34
2 Classical Field Theory In what follows we will consider rather general field theories. The only guid- ing principles that we will use in constructing these theories are (a) symme- tries and (b) a generalized Least Action Principle. 2.1 Relativistic Invariance Before we saw three examples of relativistic wave equations. They are the Maxwell equations for classical electromagnetism, the Klein-Gordon equa- tion and the Dirac equation. Maxwell’s equations govern the dynamics of a vector field, the vector potentials A μ (x)=(A 0 (x), A(x)), whereas the Klein-Gordon equation describes excitations of a scalar field φ(x) and the Dirac equation governs the behavior of the four-component spinor field ψ α (x)(α =0, 1, 2, 3). Each one of these fields transforms in a very definite way under the group of Lorentz transformations, the Lorentz group. The Lorentz group is defined as a group of linear transformations Λ of Minkowski space-time M onto itself Λ : M M such that the new coordinates are related to the old ones by a linear (Lorentz) transformation x μ = Λ μ ν x ν (2.1) The space-time components of Λ 0 i are the Lorentz boosts which relate inertial reference frames moving at relative velocity v from each other. Thus,

582-Chapter2 Classical Field Theory

Embed Size (px)

DESCRIPTION

Fradkin

Citation preview

Page 1: 582-Chapter2 Classical Field Theory

2

Classical Field Theory

In what follows we will consider rather general field theories. The only guid-ing principles that we will use in constructing these theories are (a) symme-tries and (b) a generalized Least Action Principle.

2.1 Relativistic Invariance

Before we saw three examples of relativistic wave equations. They are theMaxwell equations for classical electromagnetism, the Klein-Gordon equa-tion and the Dirac equation. Maxwell’s equations govern the dynamics ofa vector field, the vector potentials Aµ(x) = (A0(x),A(x)), whereas theKlein-Gordon equation describes excitations of a scalar field !(x) and theDirac equation governs the behavior of the four-component spinor field"!(x)(# = 0, 1, 2, 3). Each one of these fields transforms in a very definiteway under the group of Lorentz transformations, the Lorentz group. TheLorentz group is defined as a group of linear transformations ! of Minkowskispace-time M onto itself ! : M !" M such that the new coordinates arerelated to the old ones by a linear (Lorentz) transformation

x!µ = !µ"x

" (2.1)

The space-time components of !0i are the Lorentz boosts which relate

inertial reference frames moving at relative velocity $v from each other. Thus,

Page 2: 582-Chapter2 Classical Field Theory

12 Classical Field Theory

Lorentz boosts along the x1-axis have the familiar form

x0! =x0 + vx1/c!1# v2/c2

x1! =x1 + vx0/c!1# v2/c2

x2! = x2

x3! = x3

(2.2)

where x0 = ct, x1 = x, x2 = y and x3 = z (note: these are components, notpowers!). If we use the notation % = (1# v2/c2)"1/2 $ cosh#, we can writethe Lorentz boost as a matrix:

"

##$

x0!

x1!

x2!

x3!

%

&&' =

"

##$

cosh# sinh# 0 0sinh# cosh# 0 0

0 0 1 00 0 0 1

%

&&'

"

##$

x0

x1

x2

x3

%

&&' (2.3)

The space components of !ij are conventional rotations R of three-dimensional

Euclidean space.Infinitesimal Lorentz transformations are generated by the hermitian op-

erators

Lµ" = i(xµ&" # x"&µ) (2.4)

where &µ = ##xµ and µ, ' = 0, 1, 2, 3. The infinitesimal generators Lµ" satisfy

the algebra

[Lµ" , L$%] = ig"$Lµ% # igµ$L"% # ig"%Lµ$ + igµ%L"$ (2.5)

where gµ" is the metric tensor for flat Minkowski space-time (see below).This is the algebra of the group SO(3, 1). Actually any operator of the form

Mµ" = Lµ" + Sµ" (2.6)

where Sµ" are 4% 4 matrices satisfying the algebra of Eq.(2.5) satisfies thesame algebra. Below we will discuss explicit examples.

Lorentz transformations form a group, since (a) the product of two Lorentztransformations is a Lorentz transformation, (b) there exists an identitytransformation, and (c) Lorentz transformations are invertible. Notice, how-ever, that in general two transformations do not commute with each other.Hence, the Lorentz group is non-Abelian.

Page 3: 582-Chapter2 Classical Field Theory

2.1 Relativistic Invariance 13

The Lorentz group has the defining property of leaving invariant the rel-ativistic interval

x2 $ x20 # x2 = c2t2 # x2 (2.7)

The group of Euclidean rotations leave invariant the Euclidean distance x2

and it is a subgroup of the Lorentz group. The rotation group is denoted bySO(3), and the Lorentz group is denoted by SO(3, 1). This notation makesmanifest the fact that the signature of the metric has one + sign and three# signs.

We will adopt the following conventions and definitions:

1. Metric Tensor: We will use the standard (“Bjorken and Drell”) metricfor Minkowski space-time in which the metric tensor gµ" is

gµ" = gµ" =

"

##$

1 0 0 00 #1 0 00 0 #1 00 0 0 #1

%

&&' (2.8)

With this notation the infinitesimal relativistic interval is

ds2 = dxµdxµ = gµ"dxµdx" = dx20 # dx2 = c2dt2 # dx2 (2.9)

2. 4-vectors:

1. xµ is a contravariant 4-vector, xµ = (ct,x)2. xµ is a covariant 4-vector xµ = (ct,#x)3. Covariant and contravariant vectors (and tensors) are related through

the metric tensor gµ"

Aµ = gµ"A" (2.10)

4. x is a vector in R3

5. pµ =(Ec ,p

)is the energy-momentum 4-vector. Hence, pµpµ = E2

c2 #p2

is a Lorentz scalar.

3. Scalar Product:

p · q = pµqµ = p0q0 # p · q $ pµq"g

µ" (2.11)

4. Gradients: &µ $ ##xµ and &µ $ #

#xµ. We define the D’Alambertian &2

&2 $ &µ&µ $1

c2&2t #&2 (2.12)

which is a Lorentz scalar. From now on we will use units of time [T ] andlength [L] such that ! = c = 1. Thus, [T ] = [L] and we will use units likecentimeters (or any other unit of length).

Page 4: 582-Chapter2 Classical Field Theory

14 Classical Field Theory

5. Interval: The interval in Minkowski space is x2,

x2 = xµxµ = x20 # x2 (2.13)

Time-like intervals have x2 > 0 while space-like intervals have x2 < 0.

!"#$%&'()*+,-./0123456789:;<=>?@ABCDEFGHIJKLMNOPQRSTUVWXYZ[\]_abcdefghijklmnopqrstuvwxyz{|}~

time-like

space-like

x0

x1

x2

> 0

x2

< 0

Figure 2.1 The Minkowski space-time and its light cone. Events at a rel-ativistic interval with x2 = x2

0 # x2 > 0 are time-like (and are causallyconnected with the origin), while events with x2 = x2

0 # x2 < 0 are space-like and are not causally connected with the origin.

