1 Classical Field Theory

In this first section we will discuss various aspects of classical fields. We will cover only the bare minimum ground necessary before turning to the quantum theory, and will return to classical field theory at several later stages in the course when we need to introduce new ideas.

1.1 The Dynamics of Fields

A field is a quantity defined at every point of space and time (x,t). While classical particle mechanics deals with a finite number of generalized coordinates qa(t), indexed by a label a, in field theory we are interested in the dynamics of fields

ϕa(x,t) (1.3)

where both a and x are considered as labels. Thus we are dealing with a system with an infinite number of degrees of freedom — at least one for each point x in space. Notice that the concept of position has been relegated from a dynamical variable in particle mechanics to a mere label in field theory.

An Example: The Electromagnetic Field

The most familiar examples of fields from classical physics are the electric and magnetic fields, E(x,t) and B(x,t). Both of these fields are spatial 3-vectors. In a more sophisticated treatement of electromagnetism, we derive these two 3-vectors from a single 4-component field Aμ(x,t)=(ϕ,A) where μ=0,1,2,3 shows that this field is a vector in spacetime. The electric and magnetic fields are given by

E=-ϕ-At   and   B=×A (1.4)

which ensure that two of Maxwell’s equations, B=0 and dB/dt=-×E, hold immediately as identities.

The Lagrangian

The dynamics of the field is governed by a Lagrangian which is a function of ϕ(x,t), ϕ˙(x,t) and ϕ(x,t). In all the systems we study in this course, the Lagrangian is of the form,

L(t)=d3x(ϕa,μϕa) (1.5)

where the official name for is the Lagrangian density, although everyone simply calls it the Lagrangian. The action is,

S=t1t2𝑑td3x=d4x (1.6)

Recall that in particle mechanics L depends on q and q˙, but not q¨. In field theory we similarly restrict to Lagrangians depending on ϕ and ϕ˙, and not ϕ¨. In principle, there’s nothing to stop depending on ϕ, 2ϕ, 3ϕ, etc. However, with an eye to later Lorentz invariance, we will only consider Lagrangians depending on ϕ and not higher derivatives. Also we will not consider Lagrangians with explicit dependence on xμ; all such dependence only comes through ϕ and its derivatives.

We can determine the equations of motion by the principle of least action. We vary the path, keeping the end points fixed and require δS=0,

δS = d4x[ϕaδϕa+(μϕa)δ(μϕa)] (1.7)
= d4x[ϕa-μ((μϕa))]δϕa+μ((μϕa)δϕa)

The last term is a total derivative and vanishes for any δϕa(x,t) that decays at spatial infinity and obeys δϕa(x,t1)=δϕa(x,t2)=0. Requiring δS=0 for all such paths yields the Euler-Lagrange equations of motion for the fields ϕa,

μ((μϕa))-ϕa=0 (1.8)

1.1.1 An Example: The Klein-Gordon Equation

Consider the Lagrangian for a real scalar field ϕ(x,t),

=  12ημνμϕνϕ-12m2ϕ2
=  12ϕ˙2-12(ϕ)2-12m2ϕ2

where we are using the Minkowski space metric

ημν=ημν=(+1-1-1-1) (1.10)

Comparing (1.1.1) to the usual expression for the Lagrangian L=T-V, we identify the kinetic energy of the field as

T=d3x12ϕ˙2 (1.11)

and the potential energy of the field as

V=d3x12(ϕ)2+12m2ϕ2 (1.12)

The first term in this expression is called the gradient energy, while the phrase “potential energy”, or just “potential”, is usually reserved for the last term.

To determine the equations of motion arising from (1.1.1), we compute

ϕ=-m2ϕ   and   (μϕ)=μϕ(ϕ˙,-ϕ) (1.13)

The Euler-Lagrange equation is then

ϕ¨-2ϕ+m2ϕ=0 (1.14)

which we can write in relativistic form as

μμϕ+m2ϕ=0 (1.15)

This is the Klein-Gordon Equation. The Laplacian in Minkowski space is sometimes denoted by . In this notation, the Klein-Gordon equation reads ϕ+m2ϕ=0.

