Canonical Quantization - Eduardo Fradkin...

32
4 Canonical Quantization We will now begin the discussion of our main subject of interest: the role of quantum mechanical fluctuations in systems with infinitely many degrees of freedom. We will begin with a brief overview of quantum mechanics of a single particle. 4.1 Elementary Quantum Mechanics Elementary Quantum Mechanics describes the quantum dynamics of sys- tems with a finite number of degrees of freedom. Two axioms are involved in the standard procedure for quantizing a classical system. Let L(q, ˙ q) be the Lagrangian of an abstract dynamical system described by the general- ized coordinate q. In Chapter two, we recalled that the canonical formalism of Classical Mechanics is based on the concept of canonical pairs of dynam- ical variables. So, the canonical coordinate q has for partner the canonical momentum p: p = L ˙ q (4.1) In the canonical formalism, the dynamics of the system is governed by the classical Hamiltonian H (q,p)= p ˙ q L(q, ˙ q) (4.2) which is the Legendre transform of the Lagrangian. In the canonical (Hamil- tonian) formalism the equations of motion are just Hamilton’s Equations, ˙ p = H q ˙ q = H p (4.3) The dynamical state of the system is defined by the values of the canonical

Transcript of Canonical Quantization - Eduardo Fradkin...

4

Canonical Quantization

We will now begin the discussion of our main subject of interest: the roleof quantum mechanical fluctuations in systems with infinitely many degreesof freedom. We will begin with a brief overview of quantum mechanics of asingle particle.

4.1 Elementary Quantum Mechanics

Elementary Quantum Mechanics describes the quantum dynamics of sys-tems with a finite number of degrees of freedom. Two axioms are involvedin the standard procedure for quantizing a classical system. Let L(q, q) bethe Lagrangian of an abstract dynamical system described by the general-ized coordinate q. In Chapter two, we recalled that the canonical formalismof Classical Mechanics is based on the concept of canonical pairs of dynam-ical variables. So, the canonical coordinate q has for partner the canonicalmomentum p:

p =∂L

∂q(4.1)

In the canonical formalism, the dynamics of the system is governed by theclassical Hamiltonian

H(q, p) = pq − L(q, q) (4.2)

which is the Legendre transform of the Lagrangian. In the canonical (Hamil-tonian) formalism the equations of motion are just Hamilton’s Equations,

p = −∂H∂q

q =∂H

∂p(4.3)

The dynamical state of the system is defined by the values of the canonical

4.1 Elementary Quantum Mechanics 91

coordinates and momenta at any given time t. As a result of these definitions,the coordinates and momenta satisfy a set of Poisson Bracket relations

{q, p}PB = 1 {q, q}PB = {p, p}PB = 0 (4.4)

where

{A,B}PB ≡∂A

∂q

∂B

∂p− ∂A

∂p

∂B

∂q(4.5)

In Quantum Mechanics, the primitive (or fundamental) notion is the con-cept of a physical state. A physical state of a system is a represented by avector in an abstract vector space, which is called the Hilbert space H ofquantum states. The spaceH is a vector space in the sense that if two vectors|Ψ⟩ ∈ H and |Φ⟩ ∈ H represent physical states, then the linear superposition|aΨ+ bΦ⟩ = a|Ψ⟩+ b|Φ⟩, where a and b are two arbitrary complex numbers,also represents a physical state and thus it is an element of the Hilbert spacei.e., |aΨ + bΦ⟩ ∈ H. The Superposition Principle is an axiom of QuantumMechanics.

In Quantum Mechanics, the dynamical variables, i.e., the generalizedcoordinates, q, and the associated canonical momenta p, the HamiltonianH, etc. , are represented by operators that act linearly on the Hilbert spaceof states. Hence, Quantum Mechanics is linear, even though the physicalobservables obey non-linear Heisenberg equations of motion. Let us denoteby A an arbitrary operator acting on the Hilbert space H. The result ofacting on the state |Ψ⟩ ∈ H with the operator A is another state |Φ⟩ ∈ H,

A|Ψ⟩ = |Φ⟩ (4.6)

The Hilbert space H is endowed with an inner product. An inner productis an operation that assigns a complex number ⟨Φ|Ψ⟩ to a pair of states|Φ⟩ ∈ H and |Ψ⟩ ∈ H.

Since H is a vector space, there exists a set of linearly independent states{|λ⟩}, called a basis, that spans the entire Hilbert space. Thus, an arbitrarystate |Ψ⟩ can be expanded as a linear combination of a complete set of statesthat form a basis of H,

|Ψ⟩ =!

λ

Ψλ |λ⟩ (4.7)

which is unique for a fixed set of basis states. The basis states can be chosento be orthonormal with respect to the inner product,

⟨λ|µ⟩ = δλµ (4.8)

92 Canonical Quantization

In general, if |Ψ⟩ and |Φ⟩ are normalized states

⟨Ψ|Ψ⟩ = ⟨Φ|Φ⟩ = 1 (4.9)

the action of the operator A on the state |Ψ⟩ is proportional to the state |Φ⟩

A|Ψ⟩ = α|Φ⟩ (4.10)

The coefficient α is a complex number that depends on the pair of statesand on the operator A. This coefficient is the matrix element of A betweenthe state |Ψ⟩ and |Φ⟩, which we write with the notation

α = ⟨Φ|A|Ψ⟩ (4.11)

Operators that act on a Hilbert space do not generally commute witheach other. One of the axioms of Quantum Mechanics is the CorrespondencePrinciple which states that in the classical limit, !→ 0, the operators shouldeffectively become numbers, i. e. they commute in the classical limit.

The procedure of canonical quantization consists in demanding that to theclassical canonical pair (q, p), that satisfies the Poisson Bracket {q, p}PB = 1,we associate a pair of operators q and p, both of which act on the Hilbertspace of states H, and are required to obey the canonical commutation rela-tions

[q, p] = i! [q, q] = [p, p] = 0 (4.12)

Here [A, B] is the commutator of the operators A and B,

[A, B] = AB − BA (4.13)

In particular, two operators that do not commute with each other cannot bediagonalized simultaneously. Hence it is not possible to measure simultane-ously two non-commuting observables with arbitrary precision in the samephysical state. This is the the Uncertainty Principle.

By following this prescription, we assign to the classical HamiltonianH(q, p), which is a function of the dynamical variables q and p, an oper-ator H(q, p) which is obtained by replacing the dynamical variables withthe corresponding operators. Other classical dynamical quantities in Quan-tum Mechanics are similarly associated with quantum operators that act onthe Hilbert space of states. All operators associated with classical physicalquantities in Quantum Mechanics are hermitian operators relative to theinner product defined in the Hilbert space H. Namely, if A is an operatorand A† is the adjoint of A

⟨Ψ|A†|Φ⟩ ≡ ⟨Φ|AΨ⟩∗ (4.14)

4.1 Elementary Quantum Mechanics 93

and A is hermitian iff A = A† (with suitable boundary conditions).The quantum mechanical state of the system at time t, |Ψ(t)⟩, obeys the

Schrodinger Equation

i!∂

∂t|Φ(t)⟩ = H(q, p) |Ψ(t)⟩ (4.15)

The state |Ψ(t)⟩ is uniquely determined by the initial state |Ψ(0)⟩. Thus,in Quantum Mechanics, just as in Classical Mechanics, the Hamiltonian isthe generator of the (infinitesimal) time evolution of teh state of a physicalsystem.