Since a field is a function (or mapping) of Minkowski space onto some other(properly chosen) space, it is natural to require that the fields should havesimple transformation properties under Lorentz transformations. For exam-ple, the vector potential Aµ(x) transforms like 4-vector under Lorentz trans-formations, i.e., if x!µ = !µ

"x" , then A!µ(x!) = !µ"A"(x). In other words, Aµ

transforms like xµ. Thus, it is a vector. All vector fields have this property.A scalar field "(x), on the other hand, remains invariant under Lorentztransformations,

"!(x!) = "(x) (2.14)

A 4-spinor "!(x) transforms under Lorentz transformations. Namely, thereexists an induced 4% 4 linear transformation matrix S(!) such that

S(!"1) = S"1(!) (2.15)

Page 5: 582-Chapter2 Classical Field Theory

2.2 The Lagrangian, the Action and and the Least Action Principle 15

and

#!(!x) = S(!)#(x) (2.16)

Below we will give an explicit expression for S(!).

2.2 The Lagrangian, the Action and and the Least ActionPrinciple

The evolution of any dynamical system is determined by its Lagrangian. Inthe Classical Mechanics of systems of particles described by the generalizedcoordinates q, the Lagrangian L is a di$erentiable function of the coordinatesq and their time derivatives. L must be di$erentiable since, otherwise, theequations of motion would not be local in time, i.e. could not be writtenin terms of di$erential equations. An argument a-la Landau-Lifshitz enablesus to “derive” the Lagrangian. For example, for a particle in free space, thehomogeneity, uniformity and isotropy of space and time require that L beonly a function of the absolute value of the velocity |v|. Since |v| is not adi$erentiable function of v, the Lagrangian must be a function of v2. Thus,L = L(v2). In principle there is no reason to assume that L cannot be afunction of the acceleration a (or rather a2) or of its higher derivatives.Experiment tells us that in Classical Mechanics it is su%cient to specify theinitial position x(0) of a particle and its initial velocity v(0) in order todetermine the time evolution of the system. Thus we have to choose

L(v2) = const +1

2mv2 (2.17)

The additive constant is irrelevant in classical physics. Naturally, the coef-ficient of v2 is just one-half of the inertial mass.

However, in Special Relativity, the natural invariant quantity to consideris not the Lagrangian but the action S. For a free particle the relativisticinvariant (i.e., Lorentz invariant ) action must involve the invariant interval

or the invariant length ds = c*

1# v2

c2 dt, the proper length. Hence onewrites

S = mc

+ sf

si

ds = mc2+ tf

ti

dt

,1# v2

c2(2.18)

Thus the relativistic Lagrangian is

L = mc2,

1# v2

c2(2.19)

As a power series expansion, it contains all powers of v2/c2.

Page 6: 582-Chapter2 Classical Field Theory

16 Classical Field Theory

t

q

i

Figure 2.2 The Least Action Principle: the dark curve is the classical tra-jectory and extremizes the classical action. The curve with a broken tracerepresents a variation.

Once the Lagrangian is found, the classical equations of motion are de-termined by the Least Action Principle. Thus, we construct the action S

S =

+dt L

-q,&q

&t

.(2.20)

and demand that the physical trajectories q(t) leave the action S station-ary,i.e., (S = 0. The variation of S is

(S =

+ tf

ti

dt

/&L

&q(q +

&L

& dqdt

(dq

dt

0

(2.21)

Integrating by parts, we get

(S =

+ tf

ti

dtd

dt

/&L

& dqdt

(q

0

+

+ tf

ti

dt (q

1&L

&q# d

dt

/&L

& dqdt

02

(2.22)

Hence, we get

(S =&L

& dqdt

(q333tf

ti+

+ tf

ti

dt (q

1&L

&q# d

dt

/&L

& dqdt

02

(2.23)

If we assume that the variation (q is an arbitrary function of time thatvanishes at the initial and final times ti and tf (hence (q(ti) = (q(tf ) =0), we find that (S = 0 if and only if the integrand of Eq.(2.23) vanishes

Page 7: 582-Chapter2 Classical Field Theory

2.3 Scalar Field Theory 17

identically. Thus,

&L

&q# d

dt

/&L

& dqdt

0

= 0 (2.24)

These are the equations of motion or Newton’s equations. In general theequations that determine the trajectories which leave the action stationaryare called the Euler-Lagrange equations.

2.3 Scalar Field Theory

For the case of a field theory, we can proceed very much in the same way.Let us consider first the case of a scalar field "(x). The action S must beinvariant under Lorentz transformations. Since we want to construct localtheories it is natural to assume that S is given in terms of a Lagrangiandensity L

S =

+d4x L (2.25)

Since the volume element of Minkowski space d4x is invariant under Lorentztransformations, the action S is invariant if L is a local, di$erentiable func-tion of Lorentz invariants that can be constructed out of the field "(x). Sim-ple invariants are "(x) itself and all of its powers. The gradient &µ" $ #!

#xµ

is not an invariant but the D’alambertian &2" is. &µ"&µ" is also an invari-ant under a change of the sign of ". So, we can write the following simpleexpression for L:

L =1

2&µ"&

µ"# V (") (2.26)

where V (") is some potential, which we can assume is a polynomial functionof the field ". Let us consider the simple choice

V (") =1

2m2"2 (2.27)

where m = mc/!. Thus,

L =1

2&µ"&

µ"# 1

2m2"2 (2.28)

This is the Lagrangian density for a free scalar field. We will discuss lateron in what sense this field is “free”. Notice, in passing, that we could haveadded a term like &2". However this term, in addition of being odd under" " #", is a total divergence and, as such, it has an e$ect only on the

Page 8: 582-Chapter2 Classical Field Theory

18 Classical Field Theory

boundary conditions but it does not a$ect the equations of motion. In whatfollows will will not consider surface terms.

The Least Action Principle requires that S be stationary under arbitraryvariations of the field " and of its derivatives &µ". Thus, we get

(S =

+d4x

4(L("

(" +(L(&µ"

(&µ"

5(2.29)

Notice that since L is a functional of ", we have to use functional derivatives,i.e., partial derivatives at each point of space-time. Upon integrating byparts, we get

(S =

+d4x &µ

-(L(&µ"

("

.+

+d4x ("

4(L("# &µ

-(L(&µ"

.5(2.30)

Instead of considering initial and final conditions, we now have to imaginethat the field " is contained inside some very large box of space-time. Theterm with the total divergence yields a surface contribution. We will considerfield configurations such that (" = 0 on that surface. Thus, the Euler-Lagrange equations are

(L("# &µ

-(L(&µ"

.= 0 (2.31)

More explicitly, we find(L("

= #&V&"

(2.32)

and&L(&µ"

= &µ"' &µ(L(&µ"

= &µ&µ" = &2" (2.33)

By direct substitution we get the equation of motion (or field equation)

&2"+&V

&"= 0 (2.34)

For the choice

V (") =m2

2"2 ' &V

&"= m2" (2.35)

the field equation is(&2 + m2

)" = 0 (2.36)

where &2 = 1c2

#2

#t2 #(2. Thus, we find that the equation of motion for the

free massive scalar field " is

1

c2&2"

&t2#(2"+ m2" = 0 (2.37)

Page 9: 582-Chapter2 Classical Field Theory

2.3 Scalar Field Theory 19

This is precisely the Klein-Gordon equation if the constant m is identifiedwith mc

!. Indeed, the plane-wave solutions of these equations are

" = "0ei(p0x0"p·x)/! (2.38)

where p0 and $p are related through the dispersion law

p20 = p2c2 +m2c4 (2.39)

which means that, for each momentum p, there are two solutions, one withpositive frequency and one with negative frequency. We will see below that,in the quantized theory, the energy of the excitation is indeed equal to|p0|. Notice that 1

m = !

mc has units of length and is equal to the Comptonwavelength for a particle of mass m. From now on (unless it stated thecontrary) I will use units in which ! = c = 1 in which m = m.