An obvious generalization of the Klein-Gordon equation comes from considering the Lagrangian with arbitrary potential V(ϕ),

=12μϕμϕ-V(ϕ)      μμϕ+Vϕ=0 (1.16)

1.1.2 Another Example: First Order Lagrangians

We could also consider a Lagrangian that is linear in time derivatives, rather than quadratic. Take a complex scalar field ψ whose dynamics is defined by the real Lagrangian

=i2(ψψ˙-ψ˙ψ)-ψψ-mψψ (1.17)

We can determine the equations of motion by treating ψ and ψ as independent objects, so that

ψ=i2ψ˙-mψ   and   ψ˙=-i2ψ   and   ψ=-ψ (1.18)

This gives us the equation of motion

iψt=-2ψ+mψ (1.19)

This looks very much like the Schrödinger equation. Except it isn’t! Or, at least, the interpretation of this equation is very different: the field ψ is a classical field with none of the probability interpretation of the wavefunction. We’ll come back to this point in Section 2.8.

The initial data required on a Cauchy surface differs for the two examples above. When ϕ˙2, both ϕ and ϕ˙ must be specified to determine the future evolution; however when ψψ˙, only ψ and ψ are needed.

1.1.3 A Final Example: Maxwell’s Equations

We may derive Maxwell’s equations in the vacuum from the Lagrangian,

=-12(μAν)(μAν)+12(μAμ)2 (1.20)

Notice the funny minus signs! This is to ensure that the kinetic terms for Ai are positive using the Minkowski space metric (1.10), so 12A˙i2. The Lagrangian (1.20) has no kinetic term A˙02 for A0. We will see the consequences of this in Section 6. To see that Maxwell’s equations indeed follow from (1.20), we compute

(μAν)=-μAν+(ρAρ)ημν (1.21)

from which we may derive the equations of motion,

μ((μAν))=-2Aν+ν(ρAρ)=-μ(μAν-νAμ)-μFμν (1.22)

where the field strength is defined by Fμν=μAν-νAμ. You can check using (1.4) that this reproduces the remaining two Maxwell’s equations in a vacuum: E=0 and E/t=×B. Using the notation of the field strength, we may rewrite the Maxwell Lagrangian (up to an integration by parts) in the compact form

=-14FμνFμν (1.23)

1.1.4 Locality, Locality, Locality

In each of the examples above, the Lagrangian is local. This means that there are no terms in the Lagrangian coupling ϕ(x,t) directly to ϕ(y,t) with xy. For example, there are no terms that look like

L=d3xd3yϕ(x)ϕ(y) (1.24)

A priori, there’s no reason for this. After all, x is merely a label, and we’re quite happy to couple other labels together (for example, the term 3A00A3 in the Maxwell Lagrangian couples the μ=0 field to the μ=3 field). But the closest we get for the x label is a coupling between ϕ(x) and ϕ(x+δx) through the gradient term (ϕ)2. This property of locality is, as far as we know, a key feature of all theories of Nature. Indeed, one of the main reasons for introducing field theories in classical physics is to implement locality. In this course, we will only consider local Lagrangians.

1.2 Lorentz Invariance

The laws of Nature are relativistic, and one of the main motivations to develop quantum field theory is to reconcile quantum mechanics with special relativity. To this end, we want to construct field theories in which space and time are placed on an equal footing and the theory is invariant under Lorentz transformations,

xμ(x)μ=Λνμxν (1.25)

where Λνμ satisfies

ΛσμηστΛτν=ημν (1.26)

For example, a rotation by θ about the x3-axis, and a boost by v<1 along the x1-axis are respectively described by the Lorentz transformations

Λνμ=(10000cosθ-sinθ00sinθcosθ00001)  and  Λνμ=(γ-γv0 0-γvγ0000100001) (1.27)

with γ=1/1-v2. The Lorentz transformations form a Lie group under matrix multiplication. You’ll learn more about this in the “Symmetries and Particle Physics” course.