It is always possible to choose a basis in which a particular operatoris diagonal. For instance, if the operator is the canonical coordinate q, apossible set of basis states are labelled by q and are its eigenstates, i.e.,

q|q⟩ = q |q⟩ (4.16)

The bais states {|q⟩} are orthonormal and complete, i. e.

⟨q|q′⟩ = δ(q − q′), I =

"dq |q⟩⟨q| (4.17)

The state vector |Ψ⟩ can be expanded in an arbitrary basis. If the basisof states is {|q⟩}, the expansion is

|Ψ⟩ =!

q

Ψ(q) |q⟩ ≡" +∞

−∞dq Ψ(q) |q⟩ (4.18)

where we used that the eigenvalues of the coordinate q are real numbers.The coefficients Ψ(q) of this expansion

Ψ(q) = ⟨q|Ψ⟩, (4.19)

the amplitude to find the system at coordinate q in this state, are the val-ues of the wave function associated with the state |Ψ⟩ in the coordinaterepresentation.

Since the canonical momentum p does not commute with q, it is not di-agonal in this representation. In fact, just as in Classical Mechanics, themomentum operator p is the generator of infinitesimal displacements. Con-sider the states |q⟩ and exp(− i

!ap) |q⟩. It is easy to prove that the latter is

the state |q + a⟩ since

q exp#−i

!ap$|q⟩ ≡ q

∞!

n=0

1

n!

%−ia!

&n

pn |q⟩ (4.20)

94 Canonical Quantization

Using the commutation relation [q, p] = i! is easy to show that

[q, pn] = i!npn−1 (4.21)

Hence, we can write

q exp

%− i

!ap

&|q⟩ = (q + a) exp

%− i

!ap

&|q⟩ (4.22)

Thus,

exp

%− i

!ap

&|q⟩ = |q + a⟩ (4.23)

We can now use this property to compute the matrix element

⟨q| exp%i

!ap

&|Ψ⟩ ≡ Ψ(q + a) (4.24)

For a infinitesimally small, it can be approximated by

Ψ(q + a) ≈ Ψ(q) +i

!a⟨q|p|Ψ⟩+ . . . (4.25)

We find that the matrix element for p has to satisfy

⟨q|p|Ψ⟩ = !

ilima→0

Ψ(q + a)−Ψ(q)

a(4.26)

Thus, the operator p is represented by a differential operator

⟨q|p|Ψ⟩ ≡ !

i

∂qΨ(q) =

!

i

∂q⟨q|Ψ⟩ (4.27)

It is easy to check that the coordinate representation of the operator

p =!

i

∂q(4.28)

and the coordinate operator q satisfy the commutation relation [q, p] = i!.

4.2 Canonical Quantization in Field Theory

We will now apply the axioms of Quantum Mechanics to a Classical FieldTheory. The result will be a Quantum Field Theory. For the sake of simplic-ity we will consider first the case of a scalar field φ(x). We have seen beforethat, given a Lagrangian density L(φ, ∂µφ), the Hamiltonian can be foundonce the canonical momentum Π(x) is defined,

Π(x) =∂L

δ∂0φ(x)(4.29)

4.2 Canonical Quantization in Field Theory 95

On a given time surface x0, the classical Hamiltonian is

H =

"d3x [Π(x, x0)∂0φ(x, x0)− L(φ, ∂µφ)] (4.30)

We quantize this theory by assigning to each dynamical variable of theclassical theory, a hermitian operator which acts on the Hilbert space ofthe quantum states of the system. Thus, the field φ(x) and the canonicalmomentum 'Π(x) are operators acting on a Hilbert space. These operatorsobey canonical commutation relations

(φ(x), 'Π(y)

)= i!δ(x − y) (4.31)

In the field representation, the Hilbert space is the vector space of wavefunctions Ψ which are functionals of the field configurations {φ(x)},

Ψ[{φ(x)}] ≡ ⟨{φ(x)}|Ψ⟩ (4.32)

In this representation, the field is a diagonal operator

⟨{φ}|φ(x)|Ψ⟩ ≡ φ(x) ⟨{φ(x)}|Ψ⟩ = φ(x) Ψ[{φ}] (4.33)

The canonical momentum 'Π(x) is not diagonal in this representation but itacts like a functional differential operator,

⟨{φ}|'Π(x)|Ψ⟩ ≡ !

i

δ

δφ(x)Ψ[{φ}] (4.34)

What we just described is the Schrodinger Picture of quantum field theory.In this picture, as usual, the operators are time-independent but the statesare time-dependent and satisfy the Schrodinger Equation

i!∂0Ψ[{φ}, x0] = 'HΨ[{φ}, x0] (4.35)

For the particular case of a scalar field φ with the classical Lagrangian L

L =1

2(∂µφ)

2 − V (φ) (4.36)

the quantum mechanical Hamiltonian operator H is

'H =

"d3x

*1

2'Π2(x) +

1

2(▽φ(x))2 + V (φ(x))

+(4.37)

The stationary states are the eigenstates of the Hamiltonian 'H.While it is possible to proceed further with the Schrodinger picture, the

manipulation of wave functionals becomes very cumbersome rather quickly.For this reason an alternative approach has been devised. This is the Heisen-berg Picture.

96 Canonical Quantization

In the Schrodinger Picture the time evolution of the system is encodedin the time dependence of the states. In contrast, in the Heisenberg Picturethe operators are time-dependent while the states are time-independent. Theoperators of the Heisenberg Picture obey quantum mechanical equations ofmotion.

Let A be some operator that acts on the Hilbert space of states. Let usdenote by AH(x0) the Heisenberg operator at time x0, defined by

AH(x0) = ei!Hx0 A e−

i!Hx0 (4.38)

for a system with a time-independent Hamiltonian 'H. It is straightforwardto check that AH(x0) obeys the equation of motion

i!∂0AH(x0) = [AH(x0), 'H ] (4.39)

Notice that in the classical limit, the dynamical variable A(x0) obeys theclassical equation of motion

∂0A(x0) = {A(x0),H}PB (4.40)

where it is assumed that all the time dependence in A comes from the timedependence of the coordinates and momenta.

In the Heisenberg picture both φ(x, x0) and 'Π(x, x0) are time dependentoperators which obey the equations of motion

i!∂0φ(x, x0) =(φ(x, x0), 'H

), i!∂0'Π(x, x0) = ['Π(x, x0), 'H ] (4.41)

The Heisenberg field operators φ(x, x0) and 'Π(x, x0) (we will omit thesubindex “H” from now on) obey equal-time commutation relations

(φ(x, x0), 'Π(y, x0)

)= i!δ(x − y) (4.42)

4.3 Quantized elastic waves in a solid: Phonons

Let us consider the problem of the quantum dynamics of a linear array ofatoms (shown in Fig.4.1). We will see below that this problem is closelyrelated to the problem of quantization of a free scalar field. We will considerthe simple case of a one-dimensional array of atoms, a chain. Each atom hasmass M and their classical equilibrium positions are the regularly spacedlattice sites x0n = na, (n = 1, . . . , n) where a is the lattice spacing. I willassume that we have a system with N atoms and, therefore that the lengthL of the chain is L = Na. To simplify matters, I will assume that thechain is actually a ring and, thus, the N + 1-th atom is the same as the

4.3 Quantized elastic waves in a solid: Phonons 97

a

n − 1 n n + 1

un

Figure 4.1 A model of an elastic one-dimensional solid.