The Hamiltonian for a classical field is found by a straightforward gener-alization of the Hamiltonian of a classical particle. Namely, one defines thecanonical momentum &(x), conjugate to the field (the “coordinate ”) "(x),

&(x) =(L

("(x)(2.40)

where

"(x) =&"

&x0(2.41)

In Classical Mechanics the Hamiltonian H and the Lagrangian L are relatedby

H = pq # L (2.42)

where q is the coordinate and p the canonical momentum conjugate to q.Thus, for a scalar field theory the Hamiltonian density H is

H = &(x)"(x)# L

=1

2&2(x) +

1

2(!"(x))2 + V ("(x))

(2.43)

Hence, for a free massive scalar field the Hamiltonian is

H =1

2&2(x) +

1

2(!"(x))2 +

m2

2"2(x) ) 0 (2.44)

which is always a positive definite quantity. Thus, the energy of a planewave solution of a massive scalar field theory (i.e., a solution of the Klein-Gordon equation) is always positive, no matter the sign of the frequency.

Page 10: 582-Chapter2 Classical Field Theory

20 Classical Field Theory

In fact, the lowest energy state is simply " = constant. A solution made oflinear superpositions of plane waves (i.e., a wave packet) has positive energy.Therefore, in field theory, the energy is always positive. We will see that, inthe quantized theory, the negative frequency solutions are identified withantiparticle states and their existence do not signal a possible instability ofthe theory.

2.4 Classical Field Theory in the Canonical (Hamiltonian)Formalism

In Classical Mechanics it is often convenient to use the canonical formulationin terms of a Hamiltonian instead of the Lagrangian approach. For the caseof a system of particles, the canonical formalism proceeds as follows. Givena Lagrangian L(q, q), a canonical momentum p is defined to be

&L

&q= p (2.45)

The classical Hamiltonian H(p, q) is defined by the Legendre transformation

H(p, q) = pq # L(q, q) (2.46)

If the Lagrangian L is quadratic in the velocities q and separable, e.g.

L =1

2mq2 # V (q) (2.47)

then, H(pq) is simply given by

H(p, q) = pq # (mq2

2# V (q)) =

p2

2m+ V (q) (2.48)

where

p =&L

&q= mq (2.49)

The (conserved) quantity H is then identified with the total energy of thesystem.

In this language, the Least Action Principle becomes

(S = (

+L dt = (

+[pq #H(p, q)] dt = 0 (2.50)

Hence +dt

-(p q + p (q # (p &H

&p# (q &H

&q

.= 0 (2.51)

Page 11: 582-Chapter2 Classical Field Theory

2.4 Classical Field Theory in the Canonical (Hamiltonian) Formalism 21

Upon an integration by parts we get+

dt

4(p

-q # &H

&p

.+ (q

-#&H&q# p

.5= 0 (2.52)

which can only be satisfied for arbitrary variations (q(t) and (p(t) if

q =&H

&pp = #&H

&q(2.53)

These are Hamilton’s equations.Let us introduce the Poisson Bracket {A,B}qp of two functions A and B

of q and p by

{A,B}qp $&A

&q

&B

&p# &A

&p

&B

&q(2.54)

Let F (q, p, t) be some di$erentiable function of q, p and t. Then the totaltime variation of F is

dF

dt=&F

&t+&F

&q

dq

dt+&F

&p

dp

dt(2.55)

Using Hamilton’s Equations we get the result

dF

dt=&F

&t+&F

&q

&H

&p# &F

&p

&H

&q(2.56)

or, in terms of Poisson Brackets,

dF

dt=&F

&t+ {F,H, }qp (2.57)

In particular,

dq

dt=&H

&p=&q

&q

&H

&p# &q

&p

&H

&q= {q,H}qp (2.58)

since&q

&p= 0 and

&q

&q= 1 (2.59)

Also the total rate of change of the canonical momentum p is

dp

dt=&p

&q

&H

&p# &p

&p

&H

&q$ #&H

&q(2.60)

since #p#q = 0 and #p

#p = 1. Thus,

dp

dt= {p,H}qp (2.61)

Page 12: 582-Chapter2 Classical Field Theory

22 Classical Field Theory

Notice that, for an isolated system, H is time-independent. So,

&H

&t= 0 (2.62)

and

dH

dt=&H

&t+ {H,H}qp = 0 (2.63)

since

{H,H}qp = 0 (2.64)

Therefore, H can be regarded as the generator of infinitesimal time transla-tions. Since it is conserved for an isolated system, for which #H

#t = 0, we canindeed identify H with the total energy. In passing, let us also notice thatthe above definition of the Poisson Bracket implies that q and p satisfy

{q, p}qp = 1 (2.65)

This relation is fundamental for the quantization of these systems.Much of this formulation can be generalized to the case of fields. Let us

first discuss the canonical formalism for the case of a scalar field " withLagrangian density L(", &µ,"). We will choose "(x) to be the (infinite) setof canonical coordinates. The canonical momentum &(x) is defined by

&(x) =(L

(&0"(x)(2.66)

If the Lagrangian is quadratic in &µ", the canonical momentum &(x) issimply given by

&(x) = &0"(x) $ "(x) (2.67)

The Hamiltonian density H(",&) is a local function of "(x) and &(x) givenby

H(",&) = &(x) &0"(x)# L(", &0") (2.68)

If the Lagrangian density L has the simple form

L =1

2(&µ")

2 # V (") (2.69)

then, the Hamiltonian density H(",&) is

H = &"# L(", ", &j") $1

2&2(x) +

1

2(!"(x))2 + V ("(x)) (2.70)

Page 13: 582-Chapter2 Classical Field Theory

2.5 Field Theory of the Dirac Equation 23

The canonical field "(x) and the canonical momentum &(x) satisfy theequal-time Poisson Bracket (PB) relations

{"(x, x0),&(y, x0)}PB = ((x# y) (2.71)

where ((x) is the Dirac (-function and {A,B}PB is now

{A,B}PB =

+d3x

4(A

("(x, x0)

(B

(&(x, x0)# (A

(&(x, x0)

(B

("(x, x0)

5(2.72)

for any two functionals A and B of "(x) and &(x). This approach can beextended to theories other than that of a scalar field without too muchdi%culty. We will come back to these issues when we consider the problemof quantization. Finally we should note that while Lorentz invariance isapparent in the Lagrangian formulation, it is not so in the Hamiltonianformulation of a classical field.