The Lorentz transformations have a representation on the fields. The simplest example is the scalar field which, under the Lorentz transformation xΛx, transforms as

ϕ(x)ϕ(x)=ϕ(Λ-1x) (1.28)

The inverse Λ-1 appears in the argument because we are dealing with an active transformation in which the field is truly shifted. To see why this means that the inverse appears, it will suffice to consider a non-relativistic example such as a temperature field. Suppose we start with an initial field ϕ(x) which has a hotspot at, say, x=(1,0,0). After a rotation xRx about the z-axis, the new field ϕ(x) will have the hotspot at x=(0,1,0). If we want to express ϕ(x) in terms of the old field ϕ, we need to place ourselves at x=(0,1,0) and ask what the old field looked like where we’ve come from at R-1(0,1,0)=(1,0,0). This R-1 is the origin of the inverse transformation. (If we were instead dealing with a passive transformation in which we relabel our choice of coordinates, we would have instead ϕ(x)ϕ(x)=ϕ(Λx)).

The definition of a Lorentz invariant theory is that if ϕ(x) solves the equations of motion then ϕ(Λ-1x) also solves the equations of motion. We can ensure that this property holds by requiring that the action is Lorentz invariant. Let’s look at our examples:

Example 1: The Klein-Gordon Equation

For a real scalar field we have ϕ(x)ϕ(x)=ϕ(Λ-1x). The derivative of the scalar field transforms as a vector, meaning

(μϕ)(x)(Λ-1)μν(νϕ)(y)

where y=Λ-1x. This means that the derivative terms in the Lagrangian density transform as

deriv(x)=μϕ(x)νϕ(x)ημν  (Λ-1)μρ(ρϕ)(y)(Λ-1)νσ(σϕ)(y)ημν (1.29)
=  (ρϕ)(y)(σϕ)(y)ηρσ
=  deriv(y)

The potential terms transform in the same way, with ϕ2(x)ϕ2(y). Putting this all together, we find that the action is indeed invariant under Lorentz transformations,

S=d4x(x)d4x(y)=d4y(y)=S (1.30)

where, in the last step, we need the fact that we don’t pick up a Jacobian factor when we change integration variables from d4x to d4y. This follows because detΛ=1. (At least for Lorentz transformation connected to the identity which, for now, is all we deal with).

Example 2: First Order Dynamics

In the first-order Lagrangian (1.17), space and time are not on the same footing. ( is linear in time derivatives, but quadratic in spatial derivatives). The theory is not Lorentz invariant.

In practice, it’s easy to see if the action is Lorentz invariant: just make sure all the Lorentz indices μ=0,1,2,3 are contracted with Lorentz invariant objects, such as the metric ημν. Other Lorentz invariant objects you can use include the totally antisymmetric tensor ϵμνρσ and the matrices γμ that we will introduce when we come to discuss spinors in Section 4.

Example 3: Maxwell’s Equations

Under a Lorentz transformation Aμ(x)ΛνμAν(Λ-1x). You can check that Maxwell’s Lagrangian (1.23) is indeed invariant. Of course, historically electrodynamics was the first Lorentz invariant theory to be discovered: it was found even before the concept of Lorentz invariance.

1.3 Symmetries

The role of symmetries in field theory is possibly even more important than in particle mechanics. There are Lorentz symmetries, internal symmetries, gauge symmetries, supersymmetries…. We start here by recasting Noether’s theorem in a field theoretic framework.

1.3.1 Noether’s Theorem

Every continuous symmetry of the Lagrangian gives rise to a conserved current jμ(x) such that the equations of motion imply

μjμ=0 (1.31)

or, in other words, j 0/t+j=0.