1st atom. The dynamics of this system can be specified in terms of a setof coordinates {un} which represent the position of each atom relative totheir classical equilibrium positions x0n i.e., the actual position xn of thenth atom is xn = x0n + un.

The Lagrangian of the system is a function of the coordinates {un}, thedisplacements, and of their time derivatives {un}. In general the LagrangianL it will be the difference of the kinetic energy of the atoms minus theirpotential energy, i.e.,

L({un}, {un}) =

N2!

n=−N2 +1

M

2

%dundt

&2

− V ({un}) (4.43)

We will be interested in the study of the small oscillations of the system.Thus, the potential will have a minimum at the classical equilibrium posi-tions {un = 0} (which we will assume to be unique). For small oscillationsV ({un}) can be expanded in powers

V ({un}) =

N2!

n=−N2 +1

(D2(un+1 − un)

2 +K

2u2n + . . .

)(4.44)

Here D is an elastic constant (i.e., spring constant!) which represents therestoring forces that keep the crystal together. The constant K is a measureof the strength of an external potential which favors the placement of theatoms at their classical equilibrium positions. For an isolated system, K =0 but D = 0. This must be the case since an isolated system must betranslationally invariant and, therefore, V must not change under a constant,uniform, displacement of all the atoms by some amount a, un → un + a.The term proportional to u2n breaks this symmetry of uniform continuousdisplacements, although it does not spoil the discrete symmetry n→ n+m.

Let us now proceed to study the quantum mechanics of this system. Eachatom has a coordinate un(t) and a canonical momentum pn(t) which is

98 Canonical Quantization

defined in the usual way

pn(t) =∂L

∂un(T )= Mun(t) (4.45)

The quantum Hamiltonian for a chain with an even number of sites N is

H({un}, {pn}) =

N2!

n=−N2 +1

,p2n2M

+D

2(un+1 − un)

2 +K

2u2n + · · ·

-(4.46)

where the coordinates and momenta obey the commutation relations

[un, pm] = i!δn,m, [un, um] = [pn, pm] = 0. (4.47)

For this simple system, the Hilbert space can be identified with the tensorproduct of the Hilbert spaces of each individual atom. Thus, if |Ψ⟩n denotesan arbitrary state in the Hilbert space of the nth atom, the states of thechain |Ψ⟩ can be written in the form

|Ψ⟩ = |Ψ⟩1 ⊗ · · ·⊗ |Ψ⟩n ⊗ · · ·⊗ |Ψ⟩N ≡ |Ψ1, . . . ,ΨN ⟩ (4.48)

For instance, a set of basis states can be constructed by using the coordinaterepresentation. Thus, if the state |un⟩ is an eigenstate of un with eigenvalueun

un|un⟩ = un|un⟩ (4.49)

we can write a set of basis states

|u1, . . . , uN ⟩ (4.50)

of the form

|u1, . . . , uN ⟩ = |u1⟩ ⊗ · · ·⊗ |uN ⟩ (4.51)

In this basis, the wave functions are

Ψ(u1, . . . , uN ) = ⟨u1, . . . , uN |Ψ⟩ (4.52)

By inspecting the Hamiltonian it is easy to recognize that in this case it rep-resents a set of N coupled harmonic oscillators. Since the system is periodicand invariant under lattice shifts n → n + m (m integer), it is natural toexpand the coordinates un in a Fourier series

un =1

N

!

k

uk eikn (4.53)

where k is a label, that will be related with the lattice momentum, and uk

4.3 Quantized elastic waves in a solid: Phonons 99

are the Fourier components of un. The fact that we have imposed periodicboundary conditions (PBC’s), i.e., that the chain is a ring, means that

un = un+N (4.54)

This relation can hold only if the labels k satisfy

eikN = 1 (4.55)

which restricts the values of k to the discrete set

km = 2πm

N, m = −N

2+ 1, . . . ,

N

2(4.56)

where we have set the lattice spacing to unity, a0 = 1. Thus the expansionof un is

un =1

N

N2!

m=−N2 +1

ukm eikmn (4.57)

The spacing∆k between two consecutive values of the label k (km and km+1)is

∆k = km+1 − km =2π

N(4.58)

which vanishes in the thermodynamic limit, N → ∞. In particular, themomentum label kn runs over the range (−N

2 + 1)2πN ≤ km ≤ N2 . Thus, in

the limit N →∞ the momenta fill up densely the real interval (−π,π].We then conclude that, in the thermodynamic limit, N → ∞, the mo-

mentum sum converges to the integral

un = limN→∞

1

N

N2!

m=−N2 +1

ukm eikmn =

" π

−π

dk

2πu(k) eikn (4.59)

Since un is a real hermitian operator, the Fourier components u(k) mustsatisfy

u†(k) = u(−k) (4.60)

The Fourier component u(k) can be written as a linear combination of op-erators un of the form

u(k) =

N2!

n=−N2 +1

un e−ikn (4.61)

100 Canonical Quantization

where we used definition of the periodic Dirac delta function,

2πδP (k − q) ≡+∞!

m=−∞

2πδ(k − q + 2πm) =

N2!

n=−N2 +1

ei(k−q)n (4.62)

which is defined in the thermodynamic limit.The momentum operators pn can also be expanded in Fourier series. Their

expansions are

pn =

" π

−π

dk

2πp(k) eikn, p(k) =

N2!

n=− N2+1

e−ikn pn (4.63)

and satisfy

p†(k) = p(−k) (4.64)

The transformation (un, pn) → (u(k), p(k)) is a canonical transformation.Indeed, the Fourier amplitudes u(k) and p(k) obey the commutation rela-tions

[u(k), p(k′)] =

N2!

n,n′=−N2 +1

e−i(kn+k′n′) [un, pn′ ]

= i!

N2!

n=−N2 +1

e−i(k+k′)n

(4.65)

Hence, in the thermodynamic limit we find

[u(k), p(k′)] = i! 2π δP (k + k′) (4.66)

and

[u(k), p(k′)] = [p(k), p(k′)] = 0 (4.67)

We can now write the Hamiltonian H in terms of the Fourier componentsu(k) and p(k). We find

H =

" π

−π

dk

2π[1

2Mp†(k)p(k) +

M

2ω2(k)u†(k)u(k)] (4.68)

where ω2(k) is

ω2(k) =K

M+

4D

Msin2

%k

2

&(4.69)

4.3 Quantized elastic waves in a solid: Phonons 101

Thus, the system decouples into its normal modes. The frequency ω(k) isshown in Fig.4.2. It is instructive to study the long-wave length limit, k → 0.

k π

ω(k)

K = 0

K = 0!

K

M

Figure 4.2 The dispersion relation ω(k).

For K = 0 i.e., if there is no external potential, ω(k) vanishes linearly ask → 0,

ω(k) ≈.

