Advanced Quantum Mechanics
Advanced Quantum Mechanics
Peter S. Riseborough
February 19, 2015
Contents
1 Introduction 5
3 Maxwell’s Equations 14
3.1 Vector and Scalar Potentials . . . . . . . . . . . . . . . . . . . . . 15
3.2 Gauge Invariance . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
1
7 The Electromagnetic Lagrangian 54
7.1 Conservation Laws for Electromagnetic Fields . . . . . . . . . . . 58
7.2 Massive Spin-One Particles . . . . . . . . . . . . . . . . . . . . . 64
7.3 Polarizations of Massive Spin-One Particles . . . . . . . . . . . . 65
7.4 The Propagator for Massive Photons. . . . . . . . . . . . . . . . 68
9 Gravitational Interactions 74
9.1 Mathematical Structure of General Relativity . . . . . . . . . . . 78
9.2 Linearized Gravity . . . . . . . . . . . . . . . . . . . . . . . . . . 79
9.3 The Massive S = 2 Graviton . . . . . . . . . . . . . . . . . . . . 81
9.4 The Sourced Equations . . . . . . . . . . . . . . . . . . . . . . . 84
9.5 The Modes and Energies of Massive Gravitons . . . . . . . . . . 86
2
11.2.3 Raman Scattering . . . . . . . . . . . . . . . . . . . . . . 179
11.2.4 Radiation Damping and Resonance Fluorescence . . . . . 180
11.2.5 Natural Line-Widths . . . . . . . . . . . . . . . . . . . . . 183
11.3 Renormalization and Regularization . . . . . . . . . . . . . . . . 185
11.3.1 The Casimir Effect . . . . . . . . . . . . . . . . . . . . . . 186
11.3.2 The Lamb Shift . . . . . . . . . . . . . . . . . . . . . . . . 198
11.3.3 The Self-Energy of a Free Electron . . . . . . . . . . . . . 201
11.3.4 The Self-Energy of a Bound Electron . . . . . . . . . . . . 205
11.3.5 Brehmstrahlung . . . . . . . . . . . . . . . . . . . . . . . 210
3
12.16.2 Charge Conjugation . . . . . . . . . . . . . . . . . . . . . 369
4
1 Introduction
Non-relativistic mechanics yields a reasonable approximate description of phys-
ical phenomena in the range where the particles’ kinetic energies are small com-
pared with their rest mass energies. However, it should be noted that when the
relativistic invariant mass of a particle is expressed in terms of its energy E and
momentum p via
E 2 − p2 c2 = m2 c4 (1)
it implies that its dispersion relation has two branches
q
E = ±c p2 + m2 c2 (2)
+mc2
2mc2
pc
-mc2
Figure 1: The positive and negative energy branches for a relativistic particle
with rest mass m. The minimum separation between the positive-energy branch
and the negative-energy branch is 2mc2 .
can only change continuously, it is impossible that a particle with positive energy
can make a transition from the positive to negative-energy states. However, in
quantum mechanics, particles can make discontinuous transitions. Therefore, it
is necessary to consider both the positive and negative-energy branches. These
considerations naturally lead one to the concept of particles and anti-particles,
and also to the realization that one must consider multi-particle quantum me-
chanics or field theory.
5
appears to be governed by a small coupling strength
2
e 1
∼ (3)
h̄ c 137.0359979
perturbation theory does not converge. In fact, straightforward perturbation
theory is plagued by infinities. However, physics is a discipline which is aimed
at uncovering the relationships between measured quantities. The quantities
e and m which occur in quantum electrodynamics are theoretical constructs
which, respectively, describe the bare charge of the electron and bare mass of
the electron. This means one is assuming that e and m would be the results
of measurements on a (fictional) electron which does not interact electromag-
netically. That is, e and m are not physically measurable and their values are
therefore unknown. What can be measured experimentally are the renormal-
ized mass and the renormalized charge of the electron. The divergences found in
quantum electrodynamics can be shown to cancel or drop out, when one relates
different physically measurable quantities, as only the renormalized masses and
energies enter the theory. Despite the existence of infinities, quantum electrody-
namics is an extremely accurate theory. Experimentally determined quantities
can be predicted to an extremely high degree of precision.
6
A spin-one particle is described by a vector wave function. This can be
heuristically motivated as follows:
A spin-zero particle has just one state and is uniquely described by a one-
component field ψ.
A spin one-half particle has two independent states corresponding to the two
allowed values of the z-component of the intrinsic angular momentum S z =
± h̄2 . The wave function ψ of a spin one-half particle is a spinor which has two
independent components
(1)
ψ (r, t)
ψ(r, t) = (4)
ψ (2) (r, t)
These two components can be used to represent two independent basis states.
We conjecture that since a particle with intrinsic spin S has (2S + 1) inde-
pendent basis states, then the wave function should have (2S + 1) independent
components.
7
or equivalently
The above equation can be used to determine ψ 0 (r) by using the substitution
r → R̂−1 r so
ψ 0 (r) = ψ(R̂−1 r) (8)
If ê is a unit vector along the axis of rotation, the rotation of r through an
e
exr
ϕ
ex(exr)
r(e.r)
r
R̂ r = r + δϕ ê ∧ r + . . . (9)
where terms of order δϕ2 have been neglected. Hence, under an infinitesimal
ψ ψ(R-1r)
ψ(r) y-axis
r'=R-1r
x-axis
rotation, the transformation of a scalar wave function can be found from the
Taylor expansion
ψ 0 (r) = ψ(r − δϕ ê ∧ r)
= ψ(r) − δϕ ( ê ∧ r ) . ∇ ψ(r) + . . .
= ψ(r) − δϕ ( r ∧ ∇ ) . ê ψ(r) + . . .
i δϕ
= ψ(r) − ( ê . L̂ ) ψ(r) + . . .
h̄
8
i δϕ
= exp − ( ê . L̂ ) ψ(r) (10)
h̄
L̂ = − i h̄ r ∧ ∇ (11)
Therefore, locally, rotations of the scalar field are generated by the orbital an-
gular momentum operator L̂.
Since the operation R̂ is a rotation, it also rotates a vector field ψ(r). Not
only does the rotation transfer the magnitude of ψ(r) to the new point r0 but it
must also rotate the direction of the transferred vector ψ(r). The rotated vector
ψ 0 (r0 ) is denoted by R̂ ψ(r). Mathematically, the transformation is expressed
0.5
0
-0.5
-1
-1
-0.5
-0.5
0
0
0.5
0.5
Figure 4: The effect of a rotation R̂ on a vector field ψ(r). The rotation affects
both the magnitude and direction of the vector. 1
1 -2
as 0 2 1 0 -1 -2
2 ψ 0 (r0 ) = R̂ ψ(r) (12)
or equivalently
The part of the rotational operator designated by R̂ does not affect the posi-
tional coordinates (r) of the vector field, and so can be found by considering
the rotation of the vector field ψ at the origin
R̂ ψ = Iˆ + δϕ ê ∧ ψ (15)
9
That is, the operator R̂ only produces a mixing of the components of ψ. Hence,
the complete rotational transformation of the vector field can be represented as
ψ 0 (r) = R̂ ψ(r − δϕ ê ∧ r)
= ψ(r − δϕ ê ∧ r) + δϕ ê ∧ ψ(r − δϕ ê ∧ r)
= ψ(r − δϕ ê ∧ r) + δϕ ê ∧ ψ(r) + . . .
i δϕ
= ψ(r) − ( ê . L̂ ) ψ(r) + δϕ ê ∧ ψ(r) + . . .
h̄
i δϕ i δϕ
= ψ(r) − ( ê . L̂ ) ψ(r) − ( ê . Ŝ ) ψ(r) (16)
h̄ h̄
where the terms of order (δϕ)2 have been neglected and a vector operator Ŝ has
been introduced. The operator Ŝ only admixes the components of ψ µ , unlike
L̂ which only acts on the r dependence of the components. The components of
the three-dimensional vector operator S are expressed as 3 × 3 matrices1 , with
matrix elements
(Ŝ (i) )j,k = − i h̄ ξ i,j,k (18)
where ξ i,j,k is the antisymmetric Levi-Civita symbol. The Levi-Civita symbol
is defined by ξ i,j,k = 1 if the ordered set (i, j, k) is obtained by an even number
of permutations of (1, 2, 3) and is −1 if it is obtained by an odd number of
permutations, and is zero if two or more indices are repeated. Specifically, the
antisymmetric matrices are given by
0 0 0
Ŝ (1) = h̄ 0 0 −i (19)
0 i 0
and by
0 0 i
Ŝ (2) = h̄ 0 0 0 (20)
−i 0 0
and finally by
0 −i 0
Ŝ (3) = h̄ i 0 0 (21)
0 0 0
By using a unitary transform, the above three operators can be transformed
into the standard representation of spin-one operators where S (3) is chosen to
be diagonal. It is easily shown that the components of the matrix operators L̂
and Ŝ satisfy the same type of commutation relations
10
and
[ Ŝ (i) , Ŝ (j) ] = i h̄ ξ i,j,k Ŝ (k) (23)
where the repeated index (k) is summed over. The above set of operators
form a Lie algebra associated with the corresponding Lie group of continuous
rotations. Thus, it is natural to identify these operators which arise in the
analysis of transformations in classical physics with the angular momentum
operators of quantum mechanics. In terms of these operators, the infinitesimal
transformation has the form
i δϕ
ψ 0 (r) ≈ ψ(r) − ê . ( L̂ + Ŝ ) ψ(r) + . . . (24)
h̄
or
i δϕ
ψ 0 (r) = exp − ê . ( L̂ + Ŝ ) ψ(r) (25)
h̄
Thus, the transformation is locally accomplished by
i δϕ
ψ 0 (r) = exp − ( ê . Ĵ ) ψ(r) (26)
h̄
where
Ĵ = L̂ + Ŝ (27)
is the total angular momentum. The operator Ŝ is the intrinsic angular mo-
mentum of the vector field ψ. The magnitude of S is found from
which is evaluated as
1 0 0
Sˆ2 = 2 h̄2 0 1 0 (29)
0 0 1
and is the Casimir operator. It is seen that a vector field has intrinsic angular
momentum, with a magnitude given by the eigenvalue of Sˆ2 which is
E 2 − p2 c2 = 0 (31)
11
which is quantized by using the substitutions
∂
E → i h̄
∂t
p → p̂ = − i h̄ ∇ (32)
One finds that the real scalar wave function ψ(r, t) satisfies the wave equation
1 ∂2
2
− ∇ ψ = 0 (33)
c2 ∂t2
since h̄ drops out. This is not a very useful result, since it is a second-order differ-
ential equation in time, and the solution of a second-order differential equation
can only be determined if two initial conditions are given. Usually, the initial
conditions are given by
ψ(r, 0) = f (r)
∂ψ
= g(r) (34)
∂t t=0
We shall try and factorize the wave equation for the vector E into two first-
order differential equations, each of which requires one boundary condition.
This requires one to specify six quantities. Therefore, one needs to postulate
the existence of two independently measurable fields, E and B. Each of these
fields should satisfy the two wave equations
1 ∂2
2
− ∇ E = 0 (35)
c2 ∂t2
and
1 ∂2
− ∇2 B = 0 (36)
c2 ∂t2
12
The first-order equations must have the form
∂E
i h̄ = c a p̂ ∧ E + b p̂ ∧ B
∂t
∂B
i h̄ = c d p̂ ∧ B + e p̂ ∧ E (37)
∂t
since the left-hand side is a vector, the right-hand side must also be a vector
composed of the operator p̂ and the wave functions. Like Newton’s laws, these
equations must be invariant under time-reversal invariance, t → − t. The
transformation leads to the identification
a = d = 0 (38)
and
b = −e (39)
if one also requires that one of the two fields changes sign under time reversal2 .
We shall adopt the convention that the field E retains its sign, so
E → E
B → −B (40)
under time-reversal invariance. On taking the time derivative of the first equa-
tion, one obtains
2
2 ∂ E 2 2
− h̄ = − c b p̂ ∧ p̂ ∧ E (41)
∂t2
Likewise, the B field is found to satisfy
∂2B
2 2 2
− h̄ = − c b p̂ ∧ p̂ ∧ B (42)
∂t2
Thus, one has found the two equations
2
2 ∂ E 2 2 2
− h̄ = −c b − p̂ E + p̂ p̂ . E (43)
∂t2
and
∂2B
2 2 2 2
− h̄ = −c b − p̂ B + p̂ p̂ . B (44)
∂t2
On substituting the operator p̂ = − i h̄ ∇ , one obtains
∂2E
2 2 2
= − c b − ∇ E + ∇ ∇ . E (45)
∂t2
2 For the non-relativistic Schrödinger equation, time-reversal invariance implies that t →
t0 = −t and ψ → ψ 0 = ψ ∗ .
13
and
∂2B
2 2 2
= −c b − ∇ B + ∇ ∇.B (46)
∂t2
so h̄ drops out. To reduce these equations to the form of wave equations, one
needs to impose the conditions
∇.E = 0 (47)
and
∇.B = 0 (48)
On identifying the coefficients with those of the wave equation, one requires
that
b2 = 1 (49)
Thus, one has arrived at the set of the source-free Maxwell’s equations
1 ∂E
= ∇ ∧ B
c ∂t
1 ∂B
− = ∇ ∧ E
c ∂t
∇.E = 0
∇.B = 0 (50)
3 Maxwell’s Equations
Classical Field Theories describe systems in which a very large number of par-
ticles are present. Measurements on systems containing very large numbers of
particles are expected to result in average values, with only very small devia-
tions. Hence, we expect that the subtleties of quantum measurements should be
completely absent in systems that can be described as quantum fields. Classical
Electromagnetism is an example of such a quantum field, in which an infinitely
large number of photons are present.
14
1 ∂B
∇ ∧ E + = 0
c ∂t
∇.E = 4πρ
∇.B = 0 (51)
The field equations ensure that the sources j and ρ satisfy a continuity equation.
Taking the divergence of the first equation and combining it with the time
derivative of the third, one obtains
1 ∂ 4π
∇. ∇ ∧ B − ∇.E = ∇.j
c ∂t c
1 ∂ 4π
− ∇.E = ∇.j
c ∂t c
4 π ∂ρ 4π
− = ∇.j (52)
c ∂t c
Hence, one has derived the continuity equation
∂ρ
+ ∇.j = 0 (53)
∂t
which shows that charge is conserved.
One can solve the two source-free Maxwell equations, by expressing the
electric E and magnetic fields B in terms of the vector A and scalar φ potentials,
via
1 ∂A
E = − − ∇φ (54)
c ∂t
and
B = ∇ ∧ A (55)
The expressions for B and E automatically satisfy the two source-free Maxwell’s
equations. This can be seen by examining
1 ∂B
∇ ∧ E + = 0 (56)
c ∂t
15
which, on substituting the expressions for the electromagnetic fields in terms of
the vector and scalar potentials, becomes
1 ∂A 1 ∂
∇ ∧ − − ∇φ + ∇ ∧ A = 0 (57)
c ∂t c ∂t
and the terms involving A cancel since A is analytic. The remaining source-free
Maxwell equation is satisfied, since it has the form
∇.B = 0 (59)
which reduces to
∇. ∇ ∧ A = 0 (60)
Therefore, the six components of E and B have been replaced by the four
quantities A and φ. These four quantities are determined by the Maxwell equa-
tions which involve the sources, which are four in number.
The fields are governed by the set of non-trivial equations which relate A and
φ to the sources j and ρ. When expressed in terms of A and φ, the remaining
non-trivial Maxwell equations become
1 ∂ 1 ∂A 4π
∇ ∧ ∇ ∧ A + + ∇φ = j
c ∂t c ∂t c
1 ∂A
−∇. + ∇φ = 4πρ (61)
c ∂t
but since
∇ ∧ ∇ ∧ A = ∇ ∇.A − ∇2 A (62)
1 ∂2A
2 1 ∂φ 4π
− ∇ A + 2 +∇ ∇ . A + = j
c ∂t2 c ∂t c
2 1 ∂
−∇ φ − ∇.A = 4πρ (63)
c ∂t
16
3.2 Gauge Invariance
The vector and scalar potentials are defined as the solutions of the coupled
partial differential equations describing the electric and magnetic fields
1 ∂A
E = − − ∇φ (64)
c ∂t
and
B = ∇ ∧ A (65)
Hence, one expects that the solutions are only determined up to functions of
integration. That is the vector and scalar potentials are not completely deter-
mined, even if the electric and magnetic fields are known precisely. It is possible
to transform the vector and scalar potentials, in a way such that the E and B
fields remain invariant. These transformations are known as gauge transforma-
tions of the second kind3 .
A → A0 = A − ∇ Λ
1 ∂Λ
φ → φ0 = φ + (66)
c ∂t
where Λ is an arbitrary analytic function and this transformation leaves the E
and B fields invariant. The magnetic field is seen to be invariant since
B0 = ∇ ∧ A0
= ∇ ∧ A − ∇Λ
= ∇ ∧ A
= B (67)
valid for any scalar function Λ has been used. The electric field is invariant,
since the transformed electric field is given by
1 ∂A0
E0 = − − ∇ φ0
c ∂t
3 The transformation
0 iχ
ψ → ψ = ψ exp
h̄
p → p̂0 = − i h̄ ∇ − ∇ χ
used in quantum mechanics is known as a gauge transformation of the first kind.
17
1 ∂ 1 ∂Λ
= − A − ∇Λ − ∇ φ +
c ∂t c ∂t
1 ∂A
= − − ∇φ
c ∂t
= E (69)
In the above derivation, it has been noted that the order of the derivatives can
be interchanged,
∂Λ ∂
∇ = ∇Λ (70)
∂t ∂t
since Λ is an analytic scalar function.
∇.A = 0 (72)
The Lorenz gauge is manifestly Lorentz invariant, whereas the Coulomb gauge
is frequently used in cases where the electrostatic interactions are important.
For example, if the fields (φ, A) do not satisfy the Lorenz gauge condition,
since
1 ∂φ
∇.A + = χ(r, t) (73)
c ∂t
where χ is non-zero, then one can perform the gauge transformation to the new
fields (φ0 , A0 )
1 ∂φ0 1 ∂φ 1 ∂2Λ
∇ . A0 + = ∇ . A − ∇2 Λ + + 2
c ∂t c ∂t c ∂t2
2
1 ∂
= χ − ∇2 − 2 Λ (74)
c ∂t2
The new fields satisfy the Lorentz condition if one chooses Λ to be the solution
of the wave equation
1 ∂2
2
∇ − 2 Λ = χ(r, t) (75)
c ∂t2
18
This can always be done, since the driven wave equation always has a solution.
Hence, one can always insist that the fields satisfy the gauge condition
1 ∂φ0
∇ . A0 + = 0 (76)
c ∂t
Alternatively, if one is to impose the Coulomb gauge condition
∇ . A0 = 0 (77)
one can use Poisson’s equations to show that one can always find a Λ such that
the Coulomb gauge condition is satisfied4 .
In the Lorenz gauge, the equations of motion for the electromagnetic field
are given by
1 ∂2
2 4π
− ∇ + 2 2
A = j
c ∂t c
1 ∂2
− ∇2 + 2 φ = 4πρ (78)
c ∂t2
Hence, A and φ both satisfy the wave equation, where j and ρ are the sources.
The solutions are waves which travel with velocity c.
Exercise:
since in the case of the Coulomb gauge, the vector potential is only known up to the gradient
of any harmonic function Λ.
19
4 Relativistic Formulation of Electrodynamics
Physical quantities can be classified as either being scalars, vectors or tensors
according to how they behave under transformations. Scalars are invariant un-
der Lorentz transformations, and all vectors transform in the same way.
where repeated indices are summed over. The invariant length xµ xµ is related
to the proper time τ . This definition can be generalized to the scalar product
of two arbitrary four-vectors Aµ and B µ as
Aµ Bµ = gµ,ν Aµ B ν (84)
Aµ = gµ,ν Aν (85)
20
where µ labels the rows and ν labels the columns. If the four-vectors are ex-
pressed as column-vectors
(0)
A
A(1)
Aν =
A(2)
(87)
A(3)
and
A(0)
A(1)
Aν =
A(2) (88)
A(3)
then the transformation from contravariant to covariant components can be
expressed as
(0)
A(0) 1 0 0 0 A
A(1) 0 −1 0 0 A(1)
A(2) = 0 0 −1 (89)
0 A(2)
A(3) 0 0 0 −1 A(3)
The inverse transform is expressed as
Aµ = g µ,ν Aν (90)
21
involving the metric tensor gµ,ν
1 cos θ
(95)
cos θ 1
Likewise, the length |X| of the vector can be expressed as
(1)
2 (1) (2)
1 cos θ X
|X| = X X (96)
cos θ 1 X (2)
The covariant components of the vector (X(1) , X(2) ) are found from the con-
travariant components by the action of the metric tensor
(1)
X(1) 1 cos θ X
= (97)
X(2) cos θ 1 X (2)
The covariant components of any vector can be found geometrically by dropping
normals from the tip of the vector to the axes. The intersection of a normal with
its corresponding axis determines the covariant component. For a Cartesian co-
ordinate system, the covariant components are identical to the contravariant
components.
A familiar example of the Lorentz invariant scalar product involves the mo-
mentum four-vector with contravariant components pµ ≡ ( Ec , p(1) , p(2) , p(3) )
where E is the energy. The covariant components of the momentum four-vector
are given by pµ ≡ ( Ec , −p(1) , −p(2) , −p(3) ) and the scalar product defines the
invariant mass m via
2
E
p µ pµ = − p2 = m2 c2 (98)
c
Another scalar product which is frequently encountered is pµ xµ which is given
by
pµ xµ = E t − p . x (99)
This scalar product is frequently seen in the description of planes of constant
phase of waves.
22
the scalar function φ(xµ0 ) is still a scalar. Therefore, on performing a Taylor
expansion, one has
∂
φ(xµ + aµ ) = φ(xµ ) + aµ φ(xµ ) + . . . (101)
∂xµ
which is also a scalar. Therefore, the quantity
∂
aµ φ(xµ ) (102)
∂xµ
is a scalar and can be interpreted as a scalar product between the contravariant
vector displacement aµ and the covariant gradient
∂
φ(xµ ) (103)
∂xµ
The covariant gradient can be interpreted in terms of a covariant derivative
∂
∂µ = µ
∂x
1 ∂ ∂ ∂ ∂
= , , ,
c ∂t ∂x(1) ∂x(2) ∂x(3)
1 ∂
= , ∇ (104)
c ∂t
23
which is of the form of a Lorentz scalar. Likewise, if one introduces the current
density four-vector j µ with contravariant components
jµ = c ρ , j (1) , j (2) , j (3)
= cρ, j (108)
which is a Lorentz scalar. Also, the gauge transformation can also be compactly
expressed in terms of a transformation of the contravariant vector potential
Aµ → Aµ0 = Aµ + ∂ µ Λ (110)
Similarly, one can use the contravariant notation to express the quantization
conditions
∂
E → i h̄
∂t
p → − i h̄ ∇ (112)
in the form
pµ → i h̄ ∂ µ (113)
One can also express the wave equation operator in terms of the scalar product
of the contravariant and covariant derivative operators
1 ∂2
∂ µ ∂µ = − ∇2 (114)
c2 ∂t2
Hence, in the Lorenz gauge, the equations of motion for the four-vector potential
Aµ can be expressed concisely as
4π µ
∂ ν ∂ν Aµ = j (115)
c
However, these equations are not gauge invariant.
24
4.3 Lorentz Transformations
A Lorentz transform can be defined as any transformation which leaves the
scalar product of two four-vectors invariant. Under a Lorentz transformation,
an arbitrary four-vector Aµ is transformed to Aµ0 , via
Aµ0 = Λµ ν Aν (116)
where the repeated index ν is summed over. The inverse transformation is
represented by
Aµ = ( Λ−1 )µ ν Aν 0 (117)
Since the scalar product is to be relativistically invariant, one requires that
Aµ0 Bµ 0 = Aµ Bµ (118)
The left hand-side is evaluated as
Aµ0 Bµ 0 = Λµ ν Aν gµ,σ Λσ τ B τ (119)
If the scalar product is to be invariant, the transform must satisfy the condition
gν,τ = Λµ ν gµ,σ Λσ τ (120)
If this condition is satisfied, then Λµ ν is a Lorentz transformation.
25
x(3)
x'(3)
v
x(2)
x'(2)
O
O'
x(1) x'
(1)
Figure 5: Two inertial frames of reference moving with a constant relative ve-
locity with respect to each other.
x(3)
x'(2)
x(2)
x'(1)
φ
O x(1)
Figure 6: Two inertial frames of reference rotated with respect to each other.
Likewise, the rotation through an angle ϕ about the x(3) -axis represented by
1 0 0 0
0 cos ϕ sin ϕ 0
Λµ ν = 0 − sin ϕ cos ϕ 0
(126)
0 0 0 1
Since the boost velocity v and the angles of rotation ϕ are continuous, one
could consider transformations where these quantities are infinitesimal. Such
infinitesimal transformations can be expanded as
Λµ ν = δ µ ν + µ ν + . . . (127)
26
where δ µ ν is the Kronecker delta function representing the identity transforma-
tion5 and µ ν is a matrix which is first-order in the infinitesimal parameter. The
condition on µ ν required for Λµ ν to be a Lorentz transform is given by
or, on using the metric tensor to lower the indices, one has
Exercise:
Show that a Lorentz transformation from the unprimed rest frame to the
primed reference frame moving along the x(3) -axis with constant velocity v, can
be considered as a rotation through an imaginary angle θ = i χ in space-time,
where i c t plays the role of a spatial coordinate. Find the equation that deter-
mines χ.
27
shall express the six ( (16−4)
2 = 6) independent components of the antisymmetric
tensor in terms of the four-vector potential Aµ and the contravariant derivative
as
F µ,ν = ∂ µ Aν − ∂ ν Aµ (132)
so the tensor is antisymmetric
Aµ → Aµ0 = Aµ + ∂ µ Λ (134)
since
∂µ∂ν Λ − ∂ν ∂µΛ ≡ 0 (135)
µ,ν
Alternatively, explicit evaluation of F shows that the six independent com-
ponents can be expressed in terms of the electric and magnetic fields, which are
gauge invariant. Components of the field tensor are explicitly evaluated from
the definition as
1 ∂ (1) ∂
F 0,1 = A − φ
c ∂t ∂x1
1 ∂ (1) ∂
= A + φ
c ∂t ∂x(1)
= − E (1) (136)
and
∂ ∂
F 1,2 = A(2) − A(1)
∂x1 ∂x2
∂ ∂
= − A(2) + A(1)
∂x(1) ∂x(2)
= − B (3) (137)
The non-zero components of the field tensor are related to the spatial compo-
nents (i, j, k) of the electromagnetic field by
and
F i,j = − ξ i,j,k B (k) (139)
where ξ i,j,k is the Levi-Civita symbol. Therefore, the field tensor can be ex-
pressed as the matrix
28
Maxwell’s equations can be written in terms of the field tensor as
4π µ
∂ν F ν,µ = j (141)
c
For µ = i, the field equations become
1 ∂ 0,i ∂ 4 π (i)
F + F j,i = j
c ∂t ∂xj c
1 ∂ (i) ∂ 4 π (i)
− E + ξ i,j,k B (k) = j
c ∂t ∂xj c
(i)
1 ∂ (i) 4 π (i)
− E + ∇ ∧ B = j (142)
c ∂t c
since F 0,0 vanishes. The above field equations are the two Maxwell’s equations
which involve the sources of the fields. The remaining two sourceless Maxwell
equations are expressed in terms of the antisymmetric field tensor as
where the indices are permuted cyclically. These internal equations reduce to
∇.B = 0 (145)
when µ, ν and ρ are the space indices (1, 2, 3). When one index taken from the
set (µ, ν, ρ) is the time index, and the other two are different space indices, the
field equations reduce to
1 ∂B
+ ∇ ∧ E = 0 (146)
c ∂t
If two indices are repeated, the above equations are satisfied identically, due to
the antisymmetry of the field tensor.
Alternatively, when expressed in terms of the vector potential, the field equa-
tions of motion are equivalent to the wave equations
4π µ
∂ν ∂ ν Aµ − ∂ µ ∂ν Aν = j (147)
c
29
Since four-vectors Aµ and j µ transform as
Aµ0 = Λµ ν Aν
j µ0 = Λµ ν j ν (148)
and likewise for the contravariant derivative
∂ µ0 = Λµ ν ∂ ν (149)
then one can conclude that the field tensor transforms as
F µ,ν 0 = Λµ σ Λν τ F σ,τ (150)
This shows that, under a Lorentz transform, the electric and magnetic fields
(E, B) transform into themselves.
Exercise:
Show explicitly, how the components of the electric and magnetic fields
change, when the coordinate system is transformed from the unprimed refer-
ence frame to a primed reference frame which is moving along the x(3) -axis with
constant velocity v.
The Lagrangian for the string is a function of the coordinates yi and the
velocities dy
dt . The Lagrangian is given by
i
i=N 2 2
X mi dyi κi
L = − yi − yi−1 (151)
i=1
2 dt 2
The first term represents the kinetic energy of the mass elements, and the second
term represents the increase in the elastic potential energy of the section of the
string between the i-th and (i − 1)-th element as the string is stretched from its
equilibrium position. This follows since, ∆si the length of the section of string
between mass element i and i − 1 in a non-equilibrium position is given by
∆s2i = ( xi − xi−1 )2 + ( yi − yi−1 )2
= a2 + ( yi − yi−1 )2 (152)
30
y
yi+1
yi yi+1-yi
yi-1 a
xi-1 xi xi+1 x
since the x-coordinates are fixed. Thus, if one assumes that the spring constant
for the stretched string segment is κi , then the potential energy of the segment
is given by
κi
Vi = ( yi − yi−1 )2 (153)
2
We shall consider the case of a uniform string for which κi = κ for all i.
to be evaluated for arbitrary functions yi (t). The string follows the trajectory
yiex (t) which minimizes the action, which travels between the fixed initial value
yi (0) and the final value yi (T ). We shall represent the deviation of an arbitrary
trajectory yi (t) from the extremal trajectory by δyi (t), then
δyi (t) = yi (t) − yiex (t) (155)
The action can be expanded in powers of the deviations δyi as
S = S0 + δ 1 S + δ 2 S + . . . (156)
where S0 is the action evaluated for the extremal trajectories. The first-order
deviation found by varying δyi is given by
Z T i=N
X ex
dδyi dyi
δ1 S = dt mi − κ δyi yiex − yi−1
ex ex
+ κ δyi yi+1 − yiex
0 i=1
dt dt
(157)
31
in which yi (T ) and dy
dt are to be evaluated for the extremal trajectory. Since
i
the trajectory which the string follows minimizes the action, the term δ 1 S must
vanish for an arbitrary variation δyi . We can eliminate the time derivative of
the deviation by integrating by parts with respect to t. This yields
T i=N
d dyiex
Z X
1 ex ex ex ex
δ S = dt − mi δyi − κ δyi yi − yi−1 + κ δyi yi+1 − yi
0 i=1
dt dt
ex T
X dyi
+ mi δyi (t) (158)
i
dt 0
The boundary term vanishes since the initial and final configurations are fixed,
so
δyi (T ) = δyi (0) = 0 (159)
Hence the first-order variation of the action reduces to
Z T i=N
d dyiex
X
1 ex ex ex
δ S = dt δyi − mi − κ 2 yi − yi−1 − yi+1
0 i=1
dt dt
(160)
The linear variation of the action vanishes for an arbitrary δyi (t), if the term in
the square brackets vanishes
d dyiex
ex ex ex
mi + κ 2 yi − yi−1 − yi+1 = 0 (161)
dt dt
Thus, out of all possible trajectories, the physical trajectory yiex (t) is determined
by the equation of motion
d dyi
mi = − κ 2 yi − yi−1 − yi+1 (162)
dt dt
32
The Hamiltonian is only a function of the pairs of canonically conjugate mo-
menta pi and coordinates yi . This can be seen, considering infinitesimal changes
in yi , dy
dt and pi . The resulting infinitesimal change in the Hamiltonian dH is
i
expressed as
i=N
X
dyi dyi ∂L dyi ∂L
dH = dpi + pi d( ) − d( ) − dy i
i=1
dt dt ∂( dy
dt )
i dt ∂yi
i=N
X dyi ∂L
= dpi − dyi (166)
i=1
dt ∂yi
When expressed in terms of the Hamiltonian, the equations of motion have the
form
dyi ∂H
=
dt ∂pi
dpi ∂H
= − (170)
dt ∂yi
The Hamilton equations of motion reduce to
dyi pi
=
dt mi
dpi
= − κi yi − yi−1 + κi+1 yi+1 − yi (171)
dt
for each i value N ≥ i ≥ 1.
One can define the Poisson brackets of two arbitrary quantities A and B in
terms of derivatives with respect to the canonically conjugate variables
i=N
X ∂A ∂B
∂B ∂A
A, B = − (172)
i=1
∂yi ∂pi ∂yi ∂pi
33
The Poisson bracket is antisymmetric in A and B
A, B = − B, A (173)
and
pi , pj = yi , yj = 0 (175)
The increase in energy of this segment, per unit time, is clearly given by the
difference of the quantity
dyi
Pi = − κ yi+1 − yi (179)
dt
at the front end of the segment and Pi−1 at the back end of the segment. Since,
from continuity of energy, the rate of increase in the energy of the segment must
equal the net inflow of energy into the segment, one can identify Pi as the flux
of energy flowing out of the i-th into the (i + 1)-th segment.
34
5.1 The Continuum Limit
The displacement of each element of the string can be expressed as a function
of its position, via
yi = y(xi ) (180)
where each segment has length a, so that xi+1 = xi + a. The displacement
y(xi+1 ) can be Taylor expanded about xi as
a2 ∂ 2 y
∂y
y(xi+1 ) = y(xi ) + a + + ... (181)
∂x xi 2! ∂x2 xi
We intend to take the limit a → 0, so that only the first few terms of the series
need to be retained. The summations over i are to be replaced by integrations
N Z L
X 1
→ dx (182)
i=1
a 0
T = κa (183)
where 2 2
1 dy ∂y
L = ρ − κa (185)
2 dt ∂x
The equations of motion are found from the extrema of the action
Z T Z L
S = dt dx L (186)
0 0
It should be noted, that in S time and space are treated on the same footing
and that L is a scalar quantity.
35
where the Hamiltonian density H is given by
2 2
1 dy ∂y
H = ρ + κa (188)
2 dt ∂x
36
where the ck are arbitrary complex numbers that depend on k. If the time
dependence of the ψk (x) is absorbed into the complex functions ck via
ck (t) = ck (0) exp − i ωk t (198)
which is purely real. Thus, the field y(x) is determined by the amplitudes of
the normal modes, i.e. by ck (t). The time-dependent amplitude ck (t) satisfies
the equation of motion
d2 ck
= − ωk2 ck (200)
dt2
and, therefore, behaves like a classical harmonic oscillator. To quantize this
classical field theory, one needs to quantize these harmonic oscillators.
and
dc∗−k
ρa X dck
p(x) = √ + exp i k x (203)
L k dt dt
then after integrating over x, one finds that the energy has the form
dc∗k dc∗−k
ρ X dc−k dck
H = + +
2 dt dt dt dt
k
κa X 2
+ k c−k (t) + c∗k (t) ck (t) + c∗−k (t) (204)
2
k
37
but the frequency is given by the dispersion relation
κa
ωk2 = v 2 k 2 = k2 (206)
ρ
Therefore, the expression for the Hamiltonian simplifies to
X
H = ρ ωk2 c∗k (t) ck (t) + c−k (t) c∗−k (t)
k
X
= ρ ωk2 c∗k (0) ck (0) + c−k (0) c∗−k (0) (207)
k
These relations are simply the results of applying the inverse Fourier transform
to y(x) and p(x). One can find the Poisson brackets relations between ck and
c∗k from
− i ωk 0 ρ ck0 − c∗−k0 , ck + c∗−k
a X
= − p(xi ) , y(xj ) exp − i ( k xi + k 0 xj )
L i,j
a X 0
= + δi,j exp − i ( k xi + k xj )
L i,j
a X 0
= + exp − i ( k + k ) xi
L i
= + δk+k0 (211)
38
Likewise, one can obtain similar expressions for the other commutation relations.
This set of equations can be satisfied by setting
i
c∗k0 , ck = δk,k0 (212)
2 ωk ρ
and
c∗k0 , c∗k = ck0 , ck = 0 (213)
The above set of Poisson brackets can be recast in a simpler form by defining
1
ck = √ ak (214)
2 ωk ρ
and
a∗k0 , a∗k = ak0 , ak = 0 (216)
So one has
[ â†k0 , âk ] = − h̄ δk,k0 (218)
and
[ â†k0 , â†k ] = [ âk0 , âk ] = 0 (219)
To get rid of the annoying h̄ in the commutator, one can set
√
âk = h̄ b̂
†
√ k†
âk = h̄ b̂k (220)
Whether it was noted or not, b̂†k is the Hermitean conjugate of b̂k . The Her-
mitean relation can proved by taking the Hermitean conjugate of ŷ(xi ), and
39
noting that the third rule of quantization states, “Measurable quantities are to
replaced by Hermitean operators”. Therefore, the operator
s
1 X h̄
ŷ(x) = √ b̂k (t) + b̂†−k (t) exp i k x (221)
L 2 ρ ωk
k
has to be the same as ŷ(x). On setting k = −k 0 in the above equation, one has
s
† 1 X h̄ † † † 0
ŷ (x) = √ b̂−k0 (t) + ( b̂k0 (t)) exp + i k x (223)
L k0 2 ρ ωk 0
40
where the b̂k and b̂†k are to be identified as annihilation and creation operators
for the quanta.
is evaluated as
X
κa
P̂ = − h̄ k b̂†−k − b̂k b̂†k + b̂−k (229)
2ρ
k
where the plane-wave orthogonality properties have been used. This quantity
can be expressed as the sum of two terms
κa X † †
P̂ = − h̄ k b̂−k b̂k − b̂k b̂−k
2ρ
k
κa X
+ h̄ k b̂k b̂†k − b̂†−k b̂−k (230)
2ρ
k
h̄ k b̂†k b̂k
X
P̂ = v 2 (231)
k
which obviously is proportional to the sum of the momenta of the quanta. The
quantity κρa is just the square of the wave velocity v 2 . On noting that the
quanta travel with velocities given by v sign(k) and have energies given by
h̄ ωk = h̄ v | k |, one sees that P is the expressed as the total energy flux
associated with the quanta.
The eigenvalues nk are positive integers, including zero. This can be inferred
from the commutation relations
41
which has the consequence that
Therefore, since b̂k lowers the eigenvalue of the number operator by one unit,
one can write
b̂k | nk > = C(nk ) | nk − 1 > (237)
where the complex number C(nk ) has to be determined. The normalization
coefficient C(nk ) can be determined by noting that
On taking the norm of the state and its conjugate, one finds the normalization
However, on using the definition of the number operator and the normalization
condition, one finds that
| C(nk ) |2 = nk (241)
so, on choosing the phase factor, one can define
√
b̂k | nk > = nk | nk − 1 > (242)
Likewise, one can see that Hermitean conjugate operator b̂†k when acting on
an eigenstate of the number operator increases its eigenvalue by one unit
42
0
The coefficient C (nk ) is found from the normalization condition
0
< nk | b̂k b̂†k | nk > = C ∗ (nk ) C 0 (nk ) < nk + 1 | nk + 1 >
0
= | C (nk ) |2 (245)
which with
b̂k b̂†k = n̂k + 1 (246)
yields 0
| C (nk ) |2 = nk + 1 (247)
Since, the phase factor has already been determined by the Hermitean conjugate
equation, one has √
b̂†k | nk > = nk + 1 | nk + 1 > (248)
which raises the eigenvalue of the number operator.
Hence, one sees that the eigenvalues of the number operator are separated
by integers. Furthermore, the smallest eigenvalue corresponds to nk = 0, since
for nk = 0 the equation
√
b̂k | nk > = nk | nk − 1 > (249)
reduces to
b̂k | 0 > = 0 (250)
Hence, the hierarchy of states produced by the annihilation operator acting on
a number operator eigenstate terminates at nk = 0. Thus, the eigenvalues of
the number operator nk can have integer values 0 , 1 , 2 , 3 , . . . , ∞.
The repeated operation of the creation operator b̂†k creates a state with nk
bosonic excitations present in mode k and the denominator provides the correct
normalization for this state.
where the sum runs over all possible number eigenstates, and the complex coef-
ficients C({nk }) are arbitrary except that they must satisfy the normalization
condition X
| C({nk }) |2 = 1 (253)
{nk }
43
5.5 The Classical Limit
The classical limit of the quantum field theory can be characterized by the limit
in which the field operator can be replaced by a function. This requires that the
“classical” states are not only described as states with large numbers of quanta
in the excited normal modes, but also that the state is a linear superposition
of states with different number of quanta, with a reasonable well defined phase
of the complex coefficients. For a quantum state to ideally represent a given
classical state, one needs the quantum state to be composed of a coherent su-
perposition of states with different numbers of quanta.
That states which are eigenstates of the number operators ( | {nk } > ) can
not represent classical states, can be seen by noting that the expectation value
of the field operator is zero
follows from the expectation value of the creation and annihilation operators
Despite the fact that the average value of the field is zero, the fluctuation in the
field amplitude is infinite since
2 1 X h̄ † †
< {nk } | ŷ(x) | {nk } > = < {nk } | b̂k0 + b̂−k 0 b̂k + b̂−k0
0 | {nk } >
L 0 2 ρ ωk 0
k
1 X h̄
= ( 1 + 2 nk 0 ) (256)
L 0 2 ρ ωk 0
k
and the zero-point contribution diverges logarithmically at the upper and lower
limits of integration.
44
is extremal. An arbitrary field φα can be expressed in terms of the extremal
value φα α
ex and the deviation δφ as
φα = φα
ex + δφ
α
(258)
The space and time derivatives of the arbitrary field can also be expressed as
the derivatives of the sum of the extremal field and the deviation
∂ν φα = ∂ν φα
ex + ∂ν δφ
α
(259)
This set of equations determine the time dependence of the classical fields
φα
ex (x). That is, out of all possible fields with components φα , the equations
of motion determine the physical field which has the components φα ex . It is
convenient to define the field momentum density πα0 (x) conjugate to φα as
0 ν 1 ∂ β β
πα (x ) = L φ , ∂µ φ (263)
c ∂(∂0 φα )
Exercise:
45
for a real scalar field φ, determine the Euler-Lagrange equation and the Hamil-
tonian density H.
Exercise:
Exercise:
The Lagrangian density for the complex field ψ representing a charged par-
ticle is given by
h̄2
∗
h̄ ∂ψ ∂ψ
L = − ∇ ψ∗ . ∇ ψ − ψ∗ − ψ − ψ ∗ V (x) ψ
2m 2i ∂t ∂t
(267)
(i) Determine the equation of motion, and the Hamiltonian density H.
(ii) Consider the case V (x) ≡ 0, then by Fourier transforming with respect to
space and time, determine the form of the general solution for ψ.
by noting that H is only a functional of πα0 and φα . This can be seen, since as
Z X
H = d3 x c πα0 (∂0 φα ) − L (269)
α
46
but, from the Lagrangian formulation of field theory, one has
1
δL = δφα (∂0 πα0 ) + (∂0 δφα ) πα0 (271)
c
where the Euler-Lagrange equations were substituted into the first term. There-
fore, the variation in the Hamiltonian is given by
Z X
3 0 α α 0
δH = d xc δπα (∂0 φ ) − δφ (∂0 πα ) (272)
α
which does not involve the time derivative of the fields. This implies that
the Hamiltonian is a function of the fields πα0 , φα and their derivatives. On
calculating the variation of H using the independent variables πα0 and φα , and
integrating by parts, one finds that the Hamiltonian equations of motion are
given by
∂H ∂H
c ∂0 φα = − ∇
∂π 0 ∂(∇πα0 )
α
∂H ∂H
− c ∂0 πα0 = − ∇ (273)
∂φα ∂(∇φα )
The structure of these equations are similar to those of the classical mechanics of
point particles. Similar to classical mechanics of point particles, one can define
Poisson Brackets with fields. When quantizing the fields, the Poisson Bracket
relations between the fields can be replaced by commutation relations.
47
Under this combined transformation, the Lagrangian density changes by an
infinitesimal amount δL, given by
∂L ∂L
δL = α
∂µ δφα + δφα (276)
∂(∂µ φ ) ∂φα
where the field index α is to be summed over. However, the generalized mo-
mentum density παµ (x) is defined by
µ ∂L
πα (x) = (277)
∂(∂µ φα )
so
∂L
δL = παµ (∂µ δφα ) + δφα (278)
∂φα
The Euler-Lagrange equation for each field φα is given by
∂L
∂µ παµ − = 0 (279)
∂φα
where φα satisfies the appropriate boundary conditions. Thus, on adding and
subtracting a term
(∂µ παµ ) δφα (280)
to δL, one finds
∂L
δL = παµ ∂µ δφα + (∂µ παµ ) δφα + − ∂ π
µ α
µ
δφα
∂φα
∂L
= ∂µ παµ δφα + − ∂ π
µ α
µ
δφα
∂φα
= ∂µ παµ δφα (281)
since the last term in the second line vanishes if the fields φα satisfy the Euler-
Lagrange equations. If the Lagrangian is invariant under the transformation,
then δL = 0, so
∂µ παµ δφα = 0 (282)
where the field index α is to be summed over. The above equation can be
re-written as a continuity equation
∂µ j µ = 0 (283)
48
up to a constant of proportionality. The normalization of the conserved current
is arbitrary and can be chosen at will. Since it is recognized that δφα is in-
finitesimal, the normalization is chosen by introducing an infinitesimal constant
via
The conserved charge Q is defined as the integral over all space of the time
component of the current density j (0) . That is, the conserved charge is given by
Z
Q = d3 x j (0) (x) (286)
Since is a constant, the total charge Q is constant. Therefore, the total time
derivative of Q vanishes
dQ
= 0 (288)
dt
The spatial components of j µ form the current density vector.