2.5 Field Theory of the Dirac Equation

We now turn to the problem of a field theory for spinors. We will discussthe theory of spinors as a classical fuel theory. We will find that this theoryis not consistent unless it is properly quantized as a quantum field theory ofspinors. We will return to this in a later chapter.

Let us rewrite the Dirac equation

i!&#

&t=

!c

i! ·!#+ )mc2 # $ HDirac# (2.73)

in a manner in which relativistic covariance is apparent. The operator HDirac

defines the Dirac Hamiltonian.We first recall that the 4 % 4 hermitean matrices $# and ) should satisfy

the algebra

{#i,#j} = 2(ij I, {#i,)} = 0, #2i = )2 = I (2.74)

where I is the 4% 4 identity matrix.A simple realization of this algebra is given by the 2 % 2 block (Dirac)

matrices

#i =

-0 *i

*i 0

.) =

-I 00 #I

.(2.75)

where the *i’s are the three Pauli matrices

*1 =

-0 11 0

.*2 =

-0 #ii 0

.*3 =

-1 00 #1

.(2.76)

Page 14: 582-Chapter2 Classical Field Theory

24 Classical Field Theory

and now I is the 2 % 2 identity matrix. This is the Dirac representation ofthe Dirac algebra.

It is now convenient to introduce the Dirac %-matrices which are definedby the following relations:

%0 = ) %i = )#i (2.77)

Thus, the matrices %µ are

%0 = ) =

-I 00 #I

., %i =

-0 *i

#*i 0

.(2.78)

which obey the algebra

{%µ, %"} = 2gµ" I (2.79)

where I is the 4% 4 identity matrix.In terms of the %-matrices, the Dirac equation takes the much simpler

form

(i%µ&µ #mc

!)# = 0 (2.80)

where # is a 4-spinor. It is also customary to introduce the notation (knownas Feynman’s slash)

/a $ aµ%µ (2.81)

Using Feynman’s slash, we can write the Dirac equation in the form

(i/& # mc

!)# = 0 (2.82)

From now on I will use units in which ! = c = 1. Thus energy is measuredin units of (length)"1 and time in units of length.

Notice that, if # satisfies the Dirac equation, then

(i/& +m)(i/& #m)# = 0 (2.83)

Also,

/& · /& = &µ&"%µ%" = &µ&"

-1

2{%µ, %"}+ 1

2[%µ%" ]

.

= &µ&"gµ" = &2

(2.84)

where I used the fact that the commutator [%µ, %" ] is antisymmetric in theindices µ and '. Thus, # satisfies the Klein-Gordon equation

(&2 +m2

)# = 0 (2.85)

Page 15: 582-Chapter2 Classical Field Theory

2.5 Field Theory of the Dirac Equation 25

2.5.1 Solutions of the Dirac Equation

Let us briefly discuss the properties of the solutions of the Dirac equation.Let us first consider solutions representing particles at rest. Thus # mustbe constant in space and all its space derivatives must vanish. The Diracequation becomes

i%0&#

&t= m# (2.86)

where t = x0 (c = 1). Let us introduce the bispinors ! and +

# =

-!+

.(2.87)

We find that the Dirac equation reduces to a simple system of two 2 % 2equations

i&!

&t= +m!

i&+

&t= #m+

(2.88)

The solutions are

!1 = e"imt

-10

.!2 = e"imt

-01

.(2.89)

and

+1 = eimt

-10

.+2 = eimt

-01

.(2.90)

Thus, the upper component ! represents the solutions with positive energywhile + represents the solutions with negative energy. The additional two-fold degeneracy of the solutions is connected to the spin of the particle.

More generally, in terms of the bispinors ! and + the Dirac Equation takesthe form,

i&!

&t= m!+

1

i" ·!+ (2.91)

i&+

&t= #m++

1

i" ·!! (2.92)

In the limit c"*, it reduces to the Schrodinger-Pauli equation. The slowlyvarying amplitudes ! and +, defined by

! = e"imt!

+ = e"imt+ (2.93)

Page 16: 582-Chapter2 Classical Field Theory

26 Classical Field Theory

with + small and nearly static, define positive energy solutions with energiesclose to +m. In terms of ! and +, the Dirac equation becomes

i&!

&t=

1

i" ·!+ (2.94)

i&+

&t= #2m++

1

i" ·!! (2.95)

Indeed, in this limit, the l. h. s. of Eq. (2.95) is much smaller than its r. h.s. Thus we can approximate

2m+ + 1

i" ·!! (2.96)

We can now eliminate the “small component” + from Eq. (2.94) to find that! satisfies

i&!

&t= # 1

2m&2! (2.97)

which is indeed the Schrodinger-Pauli equation.

2.5.2 Conserved Current

Let us introduce one last bit of useful notation. Let us define # by

# = #† %0 (2.98)

in terms of which we can write down the 4-vector jµ

jµ = #%µ# (2.99)

which is conserved, i.e.,

&µjµ = 0 (2.100)

Notice that the time component of jµ is the density

j0 = #%0# $ #†# (2.101)

and that the space components of jµ are

j = ### = #†%0## = #†!# (2.102)

Thus the Dirac equation has an associated four vector field, jµ(x), which isconserved and hence obeys a local continuity equation

&0j0 +! · j = 0 (2.103)

Page 17: 582-Chapter2 Classical Field Theory

2.5 Field Theory of the Dirac Equation 27

However it is easy to see that in general the density j0 can be positive ornegative. Hence this current cannot be associated with a probability current(as in non-relativistic quantum mechanics). Instead we will see that that itis associated with the charge density and current.

2.5.3 Relativistic Covariance

Let ! be a Lorentz transformation, #(x) the spinor field in an inertial frameand #!(x!) be the Dirac spinor field in the transformed frame. The Diracequation is covariant if the Lorentz transformation

x!µ = !"µx" (2.104)

induces a linear transformation S(!) in spinor space

#!!(x

!) = S(!)!&#&(x) (2.105)

such that the transformed Dirac equation has the same form as the originalequation in the original frame, i.e. we will require that-i%µ

&

&xµ#m

.

!&

#&(x) = 0, and

-i%µ

&

&x!µ#m

.

!&

#!&(x

!) = 0

(2.106)Notice two important facts: (1) both the field # and the coordinate x changeunder the action of the Lorentz transformation, and (2) the %-matrices andthe mass m do not change under a Lorentz transformation. Thus, the %-matrices are independent of the choice of a reference frame. However, theydo depend on the choice of the set of basis states in spinor space.

What properties should the representation matrices S(!) have? Let usfirst observe that if x!µ = !µ

"x" , then

&

&x!µ=&x"

&x!µ&

&x"$(!"1

)"µ

&

&x"(2.107)

Thus, ##xµ is a covariant vector. By substituting this transformation law back

into the Dirac equation, we find

i%µ&

&x!µ#!(x!) = i%µ(!"1)"µ

&

&x"(S(!)#(x)) (2.108)

Thus, the Dirac equation now reads

i%µ(!"1

)"µS(!)