D

M|k| (K = 0) (4.70)

However, for non-zero K, again in the limit k → 0, we find instead

ω(k) ≈.

K

M+

D

Mk2 (4.71)

If we now restore a lattice constant a0 = 1, k = ka0 we can write ω(k) inthe form

ω(k) = vs

/m2v2s + k2 (4.72)

where vs is the speed of propagation of sound in the chain,

vs =

.D

Ma0 ≡

0Da0ρ

(4.73)

where ρ is the (linear) density. The “mass” m is

m =ω0

v2s(4.74)

where ω0 =/

KM . Thus, the waves that propagate on this chain behave

as “relativistic” particles with mass m and a “speed of light” vs. Indeed,

102 Canonical Quantization

the long wavelength limit, k → 0, the Lagrangian of the discrete chain,Eq.(4.43), can be written as an integral

L = a0!

n

1

2

%M

a0

&%∂un∂t

&2

− a02

!

n

Da0

%un+1 − un

a0

&2

− a02

!

n

K

a0u2n

(4.75)Thus, as a0 → 0 and N → ∞, with fixed total length ℓ = Na0, the sumsconverge to the integral

L = ρ

" ℓ/2

− ℓ2

dx

11

2

%∂u

∂t

&2

− 1

2v2s

%∂u

∂x

&2

− 1

2m2v4su

2(x)

2

(4.76)

Aside from the overall factor of ρ, the mass density, we see that the La-grangian for the linear chain is, in the long wavelength limit (or continuumlimit) the same as the Lagrangian for the Klein-Gordon field u(x) in one-space dimension. The last term of this Lagrangian is precisely the mass termfor the Lagrangian of the Klein-Gordon field. This explains the choice of thesymbol m. Indeed, upon the change of variables

x0 = vst x1 = x (4.77)

and by defining the rescaled field

ϕ =√ρvsu (4.78)

we see immediately that the Lagrangian density L is

L =1

2(∂0ϕ)

2 − 1

2(∂1ϕ)

2 − 1

2m2v4sϕ

2 (4.79)

This is the Lagrangian density of a free massive scalar field in 1 + 1 space-time dimension.

Returning to the quantum theory, we seek to find the eigenstates of thenormal-mode Hamiltonian. Let a†(k) and a(k) be the operators defined by

a†(k) =13

2M!ω(k)

#Mω(k)u†(k)− ip†(k)

$

a(k) =13

2M!ω(k)(Mω(k)u(k) + ip(k)) (4.80)

These operators satisfy the commutation relations

4a(k), a(k′)

5=(a†(k), a†(k′)

)= 0

(a(k), a†(k′)

)=2πδP (k + k′) (4.81)

4.3 Quantized elastic waves in a solid: Phonons 103

Up to normalization constants, the operators a†(k) and a(k) obey the algebraof creation and annihilation operators.

In terms of the creation and annihilation operators, the momentum spaceoscillator operators u(k) and p(k) are

u(k) =

0!

2Mω(k)

#a(k) + a†(−k)

$

p(k) =3

2M!ω(k)1

2i

#a(k)− a†(−k)

$(4.82)

Thus, the coordinate space operators un and pn have the Fourier expansions

un =

" π

−π

dk

0!

2Mω(k)

#a(k) eikn + a†(k) e−ikn

$

pn =

" π

−π

dk

32M!ω(k)

1

2i

#a(k) eikn − a†(k) e−ikn

$(4.83)

The normal-mode Hamiltonian has a very simple form in terms of the cre-ation and annihilation operators

H =

" π

−π

dk

!ω(k)

2

#a†(k)a(k) + a(k)a†(k)

$(4.84)

It is customary to write H in such a way that the creation operators alwaysappear to the left of annihilation operators. This procedure is called normalordering. Given an arbitrary operator A, we will denote by : A : the normalordered operator. We can see by inspection that H can be written as a sumof two terms: a normal ordered operator : H : and a complex number. Thecomplex number results from using the commutation relations. Indeed, byoperating on the last term of the Hamiltonian of Eq.(4.84), we get

a(k)a†(k) = [a(k), a†(k)] + a†(k)a(k) (4.85)

The commutator [a(k), a†(k)] is the divergent quantity

[a(k), a†(k)] = limk′→k

2πδP (k − k′) = lim

k′→k2π

+∞!

m=∞

δ(k − k′+ 2πm)

= limk′→k

N2!

n=−N2 +1

ei(k−k′)n = N

(4.86)

which diverges in the thermodynamic limit, N →∞.

104 Canonical Quantization

Using these results, we can write H in the form

H =: H : +E0 (4.87)

where : H : is the normal-ordered Hamiltonian

: H :=

" π

−π

dk

2π!ω(k) a†(k)a(k) (4.88)

and the real number E0 is given by

E0 = N

" π

−π

dk

!ω(k)

2(4.89)

We will see below that E0 is the ground state energy of this system. Thelinear divergence of E0 as N →∞ is natural since the ground state energymust be an extensive quantity, i.e., it scales like the length (volume) ofthe chain (system). Notice that the integral in Eq.(4.89) runs over the finitemomentum interval (−π,π), i.e. the first Brillouin zone. In the next subsec-tion we will encounter a similar result for the free relativistic scalar field butwith an integral running over an unbounded momentum range, which willlead to an ultraviolet (UV) divergence of the ground state energy (amongother quantities). Thus, in this simple chain system we have a cutoff at largemomentum imposed by the discrete microscopic lattice. In this case thereis no UV divergence but instead quantities such as the ground state energyhave a high sensitivity to the value of the UV cutoff, the momentum scaleπa0, where a0 is the lattice spacing.We are now ready to construct the spectrum of eigenstates of this system.

4.3.1 Ground state

Let |0⟩ be the state which is annihilated by all the operators a(k),

a(k)|0⟩ = 0 (4.90)

This state is an eigenstate with eigenvalue E0 since

H|0⟩ =: H : |0⟩+ E0|0⟩ = E0|0⟩ (4.91)

where we have used the fact that |0⟩ is annihilated by the normal-orderedHamiltonian : H :. This is the ground state of the system since the energy ofall other states is higher. Thus, E0 is the energy of the ground state. Noticethat E0 is the sum of the zero-point energy of all the oscillators.

The wave function for the ground state can be constructed quite easily.Let Ψ0 ({u(k)}) = ⟨{u(k)}|0⟩ be the wave function of the ground state. The

4.3 Quantized elastic waves in a solid: Phonons 105

condition that |0⟩ be annihilated by all the operators a(k) means that thematrix element ⟨{u(k)}|a(k)|0⟩ has to vanish identically. The definition ofa(k) yields the condition

0 = Mω(k) ⟨{u(k)}|u(k)|0⟩ + i⟨{u(k)}|p(k)|0⟩ (4.92)

The commutation relation

[u(k), p(k′)] = i! 2π δP (k + k

′) (4.93)

implies that, in the coordinate representation p(k) must bet the functionaldifferential operator

⟨{u(k)|p(k)|0⟩ = !

i2π

δ

δu(k)Ψ0 ({u(k)}) (4.94)

Thus, the wave functional Ψ0 must obey the differential equation

Mω(k) u∗(k) Ψ0 ({u(k)}) + 2π!δ

δu(k)Ψ0 ({u(k)}) = 0 (4.95)

for each value of k ∈ [−π,π). Clearly Ψ0 has the form of a product

Ψ0 ({u(k)}) =6

k

Ψ0,k (u(k)) (4.96)

where the wave function Ψ0,k (u(k)) satisfies

Mω(k)u∗(k)Ψ0,k (u(k)) + 2π!∂

∂u(k)Ψ0,k (u(k)) = 0 (4.97)

The solution of this equation is the ground state wave function for the k-thoscillator

Ψ0,k (u(k)) = N (k) exp

%−1

2

%Mω(k)

2π!