ψ0 = ψ + iψ
ψ ∗0 = ψ∗ − i ψ∗ (291)
49
where ψ and its complex conjugate ψ ∗ are regarded as independent fields. The
transformation represents a an infinitesimal constant shift of the phase of the
field8 . The conserved current is
µ ∂L ∂L ∗
j = −i ψ(x) − ψ (x) (292)
∂(∂µ ψ) ∂(∂µ ψ ∗ )
Exercise:
h̄2
∗
∗ h̄ ∗ ∂ψ ∂ψ
L = − ∇ψ . ∇ψ − ψ − ψ − ψ ∗ V (x) ψ
2m 2i ∂t ∂t
(293)
(i) Determine the conserved Noether charges.
Exercise:
δL = ∂µ Λµ (295)
8 This particular transformation is a specific example of a gauge transformations of the first
kind, in which
q
ψ 0 (x) = exp − i Λ(x) ψ(x)
h̄ c
A gauge transformation of the second kind is one in which the field changes according to
Aµ0 = Aµ + (∂ µ Λ)
q
Since p̂µ = i h̄ ∂ µ , the combination of these transformations keep the quantity (p̂µ − c
Aµ )ψ
invariant
50
for some analytic vector function with components Λµ . This type of transforma-
tion does not change the total action. If the Lagrangian changes by the above
amount for the combined transformation δφα
φα (x) → φα0 (x) = φα (x) + δφα (x) (296)
then as has been previously shown
∂L
δL = παµ (∂µ δφα ) + (∂µ παµ ) δφα + − ∂µ π µ
α δφα
∂φα
∂L
= ∂µ παµ δφα + − ∂µ π µ
α δφα
∂φα
= ∂µ παµ δφα (297)
one has
∂µ Λ µ = ∂µ παµ δφα (298)
In this case, the change in the Lagrangian density is given by the total derivative
∂L α ∂L
δL = (∂ ν δφ ) + δφα
∂(∂ν φα ) ∂φα
∂L ∂L
= µ ∂ ∂
ν µ φα
+ ∂µ φα
∂(∂ν φα ) ∂φα
= µ ∂µ L (303)
51
where the last line follows since the Lagrangian only depends implicitly on xµ
through the fields. Hence, the change in the Lagrangian is a total derivative
δL = µ ∂µ Λ (304)
φα → φα + µ (∂µ φα ) (305)
where the Euler-Lagrange equation has been used in the second line. Thus, the
fields satisfy the continuity conditions
∂L
0 = µ ∂ν ∂µ φα
− δ ν
µ L (307)
∂(∂ν φα )
where
δµ ν = 1 if µ = ν
δµ ν = 0 otherwise (308)
which is the energy-momentum density Tµν . The energy momentum tensor sat-
isfies the conservation law
∂ν T ν µ = 0 (310)
The second-rank tensor can be written in contravariant form as
∂L
T ν,µ = ∂ µ α
φ − g ν,µ
L (311)
∂(∂ν φα )
where the metric tensor has been used to raise the index µ. The component
with µ = ν = 0 is the Hamiltonian density H for the fields
0,0 ∂L α
H = T = ∂0 φ − L (312)
∂(∂0 φα )
52
so the total energy of the field is given by
Z Z
3
E = d xH = d3 x T 0,0 (313)
are related to the momentum density since the total momentum of the field is
given by Z
1
P (j) = d3 x T 0,j (316)
c
Since T 0,j is the momentum density, one expects that the components of the
orbital angular momentum density are proportional to
∂µ xρ = δµ ρ (320)
expressing the independence of the variables xρ and xµ have been used. The
divergence of the third-rank tensor vanishes if T µ,ν is symmetric. Thus, the
angular momentum tensor M µ,ν,ρ is conserved if the energy-momentum tensor
is symmetric.
It should be noted that the tensor T µ,ν is only symmetric for scalar fields.
This is related to the fact that a vector or tensor field carries a non-zero intrin-
sic angular momentum. It is possible to incorporate an additional term in the
53
momentum-energy tensor of a vector field to make it symmetric.
Exercise:
(i) Determine the momentum-energy tensor for a complex scalar field ψ governed
by the Lagrangian density
2
1 ∗ µ mc 2
L = ( ∂µ ψ ) ( ∂ ψ ) − |ψ| (321)
2 h̄
(ii) Find the forms of the energy and momentum density of the field.
(iii) Using the form of the general solution, find expressions for the total energy
and momentum of the field in terms of the Fourier components of the field.
Exercise:
(i) Determine the energy-momentum tensor for the Lagrangian density for the
complex Schrödinger field representing a charged particle given by
h̄2
∗
∗ h̄ ∗ ∂ψ ∂ψ
L = − ∇ψ . ∇ψ − ψ − ψ − ψ ∗ V (x) ψ
2m 2i ∂t ∂t
(322)
(ii) Find the forms of the energy and momentum density of the field.
(iii) Find the forms of the generalized orbital angular momentum density of the
field.
(iv) Consider the case where V (x) ≡ 0. Using the form of the general solution,
find expressions for the total energy and momentum of the field in terms of the
Fourier components of the field.
54
of the contravariant components of the field. This form allows variations to be
made directly without using the properties of the metric tensor. The first-order
variation of the action can be expressed as
Z
2 4 µ,ν µ,ν
δS = − d x ( ∂µ δAν ) F − ( ∂ν δAµ ) F
16 π c
Z
2 4 µ,ν ν,µ
= d x δAν ∂µ F − F
16 π c
Z
1
= d4 x δAν ( ∂µ F µ,ν ) (325)
4πc
where the second line has been obtained by integrating by parts and the last
line was obtained by using the antisymmetric nature of the field tensor. The
vanishing of the first-order variation of the action δS, for arbitrary δAν , yields
the Euler-Lagrange equation
∂µ F µ,ν = 0 (326)
in which the sign of the terms with mixed time and space indices have changed.
Therefore, the Lagrangian density can be expressed in terms of the electromag-
netic fields as
1
L = ( E2 − B2 ) (329)
8π
Since the Lagrangian density is completely expressed in terms of the electro-
magnetic field, it is gauge invariant.
55
In the presence of source densities, the Lagrangian density is extended to
include the interaction to become
1 1
L = − F µ,ν Fµ,ν − Aµ j µ (330)
16 π c
This interaction term is the only Lorentz scalar that one can form with the
four-vector current and the field. It should be noted that the last term is not
gauge invariant. This action yields the equation of motion
4π ν
∂µ F µ,ν = j (331)
c
as expected.
Since charge is conserved, the current density must satisfy the continuity equa-
tion
∂ µ jµ = 0 (335)
The continuity condition can be used to express the interaction as the untrans-
formed Lagrangian density and a perfect derivative
1 µ 1 µ
Lint 0 = − A jµ − ∂ ( Λ jµ ) (336)
c c
The perfect derivative term only adds a constant term to the action which does
not affect the equations of motion9 . Hence, although the Lagrangian density
is not gauge invariant in the presence of sources, the Lagrangian equations of
motion are gauge invariant.
56
which vanishes for µ = 0, indicating that the scalar potential A0 is not a
dynamic variable. This suggests that it may be appropriate to completely fix
the scalar potential by a choice of gauge, such as the Coulomb gauge which leads
to the scalar potential φ being fixed by Poisson’s equation. In the presence of
sources, the Hamiltonian density is expressed as
1
H = − ( ∂0 Aν ) F 0,ν − L
4π
1
= − ( F0,ν + ∂ν A0 ) F 0,ν − L
4π
1 1 1 µ
= − ( F0,ν + ∂ν A0 ) F 0,ν + ( B2 − E2 ) + j Aµ
4π 8π c
1 1 1 µ
= + ( E2 + B2 ) − ( ∇ . E ) A0 + j Aµ
8π 4π c
1
+ ∇ . ( A(0) E ) (338)
4π
The fourth line has been derived by noting that the non-zero components of
F 0,µ are only non-zero for space-like µ and are given by
57
the expression given in eqn(344) simplifies to
1 1
( ∇A0 ) . E = ∇ . ( A0 E ) − A0 ρ (346)
4π 4π
Therefore, the Hamiltonian density can be expressed as
1 1 µ
H = + ( E 2 + B 2 ) − ρ A0 + j Aµ
8π c
1
+ ∇ . ( A(0) E ) (347)
4π
On combining the term ρ A0 with the last term
1 µ 1
j Aµ = ρ A0 − j.A (348)
c c
which originates from the Lagrangian interaction (−Lint ), one finds that the
terms proportional to A0 ρ in the Hamiltonian density cancel. On neglecting
the total derivative term [ + 41π ∇ . ( φ E ) ], one finds that the Hamiltonian
density reduces to
1 1
H = ( E2 + B2 ) − j.A (349)
8π c
The first term is the energy density of the free electromagnetic field and the
second term represents the energy of the interaction between the electromag-
netic field and “charged particles”. It should be noted that the interaction
Hamiltonian is expressed entirely in terms of an interaction between the current
density and the vector potential, which demonstrates that the Hamiltonian is
not invariant under a Lorentz transformation
1
Hint = − j.A (350)
c
but is invariant under rotations in space. This situation is to be contrasted
with the interaction term in the Lagrangian which was Lorentz invariant as it
explicitly included an interaction between the scalar potential and the charge
density.
58
The Noetherian energy-momentum tensor T ν,µ is found from
∂L
T νµ = ∂µ Aρ
− δν µ L
∂(∂ν Aρ )
∂L
= ∂µ Aρ − δ ν µ L (353)
∂(∂ν Aρ )
The derivative of the Lagrangian density is evaluated as
∂L 1
= − F ν,ρ − F ρ,ν
∂(∂ν Aρ ) 8π
1
= − F ν,ρ (354)
4π
Therefore, the energy-momentum density is found as
ν 1 ν,ρ
T µ = − F ∂µ Aρ − δ ν µ L (355)
4π
On raising the index µ with the metric tensor, one has the contravariant second-
rank tensor
ν,µ 1 ν,ρ
T = − F ∂ Aρ − g ν,µ L
µ
(356)
4π
The energy-momentum tensor is not gauge invariant, as it explicitly involves
the fields Aµ . On using the expression for the source-free Lagrangian density
1
L = E2 − B2 (357)
8π
one finds that the time components of T µ,ν are given by
0,0 1 2 2 1
T = E + B + ∇. φE (358)
8π 4π
The expression T 0,0 is the Hamiltonian density H, in the absence of sources,
which represents the energy density of the free field. The momentum density is
given by the mixed time and space components, and is given by
1
T 0,j = − F 0,ρ ( ∂ (j) Aρ ) (359)
4π
but since F µ,ν is antisymmetric, only the terms where ρ is a spatial index are
non-zero. Hence, one has
1 X 0,i
T 0,j = − F ( ∂ (j) Ai )
4π i
1 X 0,i (j) (i) (i) (j) 1 X 0,i
= + F ∂ A − ∂ A + F ( ∂ (i) A(j) )
4π i 4π i
(360)
59
where the relation between the space-like components of the covariant and con-
travariant four-vector Ai = − A(i) has been used. Since the time component
of the field tensor is given by
and10 X
(i) (j) (j) (i)
∂ A − ∂ A = − ξ i,j,k B (k) (362)
k
∇.E = 0 (364)
The components T 0,ν , apart from the terms involving total derivatives which
integrate out to zero, are related to the total energy and the components of the
total momentum of the electromagnetic field. The components of T µ,ν satisfy
the continuity equations
∂µ T µ,ν = 0 (366)
which represent the conservation of energy and momentum. The other mixed
time and spatial components of the energy-momentum tensor are evaluated as
(j) (j)
1 1 1 ∂
T j,0 = E ∧B + ∇∧ φB − φ E (j)
4π 4π c ∂t
(367)
The components T j,0 represent the components of the energy flux.
60
is not conserved as the energy-momentum tensor is not symmetric. Additional
terms can be added to the energy-momentum tensor11 , to create a symmetric
tensor Θµ,ν . These extra terms account for the intrinsic angular momentum of
the photon.
(∂ ν Aλ ) = − F λ,ν + (∂ λ Aν ) (369)
If Θµ,ν and T µ,ν are to represent the same set of conserved quantities, the last
term in eqn(370) must be expressible as a total derivative. That this is true can
be seen by examining the asymmetric term
1 µ,ρ 1
− g Fρ,λ (∂ λ Aν ) = − F µ,λ (∂λ Aν ) (373)
4π 4π
where the index ρ was raised by using the metric tensor. On combining the
above expression with the source free Maxwell equation
where Λρ;µ,ν is an arbitrary tensor that is antisymmetric under the interchange of the first
pair of indices
Λρ;µ,ν = − Λµ;ρ,ν
will automatically satisfy the same continuity conditions as T µ,ν and leave the total energy
and momentum unaltered.
61
one obtains
1 µ,ρ 1
− g Fρ,λ (∂ λ Aν ) = − F µ,λ ν ν µ,λ
(∂λ A ) + A (∂λ F )
4π 4π
1 µ,λ ν
= − ∂λ F A (375)
4π
On comparing the right hand sides of the first and last line, one finds that they
have opposite signs and, therefore, they are zero. Thus, the difference between
continuity relations vanish
62
and the mixed temporal and spatial components are given by
(j)
0,j 1
Θ = E ∧ B (381)
4π
Noether’s theorem is purely classical, but there are generalizations for quan-
tum fields. Quantum generalizations includes the Ward-Takahashi and Taylor-
Slavnov identities.
Exercise:
Exercise:
Verify the form of the conservation laws for energy and momentum.
Exercise:
Show that the extra term added to the tensor T i,j in order that Θi,j will
be symmetric produces a contribution to the angular momentum density of the
form (j)
0,j 1
S = E ∧ A (384)
4π
which is the intrinsic spin density of the electromagnetic field.
63
7.2 Massive Spin-One Particles
The electromagnetic theory has been unified with the theory of weak interac-
tions. This generalization requires the existence of two new types of spin-one
particles in addition to the photon, which together mediate the electro-weak
interaction. These new particles have non-zero mass. The massive spin-one
particle has to satisfy the equation12
pµ pµ = m2 c2 (385)
where h̄ no longer drops out. This equation can be derived from the Lagrangian
2
1 1 mc 1 µ
L = − F µ,ν Fµ,ν + Aµ Aµ − j Aµ (388)
16 π 8π h̄ c
Neither the Lagrangian, nor the equation of motion are gauge invariant. The ap-
propriate gauge condition can be enforced by imposing conservation of charge13
∂µ j µ = 0 (390)
The first term on the left-hand side vanishes due to the definition of F µ,ν , since
F ν,µ = ∂ ν Aµ − ∂ µ Aν (392)
one finds
∂ν F ν,µ = ∂ν ∂ ν Aµ − ∂ µ ∂ν Aν (393)
12 A. Proca, J. Phys. et Radium 7, 147 (1936).
13 Note that, unlike the massless photon, charge conservation has to be imposed as an
additional assumption.
64
therefore
∂µ ∂ν F ν,µ = ∂ν ∂ ν ∂µ Aµ − ∂µ ∂ µ ∂ν Aν
= 0 (394)
The term on the right-hand side of eqn(391) also vanishes, because it was chosen
to impose charge conservation. Hence, one finds that Aµ for a massive spin-one
particle must satisfy the Lorenz gauge condition
∂µ Aµ = 0 (395)
Exercise:
which results in four components associated with four polarization vectors which
are denoted by êµ (k). For massive photons, if one assumes charge conservation,
the gauge fields must satisfy the Lorentz gauge condition. The Lorentz gauge
condition
∂µ Aµ (x) = 0 (397)
results in the Fourier components satisfying the condition
kµ Aµ (k) = 0 (398)
For a photon with mass m, one can apply the Lorentz gauge condition in
the laboratory frame. We shall choose the z-axis as the direction of the pho-
ton’s three-momentum k. With this choice, the photon’s four-momentum has
65
q
2
components (k (0) , 0, 0, k (3) ) where k (0) = ( mc 2
h̄ ) + k . The Lorentz condition
becomes
k (0) A(0) (k) − k (3) A(3) (k) = 0 (399)
(3)
Hence, A(0) (k) = kk(0) A(3) (k). This implies that in the photon’s rest frame, the
four-vector potential only has three spatial components and the scalar potential
is zero. In any case, there are only three-independent components of the four-
vector potential. The four-vector potential can be expressed in terms of the
three independent components as
1 X k (3)
A(x) = √ A(3) (k) ( (0) ê(0) (k) + ê(3) (k) )
2V k k
(1) (1) (2) (2) ν
+ A (k) ê (k) + A (k) ê (k) exp[ i k xν ] + c.c.
(400)
The longitudinal component of the electric field has a Fourier amplitude given
by
(3) (0) (3) (3) (0)
E (k) = − i k A (k) − k A (k)
k (3)2
= −i k (0) − A(3) (k)
k (0)
2
m c
h̄
= −i A(3) (k) (402)
k (0)
and the longitudinal component of the magnetic field B (3) (k) is zero since
(3) (3)
0 = − i k ê ∧ A(k) . ê(3) (k) (403)
Thus, the component A(3) (k) contributes to the E field but not the B field.
For massive photons, the unsymmmetrized energy-momentum tensor T µ,ν is
expressed in terms of contributions from, not only the E and B fields, but also
66
from the gauge field Aµ . The tensor is evaluated as
2
g µ,ν
j ρ Aρ
µ,ν 1 µ,ν ρ,σ µ,ρ ν mc
T = g Fρ,σ F −4F ∂ Aρ − Aρ Aρ + g µ,ν
16 π 8π c h̄
(404)
The energy-momentum tensor can be “symmetrized” by adding and subtracting
a term
4
+ F µ,ρ ∂ρ Aν (405)
16 π
to express the second term in a gauge-invariant form. We then note that that
remaining term (with the negative sign) could be put into the form of a diver-
gence of an antisymmetric third rank-tensor, if we could combine it with a term
of the form
1 1
− ( ∂ρ F µ,ρ ) Aν = + ( ∂ρ F ρ,µ ) Aν (406)
4π 4π
This can be accomplished by adding zero in the form of Aν times the Euler-
Lagrange equation
2
mc 4π µ
+ ∂ρ F ρ,µ + Aµ − j = 0 (407)
h̄ c
therefore, completing the divergence at the expense of adding extra mass and
source terms. The divergence of third-rank antisymmetric tensor can be dropped,
leading to the “symmetrized” energy-momentum tensor given by
µ,ν 1 µ,ν ρ,σ µ,ρ ν
Θ = g Fρ,σ F + 4F Fρ
16 π
2 2
g µ,ν m c
1 mc
− Aρ Aρ + Aµ Aν
8π h̄ 4π h̄
j ρ Aρ j µ Aν
+ g µ,ν − (408)
c c
It is seen that the coupling of the electromagnetic fields to the current densities
spoils the symmetry of the energy-momentum tensor. Hence, the charge cur-
rents act as sources of angular momentum and results in the electromagnetic
field’s angular momentum not being conserved. The energy-density H is given
by Θ0,0 so
2
j.A
1 2 2 1 mc (0)2 2
H = E + B + A + A − (409)
8π 8π h̄ c
67
Thus, the longitudinal photon with components A(3) (k) and A(0) (k) does have
physical effects, as do the two transverse photons A(1) (k) and A(2) (k). There-
fore, the four-vector potential of a massive electromagnetic field has three phys-
ical components.
Exercise:
On Fourier transforming the linear equation with respect to space and time, one
obtains
2
mc 4π ν
− kµ k µ + Aν (k) + kµ k ν Aµ (k) = j (k) (414)
h̄ c
68
which can be re-written as
2
ρ mc ν ν 4π ν
− kρ k + δµ + kµ k Aµ (k) = j (k) (415)
h̄ c
The photon propagator is defined via
and then solving for the unknown quantities B(k 2 ) and A(k 2 ). This leads to
the following expression for the photon propagator
µ ν
g µ,ν − (km ck)2
4π h̄
Dµ,ν (k) = 2 (419)
c
m c ρ k
h̄ − k ρ
Hence, Aµ (x) can be found from the inverse Fourier transform. It should be
noted that, if one imposes current conservation
k ν jν (k) = 0 (421)
the second term in the numerator of the photon propagator has no physical
effect. We also note that
2
4π h̄
kµ Aµ (k) = k ν jν (k) (422)
c mc
So, once again we see that assumption of charge conservation enforces the
Lorentz gauge condition, even if the massive photon is involved in a virtual
process.
69
8 Symmetry Breaking and Mass Generation
We shall first look at an example of Goldstone’s theorem which states that,
if a system described by a Lagrangian which has a continuous symmetry (and
only short-ranged interactions) has a broken symmetry state then the system
supports a branch of small amplitude excitations with a dispersion relation ωk
that vanishes at k = 0. We shall then examine the situation in which the system
is coupled by long-ranged interactions, as modelled by an electromagnetic field.
As was first pointed out by Anderson, the long-ranged interactions alter the
excitation spectrum of the symmetry broken state by removing the Goldstone
modes and generating a branch of massive excitations.
for any real constant α. The static or minimum energy solution corresponds to
| ψ | = φ0 (425)
ψ = φ1 + i φ2
ψ∗ = φ1 − i φ2 (426)
then the Lagrangian can be written as a Lagrangian density involving the two
real scalar fields φ1 and φ2 . The Lagrangian density has a U (1) symmetry which
corresponds to the rotation of ψ around a circle about the origin in the (φ1 , φ2 )
plane.
We shall assume the field ψ representing the physical ground state corre-
sponds to only one of the infinite number of possible candidates. The physical
state must have a phase, which shall be defined as zero. That is, one starts with
a ground state ψ = φ0 , and then consider the small amplitude excitations. A
low-energy excited state corresponds to the complex field
ψ = φ0 + δψ (427)
70
v(Ψ)
Re [Ψ]
Im [Ψ] φ0
where δψ is static and uniform and can be considered to be very small. The
small amplitude complex field δψ can be expressed in terms of its real and
imaginary parts
δψ = χ1 + i χ2 (428)
The Lagrangian density takes the form
2 2
µ µ mc
L = ( ∂µ χ1 ) ( ∂ χ1 ) + ( ∂µ χ2 ) ( ∂ χ2 ) − 2 φ0 χ1 + χ21 + χ22
2 h̄ φ0
(429)
If one only considers infinitesimally small amplitude oscillations, one only needs
consider terms quadratic in the fields. The quadratic Lagrangian density LF ree
describes non-interacting fields. The quadratic Lagrangian density is given by
2
mc
LF ree = ( ∂µ χ1 ) ( ∂ µ χ1 ) − χ21 + ( ∂µ χ2 ) ( ∂ µ χ2 ) (430)
h̄
The symmetry breaking has resulted in the complex field breaking up into two
fields: The first field χ1 describes massive excitations m and the second field χ2
describes massless excitations. The first field χ1 has plane-wave solutions if the
energy and momentum are related via the dispersion relation
2
m c2
ω 2 = c2 k 2 + (431)
h̄
and represents excitations which corresponds to a “stretching” of φ0 . It is
massive since this excitation moves the field away from the minimum of the
potential. The second excitation χ2 represents δψ which is transverse to φ0 in
71
the (φ1 , φ2 ) plane. This last excitation is known as a Goldstone boson14 . The
Goldstone boson has a dispersion relation
ω 2 = c2 k 2 (432)
(1963).
17 P.W. Higgs, “Broken Symmetries, Massless Particles and Gauge Fields”, Physics Letters,
12, 132 (1964): P.W. Higgs, “Broken Symmetries and the Masses of Gauge Bosons”. Physical
Review Letters 13, 508 (1964):P.W. Higgs, “Spontaneous Symmetry Breaking without mass-
less Bosons”, Physical Review 145, 1156 (1966)
18 G.S. Guralnik, C.R. Hagen, and T.W.B. Kibble, “Global Conservation Laws and Massless
Particles”. Physical Review Letters 13, 585-587 (1964): T.W.B. Kibble, “Symmetry Breaking
in non-Abelian Gauge Theories”, Physical Review, 155, 1554 (1967).
72
The Lagrangian density is invariant under the local gauge transformation19
0 q
ψ → ψ = exp − i Λ ψ
h̄ c
Aµ → Aµ0 = Aµ + ∂ µ Λ (436)
| ψ | = φ0 (437)
and the Aµ vanish. Any local gauge transformation leads to a state with the
same energy, therefore, the ground state is infinitely degenerate.
ψ = φ0 + δψ (438)
δψ = χ1 (439)
and on substituting in the Lagrangian and collecting the quadratic terms, one
obtains
2
mc
LF ree = ( ∂µ χ1 ) ( ∂ µ χ1 ) − χ21
h̄
2
1 q φ0
− F µ,ν Fµ,ν + Aµ Aµ (440)
16 π h̄ c
Therefore, one finds that the charged boson field has a mass m and the gauge
field has acquired a mass mA given by
2
q φ0
m2A = 8π (441)
c2
73
of the gauge field. More specifically, the field χ2 which initially corresponded
to the Goldstone mode became unphysical when the massless vector field was
introduced as it could be gauged away. This is seen by writing
χ2
ψ = φ0 + χ1 + i χ2 ≈ φ0 + χ1 exp i (442)
φ0
9 Gravitational Interactions
We introduce the field theory of the Gravitational Interaction, in the weak field
limit, by analogy with electromagnetism. More specifically, we shall develop the
classical field theory of the massive graviton in parallel with the Proca’s theory
of a massive spin-one particle that we discussed previously. This approach has
the disadvantage that it does not have the same beautiful geometric basis as
Einstein’s field equations.
74
quantum propagators.
where
F µ,ν = ∂ µ Aν − ∂ ν Aµ (447)
The mass term spoils the gauge invariance of the source-free Lagrangian. The
action S is given by Z
S = d4 x L (448)
If charge is conserved
∂µ j µ = 0 (451)
On operating on the equation of motion with ∂µ , one finds that because of the
finite mass the field Aν must satisfy the condition
∂ ν Aν = 0 (452)
Thus, conservation of charge removes the choice of gauge. In the rest frame
of a photon, k = ( mh̄ c )(1, 0, 0, 0), so one has A0 (k) = 0 and so one finds that
the massive photon has three components. This is expected for a particle with
spin S = 1. The electromagnetic propagator can be defined as the kernel of the
integral relation
Z
4π
Aν (x) = d4 x0 Dν,λ (x, x0 ) j λ (x0 ) (453)
c
Thus, the propagator satisfies the partial differential equation
2
mc
∂ρ ∂ ρ
+ g µ,ν
− ∂ ∂µ ν
Dν,λ (x, x0 ) = δλµ δ 4 (x − x0 ) (454)
h̄
20 A. Proca, C.R. Acad. Sci., Paris 203, 709-711 (1936), J. de Physique et Radium, 8, 23-28
(1937).
75
which can be solved by Fourier transforming. On performing the Fourier trans-
formation, one finds the algebraic equation
2
mc
− kρ k ρ + g µ,ν + k µ k ν Dν,λ (k) = δλµ (455)
h̄
which can be solved and has the solution
− g kν k λ
ν,λ + ( m c )2
h̄
Dν,λ (k) = (456)
k ρ kρ − ( mh̄ c )2
kν j ν (k) = 0 (460)
76
one has
d3 k ρ(k)∗ ρ(k)
Z
4π
Lint = (463)
c2 ( 2 π )3 − k 2 − ( mh̄ c )2
Thus, the interaction Lagrangian between two like charges is negative and this
corresponds to a positive (replusive) interaction potential V .
k ν Aν (k) = 0 (464)
The numerator originates from the product of polarization vectors α (k) for the
massive spin-one particle. The polarization vectors can can be chosen as
(1) = (0, 1, 0, 0)
(2) = (0, 0, 1, 0)
(3) = (0, 0, 0, 1) (466)
k ν α
ν (k) = 0 (470)
77
the propagator must satisfy the condition
A kλ + B k ν kν kλ = 0 (473)
or 2
mc
A + B = 0 (474)
h̄
Therefore, the numerator of the propagator can be expressed as
kν kλ
A gν,λ − m c 2 (475)
( h̄ )
78
where the Ricci scalar R is defined as
The Ricci tensor is obtained from the Riemann tensor via the contraction
λ
Rµ,ν = Rµ,λ,ν (481)
and the Riemann tensor is given in terms of the Christoffel symbols via
ρ
Rσ,µ,ν = ∂µ Γρν,σ − ∂ν Γρµ,σ + Γρµ,λ Γλν,σ − Γρν,λ Γλµ,σ (482)
where, in turn, the Christoffel symbols Γσρ,µ are given in terms of the metric
tensors by
1
Γσρ,µ = g σ,ρ ∂µ gν,ρ + ∂ν gρ,µ − ∂ρ gµ,ν (483)
2
In the absence of a source, the Lagrangian density of the source-free gravitational
field is given by √
L = −g R (484)
where g = det gµ,ν is negative.
xµ → x0µ = xµ − Λµ (487)
0 ∂xσ ∂xτ
gµ,ν → gµ,ν = gσ,τ (488)
∂x0 µ ∂x0 ν
so the linearized metric transforms as
79
Thus, both metrics describe the same physical gravitational field and so gravi-
tational theory has a gauge invariance. The gravitational gauge transformation
is very similar to the electromagnetic gauge transformation
Aµ → A0µ = Aµ + ∂µ Λ (490)
We can use this similarity to motivate the expectation that a graviton has S = 2.
On introducing a mass for the photon, we found that charge is not automatically
conserved. Furthermore, we found that if one enforces charge conservation, then
the electromagnetic field must satisfy the gauge condition
∂ µ Aµ = 0 (491)
In the Lorentz gauge, the linearized Einstein equations reduce to the form
ρ 1 ρ σ,τ 16 π G
∂ρ ∂ hµ,ν − ηµ,ν ∂ρ ∂ η hσ,τ = − Tµ,ν (494)
2 c3
where d = 4. This equation can be used to eliminate the second term in the
linearized equation of motion, leading to
16 π G 1
∂ρ ∂ ρ hµ,ν = − T µ,ν − ηµ,ν η σ,τ
T σ,τ (496)
c3 2
The second term in the source projects out unwanted components of the energy-
momentum tensor. For the vacuum, where Tµ,ν = 0, the linearized equations
take the form
∂ρ ∂ ρ hµ,ν = 0 (497)
80
which is just the relativistic wave equation for a massless spin-two particle. For
a graviton of mass m, the wave equation could be expected to have the form
2
ρ mc
∂ ∂ρ + hµ,ν = 0 (498)
h̄
up to a constant factor. The form of the mass term is not enforced by any known
symmetry. In fact, like the Proca Lagrangian for the massive photon, the mass
term in the Fierz-Pauli Lagrangian density violates the gauge symmetry. The
principle of extremal action leads to the equations of motion
On operating on the equation with ∂µ and on noting that only the mass term
remains, one finds the equation
81
existence of the mass. Substituting this mandatory condition back into the
equations of motion yields
2
ρ σ,τ mc σ,τ
∂ ∂ρ hµ,ν − ∂µ ∂ν η hσ,τ + hµ,ν − ηµ,ν η hσ,τ = 0 (503)
h̄
On taking the trace of the above equation, if m 6= 0, one recovers the condition
that hµ,ν must be traceless
η µ,ν hµ,ν = 0 (504)
as anticipated. Furthermore, this also implies the four mandatory “gauge con-
ditions”
∂ µ hµ,ν = 0 (505)
must be satisfied. On applying the above five constraints to the equation of
motion, one finds
2
mc
∂ ρ ∂ρ + hµ,ν = 0 (506)
h̄
Thus, the Fierz-Pauli Lagrangian for hµ,ν does indeed describe a massive particle
with S = 2, and we have confirmed the form of the five conditions on the field.
In the presence of a source, the equation could be expected to have the form
2
mc 16 π G
∂ ρ ∂ρ + hµ,ν = − T̃µ,ν (507)
h̄ c3
The above analysis implies that the Fourier transformed propagator for a
graviton with mass m must have the form
Bµ,ν;σ,τ (k)
Dµ,ν;σ,τ (k) = (510)
k ρ kρ − ( mh̄ c )2
82
The graviton’s field is normalized according to
X
α α ˆ
µ,ν (k) µ,ν (k) = I (511)
α
∂µ T µ,ν = 0 (515)
The above condition eliminates the the k-dependent terms in the propagator
when coupled to the energy-momentum tensor, so one can replace the Gµ,ν (k)
by − ηµ,ν . It should be noted that the energy density is positive
d3 k 1 + 1 − 23
Z
8πG 0,0∗
Lint = − T (0, k) T 0,0 (0, k)
c2 ( 2 π )3 − k 2 − ( mh̄ c )2 + i η
(518)
so for static mass densities, the interaction Lagrangian is positive. Thus, the
gravitational interaction between two masses is attractive. However, after we
83
have introduced the factor of − 12 to obtain the static interaction potential, we
that the result differs from the result expected from Newton’s law of gravity
by a factor of 43 . That is, one expects that the factor of 32 in the last term of
the propagator should be replaced by unity. The propagator corresponding to
the linearized theory of gravity (for which m = 0) does have the 23 replaced
by 1. Due to this discrepancy, we shall examine the connection between the
propagator and the source in more detail below.
c4
L = + ∂λ hµ,ν ∂ λ hµ,ν − ∂λ ηµ,ν hµ,ν ∂ λ ησ,τ hσ,τ
32 π G
− 2 ∂µ hν,λ ∂ ν hµ,λ + 2 ∂µ hµ,ν ∂ν η σ,τ hσ,τ
2
mc
− hµ,ν hµ,ν − η µ,ν hµ,ν η σ,τ hσ,τ
h̄
− c hµ,ν T µ,ν (519)
On operating on the equation with ∂µ and on noting that only the mass term
remains, one finds the equation
2
mc µ σ,τ 16 π G
∂ hµ,ν − ∂ν η hσ,τ = − ∂µ T µ,ν (521)
h̄ c3
We see that in the absence of a graviton mass, the Lagrangian theory of the
graviton requires that the energy-momentum tensor must be conserved. For the
massive graviton, the conservation of energy and momentum is not automat-
ically ensured, but is an independent assumption. We shall assume that the
energy-momentum tensor remains a conserved quantity.
84
into the equation of motion, yielding
2
mc 16 π G
∂ ρ ∂ρ hµ,ν − ∂µ ∂ν η σ,τ hσ,τ + hµ,ν − ηµ,ν η σ,τ hσ,τ = − Tµ,ν
h̄ c3
(523)
Taking the trace, we find that
2
mc 16 π G
( d − 1 ) η σ,τ hσ,τ = η σ,τ Tσ,τ (524)
h̄ c3
85
Since hµ,ν is symmetric in the indices, one should also symmetrize the propa-
gator. The symmetrized propagator is given by
8 π G
c3
2
hµ,ν (k) = ρ G µ,σ (k) G ν,τ (k) + G µ,τ (k) G ν,σ (k) − G µ,ν (k) G σ,τ (k) T σ,τ (k)
k kρ − ( mh̄ c )2 d − 1
(531)
2
in agreement with the result of our previous calculation. The factor of d−1 is a
2
real discrepancy with the m = 0 result which contains d−2 . This discrepancy
is due to the fact that in the limit m → 0, the 5 excitations of the Fierz-Pauli
theory decouple into helicity ± 2 excitations, helicity ± 1 vector excitations and
a scalar particle. The vector particles do not mediate an interaction, but the
scalar particle does. The scalar particle is the so-called longitudinal graviton.
However, the theory with m exactly equal to zero only has two helicity ± 2
excitations and does not include the longitudinal graviton. The discontinuity
between the m = 0 theory and the limit m → 0 was investigated by Vainshtein22 .
86
Hence, the energy of the field can be expressed as
2
c3
Z
3 µ,ν µ,ν mc µ,ν
P0 = d r ∂0 hµ,ν ∂0 h + ∇ hµ,ν . ∇ h + hµ,ν h
32 π G h̄
(535)
or, on Fourier transforming,
2
c3 π 2
Z
mc
P0 = d3 k k02 + k 2 + hµ,ν (k) hµ,ν (k) (536)
4G h̄
and
X ki kj
h0,0 (k) = hi,j (k) (540)
i,j
k0 k0
87
For k directed along the z-direction, one has
2 2 2
µ,ν k 2 k 2 2
hµ,ν (k) h (k) = −1 |h3,3 (k)| − 2 −1 |h3,1 (k)| + |h3,2 (k)|
k02 k02
+ 2 |h1,2 (k)|2 + |h1,1 (k)|2 + |h2,2 (k)|2
2
3 k2
2
2 k 2 2
= − 1 |h 3,3 (k)| − 2 − 1 |h 3,1 (k)| + |h 3,2 (k)|
2 k02 k02
1
+ 2 |h1,2 (k)|2 + | h1,1 (k) − h2,2 (k) |2 (544)
2
Thus, the massive graviton field has five independent components.
under a rotation of ϕ around the z-axis. Thus, these combinations have angular
momenta of ± 1 around the z-axis. The physical effects of the modes h3,1 (k)
and h3,2 (k) vanish when m → 0.
The two modes h1,1 (k) − h2,2 (k) and h1,2 (k) carry energy in the limit of zero
mass, have the helicity S z = ± 2 and correspond to the graviton. That is, the
linear combination
h1,1 (k) − h22 (k)
± i h1,2 (k) (549)
2
transform as
h1,1 (k) − h22 (k)
exp[ ∓ i 2 ϕ ] ± i h1,2 (k) (550)
2
under a rotation of ϕ around the z-axis and, thus, correspond to a helicity
S z = ± 2.
88
The remaining mode h3,3 (k) is the longitudinal graviton which corresponds
to S z = 0. The amplitude of the longitudinal graviton is expected to diverge
in the limit m → 0. This can be seen by examining the equation for the source
of the gravitational radiation
ρ mc 2 16 π G 1 kµ kν
k kρ − ( ) hµ,ν (k) = ηµ,σ ην,τ − ηµ,ν − m c 2 ησ,τ T σ,τ (k)
h̄ c3 d − 1 ( h̄ )
(551)
and noting that, for k along the z-direction, the term in the source which di-
verges when m → 0 only couples to h3,3 (k). Thus, we have
1
h3,3 (k) ∼ (552)
( mh̄ c )2
and, therefore, this mode yields a finite contribution to the energy. It is this
longitudinal mode which gives rise to the discontinuity at m = 0.
In the absence of sources, the (classical) wave equation for the vector poten-
tial has the form
1 ∂2
2
− ∇ + 2 A = 0 (553)
c ∂t2
when the Coulomb gauge condition is imposed
∇.A = 0 (554)
In this paper Dirac uses two different approaches to quantizing electromagnetism. In one
approach he treated a single photon as satisfying a single-particle Schrödinger equation, that
has a similar form to Maxwell’s equations. The other approach treated the fields as dynamical
variables and then quantized them. Dirac then showed that these two methods produce
equivalent results. By doing this, Dirac created second quantization.
89
where V is the volume of the system. The inverse Fourier Transform is given
by
1 X
A(r, t) = √ exp − i k . r A(k, t) (556)
V k
On Fourier transforming the wave equation with respect to space and time, one
finds the equation of motion
2
2 1 ∂
k + 2 A(k, t) = 0 (557)
c ∂t2
k . A(k, t) = 0 (558)
We shall look for solutions for A(k, t) that have a time dependence given by
linear superpositions of the terms proportional to
exp ∓ i ωk t (559)
By substituting the above terms into the wave equation, it is found that linear
superpositions of plane-waves are solutions of Maxwell’s equation but only if
the frequency ωk and wave vector k are related via the dispersion relation
ωk2 = c2 k 2 (560)
The gauge condition also requires that the vector potential is oriented perpen-
dicular to the direction of propagation. Therefore, an arbitrary plane-wave
solution can be represented as a linear superposition of two polarized waves
with polarizations described by two mutually orthogonal unit vectors denoted
by ˆα (k). The polarization vectors satisfy
k . ˆα (k) = 0
ˆα (k) . ˆβ (k) = δα,β (561)
We shall assume that three vectors k, ˆ1 (k), ˆ2 (k) form a mutually orthogonal
coordinate system. We shall define
The algebraic equations for A(k) can be solved trivially. One can express the
vector potential as a linear superposition
1 X
A(r, t) = √ ˆα (k) exp − i k . r Φα (k, t) (563)
V k,α
90
E
B
k
Figure 9: The normal modes of the classical electromagnetic field are plane-
polarized waves, in which E and B are transverse to the direction of propagation
k, and oscillate in phase.
91
The Lagrangian is given by the space integral of the Lagrangian density
Z
L = d3 r L (569)
On substituting A(r, t) in the form of eqn(563) and integrating over r and using
the identity Z
1 3 0
d r exp i ( k + k ) . r = δk+k0 (570)
V
one finds the Lagrangian is given by
1 X X
L = δk+k0
8π
k,k0 α,β
∂Φβ (k 0 )
0 1 ∂Φα (k)
× ˆα (k) . ˆβ (k ) 2
c ∂t ∂t
0 0 0
+ ( k ∧ α (k) ) . ( k ∧ β (k ) ) Φα (k) Φβ (k )
1 ∂Φ∗α (k)
1 X ∂Φα (k) 2 ∗
= − k Φα (k) Φα (k)
8π c2 ∂t ∂t
k,α
(571)
In the above expression, the summation over k is unrestricted. If the Lagrangian
is to be expressed in terms of the independent components, then the summation
over k must be restricted to half the set of allowed values. With this restriction,
-k
Figure 10: A possible partition of k-space, which does not contain both k and
its inverse −k.
one obtains
∗
2 X0 1 ∂Φα (k) ∂Φα (k) 2 ∗
L = − k Φ α (k) Φ α (k)
8π c2 ∂t ∂t
k,α
(572)
92
where the prime over the summation denotes the restriction of k to values
in the “positive” half volume of k-space. Since there are half the number of
independent normal modes, their contributions are twice as big. The Lagrangian
is a function of the six generalized variables Φα (k) and Φ∗α (k) for the independent
k values. The generalized momenta variables are found as
∗
2 ∂Φα (k)
Πα (k) =
8 π c2 ∂t
∗ 2 ∂Φ α (k)
Πα (k) = (573)
8 π c2 ∂t
k2
∂ 1 ∂Φα (k)
= − Φα (k) (574)
∂t 8 π c2 ∂t 8π
or
∂ 2 Φα (k)
= − ωk2 Φα (k) (575)
∂t2
where ωk = c k. Thus, the classical field Φα (k) has a time-dependent ampli-
tude which resembles that of a harmonic oscillator with frequency ωk = c k.
The Hamiltonian can be obtained from the Lagrangian, via the Legendre Trans-
formation
X ∂Φ∗α (k)
0 ∂Φα (k)
H = Π∗α (k) + Πα (k) − L (576)
∂t ∂t
k,α
where the summation over (k, α) runs over the independent normal modes.
Hence, the k summation only runs over the set of points in k space which are
not related via the inversion operator. The Hamiltonian is related to the energy
of the electromagnetic field, as shall be seen below.
in the Coulomb gauge. The energy density can be written in terms of the vector
potential as
2 2
1 1 ∂A
H = + ∇ ∧ A (579)
8 π c2 ∂t
93
The energy is the integral of the energy density over all space
Z
H = d3 r H (580)
in which the summation over k is unrestricted. Thus, the above expression for
the energy is identical to the Hamiltonian for the electromagnetic field. Fur-
thermore, the Hamiltonian has been expressed in terms of a set of the normal
modes labeled by (k, α).
94
known as creation operators. The commutation relations for the creation and
annihilation operators can be obtained directly from the commutation relations
of the field operators Φ̂α (k) and Π̂α (k) which are shown in eqn(584). It can
be shown that the creation and annihilation operators satisfy the commutation
relations
The field operators can be expressed in terms of the creation and annihilation
operators. Starting with
s s
8 π c2 2 k2
1 †
âk,α = √ i Π̂α (k) + Φ̂ (k) (589)
2 2 h̄ ωk 8 π h̄ ωk α
transforming k → −k and then by noting that Π̂α (−k) = Π̂†α (k) and Φ̂†α (−k) =
Φ̂α (k), one finds
s s
8 π c2 † 2 k2
1
â−k,α = √ i Π̂α (k) + Φ̂α (k) (590)
2 2 h̄ ωk 8 π h̄ ωk
One can eliminate Π̂†α (k) by adding the expression for the creation operator
given by eqn(587) and the expression for the annihilation operator with mo-
mentum −k given by eqn(590). This process yields the expression for the field
component operators Φ̂α (k) in the form
r
2 π h̄ ωk †
Φ̂α (k) = âk,α + â−k,α (591)
k2
and, by an analogous procedure, the Hermitean conjugate operator is found to
be given by r
† 2 π h̄ ωk †
Φ̂α (k) = âk,α + â−k,α (592)
k2
which is identical to Φ̂α (−k). Likewise, the canonically conjugate momenta
operators are given by
r
h̄ ωk †
Π̂α (k) = i â−k,α − âk,α (593)
8 π c2
and their Hermitean conjugates are
r
h̄ ωk
Π̂†α (k) = − i â−k,α − â†k,α (594)
8 π c2
as was anticipated.