&#

&x"#mS (!)# = 0 (2.109)

Page 18: 582-Chapter2 Classical Field Theory

28 Classical Field Theory

Or, equivalently

S"1 (!) i%µ(!"1

)"µS (!)

&#

&x"#m# = 0 (2.110)

This equation is covariant provided that S (!) satisfies

S"1 (!) %µS (!)(!"1

)"µ= %" (2.111)

Since the set of Lorentz transformations form a group, i.e., the product oftwo Lorentz transformations !1 and !2 is the new Lorentz transformation!1!2 and the inverse of the transformation ! is the inverse matrix !"1, therepresentation matrices S(!) should also form a group and obey the sameproperties. In particular,

S"1 (!) = S(!"1

)(2.112)

must hold. Recall that the invariance of the relativistic interval x2 = xµxµ

implies that ! must obey

!"µ!'" = g'µ $ ('µ (2.113)

Thus,

!"µ =(!"1

)µ" (2.114)

So we write,

S (!) %µS (!)"1 =(!"1

)µ"%" (2.115)

Eq.(2.115) shows that a Lorentz transformation induces a similarity trans-formation on the %-matrices which is equivalent to (the inverse of) a Lorentztransformation. For the case of Lorentz boosts, Eq.(2.115) shows that thematrices S(!) are hermitean. However, for the subgroup SO(3) of rotationsabout a fixed origin, the matrices S(!) are unitary.

We will now find the form of S (!) for an infinitesimal Lorentz transforma-tion. Since the identity transformation is !µ

" = gµ" , a Lorentz transformationwhich is infinitesimally close to the identity should have the form

!µ" = gµ" + ,µ

"

(!"1

)µ"= gµ" # ,µ

" (2.116)

where ,µ" is infinitesimal and antisymmetric in its space-time indices

,µ" = #,"µ

,µ" = ,µ$g

$"

(2.117)

Page 19: 582-Chapter2 Classical Field Theory

2.5 Field Theory of the Dirac Equation 29

Let us parameterize S (!) in terms of a 4 % 4 matrix *µ" which is alsoantisymmetric in its indices, i.e., *µ" = #*"µ. Then, we can write

S (!) = I # i

4*µ",

µ" + . . .

S"1 (!) = I +i

4*µ",

µ" + . . .

(2.118)

where I stands for the 4% 4 identity matrix. If we substitute back, we get

(I # i

4*µ",

µ" + . . .)%'(I +i

4*!&,

!& + . . .) = %' # ,'"%" + . . . (2.119)

Collecting all the terms linear in ,, we find

i

4

6%',*µ"

7,µ" = ,'"%

" (2.120)

Or, what is the same,

[%µ,*"'] = 2i(gµ" %' # gµ'%") (2.121)

This matrix equation is solved by

*"' =i

2[%" , %'] (2.122)

Under a finite Lorentz transformation x! = !x, the 4-spinors transform as

#!(x!) = S (!)# (2.123)

with

S (!) = exp[# i

4*µ",

µ" ] (2.124)

The matrices *µ" are the generators of the group of Lorentz transformationsin the spinor representation. However, while the space components *jk arehermitean matrices, the space-time components *0j are antihermitean. Thisfeature is telling us that the Lorentz group is not a compact unitary group,since in that case all of its generators would be hermitean matrices. Instead,this result tells us that the Lorentz group is isomorphic to the non-compactgroup SO(3, 1). Thus, the representation matrices S(!) are unitary onlyunder space rotations with fixed origin.

The linear operator S (!) gives the field in the transformed frame in termsof the coordinates of the transformed frame. However, we may also wish toask for the transformation U (!) which just compensates the e$ect of the

Page 20: 582-Chapter2 Classical Field Theory

30 Classical Field Theory

coordinate transformation. In other words we seek for a matrix U (!) suchthat

#!(x) = U (!)#(x) = S (!)#(!"1x) (2.125)

For an infinitesimal Lorentz transformation, we seek a matrix U (!) of theform

U (!) = I # i

2Jµ",

µ" + . . . (2.126)

We wish to find an expression for Jµ" . We find-I # i

2Jµ",

µ" + . . .

.# = (I # i

4*µ",

µ" + . . .)# (x$ # ,$"x" + . . .)

,=-I # i

4*µ",

µ" + . . .

.(## &$# ,$"x

" + . . .)

(2.127)

Hence

#!(x) ,=-I # i

4*µ",

µ" + xµ,µ"&" + . . .

.#(x) (2.128)

From this expression we see that Jµ" is given by the operator

Jµ" =1

2*µ" + i(xµ&" # x"&µ) (2.129)

We easily recognize the second term as the orbital angular momentum op-erator (we will come back to this issue shortly). The first term is then in-terpreted as the spin. In fact, let us consider purely spacial rotations, whoseinfinitesimal generator are the space components of Jµ" , i.e.,

Jjk = i(xj&k # xk&j) +1

2*jk (2.130)

We can also define a three component vector J( as the 3-dimensional dualof Jjk

Jjk = -jklJ( (2.131)

Thus, we get (after restoring the factors of !)

J( =i!

2-(jk(xj&k # xk&j) +

!

4-(jk*jk

= i ! -(jk xj&k +!

2

-1

2-(jk*jk

.

J( $ (x% p)( +!

2*(

(2.132)

Page 21: 582-Chapter2 Classical Field Theory

2.5 Field Theory of the Dirac Equation 31

The first term is clearly the orbital angular momentum and the second termcan be regarded as the spin. With this definition, it is straightforward to

check that the spinors

-!+

.which are solutions of the Dirac equation, carry

spin one-half.

2.5.4 Transformation Properties of Field Bilinears in the DiracTheory

We will now consider the transformation properties of a number of physicalobservables of the Dirac theory under Lorentz transformations. Let

x!µ = !µ" x" (2.133)

be a general Lorentz transformation, and S(!) be the induced transforma-tion for the Dirac spinors "a(x) (with a = 1, . . . , 4):

"!a(x

!) = S(!)ab "b(x) (2.134)

Using the properties of the induced Lorentz transformation S(!) and ofthe Dirac %-matrices, is straightforward to verify that the following Diracbilinears obey the following transformation laws:

1.

"!(x! ) "!(x! ) = "(x) "(x) (2.135)

which transforms as a scalar.2. Let let us define the Dirac matrix %5 = i%0%1%2%3. Then the bilinear

"!(x! ) %5 "!(x! ) = det! "(x) %5 "(x) (2.136)

transforms as a pseudo-scalar.3. Likewise

"!(x! ) %µ "!(x! ) = !µ" "(x) %

" "(x) (2.137)

transforms as a vector,4.