A Comment: A conserved current implies a conserved charge Q, defined as

Q=𝐑3d3xj 0 (1.32)

which one can immediately see by taking the time derivative,

dQdt=𝐑3d3xjt0=-𝐑3d3xj=0 (1.33)

assuming that j0 sufficiently quickly as |x|. However, the existence of a current is a much stronger statement than the existence of a conserved charge because it implies that charge is conserved locally. To see this, we can define the charge in a finite volume V,

QV=Vd3xj 0 (1.34)

Repeating the analysis above, we find that

dQVdt=-Vd3xj=-Aj𝑑S (1.35)

where A is the area bounding V and we have used Stokes’ theorem. This equation means that any charge leaving V must be accounted for by a flow of the current 3-vector j out of the volume. This kind of local conservation of charge holds in any local field theory.

Proof of Noether’s Theorem: We’ll prove the theorem by working infinitesimally. We may always do this if we have a continuous symmetry. We say that the transformation

δϕa(x)=Xa(ϕ) (1.36)

is a symmetry if the Lagrangian changes by a total derivative,

δ=μFμ (1.37)

for some set of functions Fμ(ϕ). To derive Noether’s theorem, we first consider making an arbitrary transformation of the fields δϕa. Then

δ = ϕaδϕa+(μϕa)μ(δϕa) (1.38)
= [ϕa-μ(μϕa)]δϕa+μ((μϕa)δϕa)

When the equations of motion are satisfied, the term in square brackets vanishes. So we’re left with

δ=μ((μϕa)δϕa) (1.39)

But for the symmetry transformation δϕa=Xa(ϕ), we have by definition δ=μFμ. Equating this expression with (1.39) gives us the result

μjμ=0   with   jμ=(μϕa)Xa(ϕ)-Fμ(ϕ) (1.40)

1.3.2 An Example: Translations and the Energy-Momentum Tensor

Recall that in classical particle mechanics, invariance under spatial translations gives rise to the conservation of momentum, while invariance under time translations is responsible for the conservation of energy. We will now see something similar in field theories. Consider the infinitesimal translation

xνxν-ϵν      ϕa(x)ϕa(x)+ϵννϕa(x) (1.41)

(where the sign in the field transformation is plus, instead of minus, because we’re doing an active, as opposed to passive, transformation). Similarly, once we substitute a specific field configuration ϕ(x) into the Lagrangian, the Lagrangian itself also transforms as

(x)(x)+ϵνν(x) (1.42)

Since the change in the Lagrangian is a total derivative, we may invoke Noether’s theorem which gives us four conserved currents (jμ)ν, one for each of the translations ϵν with ν=0,1,2,3,

(jμ)ν=(μϕa)νϕa-δνμTνμ (1.43)

Tνμ is called the energy-momentum tensor. It satisfies

μTνμ=0 (1.44)

The four conserved quantities are given by

E=d3xT00  and   Pi=d3xT0i (1.45)

where E is the total energy of the field configuration, while Pi is the total momentum of the field configuration.

An Example of the Energy-Momentum Tensor

Consider the simplest scalar field theory with Lagrangian (1.1.1). From the above discussion, we can compute

Tμν=μϕνϕ-ημν (1.46)

One can verify using the equation of motion for ϕ that this expression indeed satisfies μTμν=0. For this example, the conserved energy and momentum are given by

E = d3x12ϕ˙2+12(ϕ)2+12m2ϕ2 (1.47)
Pi = d3xϕ˙iϕ (1.48)

Notice that for this example, Tμν came out symmetric, so that Tμν=Tνμ. This won’t always be the case. Nevertheless, there is typically a way to massage the energy momentum tensor of any theory into a symmetric form by adding an extra term

Θμν=Tμν+ρΓρμν (1.49)

where Γρμν is some function of the fields that is anti-symmetric in the first two indices so Γρμν=-Γμρν. This guarantees that μρΓρμν=0 so that the new energy-momentum tensor is also a conserved current.

A Cute Trick

One reason that you may want a symmetric energy-momentum tensor is to make contact with general relativity: such an object sits on the right-hand side of Einstein’s field equations. In fact this observation provides a quick and easy way to determine a symmetric energy-momentum tensor. Firstly consider coupling the theory to a curved background spacetime, introducing an arbitrary metric gμν(x) in place of ημν, and replacing the kinetic terms with suitable covariant derivatives using “minimal coupling”. Then a symmetric energy momentum tensor in the flat space theory is given by

Θμν=-2-g(-g)gμν|gμν=ημν (1.50)

It should be noted however that this trick requires a little more care when working with spinors.