&|u|2(k)

&(4.98)

whereN (k) is a normalization factor. The total wave function for the groundstate is

Ψ0({u(k)}) = N exp(− 1

2

" π

−π

dk

%Mω(k)

2π!

&|u|2(k)

)(4.99)

where N is another normalization constant. Notice that this wave functionis a functional of the oscillator variables {u(k)}.

106 Canonical Quantization

4.3.2 One-Particle States

Let the ket |1k⟩ denote the one-particle state

|1k⟩ ≡ a†(k)|0⟩ (4.100)

This state is an eigenstate of H with eigenvalue E1(k)

H|1k⟩ =: H : |1k⟩+ E0|1k⟩ (4.101)

The normal-ordered term now does give a contribution since

: H : |1k⟩ =%" π

−π

dk′

2π!ω(k′) a†(k′)a(k′)

&a†(k)|0⟩

=

" π

−π

dk′

2π!ω(k

′)#a†(k′)[a(k′), a†(k)]|0⟩ + a†(k′)a†(k)a(k′)|0⟩

$

(4.102)

The result is

: H : |1k⟩ =" π

−π

dk′

2π!ω(k′) a†(k′) 2πδP (k

′ − k) |0⟩ (4.103)

since the last term vanishes. Hence

: H : |1k⟩ = !ω(k) a†(k) |0⟩ ≡ !ω(k) |1k⟩ (4.104)

Therefore, we find

H|1k⟩ = (!ω(k) + E0) |1k⟩ (4.105)

Let ε(k) be the excitation energy

ε(k) ≡ E1(k)− E0(k) = !ω(k) (4.106)

Thus the one-particle states represent quanta with energy !ω(k) above thatof the ground state.

4.3.3 Many Particle States

If we define the occupation number operator n(k) by

n(k) = a†(k)a(k) (4.107)

i.e., the quantum number of the k-th oscillator, we see that the most generaleigenstate is labelled by the set of oscillator quantum numbers {n(k)}. Thus,the state |{n(k)}⟩ defined by

|{n(k)}⟩ =6

k

4a†(k)

5n(k)3

n(k)!|0⟩ (4.108)

4.4 Quantization of the Free Scalar Field Theory 107

has energy E[{n(k)}]

E[{n(k)}] =" π

−π

dk

2πn(k)!ω(k) + E0 (4.109)

It is clear that the excitations behave as free particles since their energies areadditive. These excitations are known as phonons. They are the quantizedfluctuations of the array of atoms.

4.4 Quantization of the Free Scalar Field Theory

We now return to the problem of quantizing a relativistic scalar field φ(x).In particular, we will consider a free real scalar field φ whose Lagrangiandensity is

L =1

2(∂µφ)(∂

µφ)− 1

2m2φ2 (4.110)

This system can be studied using methods which are almost identical to theones we used in our discussion of the chain of atoms of the previous section.

The quantum mechanical Hamiltonian 'H for a free real scalar field is

'H =

"d3x

,1

2'Π2(x) +

1

2

#▽φ(x)

$2+

1

2m2φ2(x)

-(4.111)

where φ and 'Π satisfy the equal-time commutation relations (in units with! = c = 1)

[φ(x, x0), 'Π(y, x0)] = iδ(x− y) (4.112)

In the Heisenberg representation, φ and 'Π are time dependent operatorswhile the states are time independent. The field operators obey the equationsof motion

i∂0φ(x, x0) =[φ(x, x0), 'H ]

i∂0'Π(x, x0) =['Π(x, x0), 'H ] (4.113)

These are operator equations. After some algebra, we find the Heisenbergequations of motion for the field and for the canonical moemntum

∂0φ(x, x0) ='Π(x, x0) (4.114)

∂0'Π(x, x0) =▽2 φ(x, x0)−m2φ(x, x0) (4.115)

Upon substitution we derive the field equation for the scalar field operator7∂2 +m2

8φ(x) = 0 (4.116)

Thus, the field operators φ(x) satisfy the Klein-Gordon equation.

108 Canonical Quantization

Field Expansion

Let us solve the field equation of motion by a Fourier transform

φ(x) =

"d3k

(2π)3φ(k, x0) e

ik·x (4.117)

where φ(k, x0) are the Fourier amplitudes of φ(x). We now demand thatthe φ(x) satisfies the Klein-Gordon equation, and find that φ(k, x0) shouldsatisfy

∂20 φ(k, x0) + (k2 +m2)φ(k, x0) = 0 (4.118)

Also, since φ(x, x0) is a real hermitian field operator, φ(k, x0) must satisfy

φ†(k, x0) = φ(−k, x0) (4.119)

The time dependence of φ(k, x0) is trivial. Let us write φ(k, x0) as the sumof two terms

φ(k, x0) = φ+(k)eiω(k)x0 + φ−(k)e

−iω(k)x0 (4.120)

The operators φ+(k) and φ†+(k) are not independent since the reality con-

dition implies that

φ+(k) = φ†−(−k) φ†+(k) = φ−(−k) (4.121)

This expansion is a solution of the equation of motion, the Klein-Gordonequation, provided ω(k) is given by

ω(k) =3

k2 +m2 (4.122)

Let us define the operators a(k) and its adjoint a†(k) by

a(k) = 2ω(k)φ−(k) a†(k) = 2ω(k)φ†−(k) (4.123)

The operators a†(k) and a(k) obey the (generalized) creation-annihilationoperator algebra

[a(k), a†(k′)] = (2π)32ω(k) δ3(k − k′) (4.124)

In terms of the operators a†(k) and a(k) field operator is

φ(x) =

"d3k

(2π)32ω(k)

(a(k)e−iω(k)x0+ik·x + a†(k)eiω(k)x0−ik·x

)(4.125)

where we have chosen to normalize the operators in such a way that the

phase space factor takes the Lorentz invariant formd3k

2ω(k).

4.4 Quantization of the Free Scalar Field Theory 109

The canonical momentum also can be expanded in a similar way

'Π(x) = −i"

d3k

(2π)32ω(k)ω(k)

(a(k)e−iω(k)x0+ik·x − a†(k)eiω(k)eiω(k)x0−ik·x

)

(4.126)Notice that, in both expansions, there are terms with positive and negativefrequency and that the terms with positive frequency have creation operatorsa†(k) while the terms with negative frequency have annihilation operatorsa(k). This observation motivates the notation

φ(x) = φ+(x) + φ−(x) (4.127)

where φ+ are the positive frequency terms and φ− are the negative frequencyterms. This decomposition will turn out to be very useful.