95
10.2.1 The Energy of the Field
The Hamiltonian of the electromagnetic field
X 8 π c2 1 2 †
Ĥ = Π̂α (k) Π̂†α (k) + k Φ̂α (k) Φ̂α (k) (595)
4 8π
k,α
(596)
If one sets k → −k in the second set of terms, then one finds the Hamiltonian
becomes the sum over independent harmonic oscillators for each k value and
polarization
X h̄ ωk † †
Ĥ = âk,α âk,α + âk,α âk,α (597)
2
k,α
and has integer eigenvalues denoted by nk,α . Hence, the energy eigenvalues E
are given by
X 1
E = h̄ ωk nk,α + (599)
2
k,α
It should be noted that the contributions to the total energy from the zero-
point energy terms h̄ω2 k diverge. However, in most circumstances, only the
excitation energy of the field is measurable, hence the divergence is mainly ir-
relevant. The zero-point energy does have physical consequences, and can be
observed if the volume or boundary conditions of the field are changed. The
change in the zero-point energy of the field due to change in volume or boundary
conditions is known as the Casimir effect24 .
96
10.2.2 The Electromagnetic Field
The quantized vector potential is given by the operator Â(r), given by
s
2 π h̄ c2
†
X
Â(r) = ˆα (k) âk,α + â−k,α exp − i k . r (600)
ωk V
k,α
(603)
The above equation was obtained by noting that, in the basis composed of
eigenstates of the number operators |nk,α >, one has
âk,α (t) |nk,α > = exp + iωk t (â†k,α âk,α + 1/2) âk,α (0) |nk,α > exp − iωk t (nk,α + 1/2)
√
= exp + iωk t (nk,α − 1/2) nk,α |nk,α − 1 > exp − iωk t (nk,α + 1/2)
= exp − i ωk t âk,α |nk,α > (604)
and that the time-dependent creation operator is given by the Hermitean con-
jugate expression. Thus, the explicit form of time dependence of the vector
potential is a consequence of the explicit time dependence of the creation and
annihilation operators in the Heisenberg representation. Alternatively, one can
find the time dependence of the creation and annihilation operators directly
from the Heisenberg equations of motion without invoking a privileged set of
basis states. The equation of motion for the creation operator is given by
∂â†k,α
i h̄ = [ â†k,α , Ĥ ] (605)
∂t
and the commutator is evaluated as
97
so the equation of motion simplifies to
∂â†k,α
i h̄ = − h̄ ωk â†k,α (607)
∂t
Therefore, one finds the result
â†k,α (t) = â†k,α exp i ωk t (608)
which is just the Hermitean conjugate of the â†k,α (t) that was found previously.
Therefore, the time-dependence of the vector potential is entirely due to the
time-dependence of the Heisenberg representation of the creation and annihila-
tion operators.
This will be evaluated by expressing the Ê and B̂ field operators in terms of the
vector potential A operator via
1 ∂ Â
Ê = −
c ∂t
B̂ = ∇ ∧ Â (614)
98
The vector potential operator can be written in terms of the creation and anni-
hilation operators for the normal modes as
s
2 π h̄ c2
†
X
Â(r, t) = ˆα (k) âk,α (t) + â−k,α (t) exp − i k . r (615)
ωk V
k,α
and
s
2 π h̄ c2
â†k,α + â−k,α
X
B̂(r) = − i ( k ∧ ˆα (k) ) exp − ik.r
ωk V
k,α
(617)
For a fixed k, the polarization vectors ˆα (k) and k are mutually orthogonal.
Therefore, one has
ˆα (k) ∧ ( k ∧ ˆβ (k) ) = k ( ˆα (k) . ˆβ (k) ) − ˆβ (k) ( k . ˆα (k) )
= k δα,β (618)
Hence, the total momentum of the electromagnetic field is determined from
h̄ X
P̂ = ˆα (k) ∧ ( k ∧ ˆα (k) ) â†k,α − â−k,α â†−k,α + âk,α
2
k,α
h̄
k â†k,α − â−k,α â†−k,α + âk,α
X
= (619)
2
k,α
It should be noted that the momentum from each normal mode of the field is
parallel to its direction of propagation. Since the creation operators commute
â†k,α â†−k,α = â†−k,α â†k,α (620)
and that the annihilation operators also commute
â−k,α âk,α = âk,α â−k,α (621)
one finds that the part of the momentum represented by the summation over k
given by
X † †
h̄ k âk,α â−k,α − â−k,α âk,α = 0 (622)
k,α
vanishes since the summand is odd under inversion symmetry. Thus, the mo-
mentum of the electromagnetic field is given by
h̄ X † †
P̂ = k âk,α âk,α − â−k,α â−k,α
2
k,α
1 X
= h̄ k â†k,α âk,α − h̄ k â†−k,α â−k,α − h̄ k (623)
2
k,α
99
where the commutation relations for the creation and annihilation operators
were used to obtain the last line. The last term vanishes when summed over k,
due to inversion symmetry. Hence, the momentum of the field is given by the
operator
1 X † †
P̂ = h̄ k âk,α âk,α − h̄ k â−k,α â−k,α (624)
2
k,α
Finally, on transforming −k to k in the last term of the summand, one finds the
total momentum of the field is carried by the excitations since
h̄ k â†k,α âk,α
X
P̂ = (625)
k,α
100
However, due to the identity
ξ k,l,m ξ m,n,p = δ k,n δ l,p − δ k,p δ l,n (629)
one finds
∂ Â(l) (k)
Z
(i) 1 (l) ∂ Â
JˆEM = d3 r ξ i,j,k x(j) Ê (l) − x(j)
Ê (630)
4πc ∂x(k) ∂x(l)
On integrating by parts in the last term, one has
(l)
Z
ˆ(i) 1 3 i,j,k (j) (l) ∂ Â ∂ (j) (l) (k)
JEM = d rξ x Ê + x Ê Â
4πc ∂x(k) ∂x(l)
(631)
Since the divergence of the electric field vanishes26 ,
∂ Ê (l)
= 0 (634)
∂x(l)
and since
∂x(j)
= δ j,l (635)
∂x(l)
the total angular momentum can be re-written as
∂ Â(l)
Z
(i) 1
JˆEM = d3 r ξ i,j,k Ê (l) x(j) + Ê (j)
Â(k)
(636)
4πc ∂x(k)
The first term can be recognized as the orbital angular momentum of the field.
The orbital angular momentum operator L̂(i) is given by
∂
L̂(i) = − i h̄ ξ i,j,k x(j) (637)
∂x(k)
so the total angular momentum of the field is given by
Z
ˆ (i) i 3 (l) (i) (l) i,j,k (j) (k)
JEM = d r Ê L̂ Â − i h̄ ξ Ê Â
4 π h̄ c
Z
i
= d3 r Ê (l) L̂(i) Â(l) + Ê (j) ( Ŝ (i) )j,k Â(k) (638)
4 π h̄ c
26 In the presence of a charge density q |ψ(r)|2 , the angular momentum of the EM field will
101
where the definition
( Ŝ (i) )j,k = − i h̄ ξ i,j,k (639)
for Ŝ, the intrinsic spin operator for the photon, has been used in obtaining the
second line. The total vector angular momentum operator can be expressed as
Z
i 3 (j) j,k j,k
Ĵ EM = d r Ê L̂ δ + ( Ŝ ) Â(k) (640)
4 π h̄ c
which shows that the orbital angular momentum is diagonal with respect to
the field components and the spin angular momentum mixes the different field
components.
The total spin component of the angular momentum operator for the elec-
tromagnetic field is given by
Z
(i) i 3 (j) (i) j,k (k)
ŜEM = d r Ê ( Ŝ ) Â
4 π h̄ c
Z
1
= d3 r ξ i,j,k Ê (j) Â(k)
4πc
Z (i)
1
= d3 r Ê ∧ Â (641)
4πc
This can be expressed in terms of the photon creation and annihilation operators
as
(i) h̄ X (j) i,j,k (k)
ŜEM = − i ˆβ (k) ξ ˆα (k)
2
k,α,β
† †
× â−k,β − âk,β âk,α + â−k,α (642)
The first term in parenthesis is recognized as the i-th component of the vector
product
ˆβ (k) ∧ ˆα (k) (643)
and, therefore, it is antisymmetric in the polarization indices α and β and the
non-zero contributions are restricted to the case α 6= β. Since the creation
and annihilation operators corresponding to different polarizations commute,
the product of the two remaining parenthesis can be re-arranged as the sum of
two terms
(i)
(i) h̄ X
ŜEM = − i ˆβ (k) ∧ ˆα (k)
2
k,α,β
† †
× â−k,β âk,α − âk,β â−k,α
† †
+ â−k,β â−k,α − âk,α âk,β (644)
102
On transforming the summation variable k → −k and commuting the operators,
one finds that the first term in the square brackets is symmetric under the
interchange of α and β whereas the second term is antisymmetric. Hence, on
summing over the polarization indices, the contribution from the first term
vanishes, as it is the product of a symmetric and the antisymmetric (vector
product) term. Therefore, the total spin operator of the electromagnetic field is
expressed as
(i)
(i) h̄ X † †
ŜEM = i ˆβ (k) ∧ ˆα (k) âk,α âk,β − â−k,β â−k,α
2
k,α,β
(645)
On defining the sense of the polarization vectors relative to k̂ (≡ ê3 (k) the unit
vector in the direction of propagation) via
ˆ1 (k) ∧ ˆ2 (k) = k̂ (646)
On setting −k → k in the second part of the summation, the spin of the elec-
tromagnetic field is found as
X †
Ŝ EM = i h̄ k̂ âk,2 âk,1 − â†k,1 âk,2 (648)
k
It should be noted that in this expression, the indices (1, 2) refer to directions
in three-dimensional space and do not refer to the z-component of the spin an-
gular momentum. Therefore, the above equation shows that a plane-polarized
photon is not an eigenstate of the single-particle spin operator quantized along
the k-axis27 .
103
e (k), where
are given by Φ m
1
1
e (k)
Φ +1 = − √ i
2 0
0
e (k)
Φ 0 = 0
1
1
1
e (k)
Φ −1 = √ −i (650)
2 0
and where the subscript m refers to the eigenvalue of Ŝ (3) , in units of h̄. From
this, it follows that an arbitrary transverse vector wave function Φ(k) can only
be expressed as a linear superposition of states involving m = ±1, and that the
m = 0 component is absent. On expressing an arbitrary (non-transverse) vector
wave function Φ(k) with components Φ(1) (k), Φ(2) (k) and Φ(3) (k) in terms of
its components referred to the helicity eigenstates Φm (k) one has
(1)
Φ+1 (k) − 1 i √0 Φ (k)
1
Φ0 (k) = √ 0 0 2 Φ(2) (k) (651)
Φ−1 (k) 2 1 i 0 Φ(3) (k)
This relation between the two bases can be expressed in the alternate form
m=1
X
Φ(k) = êm Φm (k) (652)
m=−1
The above relations allow one to define the circularly-polarized creation and
annihilation operators via their relation to the quantum fields. This procedure
yields
i=3
X m=1
X
ˆi (k) âk,i = êm (k) âk,m (655)
i=1 m=−1
104
Hence, the photon annihilation operators corresponding to a definite helicity
are related to the annihilation operators for plane-polarized photons via
1
âk,m=+1 = − √ ( âk,1 − i âk,2 )
2
âk,m=0 = âk,3
1
âk,m=−1 = √ ( âk,1 + i âk,2 ) (656)
2
and the inverse relations are given by
1
âk,1 = − √ ( âk,m=1 − âk,m=−1 )
2
i
âk,2 = − √ ( âk,m=1 + âk,m=−1 )
2
âk,3 = âk,m=0 (657)
When expressed in terms of the circularly-polarized unit vectors, the spin oper-
ator for the electromagnetic field becomes
X † †
Ŝ EM = h̄ k̂ âk,m=1 âk,m=1 − âk,m=−1 âk,m=−1 (658)
k
which is expressed in terms of photons with definite helicity. Within the man-
ifold of single-photon states with momentum h̄ k, the spin operator has eigen-
values of ±h̄ when measured along the direction k̂. It is seen that the photon
has helicity m = ±1 but does not involve the helicity state with m = 0 since
the electromagnetic field is transverse. The transverse nature of the field is due
to the photon being massless. In general, a massive particle with spin S should
have (2S + 1) helicity states. However, a massless particle can only have the
two helicity states corresponding to m = ±S.
105
e1
e2 k
→ ∞ (662)
The fluctuations in the field diverge because the zero-point energy fluctuations
diverge.
106
The commutation relations between the x-component of the E field and the
B field at the same instant of time are non-zero30 . That is,
2π X
[ Êx (r) , B̂y (r0 ) ] = h̄ ωk ˆα (k)x ( k̂ ∧ ˆα (k) )y exp i k . ( r0 − r )
V
k,α
2π X
− h̄ ωk ˆα (k)x ( k̂ ∧ ˆα (k) )y exp − i k . ( r0 − r )
V
k,α
4 π h̄ c X 0
= − kz exp − i k . ( r − r )
V
k
4 π c h̄ ∂ X
= i exp − i k . ( r0 − r )
V ∂z
k
c h̄ ∂ 3 0
= i δ (r − r) (663)
2 π 2 ∂z
The fact that the two polarizations are transverse to the unit vector k̂ has been
used to obtain the third line. Since Ê and B̂ do not commute, it follows that E
and B obey an uncertainty relation in that the values of E and B cannot both
be specified to arbitrary accuracy at the same point.
However, if two points in space time x and x0 are not causally related, i.e.
| r0 − r | =
6 c | t0 − t | (664)
Thus, if the two points in space-time are not connected by the propagation of
light, then the Ex and By fields can both be determined to arbitrary accuracy.
For example, the vacuum state or ground state is an eigenstate of the annihila-
tion operator, in which case aϕ = 0.
107
The coherent state31 can be found as a linear superposition of eigenstates of
the number operator with eigenvalues n
∞
X
| aϕ > = Cn | n > (667)
n=0
On taking the matrix elements of this equation with the state < m |, and using
the orthonormality of the eigenstates of the number operator, one finds
√
Cm+1 m + 1 = aϕ Cm (670)
108
0.2
0.15
Pn
0.1
0.05
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Figure 12: The probability of finding n photons P (n) in a normal mode repre-
sented by a coherent state.
From this, it can be shown that if the number of photons in a coherent state
are measured, the result n will occur with a probability given by
( a∗ϕ aϕ )n
∗
P (n) = exp − aϕ aϕ (676)
n!
Thus, the photon statistics are governed by a Poisson distribution. Furthermore,
the quantity a∗ϕ aϕ is the average number of photons n present in the coherent
state.
The coherent states can be written in a more compact form. Since the state
with occupation number n can be written as
( ↠)n
|n > = √ |0 > (677)
n!
the coherent state can also be expressed as
∞
( aϕ ↠)n
X
1 ∗
| aϕ > = exp − aϕ aϕ |0 > (678)
2 n=0
n!
109
The above equation represents a transformation between number operator
states and the coherent states. The inverse transformation can be found by
expressing aϕ as a magnitude a and a phase ϕ
aϕ = a exp i ϕ (680)
The number states can be expressed in terms of the coherent states via the
inverse transformation
√ Z 2π
n! 1 2 dϕ
|n > = exp + a exp − i n ϕ | aϕ >
an 2 0 2π
(681)
by integrating over the phase ϕ of the coherent state. Since the set of occupa-
tion number states is complete, the set of coherent states must also span Hilbert
space. In fact, the set of coherent states is over-complete.
The coherent state | aϕ > can be represented by the point aϕ in the Argand
plane. The overlap matrix elements between two coherent states is calculated
1
Im z
aφ
0.5
a
φ Re z
0
-1 -0.5 0 0.5 1
-0.5
-1
as
| < a0ϕ0 | aϕ > | 2
= exp − | aϕ − a0ϕ0 2
| (682)
Hence, coherent states corresponding to different points are not orthogonal. The
coherent states form an over complete basis set. The over completeness relation
can be expressed as
d <e aϕ d =m aϕ
Z
| aϕ > < aϕ | = Iˆ (683)
π
110
This relation can be proved by taking the matrix elements between the occupa-
tion number states < n0 | and | n >, which leads to
d <e aϕ d =m aϕ
Z
< n0 | aϕ > < aϕ | n > = δn0 ,n (684)
π
which can be evaluated as
d <e aϕ d =m aϕ
Z
= < n0 | aϕ > < aϕ | n >
π
0
dϕ a∗ϕ n anϕ
Z ∞ Z 2π
= da a √ exp − | aϕ |2
0 0 π n0 ! n!
Z ∞ n+n0
Z 2π
a 2 dϕ 0
= da a √ exp − | aϕ | exp i ( n − n ) ϕ
0 n0 ! n! 0 π
Z ∞ 0
an+n
= da a √ exp − a2 2 δn,n0 (685)
0 n0 ! n!
On changing variable to s = a2 , one proves the completeness relation by noting
that Z ∞
n
ds s exp − s = n! (686)
0
Hence, the coherent states form a complete basis set.
The effect of the creation operator on the coherent state can be expressed
as
† † 1 ∗ †
â | aϕ > = â exp − a aϕ exp aϕ â |0 >
2 ϕ
1 ∗
= exp − a aϕ ↠exp aϕ ↠| 0 >
2 ϕ
1 ∗ ∂ †
= exp − a aϕ exp aϕ â |0 >
2 ϕ ∂aϕ
1 ∗ ∂ 1 ∗
= exp − a aϕ exp + a aϕ | aϕ >
2 ϕ ∂aϕ 2 ϕ
(687)
The coherent state is not an eigenstate of the creation operator, since the re-
sulting state does not include the zero-photon state.
The expectation value of the field operators between the coherent states
yields the classical value, since
< aϕ | ( ↠+ â ) | aϕ > = ( a∗ϕ + aϕ ) (688)
In deriving the above equation, the definition
â | aϕ > = aϕ | aϕ > (689)
111
has been used in the term involving the annihilation operator and the term orig-
inating from the creation operator is evaluated using the Hermitean conjugate
equation
< aϕ | ↠= < aϕ | a∗ϕ (690)
One also finds that that the expectation value of the number operator is given
by
< aϕ | ↠â | aϕ > = a∗ϕ aϕ (691)
so the magnitude of aϕ is related to the average number of photons in the
coherent state n. This identification is consistent with the Poisson distribution
of eqn(676) which governs the probability of finding n photons in the coherent
state. The coherent state is not an eigenstate of the number operator since
there are fluctuations in any measurement of the number of photons. The rms
fluctuation ∆n can be evaluated by noting that
< aϕ | n̂2 | aϕ > = < aϕ | ↠â ↠â | aϕ >
= < aϕ | ↠↠â â | aϕ > + < aϕ | ↠â | aϕ >
= ( a∗ϕ )2 ( aϕ )2 + a∗ϕ aϕ (692)
where the boson commutation relations have been used in the second line. Thus,
the mean squared fluctuation in the number operator is given by
< aϕ | ∆n̂2 | aϕ > = a∗ϕ aϕ (693)
The rms fluctuation of the photon number is only negligible when compared to
the average value if aϕ has a large magnitude
a∗ϕ aϕ 1 (694)
Exercise:
Determine the expectation values for the electric and magnetic field opera-
tors in a coherent state which represents a plane-polarized electromagnetic wave.
Exercise:
Determine the expectation values for the electric and magnetic field opera-
tors in a coherent state which represents a left circularly-polarized electromag-
netic wave composed of photons with a helicity of +1.
112
10.4.1 The Phase-Number Uncertainty Relation
From the discussion of coherent states, it is seen that the coherent state has a
definite phase, but does not have a definite number of quanta. In general, it is
impossible to know both the phase of a state and the number of a state. This
is formalized as a phase - number uncertainty relation.
and the Hermitean conjugate operator, the creation operator can be expressed
as
√
â†k,α = nˆk,α exp − i (ϕ̂k,α − ωk t) (697)
√
since it has been required that ˆn and ϕ̂ are Hermitean. Furthermore, the
√
operator ˆn must have the property
√ˆ √ˆ
nk,α nk,α = n̂k,α (698)
√ √ˆ
− nˆk0 ,β exp − i (ϕ̂k0 ,β − ωk0 t) exp + i (ϕ̂k,α − ωk t) nk,α
(700)
113
This relationship is satisfied, if the phase and number operators satisfy the
commutation relation
[ n̂k,α , ϕk,α ] = i (702)
If one can construct the Hermitean operators that satisfy this commutation
relation, then one can show that the rms uncertainties phase and number must
satisfy the inequality
It should be noted that only the relative phase can be measured33 . Thus, if the
phase difference of any two components (k, α) and (k 0 , α0 ) is specified precisely,
then the occupation number of either component can not be specified.
Exercise:
Express the vector potential and the electric and magnetic field operators in
terms of the amplitude and phase operators.
Hence, coherent states are not orthogonal. In fact, their overlap decreases expo-
nentially with large “separations” between the points aϕ and a0ϕ0 in the Argand
plane. We shall denote | aϕ | by a. Two states separated by distances a ∆ϕ or
∆a such that a ∆ϕ ≥ 1 and ∆a ≥ 1 are effectively orthogonal or independent.
However, states within an area given by ∆a × a ∆ϕ ≈ 1 have significant overlap
and so can represent the same state. Therefore, the minimum uncertainty state
occupies an area ∆a × a ∆ϕ ≈ 1. We note that 2 a ∆a can be interpreted
as a measure of the uncertainty ∆nϕ in the particle number for the state, and
∆ϕ is the uncertainty in the phase of the state. Hence, the phase - number
uncertainty relation sets the area of the Argand diagram that can be associated
with a single state as
a ∆a ∆ϕ ∼ 1 (705)
114
Im z
a ∆φ
∆a
∆φ
Re z
Figure 14: Due to the phase-number uncertainty principle, the minimum area
of the Argand diagram needed to represent a minimum uncertainty state has
dimensions such that a ∆a ∆ϕ ∼ 1.
p̂2 q2
q 2
Ĥ = + q φ(r) − p̂ . Â(r) + Â(r) . p̂ + Â (r)
2m 2mc 2 m c2
2 0 2
Ê (r ) + B̂ (r0 )
Z
+ d3 r 0 (706)
8π
when the vector potential is chosen to satisfy the Coulomb gauge. The second,
third and fourth terms are to be evaluated at the location of the charged point
particle, r, and the last term is evaluated at all points in space. The Hamiltonian
can be expressed as
Ĥ = Ĥ0 + Ĥrad + Ĥint (707)
where Ĥ0 is the Hamiltonian for the charged particle in the electrostatic poten-
tial φ
p̂2
Ĥ0 = + q φ(r) (708)
2m
and Ĥrad is the Hamiltonian for the electromagnetic radiation and Hint is the
interaction
q2
q 2
Ĥint = − p̂ . Â + Â . p̂ + Â (709)
2mc 2 m c2
115
field is quantized, the radiation Hamiltonian has the form
X h̄ ωk †
Ĥrad = âk,α âk,α + âk,α â†k,α (710)
2
k,α
(k,α) (k,α)
p p
p'
p'
For charged particles with spin one-half, analysis of the non-relativistic Pauli
equation shows that there is another interaction term involving the particles’
spins. This interaction can be described by the anomalous Zeeman interaction
q h̄
ĤZeeman = − σ.B (714)
2mc
where
B = ∇ ∧ A(r) (715)
116
and σi are the three Pauli matrices.
Generally, the paramagnetic interaction has a greater strength than the Zee-
man interaction. This can be seen by examining the magnitudes of the interac-
tions. The paramagnetic interaction has a magnitude given by
e
p.A (716)
mc
and for an atom of size a , the uncertainty principle yields
h̄
p ∼ (717)
a
The Zeeman interaction has a magnitude given by
e h̄
σ.(k ∧ A) (718)
mc
but since k is the wavelength of light
1
k ∼ (719)
λ
Hence, since the wave length of light is larger than the linear dimension of an
atom, λ > a, one finds the inequality between the magnitude of the paramag-
netic interaction and the Zeeman interaction
e h̄ 1 e h̄ 1
A > A (720)
mc a mc λ
Both the paramagnetic and Zeeman coupling strengths are proportional to the
magnitude of the vector potential A, hence the ratio of the strengths of the
interactions are independent of A. Therefore, there magnitudes satisfy the in-
equality
1 1
> (721)
a λ
so the Zeeman interaction can frequently be neglected in comparison with the
paramagnetic interaction.
117
11.1.1 The Emission of Radiation
We shall consider a state | (nlm) {nk0 ,β } > which is an energy eigenstate
of the unperturbed Hamiltonian Ĥ0 and the radiation Hamiltonian Ĥrad . The
interaction Ĥint causes the system to make a transition from the initial state to
a final state. In the initial state, the electron is in an energy state designated
by the quantum numbers (n, l, m) and the electromagnetic field is in a state
specified by the number of photons in each normal mode. That is, the photon
field is in an initial state which is specified by the set of photon quantum num-
bers, {nk0 ,β 0 }. We shall consider the transition in which the electron makes a
transition from the initial state to a final state denoted by (n0 , l0 , m0 ). Since the
(hω'/c,hk')
0
Enlm
-4
E
-8
e-
-12
En'l'm'
-16
Figure 16: An electron in the initial atomic state with energy En,l,m makes a
transition to the final atomic state with energy En0 ,l0 ,m0 , by emitting a photon
with energy h̄ωk0 .
photon is emitted, the final state of the photon field described by the set {n0k0 ,β }
where
n0k0 ,β = nk0 ,β for (k 0 , β) 6= (k, α) (722)
and the number of photons in a normal mode (k, α) is increased by one
n0k,α = nk,α + 1 (723)
The transition rate for the electron to make a transition from (n, l, m) to
(n0 , l0 , m0 ) can be calculated34 from the Fermi-Golden rule expression
1 2π X
= | < n0 l0 m0 {n0k0 ,β } | Ĥint | nlm {nk0 ,β } > |2 δ( Enlm − En0 l0 m0 − h̄ ωk,α )
τ h̄
k,α
(724)
34 P. A. M. Dirac, Proc. Roy. Soc. A 112, 661 (1926), A 114, 243 (1927).
118
The delta function expresses the conservation of energy. The energy of the
initial state is given by
X 1
Enlm + h̄ ωk0 ,β ( nk0 ,β + ) (725)
0
2
k ,β
The difference in the energy of the initial state and final state is evaluated as
Enlm − En0 l0 m0 − h̄ ωk,α (727)
which is the argument of the delta function and must vanish if energy is con-
served. The sum over k can be evaluated by assuming that the radiation field is
confined to a volume V . The allowed k values for the normal modes are deter-
mined by the boundary conditions. In this case, the sum over k is transformed
to an integral over k-space via
Z
X V
→ d3 k (728)
( 2 π )3
k
(729)
since only the paramagnetic part of the interaction has non-zero matrix ele-
ments. For the photon emission process, the matrix elements of the creation
operator between the initial and final states of the electromagnetic cavity is
evaluated as
< {n0k0 ,β } | â†k,α | {nk0 ,β } > =
p
nk,α + 1 (730)
hence, the matrix elements of the interaction are given by
s
q 2 π h̄ c2 p
< n0 l0 m0 {n0k0 ,β } | Ĥint | nlm {nk0 ,β } > = − nk,α + 1
mc V ωk
× < n0 l0 m0 | ˆα (k) . p̂ exp − i k . r | nlm >
(731)
119
Therefore, the transition rate for photon emission can be expressed as
2
2 π h̄ c2 X
Z
1 2π q V 3
= d k ( nk,α + 1 )
τ h̄ mc ( 2 π )3 V ωk α
× | < n0 l0 m0 | ˆα (k) . p̂ exp − i k . r | nlm > |2 δ( Enlm − En0 l0 m0 − h̄ ωk )
(732)
The above expression shows that the rate for emitting a photon into state (k, α)
is proportional to a factor of nk,α + 1, which depends on the state of occupation
of the normal mode. The term proportional to the photon occupation number
describes stimulated emission. However, if there are no photons initially present
in this normal mode, one still has a non-zero transition rate corresponding to
spontaneous emission. These factors are the result of the rigorous calculations35
based on Dirac’s quantization of the electromagnetic field, but were previously
derived by Einstein36 using a different argument. From the above expression, it
is seen that the number of photons emitted into state (k, α) increases in propor-
tional to the number of photons present in that normal mode. This stimulated
emission increases the number of photons and can lead to amplification of the
number of quanta in the normal mode, and leads to the phenomenon of Light
Amplification by Stimulated Emission of Radiation (LASER).
The first term in the expression produces results that are equivalent to the ra-
diation from an oscillating classical electric dipole. If only the first term in the
expansion is retained, the resulting approximation is known as the dipole ap-
proximation. The second term in the expansion yields results equivalent to the
radiation from an electric quadrupole. The dipole approximation, where only
the first term in the expansion is retained, is justified for transitions where the
35 P. A. M. Dirac, Proc. Roy. Soc. A 114, 243 (1927).
36 A. Einstein, Verh. Deutsche Phys. Ges. 18, 318 (1916), Phys. Z. 18, 121 (1917).
120
A(r)
Ψ(r)
V(r)
Figure 17: A cartoon depicting the relative length-scales assumed in the dipole
approximation.
successive terms in the expansion are successively smaller by factors of the order
of 10−3 . The dipole approximation crudely restricts consideration to the case
where the emitted photon can only have zero orbital angular momentum. This
follows from the dipole approximation’s requirement that the size of the atom is
negligible compared with the scale over which the vector potential varies. Then
the vector potential in the spatial region where the electron is located only de-
scribes photons with zero orbital angular momentum.
In the dipole approximation, the transition rate for single photon emission
is given by
2
2 π h̄ c2 X
Z
1 2π q V 3
≈ d k ( nk,α + 1 )
τ h̄ mc ( 2 π )3 V ωk α
× | ˆα (k) . < n0 l0 m0 | p̂ | nlm > |2 δ( Enlm − En0 l0 m0 − h̄ ωk )
(735)
The matrix elements of the momentum can be evaluated by noting that the
states | nlm > are eigenstates of the unperturbed electronic Hamiltonian so
Ĥ0 | nlm > = Enlm | nlm > (736)
where the unperturbed Hamiltonian is given by
p̂2
Ĥ0 = + V (r) (737)
2m
The electronic momentum operator p̂ can be expressed in terms of the commu-
tator of the Hamiltonian Ĥ0 and r through the relation
h̄
[ r , Ĥ0 ] = i p̂ (738)
m
On using this relation, the matrix elements of the momentum operator can be
written in terms of the matrix elements of the electron’s position operator r by
m
< n0 l0 m0 | p̂ | nlm > = − i < n0 l0 m0 | [ r , Ĥ0 ] | nlm >
h̄
121
m
= i ( Enlm − En0 l0 m0 ) < n0 l0 m0 | r | nlm >
h̄
(739)
where the property of the delta function has been used to set
( Enlm − En0 l0 m0 )2
δ( Enlm − En0 l0 m0 − h̄ ωk ) = ωk2 δ( Enlm − En0 l0 m0 − h̄ ωk )
h̄2
(741)
It is seen that the volume of the electromagnetic cavity has dropped out of the
expression of eqn(740) for the transition rate. We shall assume that the number
of photons nk,α in the initial state is zero. The (complex) factor
is defined as the electric dipole moment, and the electronic energy difference is
denoted by the frequency
The above expression yields the rate at which an electron makes a transition
between the initial and final electronic state, in which one photon of any polar-
ization is emitted in any direction.
122
If one is only interested in the decay rate of the electronic state via the
emission of a photon, one should sum over all polarizations and integrate over
all directions of the emitted photon. The direction of the emitted photon k̂ is
expressed in terms of polar coordinates defined with respect to an arbitrarily
chosen polar axis. The direction of the photon’s wave vector k̂ is defined as
dΩk
z e2(k)
k
e1(k)
θk
y
φk
Figure 18: A photon is emitted with wave vector k with a direction denoted
by the polar coordinates (θk , ϕk ). The polarization vector ê1 (k) is chosen to be
in the plane containing the polar-axis and k, therefore, ê2 (k) is parallel to the
x − y plane.
k̂ = (sin θk cos ϕk , sin θk sin ϕk , cos θk ). The directions of the two transverse
polarizations α are defined as
The scalar product between the polarization vectors and the dipole moment can
be expressed in terms of the Cartesian components via
(i)
X
ˆα (k) . dnlm,n0 l0 m0 = ˆ(i)
α (k) . ( dnlm,n0 l0 m0 ) (747)
i
As neither the polarization nor the direction of the outgoing photon are mea-
sured, the transition rates is determined as an integral over all directions
3
ωnl,n
X X Z
1 0 l0 (j) ∗ (i)
= dΩk ˆ(j) ˆ(i)
α (k) α (k) ( dnlm,n0 l0 m0 ) ( dnlm,n0 l0 m0 )
τ 2 π h̄ c3 i,j α
(748)
123
On using the identity
Z
1 X 2
dΩk ˆ(j) ˆ(i)
α (k) α (k) = δi,j (749)
4π α 3
one finds that the transition rate is given by the scalar product of complex
vectors ∗
3
4 ωnl,n0 l0 dnlm,n0 l0 m0 . dnlm,n0 l0 m0
1
= (750)
τ 3 h̄ c3
The electric dipole matrix elements can be shown to vanish between most pairs
of states. The selection rules determine which matrix elements are non-zero
and, therefore, which electric dipole transitions are allowed.
where Rn,l (r) is the radial wave function, and Yml (θ, ϕ) is the spherical harmonic
function quantized along the z-direction. The components of an arbitrarily
oriented electric dipole matrix elements involve matrix elements of the quantities
x = r sin θ cos ϕ
y = r sin θ sin ϕ
z = r cos θ (752)
Since the above expressions are the components of a vector, they can be re-
written as combinations of the spherical harmonics with angular momentum
l = 1, via
r
1 8π 1 1
x = r Y−1 (θ, ϕ) − Y1 (θ, ϕ)
2 3
r
i 8π 1
y = r Y−1 (θ, ϕ) + Y11 (θ, ϕ)
2 3
r
4π 1
z = r Y0 (θ, ϕ) (753)
3
Hence, the components of the vector r can be written as
r
4π êx + i êy 1 1 êx − i êy 1
r = r √ Y−1 + êz Y0 − √ Y1 (754)
3 2 2
124
The circular polarization vectors are given by
êx − i êy
êm=−1 = √
2
êm=0 = êz
êx + i êy
êm=+1 = − √ (755)
2
which are orthogonal
ê∗m0 . êm = δm,m0 (756)
Hence, the vector r can be written in the alternate forms
r
4π X ∗ 1
r = r êm Ym (θ, ϕ)
3 m
r
4π X
= r êm Ym1 (θ, ϕ)∗ (757)
3 m
37 In the dipole approximation, the photon is restricted to have zero orbital angular mo-
mentum. Therefore, the angular momentum is completely transformed to the photon’s spin.
More generally, the spatial (plane-wave) part of the vector potential should be expanded in
terms of spherical harmonics to exhibit the photon’s orbital angular momentum components.
125
The m-selection rules for electric dipole transitions.
m0 = m ± 1
m0 = m (765)
[ L̂z , x ] = i h̄ y
[ L̂z , y ] = − i h̄ x
[ L̂z , z ] = 0 (766)
On taking the matrix elements between states with definite z-components of the
angular momenta, one finds
which reduce to
126
From the last equation, it follows that either m0 = m or that
( m0 − m )2 = 1 (771)
or
< n0 l0 m0 | x | nlm > = 0 (772)
Hence, the m-selection rules for the electric dipole transitions are ∆m = ± 1, 0.
The selection rules for the magnitude of the electron’s orbital angular mo-
mentum can be found by considering the double commutator
[ L̂2 , [ L̂2 , r ] ] = 2 h̄2 r L̂2 + L̂2 r (773)
and
0 0 0 2 0 2
2 l (l + 1) + l(l + 1) = (l + l + 1) + (l − l) − 1 (776)
or
( l0 + l + 1 )2 − 1 ( l0 − l )2 − 1 = 0 (778)
127
The first factor in eqn(778) is always positive when l0 6= l, therefore, the electric
dipole selection rule becomes ∆l = ± 1.
The actual values of the matrix elements can be found from explicit calcu-
lations. The θ-dependence of the matrix elements is governed by the associated
Legendre functions through
s
( 2 l + 1 ) ( l − m )! l
Θlm (θ) = P (cos θ) (779)
2 ( l + m )! m
which obey the recursion relations
l+1 l−1
l Pm (cos θ) − Pm (cos θ)
sin θ Pm−1 (cos θ) = (780)
2l + 1
and
l−1 l+1
l ( l + m ) ( l + m + 1 ) Pm (cos θ) − ( l − m ) ( l − m + 1 ) Pm (cos θ)
sin θ Pm+1 (cos θ) =
2l + 1
(781)
appropriate for the ∆m = ± 1 transitions and
l+1 l−1
l ( l − m + 1 ) Pm (cos θ) + ( l + m ) Pm (cos θ)
cos θ Pm (cos θ) =
2l + 1
(782)
for the constant m transition, ∆m = 0. Using the recursion relations, one
finds that
s
l ( l + m + 2 ) ( l + m + 1 ) l+1
sin θ Θm (θ) = Θm+1
(2l + 1)(2l + 3)
s
( l − m ) ( l − m − 1 ) l−1
− Θm+1
(2l − 1)(2l + 1)
(783)
for ∆m = 1, while for ∆m = − 1 one finds
s
l ( l + m ) ( l + m − 1 ) l−1
sin θ Θm (θ) = Θm−1
(2l + 1)(2l − 1)
s
( l + 2 − m ) ( l + 1 − m ) l+1
− Θm−1
(2l + 1)(2l + 3)
(784)
and for constant m
s
(l + 1 + m)( l + 1 − m ) l+1
cos θ Θlm (θ) = Θm
(2l + 1)( 2l + 3)
s
(l + m)(l − m)
+ Θl−1 (785)
(2l − 1)(2 l + 1) m
128
The coefficients in the above equation have a similar form to the Clebsch-Gordon
coefficients. The dipole matrix elements can be evaluated by taking the matrix
0
elements of the above set of relations with Θlm0 (θ)∗ and then using the orthog-
onality properties. The above three relations give rise to the selection rules for
the magnitude of the orbital angular momentum l
l0 = l ± 1 (786)
Hence, not only have the selection rules on l been re-derived but the angular
integrations have also been evaluated.
What the above mathematics describes is how the spin angular momentum
of the emitted photon is combined with the orbital angular momentum of the
electron in the final state, so that total angular momentum is conserved. This
implies the selection rules which leads to the magnitude of the initial and final
electronic angular momentum l having to satisfy the triangular inequality
l0 + 1 ≥ l ≥ | l0 − 1 | (787)
l + l0 ≥ 1 ≥ | l − l0 | (790)
Exercise:
129
Table 1: Matrix Elements of the Components of the Dipole Moment
l0 m0 x y z
q q
(l+2+m)(l+1+m) (l+2+m)(l+1+m)
m0 = m + 1 1
2 (2l+1)(2l+3) − 2i (2l+1)(2l+3) -
q
(l+1+m)(l+1−m)
l0 = l + 1 m0 = m - - (2l+1)(2l+3)
q q
(l+2−m)(l+1−m) (l+2−m)(l+1−m)
m0 = m − 1 − 12 (2l+1)(2l+3) − 2i (2l+1)(2l+3) -
q q
(l−m)(l−1−m) (l−m)(l−1−m)
m0 = m + 1 − 12 (2l−1)(2l+1)
i
2 (2l−1)(2l+1) -
q
0 0 (l+m)(l−m)
l =l−1 m =m - - (2l−1)(2l+1)
q q
(l+m)(l−1+m) (l+m)(l−1+m)
m0 = m − 1 1
2 (2l−1)(2l+1)
i
2 (2l−1)(2l+1) -
Using the commutation relations for the j-th component of a vector V̂ j with
the i-th component of the orbital angular momentum L̂i ,
X
[ L̂i , V̂ j ] = i h̄ ξ i,j.k V̂ k (791)
k
(793)
and that the last term of the above expression is zero if V̂ = r.
130
The parity operator is its own inverse since for any state ψ(r)
P̂ 2 ψ(r) = P̂ ψ(−r)
= ψ(r) (795)
Therefore, the parity operator has eigenvalues p = ±1 for the eigenstates which
are defined by
P̂ φp (r) = p φp (r) (796)
so
P̂ 2 φp (r) = p2 φp (r) = φp (r) (797)
2
which yields p = 1 or p = ± 1. In polar coordinates, the parity operation is
equivalent to a reflection
θ → π − θ (798)
followed by a rotation
ϕ → ϕ + π (799)
In electromagnetic processes, parity is conserved since the Coulomb potential is
z r
r
θ
y
π−θ
π+ϕ ϕ
x
-r
Ĥ | φn > = En | φn >
P̂ | φn > = pn | φn > (801)
38 The weak interaction does not conserve parity.
131
Inversion transforms vector operators according to
P̂ r P̂ −1 = − r (802)
Hence, for any matrix elements of r between any eigenstates of the parity oper-
ator, one has
pn 0 pn = − 1 (804)
This is known as the Laporte selection rule for electric dipole transitions39 . The
validity of this selection follows from the fact that inversion commutes with the
orbital angular momentum operator. The spherical harmonics are eigenstates
of the parity operator since
132
For a photon emitted in the direction k̂
k̂ = (sin θk cos ϕk , sin θk sin ϕk , cos θk ) (809)
the polar polarization vectors are given by
ˆ1 (k) = (cos θk cos ϕk , cos θk sin ϕk , − sin θk )
ˆ2 (k) = (− sin ϕk , cos ϕk , 0) (810)
Therefore, the scalar products of the transition matrix elements of r with the
polarizations are given by
1
ˆ1 (k) . < n0 l0 m0 | r | nlm > = cos θk exp[ − i ϕk ] < n0 l0 m0 | (x + iy) | nlm >
2
1
+ cos θk exp[ + i ϕk ] < n0 l0 m0 | (x − iy) | nlm >
2
− sin θk < n0 l0 m0 | z | nlm > (811)
and
i
ˆ2 (k) . < n0 l0 m0 | r | nlm > = − exp[ − i ϕk ] < n0 l0 m0 | (x + iy) | nlm >
2
i
+ exp[ + i ϕk ] < n0 l0 m0 | (x − iy) | nlm >
2
(812)
Due to the m-selection rules
< n0 l0 m0 | (x + iy) | nlm > ∝ δm0 −m−1
< n0 l0 m0 | (x − iy) | nlm > ∝ δm0 −m+1 (813)
and
< n0 l0 m0 | z | nlm > ∝ δm0 −m (814)
the cross-terms in the square of the matrix elements are zero. Hence, on sum-
ming over the polarizations, one finds that the (θk , ϕk ) dependence of the decay
is governed by the dipole matrix elements through
X 1
| ˆα (k) . rnlm,n0 l0 m0 |2 = 1 + cos2 θk | < n0 l0 m0 | (x + iy) | nlm > |2
α
4
1
+ 1 + cos θk | < n0 l0 m0 | (x − iy) | nlm > |2
2
4
+ sin2 θk | < n0 l0 m0 | z | nlm > |2 (815)
0
For l = l + 1, the above sum is found to depend on the angular factors
m0 2 1 (l + 2 + m)(l + 1 + m)
Il0 =l+1 (θk , ϕk ) = 1 + cos θk δm0 −m−1
4 (2l + 1)(2l + 3)
1 (l + 2 − m)(l + 1 − m)
+ 1 + cos2 θk δm0 −m+1
4 (2l + 1)(2l + 3)
(l + 1 + m)(l + 1 − m)
+ sin2 θk δm0 −m (816)
(2l + 1)(2l + 3)
133
Since the z-component of the final electron’s orbital angular momentum is not
measured, m0 should be summed over. The angular distribution of the emitted
radiation for the l0 = l + 1 transition when neither the polarization nor the final
state m0 value are measured is given by
X 0 1 (l + 2)(l + 1) + m2 (l + 1)2 − m2
Ilm
0 =l+1 (θk , ϕk ) =( 1 + cos2 θk ) + sin2 θk
2 (2l + 1)(2l + 3) (2l + 1)(2l + 3)
m0
(817)
This factor determines the angular dependence of the emitted electromagnetic
radiation, which clearly depends on the value of m specifying the initial elec-
tronic state. On rearranging the expression, one finds that the anisotropy is
governed by the factor
X 0 (l + 2)(l + 1) + m2 1 l(l + 1) − 3 m2
Ilm
0 =l+1 (θk , ϕk ) = + sin2 θk
(2l + 1)(2l + 3) 2 (2l + 1)(2l + 3)
m0
(818)
which shows that for m = 0 the photons are preferentially emitted perpendicular
to the direction of quantization axis since this maximizes the overlap between
the polarization and the dipole matrix element. In the opposite case of large
values of m2 [ 3 m2 > l (l + 1) ], one finds that the photons are preferentially
emitted parallel (or anti-parallel) to the axis of quantization. On integrating
over all directions of the emitted photon, one obtains
Z
1 X 0 2 (l + 1)
dΩk Ilm
0 =l+1 (θk , ϕk ) = (819)
4π 0
3 (2l + 1)
m
The independence of the result on m follows since, in this case, there are no an-
gular correlations and the choice of direction of quantization of m is completely
arbitrary. The total decay rate for an electron in a state with fixed m due to
an l0 = l + 1 transition is given by
2
4 e2 ωnl,n
3 Z ∞
1 0 l0 (l + 1) 2 ∗
= dr r Rn0 l+1 (r) r Rnl (r) (820)
τl0 =l+1 3 h̄ c3 (2l + 1) 0
However, if the initial electronic state is unpolarized, then one should sta-
tistically average over the initial m. In this case, the emitted radiation becomes
isotropic
l
1 X X 0 2 (l + 1)
Ilm
0 =l+1 (θk , ϕk ) = (821)
2l + 1 0
3 (2l + 1)
m=−l m
since
l
1 X 1
m2 = l(l + 1) (822)
2l + 1 3
m=−l
134
Hence, if the initial electronic state is unpolarized, the electromagnetic radiation
is isotropic. The decay rate for the l0 = l +1 transition starting with a statistical
distribution of m values is given by
2
4 e2 ωnl,n
3 Z ∞
1 0 l0 (l + 1) 2 ∗
= dr r Rn0 l+1 (r) r Rnl (r) (823)
τl0 =l+1 3 h̄ c3 (2l + 1) 0
for l0 = l + 1. This is the same result that was previously obtained for the decay
rate of a level with a specific m value, when the m0 value of the final state and
the polarization or direction of the emitted photon are not measured.