"!(x! ) %5%µ "!(x! ) = det! !µ

" "(x) %5%" "(x) (2.138)

transforms as a pseudo-vector,5. and

"!(x! ) *µ" "!(x! ) = !µ! !"& "(x) *

!& "(x) (2.139)

transforms as a tensor

Page 22: 582-Chapter2 Classical Field Theory

32 Classical Field Theory

Above we have denoted by !µ" a Lorentz transformation and det! is its

determinant. We have also used that

S"1(!)%5S(!) = det! %5 (2.140)

Thus, the Dirac algebra provides for a natural basis of the space of 4 % 4matrices, which we will denote by

'S $ I, 'Vµ $ %µ, 'T

µ" $ %µ" , 'Aµ $ %5%µ, 'P = %5 (2.141)

where S, V , T , A and P stand for scalar, vector, tensor, axial vector (orpseudo-vector) and parity respectively. For future reference we will note herethe following useful trace identities obeyed by products of Dirac %-matrices

1.

trI = 4, tr%µ = tr%5 = 0, tr%µ%" = 4gµ" (2.142)

2. If we denote by aµ and bµ two arbitrary 4-vectors, then

/a /b = aµbµ # i*µ" aµb" , and tr /a /b = 4a · b (2.143)

2.5.5 Lagrangian for the Dirac Equation

We now seek a Lagrangian density L for the Dirac theory. It should be alocal di$erentiable Lorentz-invariant functional of the spinor field #. Sincethe Dirac equation is first order in derivatives and it is Lorentz covariant,the Lagrangian should be Lorentz invariant and first order in derivatives. Asimple choice is

L = #(i/& #m)# $ 1

2#i/&

###m## (2.144)

where #/&# $ #(/&#)# (&µ#)%µ#. This choice satisfies all the requirements.The equations of motion are derived in the usual manner, i.e., by demandingthat the action S =

8d4x L be stationary

(S = 0 =

+d4x [

(L(#!

(#! +(L

(&µ#!(&µ#! + (#- #)] (2.145)

The equations of motion are

(L(#!

# &µ(L

(&µ#!= 0

(L(#!

# &µ(L

(&µ#!= 0

(2.146)

Page 23: 582-Chapter2 Classical Field Theory

2.6 Classical Electromagnetism as a field theory 33

By direct substitution we find

(i/& #m)# = 0, #(i.#/& +m) = 0 (2.147)

Here,.#/& indicates that the derivatives are acting on the left.

Finally, we can also write down the Hamiltonian density that follows fromthe Lagrangian of Eq.(2.144). As usual we need to determine the canonicalmomentum conjugate to the field #, i.e.,

&(x) =(L

(&0#(x)= i#(x)%0 $ i#†(x) (2.148)

Thus the Hamiltonian density is

H = &(x)&0#(x)# L = i#%0&0## L= #i# ·!#+m##

= #† (i! ·!+m))9 :; <HDirac

#

(2.149)

Thus we find that the “one-particle” Dirac Hamiltonian HDirac of Eq.(2.73)appears naturally in the field theory as well. Since this Hamiltonian is firstorder in derivatives (i.e., in the “momentum), unlike its Klein-Gordon rel-ative, it is not manifestly positive. Thus there is a question of the stabilityof this theory. We will see below that the proper quantization of this theoryas a quantum field theory of fermions solves this problem. In other words, itwill be necessary to impose the Pauli Principle for this theory to describe astable system with an energy spectrum that is bounded from below. In thisway we will see that there is natural connection between the spin of the fieldand the statistics. This connection is known as the Spin-Statistics Theorem.

2.6 Classical Electromagnetism as a field theory

We now turn to the problem of the electromagnetic field generated by aset of sources. Let .(x) and j(x) represent the charge density and currentat a point x of space-time. Charge conservation requires that a continuityequation has to be obeyed.

&.

&t+! · j = 0 (2.150)

Given an initial condition, i.e., the values of the electric field E(x) and themagnetic field B(x) at some t0 in the past, the time evolution is governed

Page 24: 582-Chapter2 Classical Field Theory

34 Classical Field Theory

by Maxwell’s equations

! ·E =. ! ·B =0 (2.151)

!%B # 1

c

&E

&t=j !%E +

1

c

&B

&t=0 (2.152)

It is possible to recast these statements in a manner in which (a) the rel-ativistic covariance is apparent and (b) the equations follow from a LeastAction Principle. A convenient way to see the above is to consider the elec-tromagnetic field tensor Fµ" which is the (contravariant) antisymmetric realtensor

Fµ" = #F "µ =

"

##$

0 #E1 #E2 #E3

E1 0 #B3 B2

E2 B3 0 #B1

E3 #B2 B1 0

%

&&' (2.153)

In other words

F 0i = #F i0 = #Ei

F ij = #F ji = -ijkBk

(2.154)

where -ijk is the third-rank Levi-Civita tensor:

-ijk =

=>

?

1 if (ijk) is an even permutation of (123)#1 if (ijk) is an odd permutation of (123)0 otherwise

(2.155)

The dual tensor F is defined by

Fµ" = #F "µ =1

2-µ"$%F$% (2.156)

where -µ"$% is the fourth rank Levi-Civita tensor. In particular

Fµ" =

"

##$

0 #B1 #B2 #B3

B1 0 E3 #E2

B2 #E3 0 E2

B3 E2 #E2 0

%

&&' (2.157)

With these notations, we can rewrite Maxwell’s equations in the mani-festly covariant form

&µFµ" =j" (Equation of Motion) (2.158)

&µFµ" =0 (Bianchi Identity) (2.159)

&µjµ =0 (Continuity Equation) (2.160)

Page 25: 582-Chapter2 Classical Field Theory

2.6 Classical Electromagnetism as a field theory 35

By inspection we see that the field tensor Fµ" and the dual field tensor @Fµ"

map into each other by exchanging the electric and magnetic fields with eachothers. This electro-magnetic duality would be an exact property of electro-dynamics if in addition to the electric charge current jµ the Bianchi IdentityEq.(2.159) included a magnetic charge current (of magnetic monopoles).

At this point it is convenient to introduce the vector potential Aµ whosecontravariant components are

Aµ(x) =

-A0

c,A

.$-"

c,A

.(2.161)

where " is the scalar potential, and the current 4-vector jµ(x)

jµ(x) = (.c, j) $ (j0, j) (2.162)

The electric field strength E and the magnetic field B are defined to be

E = #1

c!A0 # 1

c

&A

&tB = !%A

(2.163)

In a more compact, relativistically covariant, notation we write

Fµ" = &µA" # &"Aµ (2.164)

In terms of the vector potential Aµ, Maxwell’s equations have the followingadditional local symmetry

Aµ(x) !" Aµ(x) + &µ"(x) (2.165)

where "(x) is an arbitrary smooth function of the space-time coordinatesxµ. It is easy to check that, under the transformation of Eq.(3.48), the fieldstrength remains unchanged i.e.,

Fµ" !" Fµ" (2.166)

This property is called Gauge Invariance and it plays a fundamental rolein modern physics. By directly substituting the definitions of the magneticfield B and the electric field E in terms of the 4-vector Aµ into Maxwell’sequations, we get the wave equation. Indeed

&µFµ" = j" (2.167)

which yields

&2 A" # &"(&µAµ) = j" (2.168)

Page 26: 582-Chapter2 Classical Field Theory

36 Classical Field Theory

This is the wave equation. We can further use the gauge-invariance to furtherrestrict Aµ (without these restrictions Aµ is not completely determined).These restrictions are known as the procedure of fixing a gauge. The choice

&µAµ = 0 (2.169)

known as the Lorentz gauge, yields the simpler wave equation

&2 Aµ = jµ (2.170)

Notice that the Lorentz gauge preserves Lorentz covariance.Another “popular” choice is the radiation (or Coulomb) gauge

! ·A = 0 (2.171)

which yields (in units with c = 1)

&2 A" # &"(&0A0) = j" (2.172)

which is not Lorentz covariant. In the absence of external sources, j" = 0,we can further make the choice A0 = 0. This choice reduces the set of threeequations, one for each spacial component of A, which satisfy

&2A = 0 ! ·A = 0 (2.173)

The solutions are, as we well know, plane waves of the form

A(x) = A ei(p0x0"p·x) (2.174)

which are only consistent if p20 # p2 = 0 and p ·A = 0. This choice is alsoknown as the transverse gauge.