We will now follow the same approach that we used for the problem ofthe linear chain and write the Hamiltonian in terms of the operators a(k)and a†(k). The result is

'H =

"d3k

(2π)32ω(k)

ω(k)

2

#a(k)a†(k) + a†(k)a(k)

$(4.128)

This Hamiltonian needs to be normal-ordered relative to a ground statewhich we will now define.

Ground State

Let |0⟩ be the state that is annihilated by all the operators a(k),

a(k)|0⟩ = 0 (4.129)

Relative to this state, which we will call the vacuum state, the Hamiltoniancan be written on the form

'H =: 'H : +E0 (4.130)

where : 'H : is normal ordered relative to the state |0⟩. In other words, in : 'H :all the destruction operators appear the right of all the creation operators.Therefore : 'H : annihilates the vacuum state

: 'H : |0⟩ = 0 (4.131)

The real number E0 is the ground state energy. In this case it is equal to

E0 =

"d3k

ω(k)

2δ(0) (4.132)

110 Canonical Quantization

when δ(0) is the infrared divergent number

δ(0) = limp→0

δ3(p) = limp→0

"d3x

(2π)3eip·x =

V

(2π)3(4.133)

where V is the (infinite) volume of space. Thus, E0 is extensive and can bewritten as E0 = ε0V , where ε0 is the ground state energy density. We find

ε0 =

"d3k

(2π)3ω(k)

2≡ 1

2

"d3k

(2π)3

3k2 +m2 (4.134)

Eq. (4.134) is the sum of the zero-point energies of all the oscillators.This quantity is formally divergent since the integral is dominated by thecontributions with large momentum or, what is the same, short distances.This is an ultraviolet divergence. It is divergent because the system hasan infinite number of degrees of freedom even if the volume is finite. Wewill encounter other examples of similar divergencies in field theory. It isimportant to keep in mind that they are not artifacts of our scheme butthat they result from the fact that the system is in continuous space-timeand thus it is infinitely large.

It is interesting to compare this issue in the phonon problem with thescalar field theory. In both cases the ground state energy was found to beextensive. Thus, the infrared divergence in E0 was expected in both cases.However,the ultraviolet divergence that we found in the scalar field theoryis absent in the phonon problem. Indeed, the ground state energy density ε0for the linear chain with lattice spacing a is

ε0 =

" πa

−πa

dk

!ω(k)

2=

" πa

−πa

dk

1

2

.!2K

M+

4D!2

Msin2(

ka

2) (4.135)

Thus integral is finite because the momentum integration is limited to therange |k| ≤ π

a . Thus the largest momentum in the chain is πa and it is finite

provided that the lattice spacing is not equal to zero. In other words, theintegral is cutoff by the lattice spacing. However, the scalar field theory thatwe are considering does not have a cutoff and hence the energy density blowsup.

We can take two different points of view with respect to this problem. Onepossibility is simply to say that the ground state energy is not a physicallyobservable quantity since any experiment will only yield information onexcitation energies and in this theory, they are finite. Thus, we may simplyredefine the zero of the energy by dropping this term off. Normal ordering isthen just the mathematical statement that all energies are measured relativeto that of the ground state. As far as free field theory is concerned, this

4.4 Quantization of the Free Scalar Field Theory 111

subtraction is sufficient since it makes the theory finite without affecting anyphysically observable quantity. However, once interactions are considered,divergencies will show up in the formal computation of physical quantities.This procedure then requires further subtractions. An alternative approachconsists in introducing a regulator or cutoff. The theory is now finite butone is left with the task of proving that the physics is independent of thecutoff. This is the program of the Renormalization group. Although it is notpresently known if there should be a fundamental cutoff in these theories,i.e., if there is a more fundamental description of Nature at short distancesand high energies such as it is postulated by String Theory, it is clear that ifquantum field theories are to be regarded as effective hydrodynamic theoriesvalid below some high energy scale, then a cutoff is actually natural.

Hilbert Space

We can construct the spectrum of states by inspection of the normal orderedHamiltonian

: 'H :=

"d3k

(2π)32ω(k)ω(k) a†(k)a(k) (4.136)

This Hamiltonian commutes with the total momentum 'P

'P =

"

x0 fixedd3x 'Π(x, x0)▽φ(x, x0) (4.137)

which, up to operator ordering ambiguities, is the quantum mechanical ver-sion of the classical linear momentum P ,

P =

"

x0

d3x T 0j ≡"

x0

d3x Π(x, x0)▽φ(x, x0) (4.138)

In Fourier space 'P becomes

'P =

"d3k

(2π)32ω(k)k a†(k)a(k) (4.139)

The normal-ordered Hamiltonian : H : also commutes with the oscillatoroccupation number n(k), defined by

n(k) ≡ a†(k)a(k) (4.140)

Since {n(k)} and the Hamiltonian H commute with each other, we can usea complete set of eigenstates of {n(k)} to span the Hilbert space. We willregard the excitations counted by n(k) as particles that have energy andmomentum (and in other theories will also have other quantum numbers).

112 Canonical Quantization

Their Hilbert space has an indefinite number of particles and it is calledFock space. The states {|{n(k)}⟩} of the Fock space, defined by

|{n(k}⟩ =6

k

N (k)[a†(k)]n(k)|0⟩, (4.141)

with N (k) being normalization constants, are eigenstates of the operatorn(k)

n(k)|{n(k)}⟩ = (2π)32ω(k)n(k)|{n(k)⟩ (4.142)

These states span the occupation number basis of the Fock space.The total number operator 'N

'N ≡"

d3k

(2π)32ω(k)n(k) (4.143)

commutes with the Hamiltonian 'H and it is diagonal in this basis i.e.,

'N |{n(k)}⟩ ="

d3k n(k) |{n(k)}⟩ (4.144)

The energy of these states is

'H|{n(k)}⟩ =,"

d3k n(k)ω(k) +E0

-|{n(k)}⟩ (4.145)

Thus, the excitation energy ε({n(k)}) of this state is

ε(k) =

"d3k n(k)ω(k). (4.146)

The total linear momentum operator 'P has an operator ordering ambigu-ity. It will be fixed by requiring that the vacuum state |0⟩ be translationallyinvariant, i.e.,

'P |0⟩ = 0. (4.147)

In terms of creation and annihilation operators the linear momentum oper-ator is

'P =

"d3k

(2π)32ω(k)k n(k) (4.148)

Thus, 'P is diagonal in the basis |{n(k)}⟩ since

'P |{n(k)}⟩ =,"

d3k k n(k)

-|{n(k)}⟩ (4.149)

The state with lowest energy, the vacuum state |0⟩ has n(k) = 0, for all

4.4 Quantization of the Free Scalar Field Theory 113

k. Thus the vacuum state has zero momentum and it is translationallyinvariant.

The states |k⟩, defined by

|k⟩ ≡ a†(k)|0⟩ (4.150)

have excitation energy ω(k) and total linear momentum k. Thus, the states{|k⟩} are particle-like excitations which have an energy dispersion curve

E =3

k2 +m2, (4.151)

characteristic of a relativistic particle of momentum k and mass m. Thus,the excitations of the ground state of this field theory are particle-like. Fromour discussion we can see that these particles are free since their energiesand momenta are additive.

Causality

The starting point of the quantization procedure was to impose equal-timecommutation relations among the canonical fields φ(x) and momenta 'Π(x).In particular two field operators on different spatial locations commute atequal times. But, do they commute at different times?