For the case where l0 = l − 1, one finds that the decay rate involves the
angular factor
m0 2 1 (l − m)(l − 1 − m)
Il0 =l−1 (θk , ϕk ) = 1 + cos θk δm0 −m−1
4 (2l − 1)(2l + 1)
2 1 (l − 1 + m)(l + m)
+ 1 + cos θk δm0 −m+1
4 (2l − 1)(2l + 1)
(l + m)(l − m)
+ sin2 θk δm0 −m (824)
(2l − 1)(2l + 1)
which on summing over the final values of m0 yields the angular dependence of
the radiation field
X 0 1 l(l − 1) + m2 l2 − m2
Ilm
0 =l−1 (θk , ϕk ) = ( 1 + cos2 θk ) + sin2 θk
2 (2l − 1)(2l + 1) (2l − 1)(2l + 1)
m0
(825)
The anisotropy of the emitted radiation is determined by the factor
X 0 l(l − 1) + m2 1 l(l + 1) − 3 m2
Ilm0 =l−1 (θk , ϕk ) = + sin2 θk (826)
0
(2l − 1)(2l + 1) 2 (2l − 1)(2l + 1)
m
which shows that for m = 0 the photons are preferentially emitted perpendicular
to the direction of quantization axis since this maximizes the overlap between
the polarization and the dipole matrix element. In the opposite case of larger
m2 [ 3 m2 > l (l + 1) ], one finds that the photons are preferentially emitted
parallel (or anti-parallel) to the axis of quantization.
Therefore, the decay rate in which the photon is emitted in any direction is
given by the expression
2
4 e2 ωnl,n
3 Z ∞
1 0 l0 l 2 ∗
= dr r R n 0 l−1 (r) r Rnl (r)
(828)
τl0 =l−1 3 h̄ c3 (2l + 1)
0
135
for l0 = l − 1.
Classical Interpretation.
The quantum mechanical results for the angular distribution of the radiation
can be understood in terms of a simple classical model of the atom. In Bohr’s
model, a single electron orbits a central nucleus to which it is bound by the
attractive Coulomb potential. We shall assume that the radius of the orbit is a
and that the electron is performing a circular orbit in the x − y plane. Since the
direction of the electron’s orbital angular momentum is aligned with the z-axis,
it corresponds to the case where m ≈ l and l 1. In this case, the electron
has an oscillating dipole moment given by
d(t) = q a cos ω t êx + sin ω t êy
= q a <e êx − i êy exp i ω t (829)
This rotating dipole moment can be decomposed into two orthogonal linear
dipole moments which oscillate out of phase with each other. It should be
m=l
-
e
d(t) ωt
x
Figure 20: A classical electron orbiting in the x-y plane (m = l) can be consid-
ered as producing two perpendicular linearly-oscillating electric-dipole moments.
recalled that a classical oscillating (linear) electric dipole moment radiates power
P (ω) into a solid angle dΩk with a distribution given by
4
dP c ω
= | d |2 sin2 Θkd (830)
dΩk linear 8π c
where Θkd is the angle between the detector and the direction of the electric
dipole. On considering the radiation from the atom to be generated from two
orthogonal linear oscillating dipoles, one finds
4
dP c ω 2 2 2
= |d| sin Θkx + sin Θky (831)
dΩk dipole 8π c
136
dΩk
ez k
Θky
θk
ey
Θkx
ex
e- m=l
Figure 21: The polarization of the radiated electromagnetic field for an electron
orbiting in the x-y plane (m = l) can be comprehended in terms of the classical
radiation emanating from two linearly oscillating electric-dipole moments. The
angles Θkx and Θky , respectively, are the angles between the emitted radiation
and the x-axis and the angle subtended by the emitted radiation and the y-axis.
which on using
becomes
4
dP c ω 2 2
= |d| 1 + cos θk (833)
dΩk dipole 8π c
Since the energy of the emitted photon is given by h̄ ω, one finds the angular
dependence of the semi-classical prediction of the decay rate is given by
3
e2
1 ωa
= 1 + cos2 θk dΩk (834)
τdΩk 8 π h̄ a c
The polarization vector is parallel to the direction of the electric field, which in
turn is given by the direction of the oscillating dipole that produced it. Hence,
a detector which is arranged to accept radiation travelling in the direction k̂ will
detect polarizations that are found by projecting the electron’s orbit onto the
plane perpendicular to k̂. For example, in this case where the electron’s orbit
is in the x − y plane, so radiation along the z-axis will be circularly-polarized,
whereas radiation in the x − y plane will be linearly-polarized.
137
Circular
e-
k
Linear
The angular dependence of the decay rate follows directly from the expres-
sions of eqn(817) and eqn(825) by setting m ≈ l 1, replacing the radial matrix
elements of r by a, adding the expressions and inserting them into eqn(808). The
analysis shows that quantum mechanics reproduces the classical limit correctly,
as is expected from the correspondence principle.
for
l+1
l0 = (836)
l−1
It should be noted that, for a fixed l0 , the lifetime of the state | nlm > is inde-
pendent of the value of m. This is expected since the choice of the quantization
138
Table 2: Radial wave functions Rnl (ρ) forRa Hydrogenic-like atom, where ρ =
Zr ∞
a . The functions are normalized so that 0 dρ ρ2 Rnl Rnl = 1.
n = 1 l = 0 2 exp − ρ
√1 ρ ρ
n = 2 l = 0 2
1 − 2 exp − 2
1 ρ
l = 1 2
√
6
ρ exp − 2
2 2 2 2 ρ
n = 3 l = 0 3 1 − 3 ρ + 27 ρ exp − 3
32
5
22 ρ ρ
l = 1 7 1 − 6 ρ exp − 3
32
3
22 2 ρ
l = 2 9 √ ρ exp − 3
32 5
139
R∞
Table 3: Values of | 0
dr r2 Rnl r Rn0 l−1 |2 in atomic units.
n, l n0 , l − 1
np 1s 28 n7 (n − 1)2n−5 (n + 1)−2n−5
There are no selection rules associated with the radial integration in the
dipole matrix elements Z ∞
dr r2 Rn0 l−1 r Rnl (837)
0
The radial part of the dipole matrix element can be expressed in terms of the
hypergeometric function F (a, b, c) via
Z ∞
dr r2 Rn0 l−1 r Rnl
0
0
s 0
a (−1)n −l (n + l)!(n0 + l − 1)! (4n0 n)l+1 (n − n0 )n +n−2l−2
=
4(2l − 1) (n − l − 1)!(n0 − l)! (n0 + n)n0 +n
2
4n0 n
0
4n0 n
0 n −n 0
× F (l + 1 − n, l − n , 2l, − 0 )− F (l + 1 − n, l − n , − 0 )
(n − n)2 n0 + n (n − n)2
(838)
Simple analytic expressions for the squares of the matrix elements for small val-
ues of (n0 , l) are shown in Table(3).
140
function expansion for the Laguerre polynomials. Eckart41 and Gordon42 have
calculated these dipole matrix elements by other means. In general, the lifetime
of the hydrogenic states increases with increasing n, varying roughly as n3 for
a fixed value of l. The decrease in the dipole matrix elements with increasing n
is simply due to the increasing numbers of nodes in the radial wave functions.
so the decay time is approximately eight orders of magnitude larger than the
time taken for the photon to cross the atom. When averaged over l, the electric
dipole decay rate is given by
1 X (2l + 1) 9
∝ ∼ n− 2 (843)
τn n5
l
so, as seen in Table(4), the decay is slower for the higher energy levels.
141
Table 4: Electric Dipole Transition Rates for Hydrogen, in units of 108 sec−1 .
2p ns 6.25 - -
3s np - 0.063 -
3p ns 1.64 0.22 -
3d np - 0.64 -
4s np - 0.025 0.018
4p ns 0.68 0.095 0.030
4p nd - - 0.003
4d np - 0.204 0.070
4f nd - - 0.137
h̄2
a = (846)
m e2
The decay rate in the Fermi-Golden rule, evaluated in the dipole approximation,
is given by
1 3
4 ω1,2 d∗1s,2p . d1s,2p
= (847)
τ 3 h̄ c3
The frequency is evaluated from
142
4 2
e2
c d1s,2p
9
= (849)
128 h̄ c a ea
Therefore, the scattering rate is determined by the ratio ac but also is modified
by the fourth power of the dimensionless electromagnetic coupling strength
2
e 1
≈ (850)
h̄ c 137.0359979
The smallness of this factor allows us to only consider the Fermi-Golden rule
expression for the decay rate. The dimensionless dipole matrix elements are
expected to be non-zero, since they obey the selection rules. They are non-zero,
as can be directly verified by performing an integration. The only non-zero
dipole matrix element originates from the z-component of the dipole
Z
∗
d1s,2p = e d3 r ψ1s (r) r ψ2p (r) (851)
since only the z-component satisfies the ∆m = 0 selection rule. The angular
integration is evaluated as
√ Z 2π Z π √
3 3 2
dϕ dθ sin θ cos2 θ = 2π
4π 0 0 4 π 3
1
= √ (852)
3
and the radial integration yields
Z ∞ ∞
r4
Z
∗ 2 3 r
dr r2 R1s (r) r R2p (r) = √ dr exp −
0 2
6 0 a3 2 a
5 Z ∞
a 2 4
= √ dx x exp − x
6 3 0
5 5
4! 2 √ 2
= a √ = 4a 6 (853)
6 3 3
Hence, the magnitude of the dipole matrix element is evaluated as
5
d1s,2p √ 2
= 4 2 (854)
ea 3
Therefore, the dipole allowed decay rate is given by
8 2 4
1 2 e c
= (855)
τ 3 h̄ c a
Hence, the time scale τ is of the order of 10−10 seconds. The exact value of the
decay time is calculated to be 1.6 × 10−9 seconds.
143
11.1.7 Electric Quadrupole and Magnetic Dipole Transitions.
Consider decays such as the 3d state (with m = 0) to the 1s state in the hydrogen
atom. Since, in this transition, the change in the electron’s angular momentum
is two units, the transition is forbidden in the dipole approximation. Therefore,
the transition rate is evaluated by keeping the next order term in the expansion
exp − i k . r ≈ 1 − i k . r + ... (856)
The second term in the expansion describes electric quadrupole and magnetic
dipole transitions.
This shall be written as the sum of two terms, with different symmetries with re-
spect to interchange of r and p. These two terms will describe electric quadrupole
and magnetic dipole transitions. The matrix elements are written as the sum
of a term symmetric under the interchange of r and p̂ and a term that is anti-
symmetric
1
( k . r ) ( ˆα (k) . p̂ ) = ( k . r ) ( ˆα (k) . p̂ ) + ( k . p̂ ) ( ˆα (k) . r )
2
1
+ ( k . r ) ( ˆα (k) . p̂ ) − ( k . p̂ ) ( ˆα (k) . r )
2
(858)
The first term represents the matrix elements for the electric quadrupole tran-
sitions43 , and the second term represents the matrix elements for the magnetic
dipole transitions. The first term can be written as the scalar products of a
symmetric dyadic
k. r p̂ + p̂ r . ˆα (k) (859)
The scalar products are organized such that the left most vector outside the
parenthesis forms a scalar product with the left most vectors within the paren-
thesis, and likewise with the right most vectors. The electronic matrix elements
only involve the dyadic operator, as the wave vector and polarization vectors
are properties of the photon. The matrix elements
0 0 0
< nlm | r p̂ + p̂ r | nlm > (860)
[ r , p̂2 ] = 2 i h̄ p̂ (861)
43 J. A. Gaunt and W. H. McCrea, Proc. Camb. Phil. Soc. 23, 930 (1927).
144
which allows the momentum operator to be written as
im
p̂ = [ Ĥ , r ] (862)
h̄
Therefore, the matrix elements of the dyadic can be expressed in the form of
the matrix elements of the commutator with the dyadic
0 0 0 im
< nlm | r p̂ + p̂ r | nlm > = < n0 l0 m0 | r [ Ĥ , r ] + [ Ĥ , r ] r | nlm >
h̄
im
= < n0 l0 m0 | [ Ĥ , r r ] | nlm >
h̄
( En0 l0 m0 − Enlm )
= im < n0 l0 m0 | r r | nlm >
h̄
= i m ωn0 ,n < n0 l0 m0 | r r | nlm > (863)
The decay rate in the Fermi-Golden rule, evaluated in the electric quadrupole
approximation, is given by
2 Z
m2 c2 ωk2 X
1 e
= d3 k | k . < n0 l0 m0 | r r | nlm > . ˆα (k) |2
τ mc 8 π ωk α
× δ( Enlm − En0 l0 m0 − h̄ ωk )
5 Z
e2
ωnl,n0 l0 X
= dΩk | k̂ . < n0 l0 m0 | r r | nlm > . ˆα (k) |2
8 π h̄ c α
(864)
145
Therefore, the scattering rate is determined by the ratio ac but also is modified
by the sixth power of the dimensionless electromagnetic coupling strength
2
e 1
≈ (868)
h̄ c 137.0359979
The smallness of this factor allows us to only consider the Fermi-Golden rule
expression for the decay rate. The dimensionless quadrupole matrix elements
are expected to be non-zero, since they obey the selection rules which involve
the exchange of two units of angular momentum. They are non-zero, as can
be directly verified by performing an integration. Therefore, the quadrupole
allowed decay rate is given by
6
e2
1 c
∼ (869)
τ h̄ c a
one can add a diagonal term to the dyadic without affecting the result. A
diagonal term with a magnitude that makes the resulting dyadic traceless is
added to the dyadic, leading to the expression
1
Qi,j = e xi xj − δi,j | r |2 (871)
3
The symmetric dyadic Qi,j has six inequivalent components, which because of
the restriction that the dyadic is traceless, can be reduced to five independent
components. Due to the transformational properties of the dyadic under ro-
tation, it can be expressed as a linear combination of the spherical harmonics
Ym2 (θ, ϕ) and nothing else45 . This can be seen from rewriting the quadrupole
tensor Q̃
2
xx − r3 xy xz
Q̃ 2
= yx yy − r3 yz (873)
e
2
r
zx zy zz − 3
44 This estimate will be modified upwards by several orders of magnitude, due to the presence
146
in terms of spherical polar coordinates
2 2
1 1 2 1
Q̃ 2 sin θ cos 2ϕ − 6 (3 cos θ − 1) 2 sin θ sin 2ϕ sin θ cos θ cos ϕ
1 2 2
= 2 sin θ sin 2ϕ − 2 sin θ cos 2ϕ − 16 (3 cos2 θ − 1)
1
sin θ cos θ sin ϕ
e r2 1 2
sin θ cos θ cos ϕ sin θ cos θ sin ϕ 3 (3 cos θ − 1)
(874)
The presence of states with orbital angular momentum of only two makes the
dyadic an irreducible second rank tensor. Application of the Wigner-Eckart
theorem to an irreducible second rank tensor results in the electric quadrupole
selection rules
l + l0 ≥ 2 ≥ | l − l0 | (875)
The angular momentum carried away by the photon consists of the spin-one
carried away by the photon in addition to the component of the photon’s wave
function described by the spherical Bessel function j1 (kr) ∼ k r which carries
off one unit of orbital angular momentum. In addition to the angular momen-
tum selection rules, there are parity selection rules for the electric quadrupole
transitions. Since the parity operator satisfies
P̂ r P̂ −1 = − r (876)
147
The magnetic dipole transition should be extended from orbital angular mo-
mentum to include the spin magnetic moment which is of the same order
e h̄
σ . ( k ∧ ˆα (k) )
2mc
e
( r ∧ p ) . ( k ∧ ˆα (k) ) (883)
2mc
since orbital angular momentum is quantized in units of h̄. The angular mo-
mentum selection rule for the magnetic dipole transition is given by
∆l = 0 (884)
and
1 ≥ | ∆m | (885)
also parity does not change.
Terms with higher-order orbital angular momentum that occur in the ex-
pansion of the photon’s wave function exp[ i k . r ] can be found by using the
Rayleigh expansion. The terms with orbital angular momentum l are propor-
tional to the spherical harmonics jl (kr) which vary as (kr)l when kr → 0, as is
found from the expansion of the exponential term. The presence of the extra
factors k l in the matrix element has the result that the electric 2s -th multi-pole
transition rates are found to vary as
2s+1
1 ωn,n0
∝ a2s (886)
τ c
where s is the magnitude of the change in the electronic orbital angular mo-
mentum, which satisfies the inequality
( l + l0 ) ≥ s ≥ | l0 − l | (887)
The extra factors from the photon’s angular momentum results in an overall de-
2
crease in the electric multi-pole transition rate by a factor of ( h̄e c )2s . It should
also be noted that the relative strength of the higher-order electric multi-pole
transitions increase more rapidly with Z than the electric dipole transitions.
Therefore, it is frequently found that the quadrupole transitions cannot be ne-
glected for the heavy elements. Alternatively, higher-order multipole transitions
do become important in the x-ray region, since in this region the wavelength of
the radiation is comparable to the spatial extent of the charged particle’s wave
function.
148
transition. The electric quadrupole transition rate can be expressed as
5 Z
1 1 ω X
= dΩk | k̂ . < 1s | Q̃ | 3d > . ˆα (k) |2 (888)
τ 8 π h̄ c α
where Q̃ represents the quadrupole tensor. The frequency factor can be evalu-
ated as 2
ω 4 e
= (889)
c 9 a h̄ c
hence, the rate can be expressed in the form
5 2 6 Z
1 c 4 e X 1
e | 3d > . ˆα (k) |2
= dΩk 2 a4
| k̂ . < 1s | Q
τ 8πa 9 h̄ c α
e
(890)
We shall consider the transition from the m = 0 state of the 3d level to the 1s
state. As can be easily shown, the matrix elements of quadrupole tensor for this
transition are diagonal and are given by
− Q2zz
0 0
< 1s | Q̃ | 3d > = < 1s | 0 − Q2zz 0 | 3d > (891)
0 0 Qzz
Therefore, the transition matrix elements are of the form
Z X 1
dΩk e | 3d > . ˆα (k) |2
| k̂ . < 1s | Q
e2 a4
α
Z X | < 1s | Qzz | 3d > |2 2
1 1
= dΩk k̂z ˆα (k)z − k̂x ˆα (k)x − k̂y ˆα (k)y
α
e2 a4 2 2
(892)
The direction of the emitted photon k̂ is expressed as
k̂ = (sin θk cos ϕk , sin θk sin ϕk , cos θk ) (893)
and the polarization vectors are given by
ˆ1 (k) = (cos θk cos ϕk , cos θk sin ϕk , − sin θk )
ˆ2 (k) = (− sin ϕk , cos ϕk , 0) (894)
Thus for the m = 0 level, one finds that the integral over the angular distribution
is given by
Z X 1
dΩk e | 3d > . ˆα (k) |2
| k̂ . < 1s | Q
e2 a4
α
2
| < 1s | Qzz | 3d > |2
Z
3
= dΩk − sin θ k cos θ k
e2 a4 2
3 | < 1s | Qzz | 3d > |2
= 4π (895)
10 e2 a4
149
The scattering rate becomes
5 2 6
1 3 4 c e | < 1s | Qzz | 3d > |2
= (896)
τ 20 9 a h̄ c e2 a4
Finally, one finds the resulting expression for the quadrupole decay rate of the
3d state with m = 0 2 6
1 1 c e
= (898)
τ 3600 a h̄ c
which is evaluated as 228 sec−1 . From the above analysis, it is seen that angular
distribution for the emitted photon is governed by the factor
and the intensity is largest for the cone with θk ≈ 0.28 π or 0.72 π. This angular
dependence of the emitted radiation is the same as found by considering the
radiation from an oscillating classical quadrupole, for which the radiated power
is given by
6
dP c ω
= Q2 cos2 θk sin2 θk (900)
dΩk quad 288 π c
sin2 θk (901)
π
which is maximum for θk = 2.
150
m=0
(903)
151
Hence, the transition matrix element is given by
Z ∞
h̄ ˆα (k) . r ∗
−i d3 r ψ2s (r) exp − i k . r ψ1s (r) (906)
a 0 r
the factor ˆα (k) . r only depends on x and y and is antisymmetric with respect
to the transformations x → − x and y → − y. All other factors are even
functions of x and y. On integrating over the directions in the x − y plane, one
finds that the integral is identically zero.
The above result could have been (partially) anticipated by considering the
selection rules. The electric dipole transition is forbidden by parity. The mag-
netic dipole transition is zero in this non-relativistic treatment. All magnetic
and electric quadrupole and higher multipole transitions are forbidden by an-
gular momentum conservation.
The 2s state decays via two-photon emission which is described by the dia-
magnetic interaction and by the effect of the paramagnetic interaction taken to
second-order in time-dependent perturbation theory. Since only the part of the
paramagnetic interaction that creates a photon is involved, for our purposes the
paramagnetic interaction can be replaced by
1
q X 2 π h̄ c2 2
Ĥpara → − p̂ . ˆα (k) a†k,α exp − i k . r (908)
mc V ωk
k,α
hk
n
n'
152
q2 2 π h̄ c2 ˆα (k) . ˆα0 (k 0 ) †
†
X
0
Ĥdia → √ a a
k,α k0 ,α0 exp − i ( k + k ) . r
2 m c2 V ωk ωk 0
k,α;k0 ,α0
(909)
for two-photon emission. The system is assumed to be initially in an eigenstate
(k,α)
(k',α')
2s
1s
of the unperturbed Hamiltonian | n > but, due to the interaction Ĥint makes
transitions to states | n0 >. These states are to be considered as products
of the electronic states and the states of the electromagnetic cavity. Following
the usual procedure of time-dependent perturbation theory, the above state
| ψn > can be decomposed in terms of a complete set of non-interacting energy
eigenstates | n > via
X i
| ψn > = Cn0 (t) exp − En0 t | n 0 > (910)
0
h̄
n
where Cn0 (t) are time-dependent coefficients. The probability of finding the
system in the final state | n0 > at time t is then given by |Cn0 (t)|2 . The rate
at which the transition n → n0 occurs is then given by the time-derivative of
|Cn0 (t)|2 .
where ω = c k and ω 0 = c k 0 are the energies of the two photons in the final
state. It should be noted that, since we have evaluated the photonic part of the
153
initial and final state, the labels n and n0 , now only describe the electronic part
of the inital and final states. The small quantity η has been absorbed as a small
imaginary part to the initial state energy
En → En + i η h̄ (912)
(b) emission of a photon (k 0 , α0 ) followed by the emission of the photon (k, α).
n'' n''
n n' n n'
(913)
154
< n0 l0 m0 | ˆα0 (k 0 ) . p̂ | n00 l00 m00 > < n00 l00 m00 | ˆα (k) . p̂ | nlm >
( En − En00 − h̄ ω )
< n0 l0 m0 | ˆα (k) . p̂ | n00 l00 m00 > < n00 l00 m00 | ˆα0 (k 0 ) . p̂ | nlm >
+ (914)
( En − En00 − h̄ ω 0 )
where the matrix elements M are due to the combined effect of the diamagnetic
interaction and the paramagnetic interaction taken to second-order. That is,
155
These three terms add coherently, and it should be noted that the intermediate
state is only a virtual state and it can have a higher-energy than the 2s state46 .
In the limit η → 0 the first term in the expression for the transition rate
of eqn(918) reduces to a delta function which expresses conservation of energy
between the initial and final states.
η exp 2 η t
1
lim = δ( E2s − E1s − h̄ωk0 − h̄ωk )
η→0 π (h̄ω 0 + h̄ω + En0 − En )2 + h̄2 η 2
(920)
In the limit η → 0 the transition rate reduces to the Fermi-Golden rule
expression
1 2π X
= | M |2 δ( E2s − E1s − h̄ωk0 − h̄ωk ) (921)
τ h̄ 0 0 k,α:k ,α
The emitted photons have continuous spectra. In the expression for the matrix
elements M , the last two terms differ in the time-order that the two photons
are emitted. On inserting the expressions for the interactions into M , one can
pull out the common factors leaving a dimensionless matrix element M 0 . This
leads to the expression
q2 2 π h̄ c2
1
M = 2
√ M0 (922)
2mc V ωk ωk 0
where M 0 is the dimensionless factor given by
0 0 0
M = ˆα (k) . ˆα0 (k ) < 1s | exp − i ( k + k ) . r | 2s >
2 X < 1s | ˆα (k) . p̂ exp[ − i k . r ]| n00 l00 m00 > < n00 l00 m00 | ˆα0 (k 0 ) . p̂ exp[ − i k 0 . r ] | 2s >
+
m E2s − En00 l00 m00 − h̄ωk0
n00 l00 m00
2 X < 1s | ˆα0 (k 0 ) . p̂ exp[ − i k 0 . r ]| n00 l00 m00 > < n00 l00 m00 | ˆα (k) . p̂ exp[ − i k . r ] | 2s >
+
m E2s − En00 l00 m00 − h̄ωk
n00 l00 m00
(923
The first term is negligible, since 1 k . r and the electronic eigenstates
are orthogonal. The order of magnitude of the second term is given by the
electronic kinetic energy divided by the excitation energy. Hence, the reduced
matrix elements have a magnitude of the order of unity. The transition rate is
given by
2 2
1 e ( 2 π )3 X h̄ c2
= 2 2 0
|M 0 |2 δ( E2s − E1s − h̄ωk − h̄ωk0 )(924)
τ 2mc V 0 0
k k
k,α;k ,α
46 Due to the Lamb shift, there is a 2p state with slightly lower energy than the 2s state.
However, due to the small magnitude of the energy difference, the part of the decay process
involving any real 2p transition is negligibly small.
156
One can assume that the dipole matrix elements of the intermediate states
should be randomly oriented in space, since the initial and final electronic states
are isotropic. After summing over the polarizations, the transition rate becomes
isotropic. On setting X
| M 0 |2 ≈ 1 (925)
α,α0
one finds
2 2
h̄ c2 d3 k d3 k 0
Z Z
1 e
= δ( E2s − E1s − h̄ωk − h̄ωk0 )(926)
τ 2 m c2 ( 2 π )3 k k0
Since the integrand is independent of the direction of k and k 0 , the angular
integrations can be performed leaving
2 2 Z ∞ Z ∞
1 e h̄ c2
= dk k dk 0 k 0 δ( E2s − E1s − h̄ωk − h̄ωk0 )(927)
τ m c2 2π 0 0
Thus, the estimated decay rate is 8.75 sec−1 . The exact value calculated by
Shapiro and Breit47 is 8.266 sec−1 .
157
11.1.10 The Absorption of Radiation
If a process occurs in which only a photon with quantum numbers (k, α) is
absorbed, then the numbers of quanta in the initial and final state of the elec-
tromagnetic field are given by
n0k,α = nk,α − 1
n0k0 ,β = nk0 ,β (934)
The matrix elements of the paramagnetic interaction are given by
< n0 l0 m0 {n0k0 ,β } | Ĥpara | nlm {nk0 ,β } >
s
√ 2 π h̄ c2
X
q 0 0 0
= − nk,α < n l m | p̂ . ˆα (k) exp + i k . r | nlm >
mc V ωk
k,α
(935)
The photon absorption rate is found from the Fermi-Golden rule expression
2
2 π h̄ c2
1 2π q X
= nk,α δ( Enlm + h̄ωk − En0 l0 m0 )
τ h̄ mc V ωk
n0 l0 m0
0 0 0
× | < n l m | p̂ . ˆα (k) exp + i k . r | nlm > |2 (936)
This is related to the lifetime due to stimulated emission, if the initial and final
states are interchanged.
which simplifies to
4 π 2 e2
X
σabsorb (ωk ) = δ( Enlm + h̄ωk − En0 l0 m0 )
m2 ωk c
n0 l0 m0
× | < n0 l0 m0 | p̂ . ˆα (k) exp + ik.r | nlm > |2
(939)
158
The absorption cross-section is independent of the volume of the electromag-
netic cavity and the number of photons in the incident beam. As a function of
frequency, the Born approximation for the cross-section for photon absorption
contains delta function lines corresponding to the atomic excitation energies.
Measured absorption lines do have natural widths ∆ωnl,n0 l0 and the absorbtion
spectra can be approximated by the sums of Lorentzian functions. The widths
70
60
σ(ω) [ h /m ]
50
40
2
30
20
10
0
0.7 0.8 0.9 1
hω [ Ryd ]
Figure 27: A sketch of the photon absorption cross-section σ(ω) (in units of
h̄2
m ) as a function of photon energy h̄ω (in units of Rydbergs). The plot over-
emphasizes the role of the photon lifetimes, since the ratio of the line-width to
e2 3
the photon frequency is of the order of ( h̄c ) .
of the lines are governed by half the sum of the decay rates of the initial and
final electronic levels.
1 1 1
∆ωnl,n0 l0 = + (940)
2 τnl τ n0 l 0
This formula implies that rapidly decaying levels will yield broad lines, but does
not imply the converse48 . The spectral widths can be described by the inclu-
sion of the effects of interaction to higher orders49 . The higher-order processes
produce small shifts of the atomic energy levels and also give the energies small
imaginary parts, resulting in a Lorentzian line shape. Since a typical atomic
transition rate is of the order of 108 sec−1 and a typical photon frequency is of
the order of 1015 sec−1 , the widths of the lines can usually be neglected.
159
× | < n0 l0 m0 | p̂ . ˆα (k) | nlm > |2 (941)
For an isotropic medium, the electronic states are degenerate with respect to
the z-components of the orbital angular momentum, so the initial state (n, l, m)
should be averaged over the different values of m
l
1 X
(943)
(2l + 1)
m=−l
and the values of m0 for the final states are summed over all possible values.
This averaging process results in an isotropic absorption rate, and is equivalent
to averaging the polarization vector over all directions in space. Therefore, in
the dipole approximation, the absorption cross-section for an isotropic medium
is given by the expression
4 π 2 e2
X
σabsorb (ω) = ωn0 l0 ,nl | < n0 l0 m0 | r | nlm > |2 δ(ωn0 l0 ,nl − ω )
3 h̄ c 0 0 0
nlm
(944)
The strength of each absorption line can be found by integrating the cross-
section over a narrow frequency range centered on the frequency of the absorp-
tion line. (More specifically, the width of the interval of integration must be
greater than the natural line-width.) The integrated intensity of the transition
(nlm) → (n0 l0 m0 ) is given by
Z ωnl,n0 l0 +
4 π2
2
e
dω σabsorb (ω) = ωn0 l0 ,nl | < n0 l0 m0 | r | nlm > |2
ωnl,n0 l0 − 3 h̄ c
(945)
The intensity of each line is proportional to the “oscillator strength” fnl→n0 l0
defined as
2 m ωn0 l0 ,nl
fnl→n0 l0 = | < n0 l0 m0 | r | nlm > |2 (946)
h̄
The intensities and the frequencies of all the transitions are related via sum
rules50 . These sum rules involve quantities of the form
X
ωnp 0 l0 m0 ,nlm | < n0 l0 m0 | r | nlm > |2 (947)
n0 l0 m0
160
Table 5: Sum Rules for Dipole Transitions
3 h̄
1 2 m
2
2 m ( Enlm − < nlm | V | nlm > )
h̄
3 2 m < nlm | ∇2 V | nlm >
and have values given in the Table(5). The sum rules can be used to provide
checks of experimental data.
There exists a systematic way of deriving sum rules for the weighted inten-
sities of the dipole allowed transitions. The sum rules are of the form
X p 0 0 0
ωnl,n0 l0 | < nlm | Â | n l m > |2 (948)
n0 l0 m0
where
h̄ ωnl,n0 l0 = Enl − En0 l0 (949)
and p is a positive integer.
161
Then, on taking successive derivatives of F (t) with respect to t, one finds
∂F i
= < nlm | [ Ĥ0 , Â(t) ] † (0) | nlm > (952)
∂t h̄
and
2
∂2F
i
= < nlm | [ Ĥ0 , [ Ĥ0 , Â(t) ] ] † (0) | nlm > (953)
∂t2 h̄
etc. This process shows that the p-th derivative is expressed as p nested com-
mutators
p p
∂ F i
p
= < nlm | [ Ĥ0 , [ . . . [ Ĥ0 , [ Ĥ0 , Â(t) ] ] . . . ] ] † (0) | nlm >
∂t h̄
(954)
Alternatively, one can insert a complete set of states in the definition for F (t)
yielding
X
F (t) = < nlm | Â(t) | n0 l0 m0 > < n0 l0 m0 | † (0) | nlm > (955)
n0 l0 m0
but since the states | nlm > are eigenstates of Ĥ0 , one has
X 2
< nlm | Â | n0 l0 m0 >
F (t) = exp i ωnl,n0 l0 t (956)
n0 l0 m0
The sum rules are found by equating the two forms of the p-th time-derivative
and then setting t = 0
X p 2
0 0 0
Enl − En0 l0 < nlm | Â | n l m >
n0 l0 m0
Hence, the p-th moment of the matrix elements of  is related to the expecta-
tion value of the product of the p-th nested commutator of Ĥ0 and  multiplied
by † .
162
value is homogeneous in time. The q-th nested commutator of the operator Â
can be defined by
B̂q = [ Ĥ0 , B̂q−1 ] (959)
where
B̂0 = Â (960)
†
Likewise, Ĉq can be defined as the q-th nested commutator of  . However, for
any pair of operators B̂p−q−1 and Ĉq , one has
n0 l0 m0
< nlm | [ Ĥ0 , B̂p−q−1 ] Ĉq | nlm > = ( − 1 ) < nlm | B̂p−q−1 [ Ĥ0 , Ĉq ] | nlm >
(962)
On using the definition of the operators B̂p and Cq , the above equation reduces
to
< nlm | B̂p−q Ĉq | nlm > = ( − 1 ) < nlm | B̂p−q−1 Ĉq+1 | nlm > (963)
By induction, this shows that the nested commutators can be distributed be-
tween the two sides of the expression.
< nlm | B̂p Ĉ0 | nlm > = ( − 1 )q < nlm | B̂p−q Ĉq | nlm > (964)
163
The cross-section is given by
X 4 π 2 e2 h̄2 k 02
0 2
σ = | < k | p .
ˆα (k) exp i k . r | 1s > | δ( E 1s + h̄ ω k − )
0
m2 ωk c 2m
k
(966)
where the initial wave function is given by
1 r
ψ1s (r) = √ exp − (967)
π a3 a
As long as the emitted electron is not close to threshold, the final state wave
function can be approximated by a plane-wave
1
ψk0 (r) = √ exp i k 0 . r (968)
V
The sum over final states of the electron can be replaced by an integral over the
magnitude of its momentum and its direction
Z ∞ Z
X V 0 02
→ dk k dΩk0 (969)
0
( 2 π )3 0
k
It is seen that the factor of the volume in the density of final states cancels with
the factors from the normalization of the electron’s final state. The differential
cross-section corresponds to the part of the cross-section where the outgoing
electron is emitted into the solid angle dΩk0 . Hence,
Z ∞
e2 h̄2 k 02
dσ V 0 02 0 2
= dk k | < k | p .
ˆα (k) exp i k . r | 1s > | δ( E 1s + h̄ ω k − )
dΩ0 2 π m2 ωk c 0 2m
(970)
The integration over the magnitude of electron’s final momentum k 0 can be
performed by using the properties of the energy conserving delta function. The
magnitude of electron’s final momentum is denoted by kf
2m
kf2 = ( h̄ωk + E1s ) (971)
h̄2
The result of the integration over k 0 is
e2
dσ V 0
= kf | < kf dΩ | p . ˆα (k) exp i k . r | 1s > |2
dΩ0 2 π h̄2 m ωk c
(972)
It is assumed that the initial photon is propagating along the x-axis and is
polarized along the z-direction. The matrix elements involving the momentum
operator only yield a finite result when p̂ acts on ψ1s (r), since k . ˆα (k) = 0.
However,
h̄ cos θ
ˆα (k) . p̂ ψ1s (r) = i ψ1s (r) (973)
a
164
eα e- (Ek',k')
'
θ
(hωk,k)
Figure 28: The geometry for the photo-emission of an electron from an atom.
An electromagnetic wave, with polarization along the z-axis, is incident along
the x-axis. The photo-emitted electron propagates along the direction k 0 .
e2
dσ V 0
= kf | < kf dΩ | cos θ exp i k . r | 1s > |2 (974)
dΩ0 2 π m ωk c a2
where (θ, ϕ) are the polar coordinates of the vector r. The matrix elements are
evaluated using the dipole approximation for the photon wave function and set
exp i k . r ≈ 1 + i k . r + ... (975)
and only keep the first term of the expansion. The factor cos θ can be expressed
as a spherical harmonic through
r
4π 1
cos θ = Y0 (θ, ϕ) (976)
3
and the final state electronic wave function can be expressed in terms of the
Rayleigh expansion
X
0
exp i kf k̂ . r = 4π il jl ( kf r ) Yml (θ, ϕ) Yml ∗ (θ0 , ϕ0 ) (977)
l,m
where (θ0 , ϕ0 ) are the polar coordinates of the electron’s final momentum. The
angular integration over the polar coordinates (θ, ϕ) can be performed by using
the orthogonality relations for the spherical harmonics. The end result is
Z ∞
1 r
< kf dΩ0 | cos θ | 1s > = − 4 π i cos θ0 √ dr r2 j1 (kf r) exp −
π a3 V 0 a
(978)
where the cos θ0 dependence refers to the direction of the emitted electron’s
angular momentum. The radial integral is evaluated to yield
r
0 0 a3 2 kf a
< kf dΩ | cos θ | 1s > = − 4 π i cos θ (979)
π V ( 1 + kf2 a2 )2
165
Therefore, the differential cross-section is given by
2 2
dσ kf e a 2 0 2 kf a
= 8 cos θ (980)
dΩ0 k m c2 ( 1 + kf2 a2 )2
Using
h̄2
a = (981)
m e2
the photo-emission cross-section can be re-written as
2 2 2
dσ kf 2 e 2 0 2 kf a
= 8 a cos θ (982)
dΩ0 k h̄ c ( 1 + kf2 a2 )2
Thus, although the photon is propagating along the x-direction, the electron is
preferentially emitted along the direction of the polarization (θ0 ≈ 0). This can
be understood as being due to the effect of c being large, so that the photon’s
momentum is negligible compared with the energy, therefore, (in the dipole
approximation) only the direction of the polarization determines the angular
distribution of the emitted electron. It should be noted that in the relativistic
case, where the momentum of the photon is important, the electrons are pre-
dominantly ejected in the direction of the photon51 . This formula also breaks
down for emitted electrons with low energies. In this case, the correct electronic
wave function for the continuous spectrum of Ĥ0 should be used52 . The in-
clusion of the Coulomb attraction of the ion in the final state has the effect of
reducing the cross-section near the threshold.
166
which is evaluated as
r
√
q 2 π h̄ c
− p̂ . ˆα (k) δk+k0 −k00 nk,α (987)
mc V ωk
This shows that momentum is conserved. Furthermore, for the transition rate
p'
p''
Hence,
( pµ + pµ0 ) ( pµ + p0µ ) = pµ00 p00µ (990)
but the electron’s momenta form a Lorentz scalar which is related to the rest
mass
pµ0 p0µ = pµ00 p00µ = m2 c2 (991)
and the photon has zero mass
pµ pµ = 0 (992)
pµ p0µ = 0 (993)
167
11.2 Scattering of Light
Kramers and Heisenberg evaluated the scattering cross-section for light incident
on atomic electrons53 . The incident photon is denoted by (k, α) and the scat-
tered photon by (k 0 , α0 ). The scattering cross-section involves the paramagnetic
interaction to second-order and the diamagnetic interaction to first-order. The
matrix elements of the diamagnetic interaction are given by
e2
0 0 0 0 0
< n l m k α | Ĥdia | nlmkα > = < n0 l0 m0 k 0 α0 | Â . Â | nlmkα >
2 m c2
e2
0 0 0 0 0 † † 0
= < n l m k α | ( ak,α ak0 ,α0 + ak0 ,α0 ak,α ) exp i ( k − k ) . r | nlmkα >
2 m c2
2 π h̄ c2
× √ ˆα (k) . ˆα0 (k 0 ) (994)
ωk ωk 0 V
where it has been assumed that only the initial and final photon are present. On
making use of the long-wavelength approximation λ a, the matrix elements
simplify to
e2
< n0 l0 m0 k 0 α0 | Ĥdia | nlmkα > ≈ < n0 l0 m0 | nlm >
2 m c2
2 π h̄ c2
× √ ˆα (k) . ˆα0 (k 0 )
ωk ωk 0 V
(995)
The scattering cross-section will be expressed in terms of a transition rate and
the transition rate will be calculated using a similar procedure to that which
was used in describing two-photon decay. An arbitrary state | ψn > can be
expressed in terms of a complete set of non-interacting states | n >
X i
| ψn > = Cn0 (t) exp − En0 t | n 0 > (996)
0
h̄
n
168
where ω = c k and ω 0 = c k 0 and the long-wavelength approximation has
been used. The small quantity η has been absorbed as a small imaginary part
to the initial state energy
En → En + i η h̄ (998)
k'α'
k''α''
n n'
(b) emission of a photon (k 0 , α0 ) followed by the absorption of the photon (k, α).
(k,α) (k',α')
(k',α')
(k,α)
n'' n''
n n' n n'
169
× < n0 l0 m0 | ˆα0 (k 0 ) . p̂ | n00 l00 m00 > < n00 l00 m00 | ˆα (k) . p̂ | nlm >
X i i
+ exp[ (En0 − En00 − h̄ω) t0 ] exp[ − (En − En00 − h̄ω 0 ) t00 ]
00
h̄ h̄
n
× < n0 l0 m0 | ˆα (k) . p̂ | n00 l00 m00 > < n00 l00 m00 | ˆα0 (k 0 ) . p̂ | nlm >
(999)
( En − En00 + h̄ ω )
< n0 l0 m0 | ˆα (k) . p̂ | n00 l00 m00 > < n00 l00 m00 | ˆα0 (k 0 ) . p̂ | nlm >
+ (1000)
( En − En00 − h̄ ω 0 )
170
where the matrix elements are given by
M = ˆα (k) . ˆα0 (k 0 ) < n0 l0 m0 | nlm >
0
1 X < n0 l0 m0 | ˆα0 (k ) . p̂ | n00 l00 m00 > < n00 l00 m00 | ˆα (k) . p̂ | nlm >
+
m 00 ( En − En00 + h̄ ω )
n
0
1 X < n0 l0 m0 | ˆα (k) . p̂ | n00 l00 m00 > < n00 l00 m00 | ˆα0 (k ) . p̂ | nlm >
+
m 00 ( En − En00 − h̄ ω 0 )
n
(1005)
On taking the limit η → 0, the first factor in the decay rate reduces to an
energy conserving delta function. Therefore, one obtains the Fermi-Golden rule
expression
2 2 2
1 2π e 2 π h̄ c2
= √ M 2 δ(h̄ωk0 + En0 − En − h̄ωk ) (1006)
τ h̄ m c2 ωk ωk 0 V
The magnitudes of the final state photon quantum numbers (k 0 ) must be inte-
grated over, since these are not measured. This integration imparts a physical
meaning to the expression for the rate which contains the Dirac delta function.