We can also regard the electromagnetic field as a dynamical system andconstruct a Lagrangian picture for it. Since the Maxwell equations are lo-cal, gauge invariant and Lorentz covariant, we should demand that the La-grangian density should be local, gauge invariant and Lorentz invariant. Asimple choice is

L = #1

4Fµ"F

µ" # jµAµ (2.175)

This Lagrangian density is manifestly Lorentz invariant. Gauge invarianceis satisfied if and only if jµ is a conserved current, i.e. if &µjµ = 0, sinceunder a gauge transformation Aµ !" Aµ + &µ"(x) the field strength doesnot change but the source term does. Hence+

d4x jµAµ !"

+d4x [jµA

µ + jµ&µ"]

=

+d4x jµA

µ +

+d4x &µ(jµ")#

+d4x &µjµ " (2.176)

Page 27: 582-Chapter2 Classical Field Theory

2.7 The Landau Theory of Phase Transitions as a Field Theory 37

If the sources vanish at infinity, lim|x|$% jµ = 0, the surface term can bedropped. Thus the action S =

8d4xL is gauge-invariant if and only if the

current jµ is locally conserved,

&µjµ = 0 (2.177)

We can now derive the equations of motion by demanding that the actionS be stationary, i.e.,

(S =

+d4x

4(L(Aµ

(Aµ +(L

(&µAµ(&µAµ

5= 0 (2.178)

Once again, we can integrate by parts to get

(S =

+d4x &"

4&L

(&µAµ(Aµ

5+

+d4x (Aµ

4(L(Aµ

# &"-

(L(&"Aµ

.5(2.179)

If we demand that at the surface (Aµ = 0, we get

(L(Aµ

= &"-

(L(&"Aµ

.(2.180)

Explicitly, we find

(L(Aµ

= #jµ (2.181)

and(L

(&"Aµ= Fµ" (2.182)

Thus, we obtain

jµ = #&"Fµ" (2.183)

or, equivalently

j" = &µFµ" (2.184)

Therefore, the Least Action Principle implies Maxwell’s equations.

2.7 The Landau Theory of Phase Transitions as a Field Theory

We now turn back to the problem of the statistical mechanics of a mag-net which was introduced in the first lecture. In order to be a little morespecific, we will consider the simplest model of a ferromagnet: the classicalIsing model. In this model, one considers an array of atoms on some lattice(say cubic). Each atom is assumed to have a net spin magnetic moment $S.

Page 28: 582-Chapter2 Classical Field Theory

38 Classical Field Theory

From elementary quantum mechanics we know that the simplest interactionamong the spins is the Heisenberg exchange Hamiltonian

H = #A

<ij>

Jij $S(i) · $S(j) (2.185)

where < ij > are nearest neighboring sites on the lattice. In many situations,

i j

!S(j)!S(i)

Figure 2.3 Spins on a lattice.

in which there is magnetic anisotropy, only the z-component of the spinoperators plays a role. The Hamiltonian now reduces to that of the Isingmodel HI

HI = #JA

<ij>

*(i)*(j) $ E[*] (2.186)

where [*] denotes a configuration of spins with *(i) being the z-projectionof the spin at each site i.

The equilibrium properties of the system are determined by the parti-tion function Z which is the sums over all spin configurations {*} of the

Page 29: 582-Chapter2 Classical Field Theory

2.7 The Landau Theory of Phase Transitions as a Field Theory 39

Boltzmann weight for each state,

Z =A

{%}

exp

-#E[*]

T

.(2.187)

where T is the temperature and {*} is the set of all spin configurations.In the 1950’s, Landau developed a method (or rather a picture) to study

these type of problems which in general, are very di%cult. Landau firstproposed to work not with the microscopic spins but with a set of coarse-grained configurations. One way to do this (using the more modern versiondue to Kadano$ and Wilson) is to partition a large system of linear sizeL into regions or blocks of smaller linear size / such that a0 << / << L,where a0 is the lattice spacing. Each one of these regions will be centeredaround a site, say x. We will denote such a region by A(x). The idea is nowto perform the sum, i.e., the partition function Z, while keeping the totalmagnetization of each region M(x) fixed

M(x) =1

N [A]

A

y&A(x)

*(y) (2.188)

where N [A] is the number of sites in A(x). The restricted partition functionis now a functional of the coarse-grained local magnetizations M(x),

Z[M ] =A

{%}

exp

B#E[*]

T

C D

x

(

"

$M(x)# 1

N(A)

A

y&A(x)

*(y)

%

' (2.189)

The variables M(x) have the property that, for N(A) very large, they ef-fectively take values on the real numbers. Also, the coarse-grained configu-rations {M(x)} are much more smooth than the microscopic configurations{*}.

At very high temperatures the average magnetization /M0 = 0 since thesystem is paramagnetic. On the other hand, if the temperature is low, theaverage magnetization may be non-zero since the system may now be ferro-magnetic. Thus, at high temperatures the partition function Z is dominatedby configurations which have /M0 = 0 while at very low temperatures, themost frequent configurations have /M0 1= 0. Landau proceeded to writedown an approximate form for the partition function in terms of sums oversmooth, continuous, configurations M(x) which can be represented in theform

Z ++

DM(x) exp

4#E [M(x), T ]

T

5(2.190)

Page 30: 582-Chapter2 Classical Field Theory

40 Classical Field Theory

where DM(x) is a measure which means “sum over all configurations.” If forthe relevant configurations M(x) is smooth and small, the energy functionalE[M ] can be written as an expansion in powers of M(x) and of its spacederivatives. With these assumptions the free energy of the magnet can beapproximated by the Landau-Ginzburg form

F (M) $ELG(M)

T

=

+ddx

41

2K(T )|!M(x)|2 + 1

2a(T )M2(x) +

1

4!b(T )M4(x) + . . .