To address this question let us calculate the commutator ∆(x− y)

i∆(x− y) = [φ(x), φ(y)] (4.152)

where φ(x) and φ(y) are Heisenberg field operators for space-time points xand y respectively. From the Fourier expansion of the fields we know thatthe field operator can be split into a sum of two terms

φ(x) = φ+(x) + φ−(x) (4.153)

where φ+ contains only creation operators and positive frequencies, andφ− contains only annihilation operators and negative frequencies. Thus thecommutator is

i∆(x− y) =[φ+(x), φ+(y)] + [φ−(x), φ−(y)]

+[φ+(x), φ−(y)] + [φ−(x), φ+(y)] (4.154)

The first two terms always vanish since the φ+ operators commute among

114 Canonical Quantization

themselves and so do the operators φ−. Thus, we get

i∆(x− y) =[φ+(x), φ−(y)] + [φ−(x), φ+(y)]

=

"dk

"dk′9[a†(k), a(k′)] exp

7−iω(k)x0 + ik · x+ iω(k′)y0 − ik′ · y

8

+[a(k), a†(k′)] exp7iω(k)x0 − ik · x− iω(k′), y0 + ik′ · y

8:

(4.155)

where "dk ≡

"d3k

(2π)32ω(k)(4.156)

By using the commutation relations, we find that the operator ∆(x− y) isproportional to the identity operator and hence it is actually a function. Itis given by

i∆(x− y) =

"dk(eiω(k)(x0−y0)−ik·(x−y) − e−iω(k)(x0−y0)+ik·(x−y)

)(4.157)

With the help of the Lorentz-invariant function ϵ(k0), defined by

ϵ(k0) =k0

|k0|≡ sign(k0) (4.158)

we can write ∆(x− y) in the manifestly Lorentz invariant form

i∆(x− y) =

"d4k

(2π)3δ(k2 −m2)ϵ(k0)e−ik·(x−y) (4.159)

The integrand vanishes unless the mass shell condition

k2 −m2 = 0 (4.160)

is satisfied. Notice that ∆(x− y) satisfies the initial condition

∂0∆|x0=y0 = −δ3(x− y) (4.161)

At equal times x0 = y0 the commutator vanishes,

∆(x− y, 0) = 0 (4.162)

Furthermore, it vanishes if the space-time points x and y are separated bya space-like interval, (x − y)2 < 0. This must be the case since ∆(x − y)is manifestly Lorentz invariant. Thus if it vanishes at equal times, where(x − y)2 = (x0 − y0)2 − (x − y)2 = −(x − y2)2 < 0, it must vanish for allevents with the negative values of (x − y)2. This implies that, for events xand y, which are not causally connected ∆(x− y) = 0 and that ∆(x− y) is

4.5 Symmetries of the Quantum Theory 115

non-zero only for causally connected events, i.e., in the forward light-cone(for details see Fig.4.3).

space

time

forward light cone

backward light cone

x2

0 − x2

> 0

x2

0 − x2

> 0

x2

0 − x2

< 0x2

0 − x2

< 0

Figure 4.3 The Minkowski space-time.

4.5 Symmetries of the Quantum Theory

In our discussion of Classical Field Theory we discovered that the presence ofcontinuous global symmetries implied the existence of constants of motion.In addition, the constants of motion were the generators of infinitesimalsymmetry transformations. It is then natural to ask what role do symmetriesplay in the quantized theory.

In the quantized theory all physical quantities are represented by opera-tors that act on the Hilbert space of states. The classical statement that aquantity A is conserved if its Poisson Bracket with the Hamiltonian vanishes

dA

dt= {A,H}PB = 0 (4.163)

in the quantum theory becomes the operator identity

idAH

dt= [AH , 'H] = 0 (4.164)

116 Canonical Quantization

we are are using the Heisenberg representation. Then, the constants of mo-tion of the quantum theory are operators that commute with the Hamilto-nian, [AH , 'H ] = 0.

Therefore, the quantum theory has a symmetry if and only if the charge'Q, which is a hermitian operator associated with a classically conservedcurrent jµ(x) via the correspondence principle,

'Q =

"

x0 fixedd3x j0(x, x0) (4.165)

is an operator that commutes with the Hamiltonian 'H

[ 'Q, 'H] = 0 (4.166)

If this is so, the charges Q constitute a representation of the generators ofthe algebra of the Lie group of the symmetry transformations in the Hilbertspace of the theory. The transformations 'U(α) associated with the symmetry

'U(α) = exp(iα 'Q) (4.167)

are unitary transformations that act on the Hilbert space of the system.For instance, we saw that for a translationally invariant system the clas-

sical energy-momentum four-vector Pµ

Pµ =

"

x0

d3x T 0µ (4.168)

is conserved. In the quantum theory P 0 becomes the Hamiltonian operator'H and 'P is the total momentum operator. In the case of a free scalar fieldwe saw before that these operators commute with each other, [ 'P , 'H] = 0.Thus, the eigenstates of the system have well defined total energy and totalmomentum. Since P is the generator of infinitesimal translations of theclassical theory, it is easy to check that its equal-time Poisson Bracket withthe field φ(x) is

{φ(x, x0), P j}PB = ∂jxφ (4.169)

In the quantum theory the equivalent statement is that the field operatorφ(x) and the linear momentum operator 'P satisfy the equal-time commu-tation relation

[φ(x, x0), 'P j ] = i∂jxφ(x, x0) (4.170)

Consequently, the field operators φ(x+ a, x0) and φ(x, x0) are related by

φ(x+ a, x0) = eia·!P φ(x, x0)e

−ia· !P (4.171)

4.5 Symmetries of the Quantum Theory 117

Translation invariance of the ground state |0⟩ implies that it is a state withzero total linear momentum, 'P |0⟩ = 0. For a finite displacement a we get

eia·!P |0⟩ = |0⟩ (4.172)

which states that the state |0⟩ is invariant and belongs to a one-dimensionalrepresentation of the group of global translations.

Let us discuss now what happens to global internal symmetries. The sim-plest case is the free complex scalar field φ(x) whose Lagrangian L is in-variant under global phase transformations. If φ is a complex field, it candecomposed it into its real and imaginary parts

φ =1√2(φ1 + iφ2) (4.173)

The classical Lagrangian for a free complex scalar field φ is

L = ∂µφ∗∂µφ−m2φ∗φ (4.174)

now splits into a sum of two independent terms

L(φ) = L(φ1) + L(φ2) (4.175)

where L(φ1) and L(φ2) are the Lagrangians for the free scalar real fields φ1and φ2. Likewise, the canonical momenta Π(x) and Π∗(x) are decomposedinto

Π(x) =δLδ∂0φ

=1√2(φ1 − iφ2) Π∗(x) =

1√2(φ1 + iφ2) (4.176)

In the quantum theory the operators φ and φ† are no longer equal to eachother, and neither are 'Π and 'Π†. Still, the canonical quantization proceduretells us that φ and 'Π (and φ† and 'Π†) satisfy the equal-time canonicalcommutation relations

[φ(x, x0), 'Π(y, x0)] = iδ3(x− y) (4.177)

The theory of the complex free scalar field is solvable by the same methodsthat we used for a real scalar field. Instead of a single creation annihilationalgebra we must introduce now two algebras, with operators a1(k) and a†1(k),

and a2(k) and a†2(k). Let a(k) and b(k) be defined by

a(k) =1√2(a1(k) + ia2(k)) , a†(k) =

1√2

#a†1(k)− ia†2(k)

$

b(k) =1√2(a1(k)− ia2(k)) , b†(k) =

1√2

#a†1(k) + ia†2(k)

$(4.178)

118 Canonical Quantization

which satisfy the algebra

[a(k), a†(k′)] = [b(k), b†(k′)] = (2π)32ω(k) δ3(k − k′) (4.179)

while all other commutators vanish.The Fourier expansion for the fields now is

φ(x) =

"d3k

(2π)32ω(k)

#a(k)e−ik·x + b†(k)eik·x

$

φ†(x) =

"d3k

(2π)32ω(k)

#b(k)e−ik·x + a†(k)eik·x

$

(4.180)

where ω(k) =√k2 +m2 and k0 = ω(k).

