0.
2
g = −r2 dt⊗dt+ dr ⊗dr +r2 (dθ ⊗dθ +sin2 θ dφ⊗dφ), (10.88)
1 − 32 Λr2
1
ϕ = ± √ t + const. (10.89)
κ
This identity must hold for all v α of compact support. Integrating by parts
via the divergence theorem, with the boundary terms vanishing because v α
is of compact support, we can remove the derivatives of the vector field v α
to get
Z
1 αγ αγ αγ
α 2
α
−∇α {−G + Λg } + T vγ + ∇ ∇α ϕ − m ϕ v ∇α ϕ = 0,
M κ
(10.97)
which can be simplified (using the contracted Bianchi identity and covariant
constancy of the metric) to
Z
∇α Tβα − ∇α ∇α ϕ − m2 ϕ ∇β ϕ v β = 0.
(10.98)
M
∇α Tβα − ∇α ∇α ϕ − m2 ϕ ∇β ϕ = 0, .
(10.99)
Problem:
13. Show by direct calculation of the divergence of (10.85) that it satisfies
the identity (10.99).
10.8. THE CONTRACTED BIANCHI IDENTITY REVISITED 245
From this identity it is easy to see that any solution (g, ϕ) to the KG
equation (10.80) will obey the compatibility condition (10.90). Recall that
(10.90) was a necessary condition implied by the Einstein field equations
(10.84) and the contracted Bianchi identity. In this way the coupled Einstein-
scalar field equations are compatible. But it is even more interesting to note
that any solution (g, ϕ) to the Einstein field equations (10.84) alone, provided
∇ϕ 6= 0, will automatically satisfy the KG equation (10.80)! The logic goes
as follows. If (g, ϕ) define a solution to (10.84) then by the Bianchi identity
the energy-momentum tensor built from the solution (g, ϕ) has vanishing
covariant divergence. But then the identity (10.99) and the assumption that
ϕ is not constant means that (g, ϕ) must satisfy (10.80). So, if we opt to
work on the space of non-constant functions ϕ only, which is reasonable
since constant functions cannot satisfy the KG equation with m 6= 0, then
the equations of motion for matter are already contained in the Einstein field
equations!
The divergence condition ∇a T ab = 0 is closely related to – but not quite
the same as – conservation of energy and momentum. You will recall that
in flat spacetime conservation of energy-momentum corresponds to the fact
that the energy-momentum tensor is divergence free, which is interpreted as
providing a divergence-free collection of currents. Here we do not have an
ordinary divergence of a vector field, but rather a covariant divergence of a
tensor field:
∇α T αβ = ∂α T αβ + Γαασ T σβ + Γβασ T ασ . (10.100)
For this reason one cannot use the divergence theorem to convert (10.90) into
a true conservation law.13 Physically, this state of affairs reflects the fact that
the matter field ϕ can exchange energy and momentum with the gravitational
field; since the energy momentum tensor pertains only to the matter field
it need not be conserved by itself. In a freely falling reference frame the
effects of gravity disappear to some extent and one might expect at least
an approximate conservation of energy-momentum of matter. This idea can
be made mathematically precise by working in geodesic normal coordinates
(defined earlier). At the origin of such coordinates the Christoffel symbols
vanish and the coordinate basis vectors can be used to define divergence
13
There are actually two reasons for this. First of all, as just mentioned, the ordinary
divergence is not zero! And, even if it were, there is no useful definition of the integral of a
vector (namely, the divergence of T ) over a volume in the absence of additional structures,
such as preferred vector fields with which to take components.
246 CHAPTER 10. GRAVITATIONAL FIELD THEORY
free currents at the origin. So, in an infinitesimally small region around the
origin of such coordinate system – physically, in a suitably small freely falling
reference frame – one can define an approximate set of conservation laws for
the energy-momentum (and angular momentum) of matter. But because
gravity can carry energy and momentum this conservation law cannot be
extended to any finite region. The crux of the matter is that there is no
useful way to define gravitational energy-momentum densities – there is no
suitable energy-momentum current for gravity. Indeed, if you had such a
current, you ought to be able to make it vanish at the origin of normal
coordinates, whence it can’t be a purely geometrical quantity or it should
vanish in any coordinates. In fact, it is possible to prove that any vector field
locally constructed from the metric and its derivatives to any order which
is divergence-free when the Einstein equations hold is a trivial conservation
law in the sense of §3.20.14
The upshot of these considerations is that gravitational energy, momen-
tum, and angular momentum will not occur via the usual mechanism of
conserved currents. Thanks to the equivalence principle, gravitational en-
ergy, momentum, and angular momentum must be non-local in character.
You may recall that an analogous situation arose for the sources of the Yang-
Mills field. Despite our inability to localize gravitational energy and momen-
tum, it is possible to define a notion of conserved total energy, momentum,
and angular momentum for gravitational systems which are suitably isolated
from their surroundings. So, for example, it is possible to compute these
quantities for a star or galaxy if we ignore the rest of the universe. Based on
our preceding discussion you will not be surprised to hear that such quanti-
ties are not constructed by volume integrals of densities. Perhaps a future
version of this text will explain how all that works.
10.9 PROBLEMS
1. Show that (10.7) enforces ∇α gβγ = 0.
µ dq ν
2. Show that gµν dqds ds
is a constant of motion for (10.10).
Show that the form of the EL equations do not depend upon the choice
of parameter λ. (Hint: consider a new parameter τ = f (λ); show that
the equations in terms of τ are of the same form as those in terms of
λ, irrespective of the choice of f .)
10. With
Z
1 αβ
g ∇α ϕ∇β ϕ + m2 ϕ2
SKG [g, ϕ] = − (10.103)
M 2
2 δSKG
T αβ = p (10.104)
|g| δgαβ
11. Show that the following metric and scalar field, given in the coordinate
chart (t, r, θ, φ) define a solution to the Einstein-scalar field equations
for m = 0.
2
g = −r2 dt⊗dt+ dr⊗dr+r2 (dθ⊗dθ+sin2 θdφ⊗dφ), (10.105)
1 − 23 Λr2
1
ϕ = ± √ t + const. (10.106)
κ
12. Repeat the analysis of §10.7 and §10.8 for the case where the matter
field is the electromagnetic field defined by
1
LEM = − g αγ g βδ Fαβ Fγδ , (10.107)
4
where
Fαβ = ∇α Aβ − ∇β Aα = ∂α Aβ − ∂β Aα . (10.108)
∇α T αβ = ∇β ϕ ∇α ∇α ϕ − m2 ϕ .
(10.109)
Chapter 11
Goodbye
If you have stayed with me to the end of this exhilarating mess, I congratulate
you. If you feel like there is a lot more you would like to learn about classical
field theory, then I agree with you. I have only given a simple introduction
to some of the possible topics. Slowly but surely I hope to add more to this
text, correct errors, add problems, etc. so check back every so often to see
if the version has been updated. Meanwhile, the next section has a list of
resources which you might want to look at, depending upon your interests.
249
250 CHAPTER 11. GOODBYE
Classical Field Theory and the Stress-Energy Tensor, S. Swanson, Morgan &
Claypool Publishers