1.3.3 Another Example: Lorentz Transformations and Angular Momentum

In classical particle mechanics, rotational invariance gave rise to conservation of angular momentum. What is the analogy in field theory? Moreover, we now have further Lorentz transformations, namely boosts. What conserved quantity do they correspond to? To answer these questions, we first need the infinitesimal form of the Lorentz transformations

Λνμ=δνμ+ωνμ (1.51)

where ωνμ is infinitesimal. The condition (1.26) for Λ to be a Lorentz transformation becomes

    (δσμ+ωσμ)(δτν+ωτν)ηστ=ημν (1.52)
   ωμν+ωνμ=0

So the infinitesimal form ωμν of the Lorentz transformation must be an anti-symmetric matrix. As a check, the number of different 4×4 anti-symmetric matrices is 4×3/2=6, which agrees with the number of different Lorentz transformations (3 rotations + 3 boosts). Now the transformation on a scalar field is given by

ϕ(x)ϕ(x) = ϕ(Λ-1x) (1.53)
= ϕ(xμ-ωνμxν)
= ϕ(xμ)-ωνμxνμϕ(x)

from which we see that

δϕ=-ωνμxνμϕ (1.54)

By the same argument, the Lagrangian density transforms as

δ=-ωνμxνμ=-μ(ωνμxν) (1.55)

where the last equality follows because ωμμ=0 due to anti-symmetry. Once again, the Lagrangian changes by a total derivative so we may apply Noether’s theorem (now with Fμ=-ωνμxν) to find the conserved current

jμ = -(μϕ)ωνρxνρϕ+ωνμxν (1.56)
= -ωνρ[(μϕ)xνρϕ-δρμxν]=-ωνρTρμxν

Unlike in the previous example, I’ve left the infinitesimal choice of ωνμ in the expression for this current. But really, we should strip it out to give six different currents, i.e. one for each choice of ωνμ. We can write them as

(𝒥μ)ρσ=xρTμσ-xσTμρ (1.57)

which satisfy μ(𝒥μ)ρσ=0 and give rise to 6 conserved charges. For ρ,σ=1,2,3, the Lorentz transformation is a rotation and the three conserved charges give the total angular momentum of the field.

Qij=d3x(xiT0j-xjT0i) (1.58)

But what about the boosts? In this case, the conserved charges are

Q0i=d3x(x0T0i-xiT00) (1.59)

The fact that these are conserved tells us that

0=dQ0idt = d3xT0i+td3xT0it-ddtd3xxiT00 (1.60)
= Pi+tdPidt-ddtd3xxiT00

But we know that Pi is conserved, so dPi/dt=0, leaving us with the following consequence of invariance under boosts:

ddtd3xxiT00=constant (1.61)

This is the statement that the center of energy of the field travels with a constant velocity. It’s kind of like a field theoretic version of Newton’s first law but, rather surprisingly, appearing here as a conservation law.

1.3.4 Internal Symmetries

The above two examples involved transformations of spacetime, as well as transformations of the field. An internal symmetry is one that only involves a transformation of the fields and acts the same at every point in spacetime. The simplest example occurs for a complex scalar field ψ(x)=(ϕ1(x)+iϕ2(x))/2. We can build a real Lagrangian by

=μψμψ-V(|ψ|2) (1.62)

where the potential is a general polynomial in |ψ|2=ψψ. To find the equations of motion, we could expand ψ in terms of ϕ1 and ϕ2 and work as before. However, it’s easier (and equivalent) to treat ψ and ψ as independent variables and vary the action with respect to both of them. For example, varying with respect to ψ leads to the equation of motion

μμψ+V(ψψ)ψ=0 (1.63)

The Lagrangian has a continuous symmetry which rotates ϕ1 and ϕ2 or, equivalently, rotates the phase of ψ:

ψeiαψ   or  δψ=iαψ (1.64)

where the latter equation holds with α infinitesimal. The Lagrangian remains invariant under this change: δ=0. The associated conserved current is

jμ=i(μψ)ψ-iψ(μψ) (1.65)

We will later see that the conserved charges arising from currents of this type have the interpretation of electric charge or particle number (for example, baryon or lepton number).