The normal ordered Hamiltonian is

: 'H :=

"d3k

(2π)32ω(k)ω(k)

#a†(k)a(k) + b†(k)b(k)

$(4.181)

and the normal-ordered total linear momentum 'P is

'P =

"d3k

(2π)32ω(k)k#a†(k)a(k) + b†(k)b(k)

$(4.182)

We see that there are two types of quanta, a and b. The field φ createsb-quanta and it destroys a-quanta. The vacuum state has no quanta.

The one-particle states have now a two-fold degeneracy since the statesa†(k)|0⟩ and b†(k)|0⟩ have one particle of type a and one of type b respec-tively but these states have exactly the same energy, ω(k), and the samemomentum k. Thus for each value of the energy and of the momentumwe have a two dimensional space of possible states. This degeneracy is aconsequence of the symmetry: the states form multiplets.

What is the quantum operator that generates this symmetry? The classi-cally conserved current is

jµ = iφ∗∂µ↔φ (4.183)

In the quantum theory jµ becomes the normal-ordered operator : jµ :. Thecorresponding global charge 'Q is

'Q = :

"d3x i

#φ†∂0φ− (∂0φ

†)φ$:

=

"d3k

(2π)32ω(k)

#a†(k)a(k)− b†(k)b(k)

$

=Na − Nb (4.184)

4.5 Symmetries of the Quantum Theory 119

where Na and Nb are the number operators for quanta of type a and brespectively. Since [ 'Q, 'H ] = 0, the difference Na − Nb is conserved. Sincethis property is consequence of a symmetry, it is expected to hold in moregeneral theories than the simple free-field case that we are discussing here,provided that [ 'Q, 'H ] = 0. Thus, although Na and Nb in general may notbe conserved separately, the difference Na − Nb will be conserved if thesymmetry is exact.

Let us now briefly discuss how is this symmetry realized in the spectrumof states.

Vacuum State

The vacuum state has Na = Nb = 0. Thus, the generator 'Q annihilates thevacuum

'Q|0⟩ = 0 (4.185)

Therefore, the vacuum state is invariant (i.e., a singlet) under the symmetry,

|0⟩′ = ei!Qα|0⟩ = |0⟩ (4.186)

Because the state |0⟩ is always defined up to an overall phase factor, itspans a one-dimensional subspace of states which are invariant under thesymmetry. This is the vacuum sector and, for this problem, it is trivial.

One-particle States

There are two linearly-independent one-particle states, |+,k⟩ and |−,k⟩defined by

|+,k⟩ = a†(k) |0⟩ |−,k⟩ = b†(k) |0⟩ (4.187)

Both states have the same momentum k and energy ω(k). The 'Q-quantumnumbers of these states, which we will refer to as their charge, are

'Q|+,k⟩ = (Na − Nb)a†(k)|0⟩ = Na a†(k)|0⟩ = +|+,k⟩

'Q|−,k⟩ = (Na − Nb) b†(k)|0⟩ = −|−,k⟩

(4.188)

Hence

'Q|σ,k⟩ = σ |σ,k⟩ (4.189)

where σ = ±1. Thus, the state a†(k)|0⟩ has positive charge while b†(k)|0⟩has negative charge.

120 Canonical Quantization

Under a finite transformation 'U(α) = exp(iα 'Q) the states |±,k⟩ trans-form as follows

|+,k⟩′ = 'U(α) |+,k⟩ = exp(iαQ) |+,k⟩ = eiα |+,k⟩|−,k⟩′ = 'U(α) |−,k⟩ = exp(iαQ)|−,k⟩ = e−iα |−,k⟩

(4.190)

The field φ(x) itself transforms as

φ′(x) = exp(−iα 'Q) φ(x) exp(iα 'Q) = eiαφ(x) (4.191)

since

[ 'Q, φ(x)] = −φ(x), [ 'Q, φ†(x)] = φ†(x) (4.192)

Thus the one-particle states are doubly degenerate, and each state trans-forms non-trivially under the symmetry group.

By inspecting the Fourier expansion for the complex field φ, we see thatφ is a sum of two terms: a set of positive frequency terms, symbolized byφ+, and a set of negative frequency terms, φ−. In this case all positive fre-quency terms create particles of type b (which carry negative charge) whilethe negative frequency terms annihilate particles of type a (which carry pos-itive charge). The states |±,k⟩ are commonly referred to as particles andantiparticles: particles have rest mass m, momentum k and charge +1 whilethe antiparticles have the same mass and momentum but carry charge −1.This charge is measured in units of the electromagnetic charge −e (see theprevious discussion on the gauge current).

Let us finally note that this theory contains an additional operator, thecharge conjugation operator 'C, which maps particles into antiparticles andvice versa. This operator commutes with the Hamiltonian, [ 'C, 'H ] = 0. Thisproperty insures that the spectrum is invariant under charge conjugation. Inother words, for every state of charge Q there exists a state with charge −Q,all other quantum numbers being the same.

Our analysis of the free complex scalar field can be easily extended tosystems which are invariant under a more general symmetry group G. Inall cases the classically conserved charges become operators of the quantumtheory. Thus, there are as many charge operators 'Qa as generators are inthe group. The charge operators represent the generators of the group inthe Hilbert (or Fock) space of the system. The charge operators obey thesame commutation relations as the generators themselves do. A simple gen-eralization of the arguments that we have used here tells us that the statesof the spectrum of the theory must transform under the irreducible repre-sentations of the symmetry group. However, there is one important caveat

4.5 Symmetries of the Quantum Theory 121

that should be made. Our discussion of the free complex scalar field showsus that, in that case, the ground state is invariant under the symmetry. Ingeneral, the only possible invariant state is the singlet state. All other statesare not invariant and transform non-trivially.

But, should the ground state always be invariant? In elementary quantummechanics there is a theorem, due to Wigner and Weyl, which states thatfor a finite system, the ground state is always a singlet under the actionof the symmetry group. However, there are many systems in Nature, suchas magnets, Higgs phases, superconductors, and many others, which haveground states which are not invariant under the symmetries of the Hamil-tonian. This phenomenon, known as spontaneous symmetry breaking, doesnot occur in simple free field theories but it does happen in non-linear orinteracting theories. We will return to this important question later on.