We shall assume that the direction of the scattered photon is to be measured
and that the photon is absorbed by a detector which subtends a solid angle dΩ0
to the atom. Therefore, the scattering rate is given by
2 2 Z ∞ 2
2 π h̄ c2
1 2π e V 0 0 02
= dΩ dk k √ | M |2 δ(h̄ωk0 +En0 −En −h̄ωk )
τdΩ0 h̄ m c2 ( 2 π )3 0 ωk ωk 0 V
(1007)
Since h̄ ωk0 = h̄ c k 0 , the integration over the delta function can be performed,
yielding
2 2 2
V dΩ0 ω 02 2 π h̄ c2
1 2π e
= √ | M |2 (1008)
τdΩ0 h̄ m c2 ( 2 π )3 h̄ c3 ω ω0 V
The scattering cross-section is defined as the transition rate divided by the
photon flux. The photon flux is found by noting that it has been assumed that
there is one photon per volume V so the photon density is V1 and the speed of
light is c. Hence, the photon flux is given by Vc . Therefore, the cross-section is
determined by the Kramers-Heisenberg formula
2 2 0
dσ e ω
= | M |2 (1009)
dΩ0 m c2 ω
The magnitude of the scattering rate is determined by the quantity re which
has the dimensions of length
2
e
re = (1010)
m c2
171
This quantity is often called the classical radius of the electron. The quantity
re can be expressed as
2 2
e e h̄
re = = ≈ 2.82 × 10−15 m (1011)
m c2 h̄ c mc
but one can re-write the Kronecker delta function in terms of the commutation
relation
[ xi , p̂j ] = i h̄ δi,j (1014)
Thus, one can express the scalar product as a commutator
1 X
ˆα (k) . ˆα0 (k 0 ) = ˆα (k)i [ xi , p̂j ] ˆα0 (k 0 )j
i h̄ i,j
1
= [ ˆα (k) . r , p̂ . ˆα0 (k 0 ) ] (1015)
i h̄
Since, in the dipole approximation, the diamagnetic contribution to the matrix
elements M is proportional to the overlap integral
the initial and final states must be identical if this is non-zero. Hence, the
result is equivalent to the expectation value in the state | nlm > . On replacing
the matrix elements by the expectation value and then insert a complete set of
electronic states, one finds
172
The matrix elements of r can be expressed in terms of the matrix elements of p̂
via
1
< n0 l0 m0 | p̂ | n00 l00 m00 > = < n0 l0 m0 | [ r , p̂2 ] | n00 l00 m00 >
2 i h̄
m
= < n0 l0 m0 | [ r , Ĥ0 ] | n00 l00 m00 >
i h̄
m
= ( En00 l00 m00 − En0 l0 m0 ) < n0 l0 m0 | r | n00 l00 m00 >
i h̄
(1018)
where
En00 l00 m00 − En0 l0 m0 = h̄ ωn00 n0 (1020)
Thus, the elastic scattering term in the Kramers-Heisenberg formula is given by
1 X < n0 l0 m0 | p̂ . ˆα0 (k 0 ) | n00 l00 m00 > < n00 l00 m00 | p̂ . ˆα (k) | nlm >
−
m h̄ ωnn00
n00 l00 m00
(1021)
1 X < n0 l0 m0 | p̂ . ˆα0 (k 0 ) | n00 l00 m00 > < n00 l00 m00 | p̂ . ˆα (k) | nlm >
+
m En00 − En
n00 l00 m00
(1022)
On substituting this back into the expression for the matrix elements M , one
obtains
1 X < n0 l0 m0 | p̂ . ˆα (k) | n00 l00 m00 > < n00 l00 m00 | p̂ . ˆα0 (k 0 ) | nlm >
M =
m 00 00 00 En00 − En
n l m
1 X < n0 l0 m0 | p̂ . ˆα0 (k 0 ) | n00 l00 m00 > < n00 l00 m00 | p̂ . ˆα (k) | nlm >
+
m En00 − En
n00 l00 m00
173
1 X < n0 l0 m0 | ˆα0 (k 0 ) . p̂ | n00 l00 m00 > < n00 l00 m00 | p̂ . ˆα (k) | nlm >
+
m ( En − En00 + h̄ ω )
n0 l00 m00
< n0 l0 m0 | ˆα (k) . p̂ | n00 l00 m00 > < n00 l00 m00 | p̂ . ˆα0 (k 0 ) | nlm >
1 X
+
m ( En − En00 − h̄ ω )
n00 l00 m00
(1023)
which simplifies to
< n0 l0 m0 | ˆα0 (k 0 ) . p̂ | n00 l00 m00 > < n00 l00 m00 | p̂ . ˆα (k) | nlm >
ω X
M =
m h̄ ωn00 n ( ωnn00 + ω )
n00 l00 m00
< n0 l0 m0 | ˆα (k) . p̂ | n00 l00 m00 > < n00 l00 m00 | p̂ . ˆα0 (k 0 ) | nlm >
−
ωn00 n ( ωnn00 − ω )
(1024)
In the limit of small photon frequencies compared with the electronic energies,
one can expand the denominators of the matrix element as
1 1 ω
= 2 ∓ 3 + ... (1025)
ωnn00 ( ωnn00 ± ω ) ωnn00 ωnn00
When this low-frequency expansion is substituted into the matrix elements, the
leading term vanishes. This can be seen since the leading term is proportional
to
X 1
2 < n0 l0 m0 | ˆα0 (k 0 ) . p̂ | n00 l00 m00 > < n00 l00 m00 | p̂ . ˆα (k) | nlm >
ω nn00
n00
− < n0 l0 m0 | ˆα (k) . p̂ | n00 l00 m00 > < n00 l00 m00 | p̂ . ˆα0 (k 0 ) | nlm >
(1026)
(1027)
or, on using the completeness relation, one finds the expectation value of the
commutator is given by
174
Thus, the leading term of the low-frequency expansion vanishes. Therefore, the
scattering rate is expressed as
2 3
dσ re X
4 1
= ω
dΩ0 m h̄ ωnn00
n00 l00 m00
× < n0 l0 m0 | ˆα0 (k 0 ) . p̂ | n00 l00 m00 > < n00 l00 m00 | p̂ . ˆα (k) | nlm >
2
0 0 0 00 00 00 00 00 00 0
+ < n l m | ˆα (k) . p̂ | n l m > < n l m | p̂ . ˆα0 (k ) | nlm >
(1029)
Finally, the scattering rate can be expressed in terms of the dipole matrix ele-
ments as
2 X 3
dσ re m 4
1
= ω
dΩ0 h̄ ωnn00
n00 l00 m00
× < n0 l0 m0 | ˆα0 (k 0 ) . r | n00 l00 m00 > < n00 l00 m00 | r . ˆα (k) | nlm >
2
0 0 0 00 00 00 00 00 00 0
+ < n l m | ˆα (k) . r | n l m > < n l m | r . ˆα0 (k ) | nlm >
(1030)
ω ωnn00 (1032)
so that the photon energy is greater than the atomic binding-energy. In this
case, the second and third terms in the Kramers-Heisenberg formula can be
neglected. This is because
1
ω ∼ ω0 < n0 l0 m0 | ˆα0 (k 0 ) . p̂ | n00 l00 m00 > < n00 l00 m00 | p̂ . ˆα (k) | nlm >
m
(1033)
175
Therefore, the scattering predominantly occurs elastically and the scattering
cross-section is given by
2
dσ 2
0
= r e
ˆα (k) .
ˆα0 (k ) (1034)
dΩ0
ˆ1 (k 0 ) = ( cos θk0 cos ϕk0 , cos θk0 sin ϕk0 , − sin θk0 ) (1035)
k
e2(k')
k'
e1(k')
θk'
e2(k)
φk'
e1(k)
Figure 32: The coordinate system and polarization vectors used to describe
Thomson scattering.
vectors, the scattering cross-section for incident radiation that is polarized along
176
the x-direction takes on the form
cos2 θk0 cos2 ϕk0 for α0 = 1
dσ 2
= r (1037)
0
dΩ x−pol e
sin2 ϕk0 for α0 = 2
If the incident beam has its polarization along the x-direction, and the de-
tector is not sensitive to the polarization, then the final polarization must be
summed over. In this case of a polarized beam and a polarization insensitive
detector, the cross-section is given by
dσ 2 2 2 2
= r e cos θ k 0 cos ϕk 0 + sin ϕk 0 (1038)
dΩ0 x−pol
where the polarizations of the final state photon have been summed over.
if the polarizations of the final state photons are measured. This result is iden-
tical to that obtained by assuming that the initial beam is composed of one half
of the number photons polarized along the x-direction and the other half of the
number of photons polarized along the y-direction. That is
2
for α0 = 1
1 2 2
dσ 2 2 cos θk ( cos ϕk + sin ϕk )
0 0 0
= r 2 (1040)
0
dΩ unpol e 1 2
2 ( sin ϕk + cos ϕk )
0 0 for α0 = 2
177
m 2
these processes are smaller by factors of ( M ) . The derivation of the Thomson
scattering cross-section breaks down for photons which have energies of the
order of the electron’s rest energy
h̄ ω ∼ me c2 (1043)
For photons with these high-energies, one must describe the scattering process
relativistically. In this energy region, Compton scattering dominates.
Classical Interpretation
In the first process, an electron bound harmonically to the atom which re-
sponds to an electromagnetic field E 0 exp[ i ω t ] can be described by the
equation of motion
q
r̈ + ω02 r = E 0 <e exp i ω t (1044)
m
where ω0 is the frequency of the electron’s natural motion. In the steady state,
one finds q
m E0
r = 2 <e exp i ω t (1045)
ω0 − ω 2
The acceleration of the charged particle can be described by
q
ω2 E 0
r̈ = − m2 <e exp i ω t (1046)
ω0 − ω 2
The accelerating charged particle radiates electromagnetic energy. The emitted
power is given by the Larmor formula
2 q2 r2 ω4
P (ω) =
3 c3
2 q E 20
4
ω4
= 2 (1047)
3 m c ( ω0 − ω 2 )2
2 3
8π 2 ω4
σ = re 2 (1049)
3 ( ω0 − ω 2 )2
178
This formula has the correct frequency dependence in the limit ω ω0 in which
case the classical cross-section varies as ω 4 , as expected for Rayleigh scattering.
On the other hand, in the limit ω ω0 the cross-section becomes frequency
independent, as is expected for Thomson scattering.
Since it is most probable that the initial electron is in the ground state, one has
h̄ ω > h̄ ω 0 (1052)
Hence, the final photon has less energy than the initial photon. That is, the
Stokes
n → n'
I(ω')
anti-Stokes
n' → n
Figure 33: The schematic frequency dependence of the observed intensity ex-
pected in a Raman scattering experiment. The ratio of intensities of the Stokes
and anti-Stokes lines provides a relative measure of the initial occupation of the
low-energy state n and the higher-energy excited state n0 .
electromagnetic field has lost energy and left the electron in an excited state.
This inelastic process describes the Stoke’s line. On the other hand, if the
electron is initially in an excited state, then it is possible that the electron loses
energy and makes a transition to the ground state. In this case,
179
so the final photon is more energetic
h̄ ω < h̄ ω 0 (1054)
since γ is related to the decay rate and is of the order of 108 sec−1 , it is usually
negligible compared with the frequency of light which is estimated as ω ∼ 1015
sec−1 . Following our previous arguments, one finds that the scattering cross-
section is given by
2 2
8π q ω4
σ(ω) = (1057)
3 m c2 ( ω 2 − ω02 )2 + γ 2 ω 2
which no longer diverges when the resonance condition is satisfied, because of
the damping of the electronic states.
180
so
En = En(0) + ∆En (1060)
Hence, due to the form of the expressions for the shift and the lifetime as the
real and imaginary parts of a complex function, it is possible to consider an
unstable state as having a complex energy54 given by
h̄
En − i Γn ≈ En(0) + ∆En − i (1061)
2 τn
That is, the lifetime can be considered as giving the state an energy-width Γn .
This is the natural width of the electronic state. The factor of two in the width
can be understood by considering the time-dependence of the state | ψn (t) >
which is given by
i
| ψn (t) > = exp − ( En − i Γn ) t | ψn (0) > (1062)
h̄
Hence, the probability Pn (t) that the state has not decayed at time t is given
by
Pn (t) = | < ψn (0) | ψn (t) > |2
i
= | < ψn (0) | exp − ( En − i Γn ) t | ψn (0) > |2
h̄
2
= exp − Γn t (1063)
h̄
due to the normalization of the initial state. This time-dependence of Pn (t) is
interpreted in terms of the exponential decay of the probability for finding the
initial state
t
Pn (t) = Pn (0) exp − (1064)
τn
This leads to the identification of the relation between the energy-width and
the lifetime
h̄
Γn = (1065)
2 τn
Hence, the lifetime τn of an unstable or metastable state can be incorporated
by introducing an imaginary part Γn to the energy.
Therefore, for the case of resonant scattering, one should replace the energies
by complex numbers such that the real part represents the state’s energy and
the imaginary part describes half the state’s decay rate. In the case of resonant
scattering, the Kramers-Heisenberg formula is modified55 to
2 2 0
dσ e ω
0
= 2
| M |2 (1066)
dΩ mc ω
54 That is, the perturbation produces a complex shift of the energy-shift which related to
181
where the matrix elements are given by
M = ˆα (k) . ˆα0 (k 0 ) < n0 l0 m0 | nlm >
1 X < n0 l0 m0 | ˆα0 (k 0 ) . p̂ | n00 l00 m00 > < n00 l00 m00 | ˆα (k) . p̂ | nlm >
+
m ( En − En00 − i Γn00 + h̄ ω )
n00 l00 m00
< n0 l0 m0 | ˆα (k) . p̂ | n00 l00 m00 > < n00 l00 m00 | ˆα0 (k 0 ) . p̂ | nlm >
1 X
+
m ( En − En00 − i Γn00 − h̄ ω 0 )
n00 l00 m00
(1067)
2
Since close to resonance, the resonant denominator is given by Γ ∼ h̄ac ( h̄e c )4
2
whereas the numerator is of the order of ea . Hence, on-resonance, the matrix
2
elements can be of the order ( h̄e c )−3 larger than the non-resonant matrix ele-
ments. Therefore, on resonance, the non-resonant terms may be neglected. In
the following, it shall be assumed that the resonant state is non-degenerate
| < n0 l0 m0 | ˆα0 (k 0 ) . p̂ | n00 l00 m00 > < n00 l00 m00 | ˆα (k) . p̂ | nlm > |2
2 2 0
dσ e ω
=
dΩ0 m2 c2 ω ( En − En00 + h̄ ω )2 + Γ2n00
(1068)
This expression can be re-expressed in terms of the product of two factors
which is the probability for absorption from the ground state to the resonant
state | n00 l00 m00 > (divided by the incident flux) times the probability for its
decay via emission. On resonance, it appears that the process corresponds to
two sequential processes, first absorption and secondly emission.
The difference between a resonant process and two step process, is deter-
mined by the lifetime of the intermediate state | n0 l0 m0 > compared with the
frequency width of the photon beam. The frequency width of the photon beam
may be limited by the monochromator, or by the time-scale of the experiment
if it involves a pulsed light source. If the lifetime of the intermediate state is
sufficiently long compared with the time scale of experiment, it may be possi-
ble to observe the decay long after the incident light has been switched off. In
182
this case, the resonance can be considered to be composed of two independent
processes56 . Furthermore, it may be possible to perform further experiments
on the surviving intermediate state. In the opposite case, where the lifetime
of the intermediate state is shorter than the time-scale of the experiment, the
intermediate state will have decayed before the experiment has terminated.
where the decay includes transitions to all possible final states, one finds
1 1 1
ρn (E) = −
2πi E − En − i 2 h̄τn E − En + i 2 h̄τn
h̄
1 2 τn
= (1072)
π ( E − En )2 + ( 2 h̄τ )2
n
This can only be an approximate form of the energy-distribution since the en-
ergy must be bounded from below. The existence of a lower-bound to en-
ergy distribution implies that the width of the electronic energy level has to
be energy-dependent 2 h̄τn = Γn as this must become zero below a threshold
energy. However, it should be noted that the width of the energy-distribution
will determine the approximate exponential decay. Since the perturbations in-
troduce an energy-dependent width to the wave packet, causality requires that
the energy-shift ∆En should also be energy-dependent. Hence, the effects of
the perturbation (such as the energy-shift and lifetime) should be described in
56 V. Weisskopf, Ann. der Physik, 9, 23 (1931).
183
terms of a self-energy Σn (E)
This complex self-energy has a real and imaginary part. The imaginary part
can be thought of as occurring via amplification of the infinitesimal imaginary
term i η in the denominator, and can be seen to be non-zero when the energy
E of the component in the wave packet falls in the region when the spectral
density of the approximate En0 is finite. Hence, since the En0 are bounded from
below, then so is the energy-distribution ρn (E) since
1 =m Σn (E + iη)
ρn (E) = − (1075)
π ( E − En − <e Σn (E) )2 + ( =m Σn (E + iη) )2
The real part of the self-energy must also be energy-dependent, since it is related
to the imaginary part via the Kramer’s-Kronig relations
Z ∞
1 =m Σn (z + iη)
<e Σn (E) = − dz
π −∞ E − z
Z ∞
Pr <e Σn (z)
=m Σn (E + iη) = + dz (1076)
π −∞ E − z
Hence, the real part of the self-energy is also energy-dependent. The Kramers-
Kronig relation is an expression of causality.
Since the electronic states in the expression for the Fermi-Golden rule decay
rate
1 2π
= | < n0 l0 m0 | ĤI | nlm > |2 δ( En0 l0 − Enl − h̄ ω ) (1078)
τ nl→n0 l0 h̄
are to be interpreted as wave packets with a distribution of energies, the factor
expressing conservation of energy should be expressed in terms of the energy
184
conservation for the components of the wave packets. Hence, the decay rate
should be written as the convolution
Z ∞ Z ∞
1 2π
= | < n0 l0 m0 | ĤI | nlm > |2 dE 0 ρn0 l0 (E 0 ) dE ρnl (E) δ( E 0 − E − h̄ ω )
τ nl→n0 l0 h̄ −∞ −∞
Z ∞
2π 0 0 0 2
= | < n l m | ĤI | nlm > | dE ρn0 l0 ( E + h̄ ω ) ρnl (E) (1079)
h̄ −∞
We shall use the approximation for the energy distributions suggested by eqn(1072).
In this case, the convolution is evaluated by contour integration as
h̄ h̄
1 2 2 τn + 2 τn0 l
= | < n0 l0 m0 | ĤI | nlm > |2
τ nl→n0 l0 h̄ ( h̄ ω + Enl − En0 l0 )2 + ( 2 h̄τnl + h̄
2 τn0 l0 )
2
(1080)
since only the terms with poles on the opposite sides of the real-axis yield
non-zero contributions. From this, one can show that the optical absorption
cross-section is given by
1 1
4π
e2
X ωn0 l0 ,nl ( 2 τn0 l0 + 2 τnl )
σabsorb (ω) = | < n0 l0 m0 | r | nlm > |2 1 1
3 h̄ c
n0 l0 m0
( ωn0 l0 ,nl − ω )2 + ( 2 τn0 l0 + 2 τnl )2
(1081)
which was first derived by Weisskopf and Wigner57 . Hence, the natural width
is given by the average of the decay rates for the initial and final electronic
states. This leads to the conclusion that even weak lines can be broad, if the
final electronic state has a short lifetime.
185
measurable quantities. In Quantum Electrodynamics, the infinities cancel in
equations which only contain physical measurable quantities. This fortunate
circumstance makes the theory of Quantum Electrodynamics renormalizable.
First, it shall be shown how the infinite zero-point energy of the electro-
magnetic field can lead to a (finite) physically measurable force between its
containing walls.
L-d d
Figure 34: The geometry of the partitioned electromagnetic cavity used to con-
sider the Casimir effect.
evaluate the total energy of this configuration and then deduce the form of the
interaction between the partition and the walls of the cavity.
We shall consider the total energy due to the zero-point fluctuations in the
container. Since the zero-point energy is divergent due to the presence of arbi-
trarily large frequencies, we shall introduce a convergence factor. The introduc-
tion of a convergence factor to remove infinities is the process of regularization.
The introduction of the convergence factor can be motivated by the observation
that, in matter, electromagnetic radiation becomes exponentially damped at
58 H. B. G. Casimir, Physica 19, 846 (1953).
186
large frequencies. Hence, one can write
1 X ωk,α
E = h̄ ωk,α exp − λ (1082)
2 c
k,α
187
√
Let t = κ + 1 so
∞
2 L2 dt ∂ 3
Z
1
Ed = − h̄ c π λ
(1090)
4π 1 t ∂λ3 exp[ d t] − 1
The factor of t−1 can be eliminated by performing one of the differentials with
respect to λ.
Z ∞
h̄ c L2 ∂2 exp[ πdλ t ]
Ed = dt (1091)
2d 1 ∂λ2 ( exp[ πdλ t ] − 1 )2
We shall set
πλt
s = exp[ ] − 1 (1092)
d
therefore
∞
h̄ c L2 ∂ 2
Z
d ds
Ed = (1093)
2 d ∂λ2 πλ s0 s2
h̄ c L2 ∂ 2
d
Ed =
2 d ∂λ2 π λ s0
d
h̄ c L2 ∂ 2
π λ
=
2 d ∂λ2 exp[ πdλ ] − 1
2 π λ
h̄ c L2 ∂ 2
d d
= (1095)
2 d ∂λ2 πλ exp[ πdλ ] − 1
188
where the n = 0 term diverges as λ−4 in the limit as λ → 0 and is proportional
to the volume of the cavity d L2 . The term with n = 1 also diverges, but
diverges as λ−3 and has the form of a surface energy since it is proportional to
L2 . The terms with n = 2 and n = 3 are identically equal to zero. The term
with n = 4 remains finite in the limit λ → 0 and all the higher-order terms
vanish in this limit. Explicitly, one has
h̄ c L2 6 B0 d2 2 B4 π 2
2 B1 d
Ed = + + + O(λ) (1098)
2d π 2 λ4 π λ3 4! d2
The first term in the energy is proportional to L2 d, which is the volume of the
cavity and the second term is proportional to L2 the surface area of the walls.
The third term is independent of the cut-off and the higher order terms vanish
in the limit λ → 0.
The Casimir force is the force between two planes, which originates from
the energy of the field59 . This energy can be separated out into a volume
part and parts due to the creation of the surfaces and an interaction energy
between the surfaces. In order to eliminate both the volume dependence of the
energy and the surface energies, we are considering two configurations of the
partitions in the cavity. In one configuration the plane divides the volume into
two unequal volumes d L2 and (L − d) L2 , and the other configuration is a
reference configuration where the cavity is partitioned into two equal volumes
L3
2 . The difference of energies for these configurations is given by
∆E = Ed + EL−d − 2 E L (1099)
2
π 2 h̄ c L2
lim ∆E → − (1100)
Ld,λ→0 720 d3
The d-dependence of the energy difference leads to an attractive force between
the two plates separated by a distance d, which is the Casimir force
π2 L2
F = − h̄ c 4 (1101)
240 d
The force is proportional to L2 which is the area of the wall of the cavity. The
predicted force was measured by Sparnaay60 . A more recent experiment involv-
ing a similar force between a planar surface and a sphere has achieved greater
59 Our considerations only include the part of Fock space that corresponds to having zero
numbers of excited quanta. Hence, the Casimir force is due to the properties of the field, and
is not due to the transmission of real particles (photons) between the planes.
60 M. J. Sparnaay, Physica 24, 751 (1958).
189
ter such a large perturbation, the feedback system required 6 3 1014 Hz, Eq. (5) gives a correction of order 20%
several minutes to reestablish equilibrium. at the closest spacings; our data does not support such
Assuming that the functional form for the Casimir force a deviation. However, the simple frequency dependence
is correct, its magnitude was determined by using linear of the electrical susceptibility used in the derivation of
least squares to determine a parameter d for each sweep Eq. (5) is not correct for Au, the index of refraction of
such that which has a large imaginary component above the plasma
Fcm sai d s1 1 ddFcT sai d 1 b 0 . (9) frequency; a rough estimate using the tabulated complex
index [14] limits the conductivity correction as no larger
In this context, b 0 should be zero, and for the complete than 3%, which is consistent with our results [15].
data set, b 0 , 5 3 1027 dyn (95% confidence level). I thank Dev Sen (who was supported by the UW NASA
The average over the 216 sweeps gives d 0.01 6 0.05, Space Grant Program) for contributions to the early stages
and this is taken as the degree of precision of the of this experiment, and Michael Eppard for assistance
measurement. There was no evidence for any variation with calculations.
of d depending on the region of the plates used for the
measurement.
The most striking demonstration of the Casimir force
is given in Fig. 4. The agreement with theory, with no
*Present address: Los Alamos National Laboratory,
Neutron Science and Technology Division P-23, M.S.
H803, Los Alamos, NM 87545.
[1] H. B. G. Casimir, Koninkl. Ned. Adak. Wetenschap. Proc.
51, 793 (1948).
[2] E. Elizalde and A. Romeo, Am. J. Phys. 59, 711 (1991).
[3] V. M. Mostepanenko and N. N. Trunov, Sov. Phys. Usp.
31, 965 (1988).
[4] M. J. Sparnaay, Physica (Utrecht) 24, 751 (1958).
[5] C. I. Sukenik, M. G. Boshier, D. Cho, V. Sangdohar, and
E. A. Hinds, Phys. Rev. Lett. 70, 560 (1993).
[6] E. M. Lifshitz, Sov. Phys. JETP 2, 73 (1956).
[7] T. H. Boyer, Phys. Rev. 174, 1764 (1968).
[8] J. Blocki, J. Randrup, W. J. Swiatecki, and C. F. Tsang,
Ann. Phys. (N.Y.) 105, 427 (1977).
[9] J. Schwinger, L. L. DeRaad, Jr., and K. A. Milton, Ann.
Phys. (N.Y.) 115, 1 (1978).
[10] J. Mehra, Physica (Utrecht) 37, 145 (1967).
[11] L. S. Brown and G. J. Maclay, Phys. Rev. 184, 1272
(1969).
[12] G. Ising, Philos. Mag. 1, 827 (1926).
[13] W. R. Smythe, Static and Dynamic Electricity (McGraw-
FIG. 4. Top: All data with electric force subtracted, averaged Hill, New York, 1950), pp. 121 –122.
into bins (of varying width), compared to the expected Casimir [14] CRC Handbook of Chemistry and Physics, 76th Ed. (CRC
Figure 35: The separation-dependent force between two closely spaced metallic
force for a 11.3 cm spherical plate. Bottom: Theoretical
Casimir force, without the thermal correction, subtracted from
Press, Boca Raton, 1995), pp. 12 –-130.
[15] S. Hacyan, R. Jauregui, F. Soto, and C. Villarreal, J. Phys.
surfaces due to the modification of the zero-point energy. The lower panel shows
top plot; the solid line shows the expected residuals. A 23, 2401 (1990).
the difference between 8the experimental results and the theoretical prediction
for the Casimir Force. [After S. K. Lamoreaux, Phys. Rev. Lett. 78, 5, (1997).]
accuracy61 .
Cut-Off Independence
It is the boundary condition and not the cut-off that plays an important role
in the Casimir effect. For simplicity, one can choose zero boundary conditions.
The zero-point energy of a cylindrical electromagnetic cavity of radius R and
length d can be expressed as the sum
s 2 s
∞ Z 2
h̄ c π R2 X
2
π n z 2
π nz
Ed = 2 dkρ kρ kρ + F kρ +
2 ( 2 π ) n =1 d d
z
(1102)
where F (z) is an arbitrary cut-off function (which may depend on an arbitrary
parameter λ which is ultimately going to be set to zero). The cut-off must not
effect the low energy-modes so one can choose F (0) = 1 and all the derivatives
of F (z) to be zero for finite values of z. These assumptions are all in accord with
61 S. K. Lamoreaux, Phys. Rev. Lett. 78, 5 (1997).
62 The independence of any cut-off procedure can be shown by evaluating the divergent sums
by using the Euler-Maclaurin summation formula.
190
1.2
0.8
F(z)
0.4
0
0 0.2 0.4 0.6 0.8 1
z/N
the ideal case of no cut-off function or F (z) = 1. The energy can be written as
∞
h̄ c R2 X
Ed = f (nz ) (1103)
2 n =1 z
where
s 2 s 2
Z ∞
π nz π nz
f (nz ) = dkρ kρ kρ2 + F kρ2 + (1104)
0 d d
The summation can be performed by changing it into an integral, however the
corrections due to smoothing will be kept. This is accomplished by the Euler-
Maclaurin formula. The integral between 0 and N of a function can be roughly
expressed as a summation
Z N N −1
1 X 1
dx f (x) ≈ f (0) + f (n) + f (N ) (1105)
0 2 n=1
2
191
We shall assume that f (n) and all its derivatives vanishes in the limit of large
n, limN →∞ f (N ) → 0, due to the behavior of the cut-off function. The
corrections in the Euler-Maclaurin summation formulae can be evaluated by
noting that the first derivative of f (n) with respect to n is given by
Z ∞ s 2
π2 n
(1) kρ 2 +
πn
f (n) = dk ρ F kρ (1107)
d2
q
0 k 2 + ( π n )2 d
ρ d
since the derivatives of F (z) all vanish for finite z. The integration over the
variable kρ is re-expressed in terms of an integration over the variable z, defined
by s 2
πn
z = kρ2 + (1108)
d
so
kρ
dz = dkρ s (1109)
2
π n
kρ2 + d
π 3 n2
= − (1110)
d3
In deriving the above expression, the condition that the first-order derivative of
F (z) vanishes for finite z has been used. It immediately follows that
2 π3 n
f (2) (n) = − (1111)
d3
and
2 π3
f (3) (n) = − (1112)
d3
and all higher order derivatives vanish. Hence, one finds that at z = 0 all the
m-th order derivatives f (m) (0) vanish, except for m = 3 which is given by
2 π3
f (3) (0) = − (1113)
d3
192
Hence, on evaluating the energy of the cylindrical cavity (and using the zero
boundary conditions), one finds that the energy is composed of the sum of an
integral and a finite number of other terms. The integral part of the expression
only depends on the volume of the cavity and is proportional to a divergent
integral, and hence drops out when the energy differences are taken. The only
terms that yield non-zero contributions to the energy difference originate with
f (3) (0) and depend on d. It is these terms that give rise to the Casimir force.
This approach also showed that any particular choice made for the cut-off is
irrelevant.
Mathematical Interlude:
The Euler-Maclaurin Summation Formula.
193
The remainder R when the series is truncated after p terms is given by
Z N
Pp+1 (x)
R = (−1)p dx f (p+1) (x) (1118)
0 (p + 1)!
where Pn (x) = Bn (x − [x]) are the periodic Bernoulli polynomials. The remain-
der term can be estimated as
Z N
2
|R| ≤ dx | f 2p−1 (x) | (1119)
(2π)p 0
Derivation by Induction
First we shall examine the properties of the Bernoulli polynomials and the
Bernoulli numbers. Then we shall indicate how the Euler-Maclaurin formula
can be obtained by induction.
where Bn are the Bernoulli constants. Hence, the Bernoulli constants are the
Bernoulli polynomials evaluated at x = 0, i.e. Bn (0) = Bn . Furthermore, on
differentiating the generating function w.r.t. x, one finds
∂G(z, x)
= z G(z, x) (1122)
∂x
which implies that
∞ ∞
X ∂Bn (x) z n X zn
= z Bn (x) (1123)
n=0
∂x n! n=0
n!
On equating the coefficients of z n in the above equation, one obtains the im-
portant relation
∂Bn (x)
= n Bn−1 (x) (1124)
∂x
Therefore, by integration it easy to show that Bn (x) are polynomials of degree
n. The first few Bernoulli polynomials can be explicitly constructed from the
194
generating function expansion. The first few polynomials are given by
B0 (x) = 1
1
B1 (x) = x−
2
1
B2 (x) = x2 − x +
6
3 1
B3 (x) = x3 − x2 + x
2 2
1
B4 (x) = x4 − 2x3 + x2 −
30
5 5 1
B5 (x) = x5 − x4 + x3 − x
2 3 6
... (1125)
From the generating function expansion, one can show that the Bernoulli poly-
nomials are either even or odd functions of x − 12 . The generating function can
be expressed
∞
zn
z(x− 12 ) z X
G(z, x) = e z z = Bn (x) (1126)
e − e− 2
2
n=0
n!
where the second factor is an even function of z, thus, the generating function is
invariant under the combined transformation z → −z and (x − 12 ) → −(x − 12 ).
Therefore, one has
∞ ∞
zn zn
X 1 1 X 1 1
Bn +x− = Bn + − x ( − 1 )n (1127)
n=0
2 2 n! n=0
2 2 n!
Bn (x) = ( − 1 )n Bn (1 − x) (1128)
The generating function with x = 0 can be re-written as the sum of its even
and odd parts
z ∞
zn
2 z X
G(z, 0) = − = Bn (0) (1130)
tanh z2 2 n=0
n!
The even part has only even terms in its Taylor expansion, and there is only
one term in the odd part. Hence, the odd Bernoulli numbers vanish for n > 1,
i.e. B2n+1 (0) = 0 for n > 0. Therefore, for n ≥ 2, one has Bn (0) = Bn (1). This
equality can be used to evaluate the integrals of the Bernoulli polynomial over
195
the range from 0 to 1. On expressing the integral of Bn (x) in terms of Bn+1 (x),
one has
Z 1 Z 1
1 ∂Bn+1 (x)
dx Bn (x) = dx
0 (n + 1) 0 ∂x
Bn+1 (1) − Bn+1 (0)
=
( n+1 )
= 0 for n ≥ 1 (1131)
Hence, the Bernoulli polynomials may be defined recursively via the relation
∂Bn (x)
= n Bn−1 (x) (1132)
∂x
if the constant of integration is fixed by
Z 1
dx Bn (x) = 0 for n ≥ 1 (1133)
0
where [x] is the integral part of x. This definition of Pn (x) reproduces to the
Bernoulli polynomials on the interval (0, 1) since [x] = 0 in this interval. The
functions Pn (x) are periodic over an extended range of x with period 1.
v = P1 (x) (1138)
196
but since the periodic Bernoulli polynomial P1 (x) is given by
1
P1 (x) = (x − [x]) − (1140)
2
it has the value of 1/2 at the limits of integration. Hence, the integration reduces
to
Z n+1 Z n+1
f (n + 1) + f (n) ∂f (x)
dx f (x) = − dx P1 (x) (1141)
n 2 n ∂x
N Z N Z N
X f (1) + f (N ) ∂f (x)
f (n) = dx f (x) + + dx P1 (x) (1143)
n=1 1 2 1 ∂x
The last two terms, therefore, give the error when the sum is approximated by
an integral. The first correction is simply the end point corrections from the
“trapezoidal rule”, and the second correction has to be evaluated to yield the
Euler-Maclaurin formula. The last correction is of the form of an integral which
can be expressed in terms of the sum of the integrals
Z n+1
dx f 0 (x) P1 (x) (1144)
n
where the prime refers to the derivative of f (x) w.r.t. x. The above expression
can be evaluated by integrating by parts. The integrand is re-written as
Z n+1 Z n+1
∂v
dx f 0 (x) P1 (x) = dx u (1145)
n n ∂x
u = f 0 (x)
∂v
= P1 (x) (1146)
∂x
Since the indefinite integral is evaluated as
Z x
1
dx0 P1 (x0 ) = P2 (x) (1147)
2
197
the integration by parts yields
Z n+1 n+1 Z n+1
P2 (x) f 0 (x)
0 1
dx P1 (x) f (x) = − dx f 00 (x) P2 (x) (1148)
n 2 n 2 n
However, one has P2 (0) = P2 (1) = B2 , therefore the above expression simplifies
to
Z n+1 0 Z n+1
f (n + 1) − f 0 (n)
0 1
dx P1 (x) f (x) = B2 − dx f 00 (x) P2 (x)
n 2 2 n
(1149)
Then, on summing the above expression from n = 1 to n = N − 1, one finds
Z N 0 Z N
f (N ) − f 0 (1)
0 1
dx P1 (x) f (x) = B2 − dx f 00 (x) P2 (x) (1150)
1 2 2 1
This yields the first term in the series of end point corrections in the Euler-
Maclaurin formula, where the correction is the sum of the first derivatives at
the end points multiplied by B2 /2!. The above process can be iterated yielding
a complete proof of the Euler-Maclaurin summation formula.
In order to get bounds on the size of the error when the sum is approximated
by the integral, we note that the Bernoulli polynomials on the interval [0, 1] at-
tain their maximum absolute values at the endpoints and the value Bn (1) is the
n-th Bernoulli number.
References
198
magnitude of the shifts should be much smaller, as the magnitude varies as n−3 .
e2
V (r) = − (1153)
r
the Laplacian is related to a point charge density at the nucleus
Hence, the shift due to the fluctuations in the electron’s potential energy occurs
primarily at the origin. The effect of the electromagnetic fluctuations on the
kinetic energy are not state specific, and can be considered as a uniform shift
of all the energy levels, like the electron’s rest mass energy m c2 . Thus, the
relative energy shift of the levels is solely determined by the potential at the
origin. Therefore, the states with non-zero angular momenta do not experience
the relative energy-shift since the electronic wave functions vanish at the origin.
Thus, only the 2s state experiences a shift but the 2p state is unshifted.
199
∆r(t)
where Eω is the Fourier component of the fluctuating electric field. Hence, the
ω-component of the mean squared fluctuation66 in the particle’s position is given
by
2
q Eω2
< | ∆r2ω | > = < | 2 | > (1157)
m ( ω0 − ω 2 )2
On approximating the electromagnetic energy associated with the fluctuating
electromagnetic field < | Eω2 | > by the half the sum of the zero-point energies
of the photon modes, one has
V 1
< | Eω2 | > = 2 h̄ ω (1158)
8π 4
where the factor 2 represents the two types of polarization of the normal modes.
Therefore, on summing over the normal modes, one finds that the mean squared
deviation of the electron’s trajectory from the classical orbit is proportional to
Z ∞
Eω2
Z
V
3
dΩ dω ω 2 < | 2 | >
(2πc) 0 ( ω0 − ω 2 )2
Z ∞
ω3
Z
4 π h̄ V
= 3
dΩ dω 2 (1159)
V (2πc) 0 ( ω0 − ω 2 )2
The integration over ω can be approximated as
mc2
m c2
Z h̄ dω
= ln (1160)
ω0 ω h̄ ω0
where an upper and lower cut-off have been introduced to prevent the integral
from diverging67 . The expectation value of the second derivative of the potential
66 The average squared fluctuation of the electromagnetic field should, in principle, be cal-
culated as an average over a volume in time and space which encompasses the electron’s
trajectory.
67 The upper limit can be considered as being determined by the spatial dimension of the
volume in which the electromagnetic fluctuations are being averaged over. The divergence at
the lower limit of integration is unphysical and is caused by the neglect of the lifetimes of
the electronic states. The inclusion of the lifetimes result in the integrand being finite at the
resonance frequency ω0
200
for the 2s state is given by
1
< | ∇2 V | > = 4 π e2 (1161)
π a3
where the second factor represents the 2s electron density at the origin. The
corresponding factor for an ns level is expected to vary proportionally to n−3 .
Combining the above expressions, one finds that the 2s level is shifted by an
energy given by
2 3
m e4 m c2
4 e
∆E2s = ln (1162)
2 π h̄ c h̄2 h̄ ω0
where the frequency of the electron’s orbit ω0 has been chosen as a lower cut-off
on the frequency of the electromagnetic fluctuations. Since the logarithmic fac-
e2
tor is approximately given by ∼ −2 ln h̄c , one can see that the above estimate
is consistent with the Lamb shift having a magnitude of approximately 4.372
×10−6 eV.
The lowest-order correction to the electron’s energy comes from the diamag-
netic interaction. From first-order perturbation theory, one finds the correction
201
(k,α)
q q
Figure 38: The first-order correction to the rest mass of the electron due to the
diamagnetic interaction.
since the electronic matrix elements give rise to the condition of conservation of
momentum. Hence, the correction to the energy is found as
e2
Z
(1) V 2 π h̄ c 1
∆Eq = 2 3
2 d3 k
2mc (2π) V k
Z ∞
e2
V 2 π h̄ c
= 8π dk k
2 m c2 ( 2 π )3 V 0
2 Z ∞
e h̄
= dk k (1166)
πmc 0
X < q {0} | Ĥpara | q 0 1k,α > < q 0 1k,α | Ĥpara | q {0} >
∆Eq(2) =
0
Eq0 + h̄ ωk − Eq
q ,k,α
(1167)
202
(k,α)
q-k
q q
| h̄ q . ˆα (k) |2
2 X
2 π h̄ c2
e
∆Eq(2) = 2 2 h̄2 (q−k)2
(1169)
m c V ωk 2
h̄ q 2
− − h̄ ω
k,α 2 m 2 m k
203
one finds that the numerator is given by
X
h̄2 | q . ˆα (k) |2 = h̄2 q 2 ( 1 − cos2 θ ) (1171)
α
q . k = q k cos θ (1172)
h̄2 q 2 ( 1 − cos2 θ )
2
2 π h̄ c2
Z
(2) e V 3
∆Eq = d k h̄2 (q−k)2
m2 c2 ( 2 π )3 V ωk h̄2 q 2
2 m − 2 m − h̄ ωk
∞ π 2
e2 h̄ h̄ q 2 ( 1 − cos2 θ )
Z Z
= 2
dk k dθ sin θ h̄2 q k 2 2
2πm c 0 0 cos θ − h̄ k − h̄ c k
m 2 m
(1173)
(1174)
h̄2 q 2 8
2
(2) e h̄
∆Eq = − ln (1176)
2 m 3 π h̄ c 2 m c λ+
204
11.3.4 The Self-Energy of a Bound Electron
The Lamb shift (a quantum electrodynamic shift of the 2s level of Hydrogen
upwards by 1058 MHz) is caused the self-energy of a bound electron. The
self-energy of the state nlm can be estimated from second-order perturbation
theory using the dipole approximation, as is appropriate for a completely non-
relativistic calculation. The second-order shift is given by
2 π h̄ c2 | < n0 l0 m0 | ˆα (k) . p̂ | nlm > |2
2
(2) e X
∆Enlm =
m2 c2 0 0 0
V ωk Enlm − En0 l0 m0 − h̄ ωk
k,α,n l m
(1178)
On summing over the polarizations using the completeness relation, one obtains
(2)
2
e h̄ c
Z ∞ Z π X | < n0 l0 m0 | p̂ | nlm > |2 ( 1 − cos2 θk )
∆Enlm = dk k dθ k sin θ k
m2 c2 ( 2 π ) 0 0 Enlm − En0 l0 m0 − h̄ ωk
n0 l0 m0
(1179)
where θk is the angle subtended between k and the matrix elements of p. The
angular integration can be performed, yielding
(2)
2
e 2 h̄ c
Z ∞ X | < n0 l0 m0 | p̂ | nlm > |2
∆Enlm = dk k
m2 c2 3π 0 Enlm − En0 l0 m0 − h̄ ωk
n0 l0 m0
(1180)
In the completely non-relativistic limit, the integration over k can be shown to
be linearly divergent at the upper limit of integration.
Hans Bethe argued70 that, within the same dipole approximation, the cor-
rection to the kinetic energy of the electron in the state | nlm > is given by an
expression analogous to that of an electron in a continuum state n
2 2 Z ∞ X | < n0 | p̂ | n > |2
(2) 2 e h̄
∆Tn = dω ω (1181)
3 π h̄ c mc 0 0
En − En0 − h̄ ω
n
Since momentum is conserved for continuum states (on average), only the state
where n = n0 contribute so the denominator simplifies71 . The expression for the
mass renormalization is divergent and is given by
2 2 Z ∞ X | < n0 | p̂ | n > |2
2 e h̄
∆Tn(2) = − dω ω
3 π h̄ c mc 0 h̄ ω
n0
2 Z ∞
4 e h̄ dω ω < n | p̂2 | n >
= − 2
(1182)
3 π h̄ c mc 0 ω 2m
where the completeness relation has been used. This expression is valid if n
labels either a continuum or a discrete state, since only the mass of the electron
70 H. A. Bethe, Phys. Rev. 72, 339 (1947).
71 Since we are now using the dipole approximation, the recoil of the free electron which was
taken into account in our previous analysis is now being ignored. [See the denominator of the
first line of eqn(1173).]
205
is being altered and the expectation value of p̂ is unaltered. Thus, Bethe argued,
the kinetic energy of an electron in a bound state which has the physical mass
m∗ should be approximated as
p̂2 p̂2
< nlm | | nlm > = < nlm | | nlm >
2 m∗ 22m Z ∞
dω ω < nlm | p̂2 | nlm >
4 e h̄
−
3 π h̄ c m c2 0 ω 2m
(1183)
p̂2
Ĥ0 = + V (r) (1184)
2m
and the unperturbed energy of the state | nlm > of a hypothetical electron
with mass m is calculated in the non-relativistic Schrödinger theory as
(0) p̂2
Enlm = < nlm | | nlm > + < nlm | V (r) | nlm > (1185)
2m
However, in order to obtain a sensible numerical value for the approximate
(0)
energy, Enlm has to be expressed in terms of the observed physical mass m∗ .
Therefore, the bare Hamiltonian has to be expressed in terms of the physical
mass and compensating radiative corrections to the mass
(0) p̂2
Enlm = < nlm | ∗
| nlm > + < nlm | V (r) | nlm >
22m Z ∞
dω ω < nlm | p̂2 | nlm >
4 e h̄
+ 2
3 π h̄ c mc 0 ω 2m
2
p̂
= < nlm | ∗
| nlm > + < nlm | V (r) | nlm >
2m
2 Z ∞
dω ω X | < n0 l0 m0 | p̂ | nlm > |2
4 e h̄
+
3 π h̄ c m c2 0 ω 0 0 0
2m
nlm
(1186)
The completeness relation was used in obtaining the last line. The second term
in the above expression for the unperturbed energy is the correction due to
the mass renormalization72 which should be combined with the second-order
radiative correction. The total energy (to second-order) is given by
(0) (2)
Enlm = En,l,m + ∆En,l,m
72 Renormalization is an idea which Bethe attributed to H. A. Kramers. Kramers had
proposed that physical quantities should be expressed in terms of observable quantities, with
all mention of bare quantities removed. Kramers was advocating a classical treatment from
which Bethe created a non-relativistic quantum treatment.