5

(2.191)

Thermodynamic stability requires that the sti$ness K(T ) and the non-linearity coe%cient b(T ) must be positive. The second term has a coe%cienta(T ) with can have either sign. A simple choice of parameters is

K(T ) 2 K0, b(T ) 2 b0, a(T ) 2 a (T # Tc) (2.192)

where Tc is an approximation to the critical temperature.The free energy F (M) defines a Classical, or Euclidean, Field Theory. In

fact, by rescaling the field M(x) in the form

"(x) =3KM(x) (2.193)

we can write the free energy as

F (") =

+ddx

B1

2(!")2 + U(")

C(2.194)

where the potential U(") is

U(") =m2

2"2 +

0

4!"4 + . . . (2.195)

where m2 = a(T )K and 0 = b

K2 . Except for the absence of the term involvingthe canonical momentum &2(x), F (") has a striking resemblance to theHamiltonian of a scalar field in Minkowski space! We will see below thatthis is not an accident.

Let us now ask the following question: is there a configuration "c($x) whichgives the dominant contribution to the partition function Z? If so, we shouldbe able to approximate

Z =

+D" exp{#F (")} + exp{#F ("c)}{1 + · · · } (2.196)

This statement is usually called the Mean Field Approximation. Since theintegrand is an exponential, the dominant configuration "c must be such

Page 31: 582-Chapter2 Classical Field Theory

2.7 The Landau Theory of Phase Transitions as a Field Theory 41

!

U(!)

!0

T > T0

T < T0

!!0

Figure 2.4 The Landau free energy for the order parameter field ": forT > T0 the free energy has a unique minimum at " = 0 while for T < T0

there are two minima at " = ±"0

that F has a (local) minimum at "c. Thus, configurations "c which leaveF (") stationary are good candidates (we actually need local minima!). Theproblem of finding extrema is simply the condition (F = 0. This is the sameproblem we solved for classical field theory in Minkowski space-time. Noticethat in the derivation of F we have invoked essentially the same type ofarguments: (a) invariance and (b) locality (di$erentiability).

The Euler-Lagrange equations can be derived by using the same argu-ments that we employed in the context of a scalar field theory. In the caseat hand they are

# (F

("(x)+ &j

(F

(&j"(x)= 0 (2.197)

For the case of the Landau theory the Euler-Lagrange Equation becomesthe Landau-Ginzburg Equation

0 = #&2"c(x) + m2"c(x) +0

3!"3c(x) (2.198)

The solution "c(x) that minimizes the energy are uniform in space and thushave &j"c = 0. Hence, "c is the solution of the very simple equation

m2"c +0

3!"3c = 0 (2.199)

Since 0 is positive and m2 may have either sign, depending on whetherT > Tc or T < Tc, we have to explore both cases.

Page 32: 582-Chapter2 Classical Field Theory

42 Classical Field Theory

For T > Tc, m2 is also positive and the only real solution is "c = 0. Thisis the paramagnetic state. But, for T < Tc, m2 is negative and two newsolutions are available, namely

"c = ±,

6 | m2 |0

(2.200)

These are the solutions with lowest energy and they are degenerate. Theyboth represent the magnetized state.

We now must ask if this procedure is correct, or rather when can we expectthis approximation to work. It is correct T = 0 and it will also turn out to becorrect at very high temperatures. The answer to this question is the centralproblem of the theory of Critical Phenomena which describes the behaviorof statistical systems in the vicinity of a continuous (or second order) phasetransitions. It turns out that this problem is also connected with a centralproblem of Quantum Field theory, namely when and how is it possible toremove the singular behavior of perturbation theory, and in the processremove all dependence on the short distance (or high energy) cuto$ fromphysical observables. In Quantum Field Theaory this procedure amounts toa definition of the continuum limit. The answer to these questions motivatedthe development of the Renormalization Group which solved both problemssimultaneously.

2.8 Field Theory and Classical Statistical Mechanics.

We are now going to discuss a mathematical “trick” which will allow us toconnect field theory with classical statistical mechanics. Let us go back tothe action for a real scalar field "(x) in D = d+ 1 space-time dimensions

S =

+dDx L(", &µ#) (2.201)

where dDx is

dDx $ dx0ddx (2.202)

Let us formally carry out the analytic continuation of the time componentx0 of xµ from real to imaginary time xD

x0 !" #ixD (2.203)

under which

"(x0,x) !" "(x, xD) $ "(x) (2.204)

Page 33: 582-Chapter2 Classical Field Theory

2.8 Field Theory and Classical Statistical Mechanics. 43

where x = ($x, xD). Under this transformation, the action (or rather i timesthe action) becomes

iS $ i

+dx0 d

dx L(", &0#, &j") !"+

dDx L(",#i&D", &j") (2.205)

If L has the form

L =1

2(&µ")

2 # V (") $ 1

2(&0")

2 # 1

2(!")2 # V (") (2.206)

then the analytic continuation yields

L(",#i&D#,&") = #1

2(&D")

2 # 1

2($&")2 # V (") (2.207)

Then we can write

iS(", &µ") ########"x0 " #ixD

#+

dDx

41

2(&D")

2 +1

2(!")2 + V (")

5(2.208)

This expression has the same form as (minus) the potential energy E(")for a classical field " in D = d + 1 space dimensions. However it is alsothe same as the energy for a classical statistical mechanics problem in thesame number of dimensions i.e., the Landau-Ginzburg free energy of thelast section.

In Classical Statistical Mechanics, the equilibrium properties of a systemare determined by the partition function. For the case of the Landau theoryof phase transitions the partition function is

Z =

+D" e#E(")/T (2.209)

where the symbol “8D"” means sum over all configurations. (We will dis-

cuss the definition of the “measure” D" later on). If we choose for energyfunctional E(") the expression

E(") =

+dDx

41

2(&")2 + V (")

5(2.210)

where

(&")2 $ (&D")2 + (!")2 (2.211)

we see that the partition function Z is formally the analytic continuation of

Z =

+D" eiS(", &µ")/! (2.212)

where we have used ! which has units of action (instead of the temperature).What is the physical meaning of Z? This expression suggests that Z

Page 34: 582-Chapter2 Classical Field Theory

44 Classical Field Theory

should have the interpretation of a sum of all possible functions "($x, t) (i.e.,the histories of the configurations of the field ") weighed by the phase factorexp{ i

!S(", &µ")}. We will discover later on that if T is formally identified

with the Planck constant !, then Z represents the path-integral quantizationof the field theory! Notice that the semiclassical limit ! " 0 is formallyequivalent to the low temperature limit of the statistical mechanical system.

The analytic continuation procedure that we just discussed is called aWick rotation. It amounts to a passage fromD = d+1-dimensional Minkowskispace to a D-dimensional Euclidean space. We will find that this analyticcontinuation is a very powerful tool. As we will see below, a number of dif-ficulties will arise when the theory is defined directly in Minkowski space.Primarily, the problem is the presence of ill-defined integrals which are givenprecise meaning by a deformation of the integration contours from the realtime ( or frequency) axis to the imaginary time (or frequency) axis. Thedeformation of the contour amounts to a definition of the theory in Eu-clidean rather than in Minkowski space. It is an underlying assumption thatthe analytic continuation can be carried out without di%culty. Namely, theassumption is that the result of this procedure is unique and that, what-ever singularities may be present in the complex plane, they do not a$ectthe result. It is important to stress that the success of this procedure isnot guaranteed. However, in almost all the theories that we know of, thisassumption seems to hold. The only case in which problems are known toexist is the theory of Quantum Gravity (that we will not discuss here).