Non-Abelian Internal Symmetries

Consider a theory involving N scalar fields ϕa, all with the same mass and the Lagrangian

=12a=1Nμϕaμϕa-12a=1Nm2ϕa2-g(a=1Nϕa2)2 (1.66)

In this case the Lagrangian is invariant under the non-Abelian symmetry group G=SO(N). (Actually O(N) in this case). One can construct theories from complex fields in a similar manner that are invariant under an SU(N) symmetry group. Non-Abelian symmetries of this type are often referred to as global symmetries to distinguish them from the “local gauge” symmetries that you will meet later. Isospin is an example of such a symmetry, albeit realized only approximately in Nature.

Another Cute Trick

There is a quick method to determine the conserved current associated to an internal symmetry δϕ=αϕ for which the Lagrangian is invariant. Here, α is a constant real number. (We may generalize the discussion easily to a non-Abelian internal symmetry for which α becomes a matrix). Now consider performing the transformation but where α depends on spacetime: α=α(x). The action is no longer invariant. However, the change must be of the form

δ=(μα)hμ(ϕ) (1.67)

since we know that δ=0 when α is constant. The change in the action is therefore

δS=d4xδ=-d4xα(x)μhμ (1.68)

which means that when the equations of motion are satisfied (so δS=0 for all variations, including δϕ=α(x)ϕ) we have

μhμ=0 (1.69)

We see that we can identify the function hμ=jμ as the conserved current. This way of viewing things emphasizes that it is the derivative terms, not the potential terms, in the action that contribute to the current. (The potential terms are invariant even when α=α(x)).

1.4 The Hamiltonian Formalism

The link between the Lagrangian formalism and the quantum theory goes via the path integral. In this course we will not discuss path integral methods, and focus instead on canonical quantization. For this we need the Hamiltonian formalism of field theory. We start by defining the momentum πa(x) conjugate to ϕa(x),

πa(x)=ϕ˙a (1.70)

The conjugate momentum πa(x) is a function of x, just like the field ϕa(x) itself. It is not to be confused with the total momentum Pi defined in (1.45) which is a single number characterizing the whole field configuration. The Hamiltonian density is given by

=πa(x)ϕ˙a(x)-(x) (1.71)

where, as in classical mechanics, we eliminate ϕ˙a(x) in favour of πa(x) everywhere in . The Hamiltonian is then simply

H=d3x (1.72)

An Example: A Real Scalar Field

For the Lagrangian

=12ϕ˙2-12(ϕ)2-V(ϕ) (1.73)

the momentum is given by π=ϕ˙, which gives us the Hamiltonian,

H=d3x12π2+12(ϕ)2+V(ϕ) (1.74)

Notice that the Hamiltonian agrees with the definition of the total energy (1.47) that we get from applying Noether’s theorem for time translation invariance.

In the Lagrangian formalism, Lorentz invariance is clear for all to see since the action is invariant under Lorentz transformations. In contrast, the Hamiltonian formalism is not manifestly Lorentz invariant: we have picked a preferred time. For example, the equations of motion for ϕ(x)=ϕ(x,t) arise from Hamilton’s equations,

ϕ˙(x,t)=Hπ(x,t)   and   π˙(x,t)=-Hϕ(x,t) (1.75)

which, unlike the Euler-Lagrange equations (1.8), do not look Lorentz invariant. Nevertheless, even though the Hamiltonian framework doesn’t look Lorentz invariant, the physics must remain unchanged. If we start from a relativistic theory, all final answers must be Lorentz invariant even if it’s not manifest at intermediate steps. We will pause at several points along the quantum route to check that this is indeed the case.