206
p̂2
= < nlm | | nlm > + < nlm | V (r) | nlm >
2 m∗
2 2 Z ∞ X | < n0 l0 m0 | p̂ | nlm > |2
2 e h̄
+ dω ω
3 π h̄ c mc 0 h̄ ω
n0 l0 m0
2 2 Z ∞ X | < n0 l0 m0 | p̂ | nlm > |2
2 e h̄
+ dω ω
3 π h̄ c mc 0 0 0 0
Enlm − En0 l0 m0 − h̄ ω
nlm
(1187)
The overall (second-order) shift from the Schrödinger estimate of the energy for
the state | nlm > (as calculated with the physical mass) is given by the sum
of the last two terms, which is expressed as
2 2 Z ∞ X | < n0 l0 m0 | p̂ | nlm > |2 ( Enlm − En0 l0 m0 )
shift = 2 e h̄
∆Enlm dω ω
3 π h̄ c mc 0 0 0 0
( Enlm − En0 l0 m0 − h̄ ω ) h̄ ω
nlm
(1188)
The integration over ω is logarithmically divergent, and can be made to con-
verge by introducing an upper cut-off ω+ = c λ−1 . Therefore, the difference
of the linearly divergent self-energy of the bound electron and the linearly di-
vergent self-energy of the free electron is only logarithmically divergent. After
introducing the cut-off, one finds the result
| < n0 l0 m0 | p̂ | nlm > |2
2 X
shift 2 e
∆Enlm = −
3 π h̄ c m2 c2
n0 l0 m0
h̄cλ−1 + En0 l0 m0 − Enlm
× ( Enlm − En l m ) ln
0 0 0
En0 l0 m0 − Enlm
(1189)
The square of the mass m can safely be replaced by the square of the physical
mass m∗ in the expression for the energy shift, since we are only working to
2
first order in h̄e c . All other quantities have been expressed in terms of the
physical mass. If the rest energy of the electron is used as the upper cut-off
energy m c2 ∼ 0.5 × 106 eV, and assuming that the averaged logarithm of the
electron excitation energy corresponds to an energy of the order of 17.8 Ryd,
then the logarithm has a value of about 7.63 and is not sensitive to the precise
value of Enlm − En0 l0 m0 and, therefore, can be taken outside the summation
2 h̄2 c2
2
shift = − 2
∆Enlm
e
ln 2 4
3 π h̄ c Z e
X | < n0 l0 m0 | p̂ | nlm > |2
× ( Enlm − En0 l0 m0 )
0 0 0
m2 c2
nlm
(1190)
73 e2
As later shown by Dyson , that divergences found in any order in h̄ c can
73 F. J. Dyson, Phys. Rev. 75, 1736 (1949).
207
be removed by consistently using the ideas of mass and charge renormaliza-
tion74 . Hence, a completely consistent relativistic theory does yield a finite
shift, without the need to invoke any cut-off75 . The weighted sum over the
matrix elements can be evaluated by expressing it in terms of an expectation
value involving commutators of Ĥ0 with p̂. That is
X
| < n0 l0 m0 | p̂ | nlm > |2 ( Enlm − En0 l0 m0 )
n0 l0 m0
X
= < nlm | p̂ | n0 l0 m0 > < n0 l0 m0 | [ p̂ , Ĥ0 ] | nlm >(1191)
n0 l0 m0
Likewise,
X
| < n0 l0 m0 | p̂ | nlm > |2 ( Enlm − En0 l0 m0 )
n0 l0 m0
X
= − < nlm | [ p̂ , Ĥ0 ] | n0 l0 m0 > < n0 l0 m0 | p̂ | nlm >
n0 l0 m0
(1193)
which results in
Thus, on adding the above two equations and dividing by two, one finds
X
| < n0 l0 m0 | p̂ | nlm > |2 ( Enlm − En0 l0 m0 )
n0 l0 m0
1
= < nlm | [ p̂ , [ p̂ , Ĥ0 ] ] | nlm > (1195)
2
On substituting
p̂ = − i h̄ ∇ (1196)
and
p̂2
Ĥ0 = + V (r) (1197)
2m
74 This statement does not imply that a properly renormalized perturbation theory is con-
vergent. In fact, one may argue that if the coupling constant changed sign then systems
containing electrons would be unstable to BCS pairing. Since the radius of convergence of
any expansion is limited by the closest singularity, perturbation theory may only have a zero
radius of convergence. In this case, the theory may be expected to contain non-analytic terms
of the form exp[ − h̄ c/ e2 ].
75 F. J. Dyson, Phys. Rev. 173, 617 (1948).
208
into the expression for the matrix elements, one obtains
X
| < n0 l0 m0 | p̂ | nlm > |2 ( Enlm − En0 l0 m0 )
n0 l0 m0
2 Z
h̄
= − d3 r ψnlm (r) ∇2 V (r) ψnlm (r) (1198)
2
Substituting the expressions for the matrix elements into the expression for the
Lamb-shift yields
2 2
2 h̄2 c2
shift = 2 e h̄ 2
∆Enlm ln Z 2 e4 < nlm | ∇ V (r) | nlm >
3 π h̄ c m2 c2
(1199)
Thus, the energy-shift only occurs for bound electrons as the expectation value of
the Laplacian of the potential will vanish for extended states. For a hydrogenic-
like atom
∇2 V (r) = 4 π Z e2 δ 3 (r) (1200)
so
4 Z e2 e2 h̄2 2 h̄2 c2
∆E shift
nlm = 2
| ψnlm (0) | ln 2 4 (1201)
3 h̄ c m2 c2 Z e
Therefore, the Lamb shift only occurs for electrons with l = 0, since electronic
wave functions with l 6= 0 vanish at the origin. The atomic wave function at
the position of the nucleus is given by
3
1 Z
| ψn00 (0) |2 = (1202)
π na
This yields Bethe’s estimate for the Lamb shift as
2 3 4 4
2 h̄2 c2
shift = 4 e Z e m
∆En00 ln (1203)
3 π n3 h̄ c h̄2 Z 2 e4
The above formulae leads to the estimate of 1040 MHz which is in good agree-
ment with the experimentally determined value76 . The exact relativistic calcu-
lation77 yields the result
2 5
m c2
shift 4 4 e 2 31
∆En00 = Z mc ln
+ (1204)
3 π n3 h̄ c 2 h̄ ωn,n0 120
where the mc2 in the logarithm comes from the Dirac theory without invoking
any cut-off. The most recent experimentally measured value78 is 1057.851 MHz
which is in good agreement with the theoretical value of 1057.857 MHz.
209
11.3.5 Brehmstrahlung
Accelerating (or decelerating) charged particles radiate. We shall consider the
radiation emitted by a charged particle (such as an electron) that scatters from
a massive charged particle via the Coulomb interaction. It is assumed that the
mass M of the massive charged particle (in most cases, this is a nucleus) is
significantly greater than the electron mass, so that the recoil of the nucleus can
be neglected. The (instantaneous) Coulomb interaction between the electron
and the nucleus is given by
Z e2
V (r) = − (1205)
r
The Hamiltonian of the unperturbed electron is simply the kinetic energy. The
incident electron is assumed to have a momentum q and the scattered electron
has momentum q 0 and the cross-section for the scattering process will be calcu-
lated via low-order perturbation theory.
Rutherford Scattering
q q'
q-q'
210
2
4 π Z e2
2π V m
= q dΩ0
h̄ ( 2 π ) h̄2
3 V | q − q 0 |2
(1208)
| q − q 0 |2 = 2 q 2 ( 1 − cos θ0 )
θ0
= 4 q 2 sin2 (1209)
2
one finds
q'
2q sinθ'/2
θ'
q
Figure 41: The geometry for Rutherford scattering. For elastic scattering, the
magnitude of the initial momentum q is equal to the magnitude of the final
momentum q 0 and the scattering angle is θ0 .
2
4 π Z e2
1 2π V m
= q θ0
dΩ0
τdΩ0 Rutherford h̄ ( 2 π ) h̄2
3
V 4 q 2 sin2 2
(1210)
211
dσ/dΩ'
0 0.25 0.5 0.75 1
θ'/π
Figure 42: The scattering angle dependence of the differential scattering cross-
section.
by the extremely high potential experienced by electrons with very small im-
pact parameters. It was the large cross-section for back-scattering of charged
α-particles from atoms, found by H. Geiger and E. Marsden in 191379 , that
was instrumental in verifying Rutherford’s 1911 conjecture80 that atoms have
nuclei which are of very small spatial extent. The divergence in the scattering
cross-section at θ0 = 0 is due to the long-ranged nature of the Coulomb inter-
action, which causes electrons to undergo scattering (no matter how slight the
scattering is) at arbitrarily large distances from the nucleus.
Brehmstrahlung
(a) Scattering of an electron from the nucleus followed by the emission of a pho-
ton. The initial state of the electron is assumed to have momentum q and the
final state of the electron is given by q 0 while the emitted photon has momentum
k. Therefore, from conservation of momentum, the momentum of the electron
79 H. Geiger and E. Marsden, Phil. Mag. 25, 1798 (1913).
80 E. Rutherford, Phil. Mag. 21, 669 (1911).
81 H. A. Bethe and W. Heitler, Proc. Roy. Soc. A 146, 82 (1934).
212
(k,α)
(k,α)
q'+k q-k
q q' q q'
and
s
4 π Z e2 2 π h̄ c2
e h̄
Mb = 0 2
ˆα (k) . ( q − k )
V |q − q − k| mc V ωk
1
× (1214)
Eq − Eq−k − h̄ ωk + i η
It should be noted that the numerators of the matrix elements simplify because
the photons have transverse polarizations
α (k) . k = 0 (1215)
From the energy conserving delta function in the expression for the decay rate,
one finds
Eq = Eq0 + h̄ ωk (1216)
hence the first energy-denominator can be expressed in a similar form to the
second
Eq − Eq0 +k = Eq0 − Eq0 +k + h̄ ωk (1217)
For small k, the energy-denominators can be expanded, yielding
h̄ 0 h̄2 k 2
Eq0 − Eq0 +k + h̄ ωk = h̄ ωk − q .k − (1218)
m 2m
213
and
h̄ h̄2 k 2
Eq − Eq−k − h̄ ωk = − h̄ ωk + q.k − (1219)
m 2m
Since the energy of the photon cannot exceed the energy of the initial electron,
one must have q > k, so the third term is smaller than the second term. Due to
the large magnitude of c compared with the electron velocities h̄mq , the second
and third terms can be neglected. Therefore, the photon-energy dominates both
the energy-denominators. On substituting the above expressions in the sum of
the matrix elements, one finds
s
4 π Z e2 2 π h̄ c2
e
M a + Mb = 0 2
V |q − q − k| mc V ωk
0
ˆα (k) . q h̄ ˆα (k) . q h̄
× +
Eq0 − Eq0 +k + h̄ ωk + i η Eq − Eq−k − h̄ ωk + i η
s
4 π Z e2 2 π h̄ c2
e
≈
V | q − q 0 − k |2 mc V ωk
ˆα (k) . ( q 0 − q )
× (1220)
ωk
Using this approximation for the matrix elements, the transition rate is given
by
2 2
4 π Z e2 2 π h̄ c2
1 2π X X e
=
τ h̄ 0
V | q − q 0 − k |2 mc V ωk
q k,α
ˆα (k) . ( q 0 − q ) 2
×
δ( Eq − Eq0 − h̄ ωk ) (1221)
ωk
If the angular distributions of the emitted photon (dΩk ) and the scattered elec-
tron (dΩ0 ) are both measured, the scattering cross-section can be represented
as
d3 σ q0
dσ
=
dΩ0 dΩk dωk Brehmse q dΩ0 Rutherford
214
e2 h̄ ˆα (k) . ( q 0 − q ) 2
X
1
×
4 π 2 ωk h̄ c α
m c
(1223)
where the second factor is the probability of emitting a photon with energy h̄ ωk
into solid angle dΩk . On summing over the polarization α and integrating over
the directions of the emitted photon, one obtains
2
d2 σ q0 2 m Z e2
=
dΩ0 dωk Brehmse h̄2 | q − q 0 |2
q
h̄ ( q 0 − q ) 2
2
2 e
× (1224)
3 π ωk h̄ c mc
Hence, the scattering rate which includes the emission of a photon of energy
h̄ ωk is given by the product of the Rutherford scattering rate with a factor
0 2
q0 e2 2 q h̄ sin θ2
2
(1225)
q 3 π ωk h̄ c mc
it is expected that the classical limit of quantum theory applies so that classical electromag-
netic theory should produce exact results.
215
frequencies ω0 . That is, on introducing an infra-red cut-off λ− , one finds that
the total inelastic scattering in which a photon with frequency less than ω0 is
emitted is given by
2
dσ dσ 1 e 2 ω 0 λ−
= A ln + ... + ...
dΩ Brehmse dΩ Rutherford 2 π h̄ c c
(1226)
where the factor A depends on the initial and final momentum of the electron.
This result is logarithmically divergent as λ− → 0. On the other hand, to the
same order, the elastic scattering cross-section is found as
2
dσ dσ 1 e h̄
= 1+ A ln + ... + ...
dΩ Elastic dΩ Rutherford 2 π h̄ c λ− m c
(1227)
Hence, on combining the results, one finds that the quasi-elastic scattering cross-
section is given by
2
dσ dσ 1 e 2 h̄ ω0
= 1+ A ln + ... + ...
dΩ Quasi-Elastic dΩ Rutherford 2 π h̄ c m c2
(1228)
so the cut-off λ− cancels and the scattering cross-section does not diverge log-
arithmically. With this reasoning, Bloch and Nordsieck found that the appro-
2 2
h̄ ω0
priate expansion parameter is not h̄e c but instead is given by h̄e c ln m c2 . The
higher-order perturbations may also describe processes involving larger numbers
of emitted soft photons and results in a multiplicative exponential factor to the
quasi-elastic scattering rate
2
dσ dσ 1 e 2 h̄ ω0
≈ exp B ln + . . .
dΩ Quasi-Elastic dΩ Rutherford 2 π h̄ c m c2
(1229)
Therefore, the scattering rate from soft photons vanishes in the limit ω0 → 0.
This occurs because perturbation theory causes the normalization of the starting
approximate wave function to change, and hence the probabilities of the vari-
ous processes are changed by including higher-order processes. In other words,
since the probability of emitting an arbitrarily large number of soft-photons is
finite, the probability of emitting either zero or any fixed number of soft photons
must be zero. Bloch and Nordsieck’s calculation was restricted to the case of
emission of sufficiently low-energy photons. Pauli and Fierz86 also considered
Brehmstrahlung in a non-relativistic approximation. Pauli and Fierz showed
that the infra-red divergences, discussed above, cancel. Pauli and Fierz went on
to examine the remaining ultra-violet divergences, and showed that portions of
the ultra-violet infinities that were found in the calculations of the scattering
processes could be associated with mass renormalization. Using a relativistic
theory Ito, Koba and Tomonaga87 showed that the remaining infinities could
86 W. Pauli and M. Fierz, Nuovo Cimento, 15, 167 (1938).
87 D. Ito, Z. Koba and S-I. Tomonaga, Prog. Theor. Phys. (Kyoto), 3, 276 (1948).
216
be absorbed into a renormalization of the electron charge. Similar conclusions
were arrived at by Lewis88 and by Epstein89 . Dyson90 showed that all infinities
that appear in Quantum Electrodynamics could be cured by renormalization to
arbitrarily high-orders in perturbation theory.
ψ (0)
ψ (1)
ψ = (1231)
..
.
ψ (N −1)
The equations have to be of this form since, if the equation is a first-order partial
differential equation in time then it must also only involve the first-order partial
derivatives with respect to the spatial components for the resulting equation to
be relativistically covariant. The wave function ψ is a N -component (column)
wave function and the three as yet unknown components of α and β are three
N × N matrices. Since the Hamiltonian is the generator of time translations,
∂
then Ĥ should be equivalent to ih̄ ∂t . Hence, as the Hamiltonian operator Ĥ
must be Hermitean, then the operators α and β must be Hermitean matrices.
88 H. W. Lewis, Phys. Rev. 73, 173 (1948).
89 Saul T. Epstein, Phys. Rev. 73, 177 (1948).
90 F. J. Dyson, Phys. Rev. 75, 486 (1949).
91 P. A. M. Dirac, Proc. Roy. Soc. A 117, 610 (1928).
217
This set of equations is required to yield the dispersion relation for a relativistic
particle
2
E
− p2 = m2 c2 (1233)
c
which, following the ordinary rules of quantization, leads to the Klein-Gordon
equation
h̄2 ∂ 2
2 2
− 2 + h̄ ∇ ψ = m2 c2 ψ (1234)
c ∂t2
(which is a second-order partial differential equation in time). The requirement
that the Dirac equation is compatible with the Klein-Gordon equation imposes
conditions on the form of the matrices. On writing the Dirac equation as
h̄ ∂ψ
i = β m c − i h̄ α . ∇ ψ (1235)
c ∂t
When expressed in terms of individual matrices α(j) , the above equation be-
comes
2 2
h̄ ∂ ψ 2 2 2
X
− = β m c − i h̄ m c ( β α(j) + α(j) β ) ∇j
c ∂t2 j
h̄2 X
− ( α(i) α(j) + α(j) α(i) ) ∇i ∇j ψ (1237)
2 i,j
α(i) β + β α(i) = 0
α (i)
α(j) + α(j) α(i) = 2 δ i,j Iˆ (1239)
218
On imposing the above conditions, Dirac’s form of the relativistic Schrödinger
equation is compatible with the Klein-Gordon equation.
From eqn(1238), one concludes that if the Hermitean matrices are brought to
diagonal form then the diagonal elements are given by ± 1. The possible dimen-
sions N of the matrix can be determined by considering the anti-commutation
relations. On taking the determinant of eqn(1239), one finds
( − 1 )N = 1 (1241)
since α(i) is its own inverse. Apart from the negative sign, the form of the
left-hand side is of the form of an equivalence transformation. By using cyclic
invariance, it can be shown that the trace of a matrix is invariant under equiv-
alence transformations. Therefore, one has
or
Trace α(i) = 0 (1245)
which proves that the matrices are traceless.
β2 = Iˆ
(α(i) )2 = Iˆ (1246)
then their eigenvalues must all be ±1, as can be seen by operating on the
eigenvalue equation
β φβ = λβ φβ (1247)
with β. This process yields
β 2 φβ = λβ β φβ
= λ2β φβ (1248)
219
ˆ requires that the eigenvalues must satisfy the equation
which with β 2 = I,
λ2β = 1 (1249)
This and the condition that the matrices are traceless implies that the set of
eigenvalues of each matrix are composed of equal numbers of +1 and −1, and
it also confirms the conclusion that dimension N of the matrices must be even.
The smallest value of the dimension for which there is a representation of the
matrices is N = 4. The smallest even value of N , N = 2 cannot be used since
one can only construct three linearly independent anti-commuting 2 × 2 ma-
trices92 . These three matrices are the Pauli spin matrices σ (j) . Hence, Dirac
constructed the relativistic theory with N = 4.
If the three matrices α(i) are to anti-commute with β and be Hermitean, they
must have the off-diagonal form
A(i)
0
α(i) = (1251)
A(i)† 0
where A(i) is an arbitrary 2 × 2 matrix. We shall choose all three A(i) matrices
to be Hermitean. Since the three α(i) matrices must anti-commute with each
other, the A(i) must also anti-commute with each other. Since the three Pauli
matrices are mutually anti-commuting, one can set
σ (i)
0
α(i) = (1252)
σ (i) 0
where the σ (i) and I are, respectively, the 2 × 2 Pauli matrices and the 2 × 2
unit matrix. The Pauli matrices are given by
(1) 0 1
σ = (1253)
1 0
92 In d + 1 space-time dimensions, one can form 2d+1 matrices from products of the set of
d+1 linearly independent (anti-commuting) Dirac-matrices. We shall assume that the product
matrices are linearly independent. Since the number of linearly independent N × N matrices
is N 2 , the minimum dimension N which will yield a representation of the Dirac-matrices is
d+1
N =2 2 .
220
0 −i
σ (2) = (1254)
i 0
and
(3) 1 0
σ = (1255)
0 −1
The matrix α(0) is defined as the 4 × 4 identity matrix
(0) I 0
α = (1256)
0 I
This set of matrices form a representation of the Dirac matrices. This can
be seen by directly showing that they satisfy the appropriate relations. Many
different representations of the Dirac matrices can be found, but they are all
related by equivalence transformations and the physical results are independent
of which choice is made.
Exercise:
By direct matrix multiplication, show that the above matrices satisfy the
relations
( α(j) )2 = β 2 = Iˆ (1257)
and the anti-commutation relations
α(i) β + β α(i) = 0
α (i)
α(j) + α(j) α(i) = 2 δ i,j Iˆ (1258)
one obtains
∂ψ
† † † 2
i h̄ ψ = − i h̄ c ψ α . ∇ ψ + ψ β ψ m c (1261)
∂t
221
The Hermitean conjugate of the Dirac equation is given by
∂ψ †
† † † † 2
− i h̄ = + i h̄ c ∇ . ψ α + ψ β m c (1262)
∂t
Since α and β are Hermitean matrices, the Hermitean conjugate equation sim-
plifies to
∂ψ †
− i h̄ = + i h̄ c ∇ . ψ † α + ψ † β m c2 (1263)
∂t
Post-multiplying the Hermitean conjugate equation by the column-vector ψ,
yields
∂ψ †
− i h̄ ψ = + i h̄ c ∇ . ψ † α ψ + ψ † β ψ m c2 (1264)
∂t
On subtracting eqn(1264) from the eqn(1261) and combining terms, one obtains
∂
i h̄ ( ψ † ψ ) = − i h̄ c ∇ . ( ψ † α ψ ) (1265)
∂t
The above equation has the form of a continuity equation
∂ρ
+ ∇.j = 0 (1266)
∂t
in which the probability density is given by
ρ = ψ† ψ (1267)
Using the rules of matrix multiplication the probability density is a real scalar
quantity, which is given by the sum of squares
ρ = | ψ (0) |2 + | ψ (1) |2 + | ψ (2) |2 + | ψ (3) |2 (1268)
and so it is positive definite. Hence, unlike the Klein-Gordon equation, the
Dirac equation does not lead to negative probability densities. The probability
current density j is given by
j = c ψ† α ψ (1269)
In this case, the total probability
Z Z
3 †
Q = d xψ ψ = d3 x ρ (1270)
is conserved, since
Z
dQ ∂ρ
= d3 x
dt ∂t
Z
= − d3 x ∇ . j
Z
= − d2 S . j (1271)
222
where Gauss’s theorem has been used to represent the volume integral as surface
integral. For a sufficiently large volume, the current at the boundary vanishes,
hence the total probability is conserved
dQ
= 0 (1272)
dt
αµ p̂µ ψ = βmcψ
i h̄ αµ ∂µ ψ = βmcψ (1273)
α(0) = Iˆ (1274)
γ µ p̂µ ψ = mcψ
i h̄ γ µ ∂µ ψ = mcψ (1277)
γ µ γ ν + γ ν γ µ = 2 g µ,ν Iˆ (1278)
where Iˆ is the 4 × 4 identity matrix, and g µ,ν is the Minkowski metric. The
gamma matrices labelled by the spatial indices are Unitary and anti-Hermitean,
as shall be proved below.
It is easy to show that the matrix with the temporal index (0) is unitary
and Hermitean
223
The gamma matrices with spatial indices are anti-Hermitean as
( γ (i) )† = ( β α(i) )†
= ( ( α(i) )† β † )
= ( α(i) β )
= ( − β α(i) )
= − γ (i) (1280)
since α(i) and β are Hermitean and, in the fourth line the operators have been
anti-commuted. Now, the gamma matrices with spatial indices can be shown
to be unitary since
where, in obtaining the second line, the anti-commutation properties of α(i) and
β have been used, and the property
( α(i) )2 = β 2 = Iˆ (1282)
was used to obtain the last line. Since it has already been demonstrated that
the spatial matrices are anti-Hermitean
Hence, since
( γ (0) )2 = Iˆ (1286)
†
the Hermitean conjugate wave function ψ can be expressed in terms of the
†
adjoint spinor ψ via
†
ψ † = ψ γ (0) (1287)
The continuity equation has the Lorentz covariant form
∂j µ
= 0 (1288)
∂xµ
224
where the four-vector conserved probability current j µ is given by
j µ = c ψ † αµ ψ (1289)
By using the definition of the Dirac adjoint, the current density can be re-
expressed as the four quantities
†
j (0) = c ψ γ (0) ψ
†
j (i) = c ψ γ (i) ψ (1290)
that, respectively, represent c times the probability density and the j (i) are the
contravariant components of the probability current density.
225
Hence, the Dirac equation can be expressed as the block-diagonal matrix equa-
tion
(0) m c
− k + h̄ I k.σ A
φ
B = 0 (1295)
φ
− k (0) − mh̄ c I
k.σ
one finds the energy eigenvalues are given by the doubly-degenerate dispersion
relations s 2
mc
k (0)
= ± + k2 (1299)
h̄
Thus, the field free relativistic electron can have positive and negative-energy
eigenvalues given by q
E = ± m2 c4 + p2 c2 (1300)
Since the solutions are degenerate, solutions can be found that are simultaneous
eigenvalues of the Hamiltonian Ĥ given by
m c2 I
− i h̄ c σ . ∇
Ĥ = (1301)
− i h̄ c σ . ∇ − m c2 I
and another operator that commutes with Ĥ. It is convenient to choose the
second operator to be the helicity operator.
226
Σk = +1 Σk = −1
k k
Figure 44: A cartoon depicting the two helicity states of a spin one-half particle.
given by
σ.∇ 0
Σ̂ = − i h̄ (1302)
0 σ.∇
This is the appropriate relativistic generalization of spin valid only for free
particles93 , as the helicity is a conserved quantity since
[ Ĥ , Σ̂ ] = 0 (1303)
In the absence of electromagnetic fields, the Hamiltonian is evaluated as
m c2 I
h̄ c σ . k
Ĥ(k) = (1304)
h̄ c σ . k − m c2 I
Likewise, for the source free case, the properly normalized Helicity operator is
found as
σ . k̂ 0
Λ(k) = (1305)
0 σ . k̂
which has eigenvalues of ±1.
φA
+ = u(0) χ+
(0) 1
= u (1307)
0
93 Helicity is not conserved for spherically symmetric potentials. However, if only a time-
227
and φB
+ as
φB
+ = u(2) χ+
1
= u(2) (1308)
0
u(0) χ+
µ
ψ+ (x) = exp − i kµ x (1309)
u(2) χ+
φA
− = u(1) χ−
0
= u(1) (1310)
1
and φB
+ as
φB
− = u(3) χ−
0
= u(3) (1311)
1
On substituting the helicity eigenstates ψΛ into the Dirac equation for the
free spin one-half particle
∂
i h̄ ψΛ = Ĥ ψΛ (1314)
∂t
one finds
φA m c2 σ (3) c h̄ k (3) φA
E Λ = Λ (1315)
φB
Λ σ c h̄ k (3)
(3)
− m c2 φB
Λ
σ (3) c h̄ k (3) A
φB
Λ = φ (1316)
E + m c2 Λ
228
This equation shows that the components φBΛ are small for the positive-energy
solutions, whereas the complementary expression
σ (3) c h̄ k (3) B
φA
Λ = − φ (1317)
m c2 − E Λ
shows that φA Λ is small for the negative-energy solutions. Hence, the two
positive-energy and two negative-energy (un-normalized) solutions of the Dirac
equation can be written as
χ+
ψ+ (x) = Ne exp − i kµ xµ (1318)
c h̄ k(3)
E + m c2 χ+
χ−
µ
ψ− (x) = Ne exp − i kµ x (1319)
c h̄ k(3)
− E + m c2 χ−
c2 h̄2 k 2
2
1 = V Ne 1 +
( E + m c2 )2
E + 2 E m c2 + m2 c4 + c2 h̄2 k 2
2
2
= V Ne
( E + m c2 )2
2 E + 2 E m c2
2
2
= V Ne
( E + m c2 )2
2 2E
= V Ne
E + m c2
Hence, the normalization constant can be set as
r
E + m c2
Ne = (1321)
2EV
for positive E.
229
the lower components are the large components. In this case, it is more conve-
nient to express the negative-energy solutions as
− m cc2h̄ −k E χ+
ψ+ (x) = Np exp − i kµ xµ (1323)
χ+
for helicity -1. Furthermore, in this expression the normalization constant has
the form r
m c2 − E
Np = (1325)
−2EV
Hence, the positive and negative-energy solutions are symmetric under the in-
terchange E → − E, if Λ → − Λ and the upper and lower two-component
spinors (φA , φB ) are interchanged.
which since
Λ(k) Λ(k) = I (1327)
94
has eigenvalues Λ of ±1. The helicity eigenstates are given by the two-
component spinors χΛ± . The positive helicity state is given by
(1)
− i k (2)
k
1
χΛ+ = p
2 k ( k − k (3) ) k − k (3)
cos θ2k exp[−i ϕ2k ]
ϕk
= exp[ − i ] (1328)
2 θk ϕk
sin 2 exp[+i 2 ]
94 C. G. Darwin, Proc. Roy. Soc. A 118, 654 (1928).
230
in which (k, θk , ϕk ) are the polar coordinates of k. The negative helicity eigen-
state is given by the spinor χΛ−
− k + k (3)
1
χΛ− = p
2 k ( k − k (3) ) k (1)
+ ik (2)
Therefore, the general helicity eigenstate plane-wave solutions of the Dirac equa-
tion can be written in terms of two two-component spinors as
χΛ±
µ
ψΛ± (x) = Ne exp − i kµ x (1330)
c h̄ k Λ±
E + m c 2 χΛ±
αµ p̂µ ψ = βmcψ
i h̄ αµ ∂µ ψ = βmcψ (1332)
231
and q is the charge of the particle, the Dirac equation in the presence of an
electromagnetic field becomes
q
αµ p̂µ − Aµ ψ = β m c ψ
c
q
i h̄ αµ ∂µ + i Aµ ψ = β m c ψ (1334)
h̄ c
This process has resulted in the inclusion of the interaction with the electro-
magnetic field in a gauge invariant, Lorentz covariant manner. The appearance
of the gauge field together with the derivative results in local gauge invariance.
Sometimes it is convenient to define a covariant derivative as the gauge-invariant
combination
q
Dµ ≡ ∂µ + i Aµ (1335)
h̄ c
The concept of the covariant derivative also appears in the context of other
gauge field theories. Using this definition we can express the Dirac equation in
the presence of an electromagnetic field in the compact covariant form
i h̄ γ µ Dµ ψ = m c ψ (1336)
The presence of an electromagnetic field does not alter the form of the conserved
four-vector current
†
jµ = c ψ γµ ψ (1337)
which is explicitly gauge invariant.
The interaction Hamiltonian with the electrostatic field of the nucleus is given
by the diagonal matrix
Z e2
I 0
ĤInt = − (1340)
r 0 I
232
The flux of incident electrons is defined by
|v|
F = (1341)
V
where
∂E
v = (1342)
∂p
Therefore, the electron flux is given by
h̄ k c2
F = (1343)
V Ek
The elastic scattering cross-section in which the final state polarization is un-
measured is given by
Ek V 2 X ∞
Z
dσ 1
= dk 0 k 02 | < k 0 σ 0 | ĤInt | k, σ > |2 δ( Ek − Ek0 )
dΩ0 ( 2 π )2 h̄2 k c2 0 0
σ
(1344)
where the delta function ensures conservation of energy. Since the polarization
of the final state electron is unmeasured, the spin σ 0 is summed over. The
integration over k 0 can be performed, yielding
2 X
dσ EV
= | < k 0 σ 0 | ĤInt | k, σ > |2 (1345)
dΩ0 2 π h̄2 c2 σ 0
233
The resulting matrix elements involve the spin-dependent factor
c2 h̄2 ( σ . k ) ( σ . k 0 ) c2 h̄2 ( σ . k 0 ) ( σ . k )
T
χσ I + I + χσ
( E + m c2 )2 ( E + m c2 )2
(1349)
The products of matrix elements shown above can be evaluated with the aid of
the Pauli identity. The sum of the cross-terms can be evaluated directly using
the Pauli identity. We note that since the vector product are antisymmetric in
k and k 0 , the sum of the vector product terms cancel. That is
c2 h̄2
0 0
( σ . k ) ( σ . k ) + ( σ . k ) ( σ . k )
( E + m c2 )2
c2 h̄2
= 2 ( k . k 0 ) Iˆ (1350)
( E + m c2 )2
The remaining term is evaluated by using the Pauli identity for the inner two
scalar products, and then re-using the identity for the outer two scalar products.
Explicitly, this process yields
c4 h̄4
0 0
(σ.k)(σ.k )(σ.k )(σ.k)
( E + m c2 )4
c4 h̄4
= k 02 k 2 Iˆ (1351)
( E + m c2 )4
Hence, the cross-section is given by
2
Z e2 c4 h̄4 k 2 k 02
dσ 2 2 2 2 0
= ( E + m c ) + 2 c h̄ k . k +
dΩ0 h̄2 c2 | k − k 0 |2 ( E + m c2 )2
(1352)
It should be noted that the last two terms originated from the combined action
of the Pauli spin operators and involved the lower two-component spinors. The
last term can be simplified by using the elastic scattering condition k = k 0 and
then using the identity
c4 h̄4 k 4 = ( E 2 − m2 c4 )2 (1353)
k . k 0 = k 2 cos θ0 (1355)
234
and also in the denominator of the Coulomb interaction by
θ0
| k − k 0 |2 = 4 k 2 sin2 (1356)
2
Furthermore, the factor of m2 c4 in the square parenthesis can be replaced by
m2 c4 = E 2 − c2 h̄2 k 2 (1357)
has been introduced. The above result is the Mott scattering cross-section96 ,
which describes the scattering of charged electrons. It differs from the Ruther-
ford scattering cross-section due to the multiplicative factor of relativistic origin,
which deviates from unity due to the electron’s internal degree of freedom. The
extra contribution to the scattering is interpreted in terms of scattering from the
magnetic moment associated with the electron’s spin interacting with the mag-
netic field of the nuclear charge that the electron experiences in its rest frame.
It should be noted that even if the initial beam of electrons is un-polarized, the
scattered beam will be partially spin-polarized (due to higher-order corrections).
235
where j is the contravariant form of the current four-vector
cρ
j (1)
j = j (2)
(1362)
j (3)
We shall require that the matrices αµ are Hermitean and that they satisfy the
equation
( αµ )2 = Iˆ (1363)
On comparing with the form of Maxwell’s equations97 , one finds that the Ma-
trices are given by
1 0 0 0
0 1 0 0
α(0) = 0 0 1 0
(1364)
0 0 0 1
0 −1 0 0
−1 0 0 0
α(1) = (1365)
0 0 0 −i
0 0 i 0
0 0 −1 0
0 0 0 i
α(2) = −1 0
(1366)
0 0
0 −i 0 0
0 0 0 −1
0 0 −i 0
α(3) = 0 i 0
(1367)
0
−1 0 0 0
The matrices corresponding to the spatial indices are traceless and satisfy the
anti-commutation relations
and X
α(i) α(j) = i ξ i,j,k α(k) (1369)
k
directly from the comparison. The first rows are determined by demanding that the matrices
are Hermitean.
236
with the operator
i αν ∂ν (1371)
one obtains
4π ν
− αν αµ ∂ν ∂µ ψ = − i α ∂ν j (1372)
c
Utilizing the anti-commutation of the spatial matrices, the left-hand side sim-
plifies to
1 ∂
− − ∂µ ∂ µ + 2 αν ∂ν ψ (1373)
c ∂t
On substituting the new form of Maxwell’s equations in the second term, the
expression reduces to
8π ∂
− − ∂µ ∂ µ ψ + i 2 j (1374)
c ∂t
Thus, the equation becomes
4π 2 ∂
∂µ ∂ µ ψ = i − αν ∂ν j (1375)
c c ∂t
The zero-th component of the source term vanishes, due to conservation of
charge.
237
where
†
ψ = ψ † γ (0) (1379)
One can rewrite the current density by using the Dirac equation
q
i h̄ γ µ ∂µ + i Aµ ψ = m c ψ (1380)
h̄ c
On symmetrizing the current density and then substituting the Dirac equation
in one term and its Hermitean conjugate in the other term, one obtains
ν c † ν † ν
j = ψ γ ψ + ψ γ ψ
2
i h̄ q † µ† (0) ν † (0) ν µ q
= − ( ∂µ − i Aµ )ψ γ γ γ ψ + ψ γ γ γ ( ∂µ + i Aµ )ψ
2m h̄ c h̄ c
i h̄ q † (0) µ† (0) ν † ν µ q
= − ( ∂µ − i Aµ )ψ γ γ γ γ ψ + ψ γ γ ( ∂µ + i Aµ )ψ
2m h̄ c h̄ c
(1382)
where the partial derivatives only operate on the wave function immediately to
the right of it. The identity
γ (0) γ (0) = Iˆ (1383)
†
has been used to express ψ † in terms of ψ . However, since the γ matrices
satisfy
γ (0) γ µ† γ (0) = γ µ (1384)
the current can be further simplified to yield
ν i h̄ q † µ ν † ν µ q
j = − ( ∂µ − i Aµ )ψ γ γ ψ + ψ γ γ ( ∂µ + i Aµ )ψ
2m h̄ c h̄ c
(1385)
where, once again, the partial derivative only operates on the wave function
immediately to the right of it. Furthermore, if one sets
1
γµ γν + γν γµ = g µ,ν Iˆ
2
1
γµ γν − γν γµ = − i σ µ,ν (1386)
2
238
then the current density can be expressed as the sum of two contributions
jν = jcν + jsν
i h̄ µ,ν † † q µ,ν †
= − g ( ∂µ ψ ψ − ψ ∂µ ψ ) + 2 i g ψ Aµ ψ
2m h̄ c
h̄ ∂ †
− µ
ψ σ µ,ν ψ (1387)
2 m ∂x
where
i h̄ † † q †
jcν = − ( ∂ν ψ ψ − ψ ∂ν ψ ) + 2 i ψ Aν ψ
2m h̄ c
h̄ ∂ † µ,ν
jsν = − ψ σ ψ (1388)
2 m ∂xµ
Let us examine the first term in the probability current density. If ψ repre-
(0)
sents an energy eigenstate, then jc is given by
(0) E † q †
jc = ψ ψ − ψ A(0) ψ (1389)
mc mc
This contribution obviously yields the main contribution to (c times) the prob-
ability density
†
jc(0) ≈ c ψ ψ (1390)
in the non-relativistic limit since the rest mass energy dominates the energy
(i)
E ∼ m c2 . The spatial components of jc are given by
i h̄ † † q †
jc = (∇ψ )ψ − ψ (∇ψ) − ψ Aψ (1391)
2m mc
where the derivatives have been expressed as derivatives w.r.t. the contravariant
components x(i) of the position vector. This expression coincides with the full
non-relativistic expression for the current density j (i) .
We now examine the second term jsµ in the Gordon decomposition. For
future reference, the anti-symmetrized products of the Dirac matrices σ µ,ν will
98 W. Gordon, Zeit. für Physik, 50, 630 (1928).
239
be expressed in 2 × 2 block diagonal form. Therefore, since
(0) I 0
γ =
0 −I
σ (i)
(i) 0
γ = (1392)
−σ (i) 0
and
i
σ µ,ν = γµ γν − γν γµ (1393)
2
the matrices are found as
σ (j)
0
σ 0,j = i (1394)
σ (j) 0
and
σ (k)
i,j
X
i,j,k 0
σ = ξ (1395)
0 σ (k)
k
The two by two block diagonal matrix of Pauli spin matrices will be denoted
(0)
by σ̂. For an energy eigenstate, the time component of js is identically zero.
(i)
Hence, the spatial components of js are given by
h̄ †
js = − ∇ ∧ ( ψ σ̂ ψ ) (1396)
2m
where σ̂ is the 2 × 2 block-diagonal Pauli spin matrix
σ 0
σ̂ = (1397)
0 σ
The additional term in the current density clearly involves the Pauli spin-
matrices. To elucidate its meaning, its contribution to the energy shall be
examined. On substituting this term in the interaction Hamiltonian density,
one finds a contribution
spin q
ĤI = − j .A
c s
q h̄ †
= + A . ∇ ∧ ( ψ σ̂ ψ ) (1398)
2mc
On integrating over space, the interaction Hamiltonian density gives rise to the
interactions contribution to the total energy. By integrating by parts, it can
be shown that this energy contribution is equivalent to the energy contribution
caused by an equivalent form of the interaction Hamiltonian density
spin q h̄ †
ĤI ≡ − ( ψ σ̂ ψ ) . ( ∇ ∧ A )
2mc
q h̄ †
≡ − ( ψ σ̂ ψ ) . B (1399)
2mc
240
where B is the magnetic field. Hence, the interaction energy contains a term
which represents an interaction between the electron’s internal degree of free-
dom and the magnetic field.
The first step of the proof of the Lorentz covariance of the Dirac equation
requires that one should be able to show that under a Lorentz transformation
defined by
Aµ → Aµ0 = Λµ ν Aν (1400)
then the Dirac equation is transformed from
q
γ µ ( p̂µ − Aµ ) ψ = m c ψ (1401)
c
to an equation with an equivalent form
q 0
γ µ0 ( p̂0µ − A ) ψ0 = m c ψ0 (1402)
c µ
Furthermore, the four components of the spinor wave function ψ 0 are assumed
to be linearly related to the components of ψ by a four by four matrix R̂(Λ)
which is independent of xµ
Hence, the transformed Dirac equation can be re-written in terms of the un-
transformed spinor
q 0
γ µ0 ( p̂0µ − A ) ψ0 = m c ψ0
c µ
q 0
γ µ0 ( p̂0µ − A ) R̂(Λ) ψ = m c R̂(Λ) ψ (1404)
c µ
if such an R̂(Λ) exists. The γ µ0 matrices must satisfy the same anti-commutation
relations as the γ µ and, therefore, only differ from them by a similarity trans-
241
formation99 . The transformations of γ µ0 just results in the set of the four linear
equations that compose the Dirac equation being combined in different ways,
so this rearrangement can be absorbed in the definition of R̂(Λ). That is, one
can choose to impose the convention that γ µ0 = γ µ . The transformed Dirac
equation can be expressed as
q 0
γ µ0 ( p̂0µ − A ) R̂(Λ) ψ = m c R̂(Λ) ψ
c µ
q
γ µ0 Λµ ν ( p̂ν − Aν ) R̂(Λ) ψ = m c R̂(Λ) ψ (1405)
c
where the transformation properties of the momentum four-vector have been
used100 . On multiplying by the inverse of R̂(Λ), one has
q
R̂−1 (Λ) γ µ0 Λµ ν ( p̂ν − Aν ) R̂(Λ) ψ = mcψ (1406)
c
q
R̂−1 (Λ) γ µ0 R̂(Λ) Λµ ν ( p̂ν − Aν ) ψ = mcψ (1407)
c
where the four by four matrices R̂(Λ) have been commuted with the differential
operators and also with the components of the Lorentz transform. The condition
for covariance as
R̂−1 (Λ) γ µ0 R̂(Λ) Λµ ν = γ ν (1408)
The transformed Dirac equation has the same form as the original equation if
the transformed γ µ0 matrices satisfy the same anti-commutations and conditions
as the unprimed matrices. This can be achieved by choosing γ µ0 = γ µ . This
choice yields the condition for covariance as
Λ µ ν Λ ρ ν = δµ ρ (1410)
The above equation determines the 4 × 4 matrix R̂(Λ). If R̂(Λ) exits, the Dirac
equation has the same form in the two frames of reference and the solutions
are linearly related. Pauli’s “fundamental theorem” guarantees that a matrix
R̂(Λ) exists which does satisfy the condition. Instead of following the general
theorem, the solution will be inferred from consideration of infinitesimal Lorentz
99 This is a statement of Pauli’s fundamental theorem [W. Pauli, Ann. Inst. Henri Poincaré
6, 109 (1936).]. For a general discussion, see R. H. Good Jr. Rev. Mod. Phys. 27, 187
(1955).
100 It should be noted that the matrices Λ ν and R̂ act on totally different spaces. The
µ
matrices Λµ ν act on the components of the four-vectors xν , whereas the R̂ matrices act on
the components of the four-component Dirac spinor ψ.
242
transformations.
Λµ ν = δ µ ν + µ ν + . . . (1412)
where δ µ ν is the Kronecker delta function. The matrix R̂(Λ) for the infinitesimal
transformation can also be expanded as
i µ
R̂ = Iˆ − ν ωµ ν + . . . (1413)
4
where ωµ ν is a four by four matrix that has yet to be determined. The inverse
matrix can be written as
i µ
R̂−1 = Iˆ + ν ωµ ν + . . . (1414)
4
to first-order in the infinitesimal quantity µ ν . On substituting the matrices for
the infinitesimal transform into the equation that determines R̂, one obtains
i ρ σ µ µ σ
σ ωρ γ − γ ωρ = µ ν γ ν (1415)
4
243
The set of (as yet unknown) matrices ω ρσ that solve the above set of equations
are given by
i
ω αβ = σ αβ = [ γα , γβ ] (1420)
2
which are the six generators of the general infinitesimal Lorentz transformation.
This solution, and hence, the existence of R̂(Λ) shows that the solutions of the
Dirac equation and the transformed equation are in a one to one correspondence.
——————————————————————————————————
Proof of Solution
It can be shown that the expression for σ α,β given in eqn(1420) satisfies the
requirement of eqn(1419), by evaluating the nested commutator through repeat-
edly using the anti-commutation properties of the γ matrices. The commutator
can be expressed as a nested commutator or as the sum of two commutators
i
[ σ αβ , γ µ ] = [ [ γα , γβ ] , γµ ]
2
i i
= [ γα γβ , γµ ] − [ γβ γα , γµ ] (1421)
2 2
On using the anti-commutation relation for the γ matrices
1
γα γβ + γβ γα = g α,β Iˆ (1422)
2
one can eliminate the second term leading to
[ σ αβ , γ µ ] = i [ γ α γ β , γ µ ] + i g α,β [ Iˆ , γ µ ]
= i [ γα γβ , γµ ] (1423)
where the second line follows since the identity matrix commutes with γ µ . One
notices that if the γ µ ’s are anti-commuted to the center of each product, some
terms will cancel and there may be some simplification. On using the anti-
commutation relation in the second term of the expression
αβ µ α β µ µ α β
[σ , γ ] = i γ γ γ − γ γ γ (1424)
one finds
[ σ αβ , γ µ ] = i γ α γ β γ µ + γ α γ µ γ β − 2 g µ,α γ β (1425)
Likewise, the γ matrices in the first term can also be anti-commuted, leading to
αβ µ µ,β α α µ β α µ β µ,α β
[σ , γ ] = i 2g γ − γ γ γ + γ γ γ − 2g γ
= 2 i g µ,β γ α − g µ,α γ β (1426)
244
since the middle pair of terms cancel. Hence, one has proved that
i αβ µ µ,α β µ,β α
[σ , γ ] = g γ − g γ (1427)
2
which completes the identification of the solution of the equation for ω α,β .
Therefore, since R̂(Λ) exists, it has been shown that the form of the Dirac
equation is maintained in the primed reference frame and that there is a one
to one correspondence between the solutions of the primed and unprimed frames.
——————————————————————————————————
It remains to be shown that the ψ and ψ 0 describe the properties of the same
physical system, albeit in two different frames of reference. That is, the proper-
ties associated with ψ must be related to the properties of ψ 0 and the relation
can be obtained by considering the Lorentz transformation. The most complete
physical descriptions of a unique quantum mechanical state are related to the
probability density, which can only be inferred from an infinite set of position
measurements. The probability density, should behave similarly to the time
component of a four-vector as was seen from the consideration of the continuity
equation. Therefore, it follows that if the four-vector probability currents of ψ
and ψ 0 are related via a Lorentz transformation, then the two spinors describe
the same physical state of the system.
j µ0 = c ψ †0 γ (0) γ µ ψ 0
= c ψ † R̂† γ (0) γ µ R̂ ψ (1429)
The identity
R̂−1 = γ (0) R̂† γ (0) (1430)
will be proved below, so on using this identity together with
245
However, because the covariant condition is given by
j µ0 = c ψ † γ (0) Λµ ν γ ν ψ
†
= Λµ ν c ψ γ ν ψ
= Λµ ν j ν (1434)
Hence, the probability current densities j µ0 and j µ found in the two reference
frames are simply related via the Lorentz transformation. Therefore, the Dirac
equation gives consistent results, no matter what inertial frame of reference is
used.
——————————————————————————————————
Proof of Identity
The identity
R̂−1 (Λ) = γ (0) R̂† (Λ) γ (0) (1435)
can be proved by starting from the expression for the expression for R̂ appro-
priate for infinitesimal transformation given by
1
R̂ = Iˆ + µν [ γ µ , γ ν ] + . . . (1436)
8
Hence, the Hermitean conjugate is given by
1
R̂† = Iˆ + µν [ γ ν † , γ µ† ] + . . .
8
1
R̂† = Iˆ − µν [ γ µ† , γ ν † ] + . . . (1437)
8
since the Hermitean conjugate of a product is the product of the Hermitean
conjugate of the factors taken in opposite order. On forming the product
γ (0) R̂† γ (0) and inserting a factor of
between the pairs of four by four γ matrices in the commutator and noting that
246
The last line follows from the observation that on combining the expression for
R̂ with the expression for γ (0) R̂† γ (0) , the terms of order cancel. Hence to
the order of 2 , the product γ (0) R̂† γ (0) coincides with R̂−1 . This concludes
the discussion of the desired identity.
Finite Rotations
x(2)
x(2)'
x(1)'
ϕ
x(1)
Figure 45: A passive rotation of the coordinate system through an angle ϕ about
the ê3 -axis.
Λµ ν = δ µ ν + µ ν + . . . (1444)
247
on lowering the first index, one identifies
248
The above expression can be simplified by expanding the trigonometric functions
in series of ϕ and then using the property of the σ̂ (j) matrices
( σ̂ (3) )2 = Iˆ (1453)
( σ̂ (3) )2n = Iˆ
( σ̂ (3) )2n+1 = σ̂ (3) (1454)
Therefore, under a finite rotation through angle ϕ around the unit vector ê, a
spinor is rotated by the operator
ϕ ˆ ϕ
R̂(ϕ) = cos I + i sin ê . σ̂ (1456)
2 2
From the above equation, due to the presence of the half-angle, one notes that
a rotation ϕ and through ϕ + 2π are not equivalent, since
which changes the sign of the spinor. For spin one-half electrons, it is necessary
to rotate through 4π to return to the same state
A finite Lorentz boost by velocity v along the ê1 direction can be expressed
in terms of the transformation
cosh χ − sinh χ 0 0
− sinh χ cosh χ 0 0
Λ = (1459)
0 0 1 0
0 0 0 1
249
so
1
cosh χ =
1 − ( vc )2
p
v
c
sinh χ = (1461)
1 − ( vc )2
p
250
which can be expressed in terms of even and odd-powers of σ 0,1 via
χ 0,1 χ 0,1
R̂(χ) = cosh i σ + sinh i σ (1469)
2 2
but since
i
σ 0,1 = [ γ (0) , γ (1) ]
2
= i α(1) (1470)
( α(1) )2 = Iˆ (1472)
( α(1) )2n = Iˆ
( α(1) )2n+1 = α(1) (1473)
Exercise:
251
Determine the relationship between the rapidities for a combined Lorentz
transformation consisting of two successive Lorentz boosts with parallel veloci-
ties v0 and v1 .
Exercise:
v A
L'
Exercise:
Show that the helicity eigenvalue of a free Dirac particle can be reversed by
going to a new reference frame which is “overtaking” the particle.
{ γ µ , γ ν }+ = 2 g µ,ν Iˆ (1477)
all other products can be reduced to the above products. The order of the
matrices is irrelevant, since the different matrices anti-commute. Also, since
( γ µ )2 = ± I,ˆ one only needs to consider the products in which each matrix
enters at most one time. Hence, since each of the four matrices either appear as
a factor or do not, there are only 24 such matrices. These sixteen Γi matrices
can be constructed from I, ˆ γ µ , σ µ,ν = i γ µ γ ν , γ (4) and γ (4) γ µ , by choosing
252
Table 6: The Set of the Sixteen Matrices Γn with their Phase Factors (j > i)
Iˆ
γ (0) i γ (i)
The set of matrices Γi formed from the set of γ µ are closed under multipli-
cation, so
Γi Γj = ai,j Γk (1478)
where a4i,j = 1. The sixteen Γi matrices can be chosen as the product of the
members of the above set multiplied by a phase factor taken from the set ± 1
and ± i, such that the condition
( Γi )2 = Iˆ (1479)
Γi Γj = Iˆ only if i = j (1480)
Also, by anti-commuting the factors of γ µ in the products, one can show that
Γi Γj = ± Γj Γi (1481)
Specifically, for a fixed Γi not equal to the identity, one can always find a
specific Γk such that
Γi Γk = − Γk Γi (1482)
which on multiplying by Γk results in
Γk Γi Γk = − Γi (1483)
253
Traceless Matrices
The above facts can be used to show that the Γi matrices, other than the
identity, are traceless. This can be proved by considering
Trace Γi = 0 (1485)
Linear Independence
other than Ci ≡ 0 for all i. If the Γi are linearly independent, the only solution
of this equation is
Ci ≡ 0 for all i (1487)
This can be proved by multiplying eqn(1486) by any one Γj in the set which
leads to
X
Cj Iˆ + Ci Γi Γj = 0
i6=j
X
Cj Iˆ + Ci ai,j Γk = 0 (1488)
i6=j
since the matrices Γk are traceless. Hence, all the Cj are zero, so the matrices
are linearly independent.
254
Uniqueness of Expansions
The existence of sixteen linearly independent matrices require that the ma-
trices can be represented in a space of N × N matrices, where N ≥ 4. Any
matrix A in the space of 4 × 4 matrices can be uniquely expressed in terms of
the basis set of the Γi . For example, if
X
A = Ci Γi (1490)
i
1
Cj = Trace( A Γj ) (1492)
4
Schur’s Lemma
The uniqueness of the expansion can be used to show that the product of Γi
for fixed i with the set of Γj for leads to a different Γk for each j. This can be
shown by assuming that there exist two different (linearly independent) values
Γj and Γj 0 which lead to the same Γk
Γi Γj = ai,j Γk
Γi Γj 0 = ai,j 0 Γk
(1493)
Γj = ai,j Γi Γk
Γj 0 = ai,j 0 Γi Γk
(1494)
255
which contradicts the assumption that Γj and Γj 0 are linearly independent.
Therefore for fixed i, the product of Γi Γj leads to a different result Γk for the
different Γj .
One can also prove Schur’s lemma. Schur’s Lemma states that if a matrix A
commutes with all the γ µ ’s, then A is a multiple of the identity. If A commutes
with the γ µ ’s, it also commutes with all the Γi ’s. Schur’s lemma follows from
the expansion of A as X
A = Ci Γi + Cj Γj (1496)
j6=i
Γk Γi Γk = − Γi (1497)
Since it has been assumed that A commutes with all the Γi , for the specific Γk
one has
A = Γk A Γk
X
= Ci Γk Γi Γk + Cj Γk Γj Γk
j6=i
X
= − Ci Γi + Cj Γk Γj Γk (1498)
j6=i
Γk Γj Γk = ( ± 1 )j,k Γj (1499)
Since the expansion is unique, the coefficients of the Γj are unique and in par-
ticular
Ci = − Ci (1502)
ˆ Hence, if A commutes with all the Γi
so Ci = 0 for any i such that Γi 6= I.
then A must be proportional to the identity.
256
Pauli’s fundamental theorem states that if there are two representations of
the algebra of anti-commuting γ-matrices, say γ µ0 and γ µ , then these represen-
tations are related via a similarity transformation
γ µ0 = Ŝ γ µ Ŝ −1 (1503)
The theorem requires that one constructs a set of sixteen matrices Γi 0 from
the γ µ0 following the same rules with which the Γi were constructed from γ µ .
Then one can describe the non-singular matrix by
X
Ŝ = Γi 0 F Γi (1504)
i
since
Γ2k = Iˆ (1507)
On pre-multiplying eqn(1506) by Γj Γi , one obtains
Γj Γi Γi Γj Γi Γj = a2i,j Γj Γi (1508)
but since
Γj Γi Γi Γj = Iˆ (1509)
eqn(1508) reduces to
Γi Γj = a2i,j Γj Γi (1510)
However, as
Γi Γj = ai,j Γk (1511)
the equation becomes
Γj Γi = a3i,j Γk (1513)
257
0
The Γi matrices are constructed so that they satisfy similar relations to the Γi .
0
In particular, the Γi matrices satisfy
Γi 0 Γj 0 = ai,j Γk 0 (1514)
However, since i is fixed and j is being summed over, every Γk appears once
and only once in the product. Therefore, the sum can be performed over k
X
Γi 0 Ŝ Γi = Γk 0 F Γk = Ŝ (1520)
k
If one can show that the matrix Ŝ has an inverse, then on post-multiplying by
Ŝ −1 , one finds
Γi 0 Ŝ Γi Ŝ −1 = Iˆ (1521)
Furthermore, since Γi 0 is its own inverse, then on pre-multiplying by Γi 0 the
equation reduces to
Ŝ Γi Ŝ −1 = Γi 0 (1522)
This is a generalization of the statement of the theorem. As a particular case,
one may choose Γi = γ µ in which case the theorem becomes
γ µ0 = Ŝ γ µ Ŝ −1 (1523)
which was the initial statement of Pauli’s fundamental theorem made above.
258
The matrix Ŝ is non-singular and has an inverse. This can be shown by
using Schur’s Lemma. One can construct a matrix Ŝ 0 in a manner which is
symmetrical to the construction of Ŝ. That is
X
Ŝ 0 = Γi G Γi 0 (1524)
i
Ŝ = Γi 0 Ŝ Γi (1525)
Ŝ 0 Ŝ = Γi Ŝ 0 Γi 0 Γi 0 Ŝ Γi
= Γi Ŝ 0 Ŝ Γi (1527)
Hence, by Schur’s Lemma one sees that Ŝ 0 Ŝ commutes with all the matrices in
the space, therefore it must be a multiple of the identity
Ŝ 0 Ŝ = κ Iˆ (1528)
Ŝ 0 Ŝ = Iˆ (1529)
The probability of non-helicity flip scattering and helicity flip scattering can
be evaluated using the Born approximation. The initial beam will be considered
as having a momentum p parallel to the ê3 -axis and as having a helicity of +1.
The initial spinor is proportional to
s
Ep + m c2 p.r
χ+
ψp,+ (r) = c p exp i (1530)
2 Ep V Ep + m c2 χ+ h̄
101 N. F. Mott, Proc. Roy. Soc. A 124, 425 (1929).
259
p'
p θ'
p'
Figure 47: Helcity non-flip and helicity flip Mott scattering of an electron with
helicity +1. The scattering angle is θp0 .
The electrons are assumed to be elastically scattered to a state with final mo-
mentum p0 . The scattering is defined to occur through an angle θp0 in the z − x
plane. The final state is composed of a linear-superposition of states with dif-
ferent helicities. Since the final state helicities are specified relative to the final
momentum, the final state helicity eigenstates can be obtained by rotating the
initial state helicity eigenstates through an angle θp0 around the ê2 -axis
and
θp0 θp0 (2)
χ0− = cos I − i sin σ χ−
2 2
260
!
θ 0
− sin 2p
= θ 0 (1534)
cos 2p
which is evaluated as
2 2 2 2
4 π Z e2 c2 p2 E p + m c2
0
0†
1 + Λ χ Λ0 χ+
V | p − p0 |2 ( Ep + m c2 )2 2 Ep
2 2 2
2
( Ep + m c ) + Λ0 c2 p2
2 2
4πZ e 0†
= χ Λ0 χ+ (1537)
V | p − p0 |2 2 E p ( E p + m c2 )
261
whereas the cross-section for spin flip scattering is given by
2
2 Z e2 m c 2
dσ θp0
= sin2 (1543)
dΩ0 +,− 4 c2 p2
θ 0
sin2 2p 2
The equation can be written in 2 × 2 block diagonal form, if the wave function
is expressed in the form of two two-component spinors. We shall mainly focus
on the positive-energy solutions and recognize that, in the non-relativistic limit,
the largest component of the wave function is φA and the largest term in the
102 W. Pauli, Z, Phys. 44, 601 (1927).
262
energy is the rest mass energy m c2 . Therefore, the spinor wave function will
be expressed as
m c2
A
φ
ψ = exp − i t (1547)
φB h̄
The above form explicitly displays the rest-mass energy of the positive-energy
solution of the Dirac equation. Hence, the Dirac equation takes the form
c σ . ( p̂ − qc A ) φB
A
∂ φ
i h̄ − q A0 =
∂t φB c σ . ( p̂ − qc A ) φA
2 0
−2mc (1548)
φB
where the rest mass has been eliminated from the equation for the large com-
ponent φA of the positive-energy solution. Since the kinetic energy and the
potential energy are assumed to be smaller than the rest mass energy, the equa-
tion for the small component
∂ B q
i h̄ − q A0 φ = c σ . p̂ − A φA − 2 m c2 φB (1549)
∂t c
can be expressed as
1 q
φB = σ. p̂ − A φA (1550)
2mc c
Substituting the expression for the small component into the equation for the
large component, hence eliminating φB , one finds the equation
2
∂ A 1 q
i h̄ − q A0 φ = σ . p̂ − A φA (1551)
∂t 2m c
which is the Pauli equation. The equation can be simplified by expanding the
terms involving the Pauli spin matrices. The Pauli identity can be used to
obtain
2 2
q q q q
σ . p̂ − A = I p̂ − A + iσ. p̂ − A ∧ p̂ − A
c c c c
2
q q h̄
= I p̂ − A − σ. ∇ ∧ A (1552)
c c
where the last term originates from the non-commutativity of the components
of p̂ and A. Since the magnetic field B is given by
B = ∇ ∧ A (1553)
263
The Pauli equation103 is the non-relativistic limit of the Dirac equation. It rep-
resents the Schrödinger equation for a charged particle with spin one-half. The
two components of the spinor φA in the Pauli equation represent the internal
spin of the electron. The last term represents the anomalous Zeeman interaction
between the magnetic field and the electron’s spin.
The other contribution to the Zeeman interaction originates with the elec-
tron’s orbital angular momentum L. The ordinary Zeeman interaction occurs
between the constant magnetic field B and the orbital angular momentum and
originates from the gauge-invariant term in the Hamiltonian
2 2
1 q 1 q
p̂ − A = p̂ − B ∧ r (1555)
2m c 2m 2c
where the vector potential has been expressed in terms of the uniform magnetic
field via
1
A = B ∧ r (1556)
2
The expression for the energy term can be further simplified to
2
p̂2
1 q q
p̂ − A = − p̂ . (B ∧ r) + (B ∧ r) . p̂
2m c 2m 4mc
2
q
+ A2
2 m c2
p̂2 q2
q
= − (B ∧ r) . p̂ + A2
2m 2mc 2 m c2
p̂2 q2
q
= − B . (r ∧ p̂) + A2
2m 2mc 2 m c2
p̂2 q2
q
= − B . L̂ + A2 (1557)
2m 2mc 2 m c2
In obtaining the second line, the i-th component of p̂ has been commuted with
the i-th component of ( B ∧ r ). In obtaining the third line, the (cyclic) vector
identity
(A ∧ B).C = (B ∧ C ).A (1558)
has been used. The first term in eqn(1557) represents the usual non-relativistic
expression for the kinetic energy of the electrons, the second term represents
the ordinary Zeeman interaction which originates from the paramagnetic inter-
action. The last term represents the diamagnetic interaction.
The total Zeeman interaction is the energy of the total magnetic moment M
in the field B
ĤZeeman = − M . B (1559)
103 W. Pauli, Z, Phys. 44, 601 (1927).
264
The Dirac equation results in the Zeeman interaction of the form
q
ĤZeeman = − B . L̂ + h̄ σ
2mc
q
= − B . L̂ + 2 S (1560)
2mc
where the spin angular momentum S has been identified as
h̄
S = σ (1561)
2
It is seen that both the spin angular momentum and the orbital angular momen-
tum of the charged particle interacts with the magnetic field, therefore, both
contribute to the magnetic moment. However, it is noted that the magnetic
moment can be written in the form
q
M = L̂ + g S (1562)
2mc
where the magnitude of the magnetic moment is determined by the factor 2 qmh̄ c
which is the Bohr magneton. The Dirac equation shows that the spin angular
momentum couples with a different strength to orbital angular momentum, and
the relative coupling strength g (the gyromagnetic ratio) is given by g = 2.
The existence of spin and the value of 2 for the gyromagnetic ratio were the
first successes of Dirac’s theory. Quantum Electrodynamics104 yields a small
correction to the gyromagnetic ratio of
2
1 e
g = 2 1 + + ... (1563)
2 π h̄ c
which has been experimentally verified to incredible precision105 . Using the fea-
tures associated with spin, Dirac’s theory correctly described the fine structure
of the Hydrogen atom. The second success of the Dirac equation followed Dirac’s
physical interpretation of the negative-energy states in terms of anti-particles106 .
The second round of success came with the discovery of the positron by Ander-
son107 .
Exercise:
265
Show that the modified equation is Lorentz covariant and that the Hamiltonian
is Hermitean. Also derive the corrections to the magnetic moment due to the
spin by examining the non-relativistic limit.
L̂ = r ∧ p̂ (1565)
and
I 0
β = (1568)
0 −I
Finally, the α matrices are of off-diagonal form
0 σ
α =
σ 0
0 I
= σ̂ (1569)
I 0
266
Hence, the Heisenberg equation of motion can be expressed in the form
∂ 0 I
i h̄ L̂ = c [ L̂ , σ̂ . p̂ ]
∂t I 0
0 I
= −c σ̂ . [ p̂ , L̂ ] (1572)
I 0
However, the components of the orbital angular momentum L̂(i) and momenta
p(j) satisfy the commutation relations
X
[ L̂(i) , p(j) ] = i h̄ ξ i,j,k p(k) (1573)
k
which shows that orbital angular momentum is not conserved for a relativistic
electron with a central potential.
The spin angular momentum is also not conserved. This can be seen by
examining the Heisenberg equation of motion for the Pauli spin operator
∂
i h̄ σ̂ = [ σ̂ , Ĥ ] (1575)
∂t
The spin operator commutes with Iˆ and β but does not commute with the α
matrices. Hence,
∂
i h̄ σ̂ = c [ σ̂ , α . p̂ ]
∂t
0 I
= c [ σ̂ , σ̂ ] . p̂ (1576)
I 0
The components of the Pauli spin operators satisfy the commutation relations
X
[ σ (i) , σ (j) ] = 2 i ξ i,j,k σ (k) (1577)
k
which, clearly, have a similar form to the commutation relations for the orbital
angular momentum. Hence, spin angular momentum is not conserved since
∂ 0 I
i h̄ σ̂ = − 2 i c ( σ̂ ∧ p̂ ) (1578)
∂t I 0
267
The total angular momentum Jˆ is defined via
Jˆ = L̂ + Ŝ
h̄
= L̂ + σ̂ (1579)
2
The total angular momentum is conserved since
∂ ˆ ∂ ∂
i h̄ J = i h̄ L̂ + i h̄ Ŝ
∂t ∂t ∂t
0 I 0 I
= i h̄ c ( σ̂ ∧ p̂ ) − ( σ̂ ∧ p̂ )
I 0 I 0
= 0 (1580)
which follows from combining eqn(1574) and eqn(1578). This confirms the in-
terpretation of the quantity Ŝ defined by
h̄
Ŝ = σ̂ (1581)
2
as the spin angular momentum of the electron.
The parity transform P acting on the coordinates (t, r) has the effect
which is an inversion of the spatial coordinates. Thus, the parity reverse the
space-like components of vectors, so the effects of the parity operation on the
position and momentum vectors are given by
P̂ r P̂ −1 = −r
P̂ p P̂ −1 = −p (1583)
P̂ L P̂ −1 = L (1584)
108 The question of parity conservation in weak interactions was raised subsequently by T. D.
Lee and C. N. Yang [T. D. Lee and C. N. Yang, Phys. Rev. 104, 254 (1956).]
268
which is unchanged. This implies that spin angular momentum should also be
invariant under the parity transform
P̂ σ P̂ −1 = σ (1585)
If the Hamiltonian Ĥ is invariant under a parity transform, one requires that
Ĥ = P̂ Ĥ P̂ −1 (1586)
Imposing parity invariance of the Dirac Hamiltonian
Ĥ = c α . p̂ + β m c2 + Iˆ V (r) (1587)
yields a condition on the potential
V (r) = V (−r) (1588)
and also to conditions on the Dirac matrices
P̂ α P̂ −1 = −α
P̂ β P̂ −1 = β (1589)
The condition on the potential is the familiar condition for parity invariance
in classical mechanics. In the standard representation, in 2 × 2 block diagonal
form, the requirement of parity invariance on the Dirac matrices become the
matrix equations
0 σ 0 σ
P̂ P̂ −1 = −
σ 0 σ 0
I 0 I 0
P̂ P̂ −1 = (1590)
0 −I 0 −I
The above equation shows that, in the standard representation, the parity op-
erator can be uniquely factorized as
I 0
P̂ = P̂ (1591)
0 −I
where the operator P̂ only acts on the coordinates r. The presence of the matrix
in the parity operation on the Dirac spinor should be compared with the effect
of the parity operator on the four-vector potential of Electrodynamics Aµ (r)
which is given by the product of spatial inversion and a matrix operation
P̂ Aµ (r, t) = γ µ ν P̂ Aν (r, t)
= γ µ ν Aν (−r, t) (1592)
where the matrix γ µ ν given by
1 0 0 0
0 −1 0 0
γµν =
0 0 −1
(1593)
0
0 0 0 −1
269
reverses the direction of the spatial components of the vector field.
The effect of the parity operator on the Dirac four-component spinor wave
function can be computed from
A
φ (t, r)
P̂ ψ(t, r) = P̂
φB (t, r)
A
I 0 φ (t, −r)
=
0 −I φB (t, −r)
φA (t, −r)
= (1594)
− φB (t, −r)
Hence, in the standard representation, the parity operator changes the relative
sign of the two two-component spinors. Due to the presence of the term − I in
the lower diagonal block of the parity matrix, the lower two-component spinor
φB in the Dirac spinor is said to have a negative intrinsic parity.
P̂ ψ = ηp ψ (1595)
P̂ φA ηp φA
= (1596)
−P̂ φB ηp φB
Hence, the two-component spinors φA (r) and φB (r) have opposite parities under
spatial inversion
P̂ φA (r) = ηp φA (r)
P̂ φB (r) = − ηp φB (r) (1597)
sin θ → sin θ
cos θ → − cos θ
m
exp i m ϕ → ( − 1) exp i m ϕ (1600)
270
Hence, the spherical harmonics with m = l
r
( − 1 )l
l 2l + 1 l
Yl (θ, ϕ) = sin θ exp i l ϕ (1601)
2l l! 4π
are eigenstates of the parity operator and have parity eigenvalues of (−1)l . The
lowering operator L̂− , defined via
− ∂ ∂
L̂ = − h̄ exp − i ϕ − i cot θ (1602)
∂θ ∂ϕ
is invariant under the parity transformation
Therefore, on repeatedly operating on Yll (θ, ϕ) with the lowering operator L̂−
(l − m) times, one finds that under the parity transformation
which shows that all states with a definite magnitude of the orbital angular mo-
mentum l are eigenstates of the parity operator and have the same eigenvalue.
Exercise:
Show that under a parity transformation the positive-energy solution for the
+
free Dirac particle ψk,σ (x) transforms as
+ +
P̂ ψk,σ (x) = ψ−k,σ (x) (1605)
−
while the negative-energy solutions ψk,σ (x) transform as
− −
P̂ ψk,σ (x) = − ψ−k,σ (x) (1606)
Hence, the parity operation reverses the momentum and keeps the spin invari-
ant for the positive-energy and negative-energy solutions solution. The extra
negative sign implies that the negative-energy solution has opposite intrinsic
parity to the positive-energy solution.
Exercise:
271
where R̂(Λ) “rotates” the spinor. The covariant condition for the Dirac equation
is
R̂−1 (Λ) γ µ R̂(Λ) = Λµ ν γ ν (1609)
For a parity transformation, one has
xµ0 = xµ (1610)
µ
since the spatial components of x change sign. Hence, for a parity transforma-
tion, the transformation matrix is determined as
Λµ ν = gµ,ν (1611)
which is an improper Lorentz transformation since
det | g | = − 1 (1612)
Therefore, the covariant condition reduces to
R̂−1 (Λ) γ µ R̂(Λ) = gµ,ν γ ν (1613)
Solve for the matrix R̂(Λ) which shuffles the components of the Dirac spinor.
272
†
Table 7: The sixteen bi-linear covariants ψ Q̂ ψ for the Dirac equation.
†
ψ Q̂ ψ R̂−1 (Λ) Q̂ R̂(Λ)
†
Scalar ψ Iˆ ψ Iˆ 1
†
Vector ψ γµ ψ Λµ ν γ ν 4
†
Anti-symmetric Tensor ψ σ µ,ν ψ Λµ ρ Λν τ σ ρ,τ 6
†
Pseudo-scalar ψ γ (4) ψ det | Λ | γ (4) 1
†
Axial-Vector ψ γ (4) γ µ ψ det | Λ | Λµ ν γ (4) γ ν 4
where a factor of ( γ (0) )2 = Iˆ has been used in the third line and the identity
†
has been used in the fourth. Thus, one finds that ψ ψ transforms like a scalar.
†
Likewise, one can show that the bi-linear quantities ψ γ µ ψ transform like
the components of a four-vector. That is
†0
ψ γ µ ψ0 = ψ †0 γ (0) γ µ ψ 0
= ψ † R̂† (Λ) γ (0) γ µ R̂(Λ) ψ
= ψ † ( γ (0) )2 R̂† (Λ) γ (0) γ µ R̂(Λ) ψ
= ψ † γ (0) R̂−1 (Λ) γ µ R̂(Λ) ψ
†
= ψ R̂−1 (Λ) γ µ R̂(Λ) ψ
†
= Λµ ν ψ γ ν ψ (1619)
where the covariant condition has been used in obtaining the last line. Since
†
this relation holds for Lorentz boosts, rotations and spatial inversions, ψ γ µ ψ
is a four-vector.
273
transforms like a second-rank anti-symmetric tensor, since
†0
ψ σ µ,ν ψ 0 = ψ †0 γ (0) σ µ,ν ψ 0
= ψ † R̂† (Λ) γ (0) σ µ,ν R̂(Λ) ψ
= ψ † ( γ (0) )2 R̂† (Λ) γ (0) σ µ,ν R̂(Λ) ψ
= ψ † γ (0) R̂−1 (Λ) σ µ,ν R̂(Λ) ψ
†
= ψ R̂−1 (Λ) σ µ,ν R̂(Λ) ψ (1621)
σ µ,ν = i γ µ γ ν (1622)
where we have inserted a factor of Iˆ = R̂(Λ) R̂−1 (Λ) in the second line, and
used the covariant condition (twice) in the third line. Hence, the bi-linear quan-
†
tity ψ σ µ,ν ψ transforms like an anti-symmetric second-rank tensor.
One can define a quantity γ (4) in terms of a product of all the γ-matrices
{ γ µ , γ (4) }+ = 0 (1625)
274
= ψ † R̂† (Λ) γ (0) γ (4) R̂(Λ) ψ
= ψ † ( γ (0) )2 R̂† (Λ) γ (0) γ (4) R̂(Λ) ψ
†
= ψ γ (0) R̂† (Λ) γ (0) γ (4) R̂(Λ) ψ
†
= ψ R̂−1 (Λ) γ (4) R̂(Λ) ψ (1628)
so one has
† †
ψ 0 (x0 ) γ (4) ψ 0 (x0 ) = det | Λ | ψ (x) γ (4) ψ(x) (1632)
†
Therefore, the quantity ψ γ (4) ψ transforms as a pseudo-scalar.
†
One can also define the bi-linear axial-vector ψ γ (4) γ µ ψ. From consid-
erations similar to those used previously, one can show that these quantities
transform according to
† †
ψ 0 (x0 ) γ (4) γ µ ψ 0 (x0 ) = det | Λ | Λµ ν ψ (x) γ (4) γ ν ψ(x) (1633)
†
Hence, ψ γ (4) γ µ ψ transforms like a four-vector under proper orthochronous
Lorentz transformations. However, the spatial components do not change sign
†
under an inversion, but the time components do change sign. Therefore, ψ γ (4) γ µ ψ
transforms like an axial-vector.
Exercise:
275
is covariant under proper Lorentz transformations, but is not covariant under
improper transformations.
Show, by considering the non-relativistic limit, that the above equation de-
scribes an electron with an electric dipole moment. Determine an expression for
the electric dipole moment.
The angular momentum operator Jˆ and the parity operator P̂ commute with
the Hamiltonian Ĥ. Therefore, one can find simultaneous eigenstates of the
three operators Ĥ, Jˆ2 , Jˆz and P̂. The energy eigenstates satisfy the equation
c α . p̂ + β m c2 + Iˆ V (r) ψ = E ψ (1636)
which has a quite complicated structure. For future reference, it shall be noted
that the matrix part of the coefficient of the partial derivative w.r.t. r is simply
equal to
r.σ
(1639)
r
which is independent of the radial coordinate r. The operator ( σ . p̂ ) can be
cast in a more convenient form through the repeated use of the Pauli identity.
276
First, the 2 × 2 unit matrix can be written as
2
r.σ
I = (1640)
r
and are their own inverses. Therefore, one can express the operator ( σ . p̂ ) as
2
r.σ
( σ . p̂ ) = ( σ . p̂ )
r
r.σ
= ( r . σ ) ( σ . p̂ )
r2
r.σ
= r . p̂ + i σ . ( r ∧ p̂ )
r2
r.σ ∂
= − i h̄ r + i σ . L̂
r2 ∂r
r.σ ∂ 2i
= − i h̄ r + S . L̂ (1642)
r2 ∂r h̄
where the Pauli identity has been used in going between the second and third
lines. Therefore, the two-component spinors satisfy the set of coupled equations
2 A r.σ ∂ 2i
( E − V (r) − m c ) φ (r) = c − i h̄ r + S . L̂ φB (r)
r2 ∂r h̄
2 B r.σ ∂ 2i
( E − V (r) + m c ) φ (r) = c − i h̄ r + S . L̂ φA (r)
r2 ∂r h̄
(1643)
It is seen that, due to the effect of special relativity, the Dirac equation results
in the coupling of the spin and the orbital angular momentum.
J = L + S (1644)
Thus, the two-component spinor eigenstates of total angular momentum Ωlj,jz (θ, ϕ)
which describes the angular dependence, are formed by combining states of or-
bital angular momentum l, represented by Yml (θ, ϕ), and the spin eigenfunction
277
Table 8: The Clebsch-Gordon Coefficients for adding orbital angular momentum
(l, m) with spin quantum numbers ( 21 , sz ) to yield a state with total angular
momentum quantum numbers (j, jz ). The allowed values of m are given by
jz = m + sz .
sz = + 12 sz = − 12
q 1
q 1
1 l + jz + l − jz +
j =l+ 2 2 l + 1
2
2 l + 1
2
q 1
q
1 l − jz + l +jz + 12
j =l− 2 - 2 l + 1
2
2 l + 1
(1646)
where the coefficients are identified with the Clebsch-Gordon coefficients given
in Table(8). The functions Ωlj,jz (θ, ϕ) are the analogue of the spherical harmon-
ics Yml (θ, ϕ) in relativistic problems where spin and orbital angular momentum
are coupled.
278
where l0 = l + 1. The appropriate two-component spinor angular momentum
eigenstate with quantum numbers (j, jz ) found by combining a spin one-half
and orbital angular momentum l0 = (l + 1) is given by
s s
l+1 l + 32 − jz l+1 l + 32 + jz l+1
Ωl+ 1 ,j (θ, ϕ) = − Yj − 1 (θ, ϕ) χ+ + Yj + 1 (θ, ϕ) χ−
2 z 2l + 3 z 2 2l + 3 z 2
(1648)
0
As shall be seen later, the two-component spinors Ωll0 − 1 ,jz (θ, ϕ) and Ωll+ 1 ,jz (θ, ϕ)
2 2
have opposite parities. In fact, the two-component spinors generated by angular
momentum l and l0 = (l + 1) are related by the action of the pseudo-scalar
r.σ cos θ sin θ exp[−iϕ]
= (1649)
r sin θ exp[+iϕ] − cos θ
one finds that the inverse relationship between the two-component spinors is
also given by
r.σ
Ωl+1 l
j,jz (θ, ϕ) = − Ωj,jz (θ, ϕ) (1652)
r
Therefore, one concludes that the two angular momentum eigenstates have dif-
ferent properties under the spatial inversion transformation r → −r.
——————————————————————————————————
Mathematical Interlude:
The Action of the Operator ( r̂ . σ ) on the Spinor Spherical Harmon-
j± 1
ics Ωj,jz2 (θ, ϕ).
Here, it will be argued that the spinor spherical harmonics satisfy the equa-
tions
r.σ j+ 1 j− 1
Ωj,jz2 (θ, ϕ) = − Ωj,jz2 (θ, ϕ)
r
r.σ j− 1 j+ 1
Ωj,jz2 (θ, ϕ) = − Ωj,jz2 (θ, ϕ) (1653)
r
279
The components of the total angular momentum
Jˆ(i) = L̂(i) + Ŝ (i) (1654)
commute with ( r . Ŝ ). That is
[ Jˆ(i) , ( r . Ŝ ) ] = 0 (1655)
The complete proof of this statement immediately follows from the proof of the
relation for any one component Jˆ(i) , since ( r . Ŝ ) is spherically symmetric.
Thus, for i = 1, one has
[ Jˆ(1) , ( r . Ŝ ) ] = [ L̂(1) + Ŝ (1) , x(1) Ŝ (1) + x(2) Ŝ (2) + x(3) Ŝ (3) ]
= Ŝ (2) [ L̂(1) , x(2) ] + Ŝ (3) [ L̂(1) , x(3) ]
+ x(2) [ Ŝ (1) , Ŝ (2) ] + x(3) [ Ŝ (1) , Ŝ (3) ] (1656)
Using the commutation relations
[ Ŝ (i) , Ŝ (j) ] = i h̄ εi,j,k Ŝ (k) (1657)
and
[ L̂(i) , x(j) ] = i h̄ εi,j,k x(k) (1658)
one finds that
[ Jˆ(1) , ( r . Ŝ ) ] = i h̄ Ŝ (2) x(3) − Ŝ (3) x(2) + x(2) Ŝ (3) − x(3) Ŝ (2)
= 0 (1659)
which was to be shown. From repeated use of the above commutation relations
which involve the components Jˆ(i) , it immediately follows that
2
[ Ĵ , ( r . Ŝ ) ] = 0 (1660)
1 2
j±
Thus, since Ωj,jz2 is a simultaneous eigenstate of Ĵ and Jˆ(3) and because these
j± 1
operators commute with ( r . Ŝ ), then ( r . Ŝ ) Ωj,jz2 is also a simultaneous
eigenstate with eigenvalues (j, jz ).
1 2
j±
Since the states ( r . Ŝ ) Ωj,jz2 are simultaneous eigenstates of Ĵ and Jˆ(3)
with eigenvalues (j, jz ), and because this subspace is spanned by the basis com-
j± 1
posed of the two states Ωj,jz2 (θ, ϕ), the transformed states can be decomposed
as
r.σ j+ 1 j+ 1 j− 1
Ωj,jz2 (θ, ϕ) = C++ (j, jz ) Ωj,jz2 (θ, ϕ) + C+− (j, jz ) Ωj,jz2 (θ, ϕ)
r
r.σ j− 1 j+ 1 j− 1
Ωj,jz2 (θ, ϕ) = C−+ (j, jz ) Ωj,jz2 (θ, ϕ) + C−− (j, jz ) Ωj,jz2 (θ, ϕ)
r
(1661)
280
where the coefficients C±,± (j, jz ) will be determined below.
First, we shall show that the coefficients C±,± (j, jz ) are independent of jz .
This follows as Jˆ± commutes with ( r . Ŝ ) since all the components Jˆ(i) com-
mute with ( r . Ŝ ). Thus, one has
ˆ ± r.σ j+ 12 r . σ ˆ± j+ 12
J Ωj,jz (θ, ϕ) = J Ωj,jz (θ, ϕ) (1662)
r r
and
r.σ j+ 1 j+ 1 j− 1
Jˆ± Ωj,jz2 (θ, ϕ) = C++ (j, jz ) Jˆ± Ωj,jz2 (θ, ϕ) + C+− (j, jz ) Jˆ± Ωj,jz2 (θ, ϕ)
r
r . σ ˆ± j+ 12 j+ 1 j− 1
J Ωj,jz (θ, ϕ) = C++ (j, jz ± 1) Jˆ± Ωj,jz2 (θ, ϕ) + C+− (j, jz ± 1) Jˆ± Ωj,jz2 (θ, ϕ)
r
(1663)
Hence, on comparing the linearly-independent terms on the left-hand sides, one
concludes that
C++ (j, jz ± 1) = C++ (j, jz )
C+− (j, jz ± 1) = C+− (j, jz ) (1664)
etc. Therefore, the coefficients C±,± (j, jz ) are independent of the value of jz .
Henceforth, we shall omit the index jz in C±,± (j, jz ).
From considerations of parity, it can be determined that C++ (j) = C−− (j) =
0. Under the parity transformation r → − r, one has
j± 1 1 j± 1
Ωj,jz2 (θ, ϕ) → ( − 1 )j± 2 Ωj,jz2 (θ, ϕ) (1665)
which follows from the properties of the spherical harmonics Yml (θ, ϕ) under the
parity transformation. Also one has
r.σ r.σ
→ − (1666)
r r
under the parity transform. Thus, after the parity transform, one finds that the
transformed states have the decompositions
r.σ j+ 1 j+ 1 j− 1
Ωj,jz2 (θ, ϕ) = − C++ (j) Ωj,jz2 (θ, ϕ) + C+− (j) Ωj,jz2 (θ, ϕ)
r
r.σ j− 1 j+ 1 j− 1
Ωj,jz2 (θ, ϕ) = C−+ (j) Ωj,jz2 (θ, ϕ) − C−− (j) Ωj,jz2 (θ, ϕ)
r
(1667)
which by comparison with eqn(1661) leads to the identification
C++ (j) = C−− (j) = 0 (1668)
281
Therefore, recalling that the coefficients are independent of jz , one can express
the effect of the operator on the spinor spherical harmonics as
r.σ j+ 1 j− 1
Ωj,jz2 (θ, ϕ) = C+− (j) Ωj,jz2 (θ, ϕ)
r
r.σ j− 1 j+ 1
Ωj,jz2 (θ, ϕ) = C−+ (j) Ωj,jz2 (θ, ϕ) (1669)
r
Furthermore, since
2
r.σ
= I (1670)
r
one obtains the condition
C+− (j) C−+ (j) = 1 (1671)
r . σ
This condition can be made more restrictive as r is Hermitean, which
leads to
C+− (j) = C−+ (j)∗ (1672)
The above two equations suggest that C−+ (j) and C+− (j) are pure phase fac-
tors, such as
C+− (j) = exp + i φ(j)
C−+ (j) = exp − i φ(j) (1673)
which becomes real for ϕ = 0 since the spherical harmonics become real. Hence,
on inspecting eqn(1669) with ϕ = 0, one concludes that the phase factors are
equal and are purely real. That is
C+− (j) = C−+ (j) = ± 1 (1676)
282
Finally, by considering θ = 0, for which
r.σ 1 0
= (1677)
r 0 −1
are non-zero. The spinor spherical harmonics with θ = 0 are connected via
1 0 j± 1 j∓ 1
Ωj,±21 (0, ϕ) = − Ωj,±21 (0, ϕ) (1680)
0 −1 2 2
which holds independent of the values of θ and j, so the effect of the operator
on the spinor spherical harmonics is completely specified by
r.σ j+ 1 j− 1
Ωj,jz2 (θ, ϕ) = − Ωj,jz2 (θ, ϕ)
r
r.σ j− 1 j+ 1
Ωj,jz2 (θ, ϕ) = − Ωj,jz2 (θ, ϕ) (1682)
r
as was to be shown.
——————————————————————————————————
The Ansatz
If one only considers the spatial part of the parity operator, P̂ , the two-
0 0
component spinor states Ωll0 ± 1 ,jz (θ, ϕ) have parities (−1)l
2
0 0 0
P̂ Ωll0 ± 1 ,jz (θ, ϕ) = (−1)l Ωll0 ± 1 ,jz (θ, ϕ) (1683)
2 2
283
Furthermore, as has been seen, the upper and lower two-component spinors of
the four-component Dirac spinor must have opposite intrinsic parity. Therefore,
the desired simultaneous eigenstates for the relativistic electron can be either
−
represented by the four-component Dirac spinor ψj,j z
(r) with parity (−1)l =
1 1
(−1)(l+ 2 − 2 ) of the form
f − (r)
l
− r Ω 1
l+ ,j
(θ, ϕ)
ψl+ 1 (r) = g− (r) l+12 z (1684)
2 ,jz i r Ωl+ 1 ,j (θ, ϕ)
2 z
+
or by ψj,j z
(r)
f + (r)
+ r Ωl+1
l+ 1 ,j
(θ, ϕ)
z
ψl+ 1
,j
(r) = g + (r)
2 (1685)
2 z
i r Ωll+ 1 ,jz (θ, ϕ)
2
1 1
which has parity (−1)l+ 2 + 2 . In these expressions f ± (r) and g ± (r) are scalar
radial functions that have to be determined as solutions of the radial equation.
These states do not correspond to definite values of the orbital angular mo-
mentum since the upper and lower two-component spinors correspond to the
different values of either l or l0 = l + 1 for the orbital angular momentum.
To condense the notation, the energy eigenstates will be written in the com-
pact form
f ± (r)
!
lA
± r Ω j,jz (θ, ϕ)
ψj,jz (r) = ± (1686)
i g r(r) Ωlj,j
B
z
(θ, ϕ)
1 1
where lA = j ± 2 and lB = j ∓ 2.
Jˆ = L̂ + S (1687)
When this operator acts on the relativistic two-component spinor spherical har-
monic Ωlj,j
A
z
, one finds
h̄2
3
S . L̂ Ωlj,j
A
= j ( j + 1 ) − lA ( lA + 1) − Ωlj,j
A
(1689)
z
2 4 z
284
1
which for j = lA + 2 yields
h̄2 1
S . L̂ Ωlj,j
A
= (j − ) Ωlj,j
A
(1690)
z
2 2 z
h̄2 3
S . L̂ Ωlj,j
A
= − (j + ) Ωlj,j
A
(1691)
z
2 2 z
2
h̄
( S . L̂ ) Ωlj,j
B
= − ( 1 − κ ) Ωlj,j
B
(1693)
z
2 z
j+ 1
Therefore, if Ωlj,j
A
z
= Ωj,jz2 , i.e. j = lA − 12 , then κ = (j + 12 ).
j− 1
Otherwise, if Ωlj,j
A
z
= Ωj,jz2 , i.e. j = lA + 12 , then κ = − (j + 12 ).
285
Table 9: The Relationship between j, lA , lB , κ and Parity. The parity eigenvalue
A
is given by ηp = (−1)l and κ = ±(j + 21 ).
κ lA lB Parity
κ = (j + 12 ) j+ 1
2 j− 1
2 (−1)κ
κ = −(j + 12 ) j− 1
2 j+ 1
2 (−1)1−κ
Therefore, the Dirac radial equation consists of the two coupled first-order dif-
ferential equations for f (r) and g(r). On multiplying by a factor of r and
simplifying the derivatives of f (r)/r, one finds the pair of more symmetrical
equations
∂ κ
( E − V (r) − m c2 ) f (r) + c h̄ − g(r) = 0
∂r r
∂ κ
( E − V (r) + m c2 ) g(r) − c h̄ + f (r) = 0
∂r r
(1696)
The above pair of equations are the central result of this lecture.
(1697)
However, due to the identity
Ωlj,j
A
z
(θ, ϕ)† Ωlj,j
A
z
(θ, ϕ) = Ωlj,j
B
z
(θ, ϕ)† Ωlj,j
B
z
(θ, ϕ) = Aj,|jz | (θ) (1698)
the probability is independent of the azimuthal angle ϕ and the sign of jz (just
like in the non-relativistic case) and has a common angular factor of Aj,|jz | (θ).
Thus, the probability distribution factorizes into a radial and the angular factor
|f (r)|2 |g(r)|2
P (r) = + Aj,|jz | (θ) (1699)
r2 r2
286
Table 10: Relativistic Angular Distribution Functions
1 1 1
2 2 4 π
3 1 1
2 2 8 π ( 1 + 3 cos2 θ )
3
2
3
2
3
8 π sin2 θ
5 1 3
2 2 16 π ( 1 − 2 cos2 θ + 5 cos4 θ )
5
2
3
2
3
32 π sin2 θ ( 1 + 15 cos2 θ )
5
2
5
2
15
32 π sin4 θ
The angular distribution function for a closed shell is given by the sum over the
angular distribution functions. Due to the identity,
j
X 2j + 1
Aj,|jz | (θ) = (1700)
jz =−j
4π
one finds that closed shells are spherically symmetric, as is expected. The first
few angular dependent factors Aj,|jz | (θ) are given in Table(10) and the corre-
sponding non-relativistic angular factors are given in Table(11). On compar-
ing the relativistic angular dependent factors with the non-relativistic factors
|Yml (θ, ϕ)|2 , one finds that they are identical for |jz | = j. Since the relativistic
distribution is the sum of two generally different positive definite forms origi-
nally associated with the two spinors χ+ and χ− , it generally does not go to
zero for non-zero values of θ.
287
Figure 48: The relativistic (left) and non-relativistic (right) angular distribu-
tions Aj,|jz | (θ) for j = 21 and j = 32 .
288
Figure 49: The relativistic (left) and non-relativistic (right) angular distribu-
tions Aj,|jz | (θ) for j = 25 .
289
Table 11: Non-Relativistic Angular Distribution Functions
1
0 0 4 π
3
1 0 4 π cos2 θ
1 1 3
8 π sin2 θ
5
2 0 16 π ( 1 − 3 cos2 θ )2
2 1 15
8 π sin2 θ cos2 θ
2 2 15
32 π sin4 θ
The above equations will be written in dimensionless units, where the energy is
expressed in terms of the rest mass m c2 and lengths are expressed in terms of
the Compton wave length mh̄ c . A dimensionless energy is defined as the ratio
of E to the rest mass energy
E
= (1702)
m c2
For a bound state, m c2 > E > − m c2 so the value of the magnitude of
is expected to be a little less than unity. A dimensionless radial variable ρ is
introduced which governs the asymptotic large r decay of the bound state wave
function. The variable is defined by
p
2
rmc
ρ = 1 − (1703)
h̄
In terms of these dimensionless variables, the Dirac radial equations for the
hydrogen-like atom become
r
1 − γ ∂ κ
− + f + − g = 0
1 + ρ ∂ρ ρ
r
1 + γ ∂ κ
+ g − + f = 0 (1704)
1 − ρ ∂ρ ρ
where
Z e2
γ = (1705)
h̄ c
290
is a small number.
Boundary Conditions
The asymptotic ρ → ∞ form of the solution can be found from the asymp-
totic form of the equations
r
1 − ∂
− f + g ∼ 0
1 + ∂ρ
r
1 + ∂
g − f ∼ 0 (1706)
1 − ∂ρ
Hence, on combing these equations, one sees that the asymptotic form of the
equation is given by
∂2f
= f (1707)
∂ρ2
Therefore, one has
f ∼ A exp − ρ + B exp + ρ (1708)
f ∼ A ρs
g ∼ B ρs (1710)
where the exponent s is an unknown constant and then substitute the ansatz in
the above equations. This procedure yields the coupled algebraic equations
γA + (s − κ)B = 0
γB − (s + κ)A = 0 (1711)
291
Hence, it is found that the exponent s is determined as solutions of the indicial
equation which is a quadratic equation. The solutions are given by
p
s = ± κ2 − γ 2 (1712)
one must choose the positive solution for s. Normalizability near the origin
requires that 2 s > − 1. Hence, one may set
p
s = κ2 − γ 2 (1714)
This will be a good solution for κ = − 1 if Z does not exceed a critical value.
For values of Z greater than ≈ 172, the point charge can spark the vacuum
and spontaneously generate electron-positron pairs109 . The solution with the
negative value of s given by
p
s = − κ2 − γ 2 (1715)
could also possibly exist and be normalizable if γ is greater than a critical value
γc determined as
1 p
= 1 − γc2 (1716)
2
This critical value of γ is found from
√
3
γc = (1717)
2
which corresponds to Zc ∼ 118. The solutions corresponding to negative s
are, infact, un-physical and do not survive if the nucleus is considered to have
a finite spatial extent.
We shall use the Fröbenius method to find a solution. The solutions of the
radial equation shall be written in the form
f (r) = exp − ρ ρs F (ρ)
g(r) = exp − ρ ρs G(ρ) (1718)
H. Bokemeyer, P. Vincent, Y. Nakayama, and J. S. Greenberg, Phys. Rev. Lett. 40, 1443
(1978).
292
This form incorporates the appropriate boundary conditions at ρ → 0 and
ρ → ∞. The coupled radial equations are transformed to
r
1 − ∂
− ρ + γ F + ρ + s − κ − ρ G = 0
1 + ∂ρ
r
1 + ∂
ρ + γ G − ρ + s + κ − ρ F = 0
1 − ∂ρ
(1719)
The functions F (ρ) and G(ρ) can be expressed as an infinite power series in ρ
∞
X
F (ρ) = an ρn
n=0
X∞
G(ρ) = bn ρn (1720)
n=0
where the coefficients an and bn are constants which have still to be determined.
The coefficients are determined by substituting the series in the differential
equation and then equating the coefficients of the same power in ρ. Equating
the coefficient of ρn yields the set of relations
r
1 −
− an−1 + γ an + ( n + s − κ ) bn − bn−1 = 0
1 +
r
1 +
bn−1 + γ bn − ( n + s + κ ) an + an−1 = 0
1 −
(1721)
valid for any n. The above equation can be used to eliminate the coefficients
bn and yield a recursion relation between an and an−1 . The ensuing recursion
293
relation will enable us to explicitly calculate the wave functions G(ρ) and hence
F (ρ).
The behavior of the recursion relation for large values of n can be found by
noting that eqn(1723) yields
r
1 +
n an ∼ n bn (1724)
1 −
which when substituted back into the large n limit of the first relation of
eqn(1721) yields
n an ∼ 2 an−1 (1725)
Since the large ρ limit of the function is dominated by the highest powers of
ρ, it is seen that if the series does not terminate, the functions F (ρ) and G(ρ)
would be exponentially growing functions of ρ
F (ρ) ∼ exp + 2 ρ
G(ρ) ∼ exp + 2 ρ (1726)
Therefore, the set of recursion relations must terminate, since if the series does
not terminate, the large ρ behavior of the functions F (ρ) and G(ρ) would gov-
erned by the growing exponentials. Even when combined with the decaying
exponential term that appear in the relations
s
f (r) = ρ F (ρ) exp − ρ
s
g(r) = ρ G(ρ) exp − ρ (1727)
the resulting functions f (r) and g(r) would not satisfy the required boundary
conditions at ρ → ∞. We shall assume that the series truncate after the nr -th
terms. That is, it is possible to set
anr +1 = 0
bnr +1 = 0 (1728)
Thus, the components of the radial wave function may have nr nodes. Assuming
that the coefficients with indices nr + 1 vanish and using the first relation in
eqn(1721) with n = nr + 1, one obtains the condition
r
1 +
bn = − anr (1729)
1 − r
294
A second condition is given by the relation between an and bn
r r
1 + 1 +
γ − ( n + s + κ ) an + γ + ( n + s − κ ) bn = 0
1 − 1 −
(1730)
valid for any n. We shall set n = nr and then eliminate anr using the termination
condition expressed by eqn(1729). After some simplification, this leads to the
equation p
γ = ( nr + s ) 1 − 2 (1731)
This equation determines the square of the dimensionless energy eigenvalue 2 .
On squaring this equation, simplifying and taking the square root, one finds
( nr + s )
= ± p (1732)
( nr + s )2 + γ 2
where r
1 2
s = (j + ) − γ2 (1734)
2
This expression for the energy eigenvalue is independent of the sign of κ and,
therefore, it holds for both cases
1
j = (l + 1) −
2
1
j = l + (1735)
2
Hence, the energy eigenstates are predicted to be doubly degenerate (in addition
to the (2j + 1) degeneracy associated with j3 ), since states with the same j but
have different values of l0 have the same energy. If the positive-energy eigenvalue
is expanded in powers of γ, one obtains
1 γ2
E ≈ m c2 − m c2 1 + ... (1736)
2 ( nr + j + 2 )2
which agrees with the energy eigenvalues found from the non-relativistic Schrödinger
equation. However, as has been seen, the exact energy eigenvalue depends on
nr and (j + 21 ) separately, as opposed to being a function of the principle quan-
tum number n which is defined as the sum n = nr + j + 21 . Hence, the Dirac
110 C. G. Darwin, Proc. Roy. Soc. A 118, 654 (1928).
295
equation lifts the degeneracy between states with different values of the angular
momentum. The energy levels together with their quantum numbers are shown
in Table(12). The energy splitting between states with the same n and different
j values has a magnitude which is governed by the square of the fine structure
2
constant Z ( h̄e c ). That is
γ2
1
E ≈ m c2 1 −
2 ( nr + j + 12 )2
γ4
1 1 3
− − + . . .
2 ( nr + j + 12 )3 ( j + 12 ) 4 ( nr + j + 12 )
(1737)
The fine structure splittings for H-like atoms was first observed by Michelson111
and the theoretical prediction is in agreeement with the accurate measurements
of Paschen112 . The fine structure splitting is important for atoms with larger Z.
This observation has a classical interpretation which reflects the fact that for
large Z the electrons move in orbits with smaller radii and, therefore, the elec-
trons must move faster. Relativistic effects become more important for electrons
which move faster, and this occurs for atoms with larger values of Z. Although
the fine structure splitting does remove some degeneracy, the two states with
the same principle quantum number n and the same angular momentum j but
which have different values of l are still predicted to be degenerate. Thus, for
example, the 2Sj= 21 and the 2Pj= 12 states of Hydrogen are predicted to be de-
generate by the Dirac equation. It has been shown that this degeneracy is
removed by the Lamb shift, which is due to the interaction of an electron with
its own radiation field. The Lamb shift is smaller than the fine structure shifts
2
discussed above because it involves an extra factor of h̄e c .
The ground state wave function of the hydrogen atom is slightly singular
at the origin. This can be seen by noting that it corresponds to nr = 0 and
κ = −1. Since the dimensionless energy is given by the expression
p
= 1 − γ2 (1738)
one finds that the dimensionless radial distance ρ is simply given by
p
2
rmc
ρ = 1 −
h̄
mc
= γ r
h̄
Z e2 m
= r (1739)
h̄2
111 A. A. Michelson, Phil. Mag. 31, 338 (1891).
112 F. Paschen, Ann. Phys. 50, 901 (1916).
296
Table 12: The Equivalence between Relativistic and Spectroscopic Quantum
Numbers.
n = nr + |κ| nr κ = ±(j + 12 ) nLj Degenerate E
m c2
Partner
p
1 0 -1 1S 12 1 − γ2
r
γ 2
2 1 -1 2S 12 2P 12 1− √
2+2 1−γ 2
2 1 +1 2P 12 2S 12 --
q
1
2 0 -2 2P 32 1− 4 γ2
r
γ 2
3 2 -1 3S 12 3P 12 1− √
5+4 1−γ 2
3 2 +1 3P 12 3S 12 --
r
γ 2
3 1 -2 3P 32 3D 32 1− √
5+2 4−γ 2
3 1 +2 3D 32 3P 32 --
q
1
3 0 -3 3D 52 1− 9 γ2
297
where the characteristic length scale is just the non-relativistic Bohr radius
divided by Z. The wave functions are written as
f (ρ) = exp − ρ ρs F (ρ)
g(ρ) = exp − ρ ρs G(ρ) (1740)
γ a0 + ( s + 1 ) b0 = 0
γ b0 − ( s − 1 ) a0 = 0 (1741)
and p
b0 s + κ 1 − γ2 − 1
= = (1743)
a0 γ γ
This shows that the lower component is smaller than the upper constant by
approximately γ, which has the magnitude of vc where v is the velocity in Bohr’s
theory. The ratio of b0 to a0 determines the radial functions as
f (r) s−1
= a0 ρ exp − ρ
r
p
1 − γ 2 − 1 s−1
g(r)
= a0 ρ exp − ρ (1744)
r γ
Since Y00 (θ, ϕ) = √1 , the angular spherical harmonics for the upper com-
4 π
ponents are just
1
ΩA (θ, ϕ) = √ χσ (1745)
4π
and the lower components are given by
r.σ 1
ΩB (θ, ϕ) = − √ χσ
r 4π
1 cos θ sin θ exp[−iϕ]
= − √ χσ (1746)
4π sin θ exp[+iϕ] − cos θ
Thus, apart from an over all normalization factor, the four-component spinor
Dirac wave function ψ is given by
√ 2 χσ
N
ρ 1−γ −1 exp − ρ
ψ = √ r . σ (1747)
4π −i r χσ
298
0.8
1S1/2
0.6
0.4
f(r)
0.2
g(r) x 100
0
0 0.5 1 1.5 2 2.5 3
-0.2
-0.4
mcγr/h
Figure 50: The large f (r) and small component g(r) radial wave functions for
the 1S 12 ground state of Hydrogen.
Hence, it is seen that as ρ approaches the origin, at first the wave function is
slowly varying since
√ 2
γ2
γ2
ρ 1−γ −1 ∼ exp − ln ρ ∼ 1 − ln ρ (1748)
2 2
the wave function exhibits a slight singularity. This length scale is much smaller
that the nuclear radius so, due to the spatial distribution of the nuclear charge,
the singularity is largely irrelevant. This singularity is not present in the non-
relativistic limit, since in this limit one assumes that the inequality | V (r) |
m c2 always holds, although this assumption is invalid for r ∼ 0. Therefore, one
concludes that the relativistic theory differs from the non-relativistic theory at
small distances, which could have been discerned from the use of the Heisenberg
uncertainty principle.
299
0.4 0.2
2P1/2
2S1/2
0.2 0
0 2 4 6 8 10
0 f(r)
0 2 4 6 8 10 -0.2
f(r) g(r) x 100
-0.2
g(r) x 100
-0.4
-0.4
-0.6
-0.6 mcγr/h
mcγr/h
Figure 51: The radial wave functions for the 2S 12 and 2P 12 states of Hydrogen.
300
involving the radial wave functions. The integral is evaluated with the aid of
the identity
Z ∞
dρ ρa+b exp − 2 ρ = 2−(a+b+1) Γ(a + b + 1) (1755)
0
The coefficients cn and dn in the above expansion of the radial functions differ
Table 13: Parameters specifying the Radial Functions for the Hydrogen atom.
E γ m c
State s m c2 a h̄ N
p p s+ 1
κ = −1 1 − γ2 1 − γ2 1 √1 √2 2
a 2Γ(2s+1)
1S 12
q √ r
2 E
p 1 + 1 − γ2 ( −1) s+ 1
E 1 m c2 √1 √2 2
κ = −1 1 − γ2 2 2 m c2 2 ( 2 E2 ) a
m c Γ(2s+1)
2S 32
q √ r
2 E
p 1 + 1 − γ2 ( +1) s+ 1
E 1 m c2 √1 √2 2
κ=1 1 − γ2 2 2 m c2 2 ( 2 E2 ) a
m c Γ(2s+1)
2P 12
s+ 1
q
γ2
p
κ = −2 4 − γ2 1 − 2 √1 √2 2
4 a 2Γ(2s+1)
2P 32
from the coefficients an and bn that occur in the Frobenius expansion, since the
values of the ratio cnr /dnr has been chosen to simplify in the limit of large n. In
particular at the value of nr (at which the series terminates), the ratio is chosen
to satisfy
cnr
= 1 (1756)
d nr
instead of the condition
s
anr 1 + ( mEc2 )
= − (1757)
b nr 1 − ( mEc2 )
301
The relative negative sign and the square root factors in the coefficients have
been absorbed into the expressions for the upper and lower components f (r) and
g(r). The square root factors are responsible for converting the upper and lower
components, respectively, into the large and small components for positive E,
and vice versa for negative E. The expansion coefficients are given in Table(14).
Since the ratio of the magnitudes of the polynomial factors is generally of the
Table 14: Coefficients for the Polynomial in the Hydrogen atom Radial Wave-
functions.
State c0 c1 d0 d1
κ = −1 1 0 1 0
1S 12
2 E
2 E
E ( )+1 E ( )+1
m c2 m c2
κ = −1 2 m c2 2 s + 1 2 m c2 + 1 2 s + 1
2S 32
2 E
2 E
E ( )−1 E ( )−1
m c2 m c2
κ=1 2 m c2 − 1 2 s + 1 2 m c2 2 s + 1
2P 12
κ = −2 1 0 1 0
2P 32
order of unity, the ratio of the magnitudes of the small to large components is
found to be of the order of γ.
302
decisive role in compelling Pauli to reluctantly accept Dirac’s theory.
where φA and φB are, respectively, the upper and lower two component spinors
of the four-component Dirac spinor ψ. The energy eigenvalues of these equations
are sought, so to this end the explicit time-dependence of the energy eigenstates
will be separated out via
A
φ i
ψ = exp − Et (1759)
φB h̄
Also the non-relativistic energy will be defined as the energy referenced with
respect to the rest-mass energy
E = m c2 + (1760)
The pair of equations will be expanded in powers of ( mp c )2 and only the first-
order relativistic corrections will be retained. One can express φB as
− i h̄ c ( σ . ∇ ) φA
φB =
− V + 2 m c2
−1
1 − V
= 1 + ( σ . p̂ ) φA
2mc 2 m c2
1 − V
≈ 1 − + ... ( σ . p̂ ) φA (1762)
2mc 2 m c2
303
probability density associated with the four-component Dirac spinor depends
on both φA and φB ,
P (r) = φA† φA + φB † φB (1763)
p̂2
ψS = I + + ... φA (1766)
8 m2 c2
p̂2
A
φ ≈ I − + ... ψS (1767)
8 m2 c2
p̂2
B 1 − V
φ ≈ 1 − ( σ . p̂ ) I − ψS
2mc 2 m c2 8 m2 c2
p̂2
1 − V
≈ ( σ . p̂ ) I − − ( σ . p̂ ) ψS
2mc 8 m2 c2 2 m c2
(1768)
On substituting φB and ψS into the equation for φA , one finds the (two-
component) energy eigenvalue equation
p̂2
− V I − ψS
8 m2 c2
( σ . p̂ ) p̂2
− V
= ( σ . p̂ ) I − − ( σ . p̂ ) ψS
2m 8 m2 c2 2 m c2
(1769)
304
or
p̂2 p̂2
− V − + V ψS
8 m2 c2 8 m2 c2
2
p̂2
p̂ − V
= I − − ( σ . p̂ ) ( σ . p̂ ) ψS
2m 8 m2 c2 4 m2 c2
(1770)
The above energy eigenvalue equation can be expressed as
p̂2 p̂2 p̂2
− V − + + V ψS
2m 8 m2 c2 8 m2 c2
p̂4
V
= − + ( σ . p̂ ) ( σ . p̂ ) ψS (1771)
16 m3 c2 4 m2 c2
The term proportional to the product of the energy eigenvalue and the kinetic
energy can be re-written as
p̂2 p̂2
ψS = ψS
8 m2 c2 8 m2 c2
p̂2 p̂2
≈ ( V + ) ψS (1772)
8 m2 c2 2m
to the required order of approximation. On substituting the above expression
into the energy eigenvalue equation (1771), one finds
p̂2 p̂4 p̂2 p̂2
− V − + + V + V ψS
2m 8 m3 c2 8 m2 c2 8 m2 c2
V
= ( σ . p̂ ) ( σ . p̂ ) ψS (1773)
4 m2 c2
The above equation will be interpreted as the non-relativistic energy eigenvalue
equation for the two-component wave function ψS , which contains relativistic
corrections of order ( vc )2 . The energy eigenvalue equation (1773) will be written
in the form
2
p̂4
p̂
ψS = + V ψS − ψS
2m 8 m3 c2
2
p̂ V + V p̂2
V
− ψ S + ( σ . p̂ ) ( σ . p̂ ) ψS
8 m2 c2 4 m2 c2
(1774)
where the relativistic corrections are symmetric in p2 and V . This represents the
energy eigenvalue equation for a two-component wave function ψS , similar to
the Schrödinger wave function, but the above equation does include relativistic
corrections to the Hamiltonian. The first correction term is
p̂4
ĤKin = − (1775)
8 m3 c2
305
which is recognized as the relativistic kinematic energy correction, that origi-
nates from the expansion of the kinetic energy
p
= m2 c2 + p2 c2 − m c2
p2 p4
≈ − + ... (1776)
2m 8 m3 c2
The remaining two correction terms
2
p̂ V + V p̂2
V
( σ . p̂ ) ( σ . p̂ ) − (1777)
4 m2 c2 8 m2 c2
will be interpreted as the sum of the spin-orbit interaction and the Darwin term.
It should be noted that the sum of these two terms would identically cancel in
a purely classical theory. This cancellation can be shown to occur since, in the
classical limit, V and p commute, and then the Pauli-identity can be used to
show that the resulting pairs of terms cancel.
The factor
2 ( σ . p̂ ) V ( σ . p̂ ) − p̂2 V + V p̂2 (1778)
can be evaluated as
2 2
2 p̂ . V p̂ − p̂ V + V p̂ + 2iσ. p̂ ∧ V p̂ (1779)
The first two terms can be combined to form a double commutator, yielding
− [ p̂ , [ p̂ , V ] ] + 2 i σ . p̂ ∧ V p̂ (1780)
or
+ h̄2 ∇2 V + 2 i σ . p̂ ∧ V p̂ (1781)
since
p̂ ∧ p̂ ≡ 0 (1783)
Using these substitutions, the remaining interactions can be expressed as the
sum of the spin-orbit interaction and the Darwin interaction
h̄2
h̄
ĤSO + ĤDarwin = + 2 2
σ . ∇ V ∧ p̂ + ∇2 V (1784)
4m c 8 m2 c2
306
The first term is the spin-orbit interaction term, and the second term is the
Darwin term. For central potentials, the Darwin term is only important for
electrons with l = 0. The evaluation and the physical interpretation of the
energy shifts due to the three fine-structure interactions will be discussed sepa-
rately.
307
which can be identified with a factor of ( vc )2 as can be inferred from an analysis
based on the Bohr model of the atom. One sees that the relativistic corrections
become more important for atoms with larger Z, since the correction varies as
Z 4 . This occurs because for larger Z the electrons are drawn closer to the nu-
cleus and, hence have higher kinetic energies, so the electron’s velocities draw
closer to the velocity of light.
308
However, the electron is bound to the nucleus and is orbiting with angular
momentum L. Therefore, one has to consider the corrections to the precession
rate (and the interaction) caused by the acceleration of the electron’s rest frame.
Thomas Precession
In the electron’s rest frame, the gyromagnetic ratio due to the orbital mag-
netic field B 0 (caused by the charged nucleus) is given by gs = 2. This gyro-
magnetic ratio yields a spin precession rate in the electron’s rest frame of
e
ωrest = gS B 0 (1796)
2mc
The spin precession rate observed in the lab frame will be calculated later. The
rate of precession as observed in the electron’s rest frame has to be corrected
by taking into account the motion of the electron. The correction is due to
the non-additivity of velocities in successive Lorentz transformations. First,
the transformation properties of Dirac spinors under infinitesimal rotations and
boosts will be re-examined. Secondly, infinitesimal transformations will be suc-
cessively applied to describe the particle’s instantaneous rest frame and the
Thomas precession.
309
where
i
σ i,j = [ γ (i) , γ (j) ]
2 (k)
X
i,j,k σ 0
= ξ (1800)
0 σ (k)
k
Hence, for a passive rotation through an infinitesimal angle δϕ, the four-component
Dirac spinor is rotated by
(k)
δϕ X i,j,k σ 0
R̂(δϕ) = Iˆ + i ξ + ... (1801)
2 0 σ (k)
k
which can be expressed in terms of the projection of the (block diagonal) spin
operator Ŝ = h̄2 σ̂ on the axis of rotation ê as
i δϕ
R̂(δϕ) = Iˆ + ( ê . σ̂ ) + . . .
2
i δϕ
= Iˆ + ( ê . Ŝ ) + . . . (1802)
h̄
If the primed frame of reference has a velocity v along the k-axis relative
to the un-primed frame, the infinitesimal Lorentz transform has the non-zero
elements
v
0,k = − k,0 = − (1803)
c
A Lorentz boost along the k-axis corresponds to a rotation in the 0 - k plane
through an “angle” χ
χ 0,k
R̂(χ) = exp + i σ (1804)
2
310
The infinitesimal transformation is guaranteed to be consistent with the source
free solution of the Dirac equation. For example, if the above transformation is
applied to the solution of the Dirac equation describing a positive-energy parti-
cle at rest, the transformed solution describes a particle moving with momentum
p = − m v when viewed from the moving frame of reference.
ωRest
a q ωΤ E
Figure 52: A cartoon depicting a rotating charged spin one-half particle, along
with the precession of the spin due to the external field in the particle’s rest
frame and the Thomas precession.
v 0 = v + a δt (1810)
On performing a second Lorentz transform with the boost a δt, one finds the
rotation
1
R̂2 = Iˆ − α . a δt + . . . (1811)
2c
The combined Lorentz transform is given by
1 1
R̂ = R̂2 R̂1 = Iˆ − α . a δt Iˆ − α.v + ...
2c 2c
1 1
= Iˆ − α . ( v + a δt ) + ( α . a ) ( α . v ) δt + .(1812)
..
2c 4 c2
The Pauli identity can be used to evaluate the last term
(α.a)(α.v) = ( σ̂ . a ) ( σ̂ . v )
= a . v Iˆ + i σ̂ . ( a ∧ v ) (1813)
311
where, since the product of the two α’s yields a two by two block diagonal form
which involves the four by four matrices Iˆ and σ̂. Hence, the right-hand side acts
equally on both the upper and lower two-component spinors. Furthermore, since
the orbit is circular, the acceleration is perpendicular to the velocity, therefore
a.v = 0 (1814)
On combing the two precession frequencies, one finds that in the lab frame
the spin’s precession rate is given by
ωLab = ωrest − ωT
e
= ( gS − 1 ) B 0 (1819)
2mc
It is clear that the moving spin experiences an effective interaction which is
reduced by the factor
gS − 1
(1820)
gS
when compared to the interaction in the electron’s rest frame. Hence, the gyro-
magnetic ratio that enters the spin-orbit coupling should not be gS but should
312
be given by ( gS − 1 ).
In the lab frame, the interaction between the moving electron’s spin S mag-
netic moment and its field is inferred to be
Lab = − q ( g − 1 ) B 0 . S
ĤInt (1821)
S
2mc
where gS is the gyromagnetic ratio. Since the magnetic induction field is given
by
0 1 ∂φ
B = − L (1822)
m c r ∂r
where the electrostatic potential is given by
q0
φ(r) = (1823)
r
the spin-orbit interaction can be expressed as
q q0
ĤSO = − ( gS − 1 ) L.S (1824)
2mc m c r3
Hence, the spin-orbit interaction is found to be given by
Z e2
ĤSO = ( gS − 1 ) L . S (1825)
2 m2 c2 r3
The spin-orbit coupling is a relativistic coupling which, apart from the Thomas
precession factor, indicates that the electron’s spin interacts with a magnetic
field in its rest frame via the gyromagnetic ratio of 2. The magnitude of the
interaction agrees precisely with the interaction found from the perturbative
treatment of the Dirac equation.
313
for l 6= 0. So the first-order energy-shift due to the spin-orbit coupling can be
expressed as
4 1
Z e2 ±(l + ) − 12
∆ESO = m c2 2
1 (1829)
h̄ c 4 n3 l ( l + 2 )(l + 1)
Therefore, the spin-orbit interaction lifts the degeneracy between states with
different j = l ± 12 values. For l = 0, the numerator vanishes since the total
angular momentum can only take the value
1
j = + (1830)
2
The energy shift produced by the spin-orbit coupling is about a factor of the
square of the fine structure constant
2 2
e2
1
∼ ∼ 10−4 (1831)
h̄ c 137
π Z e2 h̄2 3
ĤDarwin = δ (r) (1833)
2 m2 c2
which produces the first-order shift
π Z e2 h̄2 †
∆EDarwin = ψS (0) ψS (0) (1834)
2 m2 c2
314
Hence, the shift only occurs for electrons with l = 0. Furthermore, since the
probability density for finding the electron at the origin is given by
3
1 Z
ψS† (0) ψS (0) = δl,0 (1835)
π na
which shifts the energies of s states upwards. The Darwin term reflects the fact
that the relativistic corrections are important for small r since the inequality
Z e2
m c2 (1837)
r
required for the non-relativistic treatment to be reasonable is violated in this
region.
E S P
Kinematic 2P3/2
Spin Orbit
2S1/2 2P1/2
Kinematic
Darwin
n=2
Figure 53: The Grotarian energy level diagram for the n = 2 shell of hydrogen
(blue). The diagram shows the magnitude and sign of the various relativistic
corrections. It should be noted that states with the same j are degenerate.
When the various relativistic corrections are combined, for l = 0, the Darwin
term exactly compensates for the absence of the spin-orbit interaction. There-
fore, the energy shifts combine to yield one formula in which l drops out. This
implies that the energy levels only depend on the principle quantum number n
315
and the total angular momentum j. States with different orbital angular mo-
menta are degenerate, even though the individual interactions appear to raise
the degeneracy. The relativistic corrections inherent in Dirac’s theory of hydro-
gen yields energy shifts and line-splittings which are described as fine structure.
The energy levels are described by
1 Z 2 α2 1 Z 4 α4
1 3
E ≈ m c2 1 − − − + . . .
2 n2 2 n3 ( j + 12 ) 4n
(1838)
where
e2
α = (1839)
h̄ c
is the fine structure constant. Generally, states with larger j values have higher
energies. The fine structure splittings decrease with increasing n like n−3 , but
increase with increasing Z like Z 4 . The splitting of the lower energy levels are
largest, for example
m c2 α 4
1 1
E2P 3 − E2P 1 = − − ≈ 4.533 × 10−5 eV (1840)
2 2 16 2 1
This splitting corresponds to a frequency of 10.96 GHz. The energy levels are
predicted to be doubly degenerate (in addition to the degeneracy associated
with j3 ), the degeneracy is just the number of states with different l values that
yield the same value of j. Since j is found by combining l with the electronic
spin s = 21 , there are two possible l values for each energy level which are given
by the solutions of either
1
j = l + (1841)
2
or
1
j = l − (1842)
2
The higher-order relativistic corrections do not alter the conclusion that the
states labeled by (n, j) are degenerate, as the energy levels found from the exact
solution of the Dirac equation only depend on n and j. For j = 12 the energy
levels, although predicted to be degenerate by Dirac’s theory, are experimentally
observed as being non-degenerate. The first experiments that revealed this split-
ting were performed by Lamb and Retherford114 . These scientists found that
the 2S 12 was shifted by about 1057 MHz to higher energies relative to the 2P 12 .
The relative shift of the nS 12 level of hydrogen with respect to the nP 12 level is
known as the Lamb shift.
114 W. E. Lamb Jr. and R. E. Retherford, Phys. Rev. 72, 241 (1947).
316
E S P D
3D5/2
Kinematic
Kinematic Spin Orbit
3P3/2 3D3/2
Spin Orbit
3S1/2
3P1/2
Kinematic
Darwin
n=3
Figure 54: The Grotarian energy level diagram for the n = 3 shell of hydrogen
(blue). The diagram shows the magnitude and sign of the various relativistic
corrections. It should be noted that states with the same j are degenerate.
I
-
e
e-
Oven EM
H1 Cavity H1
317
in an oven. The thermal beam of hydrogen atoms was then cross-bombarded
with electrons, which excited some of the hydrogen atoms out of the ground
state. Since the electron-atom scattering doesn’t obey the radiation selection
rules, a finite population of atoms (about 1 in 108 ) were excited to the long-lived
2S 12 state. Subsequently, the other excited electronic states rapidly decayed to
the ground state by the emission of radiation. The beam of hydrogen atoms was
then passed through a tuneable (microwave) electromagnetic resonator, which
could cause the hydrogen atoms in the meta-stable level to make transitions
to selected nearby energy levels. Again, any non-2S excited state of hydrogen
produced by the action of the resonator rapidly decayed to the ground state.
The resulting beam of hydrogen atoms was incident on a Tungsten plate, and
the collision could result in electron emission if the atoms were in an excited
state, but no emission would take place if the hydrogen atom was in the ground
state. Therefore, the current due to the emitted electrons was proportional to
the number of meta-stable hydrogen atoms that survived the passage through
the resonator. Hence, analysis of the experiment yielded the number of transi-
tions undergone in the electromagnetic resonator.
Figure 56: The dependence of the current emitted from the tungsten plate on
the applied magnetic field. The resonance frequency was set to 9487 Megacycles.
[W. E. Lamb Jr. and R. C. Retherford, Phys. Rev. 72, 241 (1947).]
In the resonator, an applied magnetic field Zeeman split the excited levels of
hydrogen and, when the oscillating field was on-resonance with the splitting of
the energy levels, the hydrogen atom made transitions out from the meta-stable
2S 12 state. At resonance, the frequency of the oscillating electromagnetic field
is equal to the energy splitting. Therefore, for fixed frequency, knowledge of the
resonance magnetic field allowed the splitting of the energy levels to be accu-
318
rately determined. The field dependence of the resonance frequency indicated
Figure 57: The observed dependence of the resonance frequencies on the applied
magnetic field. The solid lines are the predictions of the Dirac theory and the
dashed lines are the result of Dirac’s theory if the energy of the 2S state is
simply shifted. [W. E. Lamb Jr. and R. C. Retherford, Phys. Rev. 72, 241
(1947).]
that at zero field the degeneracy between the 2S 21 and 2P 12 states were lifted,
with the 2S 21 state having the higher energy.
319
We shall examine the case of an attractive central square well potential V (r)
which is defined by
− V0 for r < a
V (r) = (1844)
0 for r > a
In the region r < a where the potential is finite, the Dirac radial equation
0.5
0
V(r)/V0
-0.5
-1
-1.5
0 0.5 1 1.5 2
r/a
becomes
2 ∂ κ
( E + V0 − m c ) f (r) + c h̄ − g(r) = 0
∂r r
∂ κ
( E + V0 + m c2 ) g(r) − c h̄ + f (r) = 0
∂r r
(1845)
320
By using a similar procedure, starting from the second equation, one can find
the analogous equation for g(r)
2
2 2 ∂ κ(κ − 1) 2 2 4
c h̄ − g(r) = − ( E + V0 ) − m c g(r)
∂r2 r2
(1848)
Real Momenta
321
of half-integer order. The spherical Bessel functions and spherical Neumann
functions of order n are defined in terms of the Bessel functions via
r
π
jn (ρ) = J 1 (ρ)
2 ρ n+ 2
r
π
ηn (ρ) = N 1 (ρ) (1855)
2 ρ n+ 2
Therefore, the general solutions of each of the radial equations can be expressed
as
f (r)
= A0 j|κ+ 12 |− 12 (k0 r) + A1 η|κ+ 12 |− 12 (k0 r) (1856)
r
and
g(r)
= B0 j|κ− 12 |− 12 (k0 r) + B1 η|κ− 12 |− 12 (k0 r) (1857)
r
However, since the functions f (r) and g(r) in the upper and lower components
are related by the differential equations
( E + V0 + m c2 ) g(r)
∂ 1 + κ f (r)
+ = (1858)
∂ρ ρ r c h̄ k0 r
and
( E + V0 − m c2 )
∂ 1 − κ g(r) f (r)
+ = − (1859)
∂ρ ρ r c h̄ k0 r
the two sets of coefficients (A0 , A1 ) and (B0 , B1 ) must also be related. The
explicit relations can be found by using the recurrence relations for the spherical
Bessel functions jn (ρ)
∂ n+1
ρ jn (ρ) = ρn+1 jn−1 (ρ) (1860)
∂ρ
and
∂
ρ−n jn (ρ) = − ρ−n jn+1 (ρ) (1861)
∂ρ
The spherical Neumann functions ηn (ρ) satisfy identical recurrence relations.
This yields the relations
E + V 0 + m c2
A0 = sign κ B0
c h̄ k0
E + V 0 + m c2
A1 = sign κ B1 (1862)
c h̄ k0
Hence, for positive-energy solutions. the upper components are the large com-
ponents and the lower components are the small components. In the inner
region, one must set A1 = B1 = 0, since the wave function are required to be
322
normalizable near the origin and the spherical Neumann functions ηn (ρ) diverge
as ρ−(n+1) as ρ → 0.
Imaginary Momenta
Bound States
323
For asymptotically large ρ, these functions are complex conjugates and represent
out-going or incoming spherical waves
1 1 π
lim h±n (ρ) → exp ± i ρ − ( n + ) (1871)
ρ→∞ ρ 2 2
The factor of ρ−1 reflects the fact that the intensity of an outgoing wave-packet
decreases in proportion to ρ−2 in order to conserve energy and probability. From
the asymptotic variation, it is seen that the spherical Hankel functions h± n (iρ)
with imaginary arguments, respectively, represent exponentially attenuating or
growing spherical waves. In the exterior region, the solutions are represented
by
f (r)
= C00 h+ |κ+ 12 |− 12
(iκ1 r) + C10 h− |κ+ 12 |− 12
(iκ1 r) (1872)
r
and
g(r)
= D00 h+ |κ− 12 |− 12
(iκ1 r) + D10 h− |κ− 12 |− 12
(iκ1 r) (1873)
r
The coefficients of the upper and lower components are related via
E + m c2
C00 = − D00
c h̄ κ1
E + m c2
0
C1 = D10 (1874)
c h̄ κ1
as can be seen by substituting the asymptotic form of the Hankel functions given
by eqn(1871) in the asymptotic form of the differential equations relating f (r)
and g(r) with V0 = 0. If this wave function is to be normalizable at ρ → ∞,
one must set C10 = D10 = 0.
The solutions for the wave functions have been found in the inner and outer
regions of the potential. The solution must also hold at r = a. This is achieved
by demanding that the upper and lower components of the wave function are
continuous at r = a. These conditions are demanded due to charge conservation
∂µ j µ = 0, since the current j µ only depends on the components of ψ and does
not (explicitly) depend on their derivatives.
Since the wave function at the origin must be normalizable, and since the
wave function must be exponentially decaying, when r → ∞, the matching
condition for the upper component becomes
324
By eliminating the amplitudes from the two matching conditions by using
eqn(1874), one can arrive at the equation
h+ (iκ1 a)
E + V 0 + m c2 j|κ+ 21 |− 12 (k0 a) E + m c2
|κ+ 12 |− 12
sign(κ) = −
c h̄ k0 j|κ− 12 |− 12 (k0 a) c h̄ κ1 h+
|κ− 1 |− 1
(iκ1 a)
2 2
(1877)
In the above expression, the quantities k0 and κ1 are defined by
h̄2 c2 k02 = ( E + V0 )2 − m2 c4 (1878)
and
h̄2 c2 κ21 = m2 c4 − E 2 (1879)
These equations determine the allowed values for the energy. The above set of
equations have to be solved numerically to find the energy eigenvalues. We note
that for the Dirac particle, the spin effectively results in the formation of a cen-
trifugal barrier (either for the upper or the lower component) even for electrons
in s states. As a result, the potential V0 must exceed a critical strength if it is
to yield a bound state.
The MIT bag model115 is a simple purely phenomenological model for the
structure of strongly interacting particles (hadrons). The model is based on
the spherically symmetric potential of radius a, but it will be assumed that the
quark mass can have one or the other of two values. The quark is assumed to
have a small mass (approximately zero) if it is located within a sphere of radius
a, and the mass is assumed to be very large (or infinite) if r > a. To be sure,
the quark mass is assumed to be a function of r such that
m = 0 if r < a
m → ∞ if r > a (1880)
115 A. Chodos, R. L. Jaffe, K. Johnson, C. B. Thorn, and V. F. Weisskopf, Phys. Rev. D 9,
3471, (1974).
325
It is the infinite mass of the quark for r > a that results in the confinement of
the quark to within the hadron. That is, in the exterior region, the infinite rest
mass energy exceeds the bound state energy so the exterior region is classically
forbidden, therefore, the particle is confined to the interior.
Inside the hadron, where both the potential energy and the mass m are zero,
the kinetic energy parameter k0 can be expressed entirely in terms of the energy
via E = h̄ c k0 since the potential is assumed to be zero. Therefore, the radial
components of the Dirac wave function can be expressed as
f (r)
= A0 j|κ+ 12 |− 12 (k0 r)
r
g(r)
= sign(κ) A0 j|κ− 12 |− 12 (k0 r) (1881)
r
where the amplitudes of the upper and lower components are the same, since
the potential and mass are zero for r < a.
Outside the hadron, where r > a, the energy E is assumed to be much less
than the rest mass energy, m c2 E, therefore, the momentum parameter is
imaginary and one can set h̄ c κ1 ≈ m c2 . In the exterior region, the radial
functions can be expressed as
f (r)
= C00 h+
|κ+ 1 |− 1
(iκ1 r)
r 2 2
g(r)
= − C00 h+
|κ− 1 |− 1
(iκ1 r) (1882)
r 2 2
Due to the asymptotic properties of the spherical Hankel functions, their ratio
is unity for large κ1 . This leads to the energies of the quarks being governed by
the simplified matching condition
where
E = c h̄ k0 (1885)
The above equation governs the ground state and excited state energies of the
individual quarks inside the hadron. Since the spherical Bessel functions oscil-
late in sign, the above equations will result in a set of solutions for k0 with fixed
326
κ. From the structure of the equations, it is seen that the solutions k0 will only
depend on the integer number κ and the value of a. Since another boundary
condition should also be imposed at the bag’s surface, only states with angular
momentum j = 21 should be retained. This extra condition restricts the interest
to states with κ = − 1.
Since
sin ρ
j0 (ρ) = (1887)
ρ
and
sin ρ − ρ cos ρ
j1 (ρ) = (1888)
ρ2
the energy eigenvalues are determined by the solutions of
1
ρ = (1889)
1 + cot ρ
which has an infinite number of solutions which, asymptotically, are spaced by
π. The smallest solution corresponds to k0 a = 2.04. Hence, the energy of the
1
P(ρ)/P(0)
0.5
0
0 1 2 3 4
ρ
Figure 59: The radial dependence of the quark-distribution in the ground state
of the MIT bag.
327
Table 15: The lowest single-particle energies (in units of Eκ,nr a/c h̄) of the MIT
Bag Model.
nr κ = −1 κ = +1 κ = −2 κ = +2
single particle levels. This could allow one to calculate the excitation energies
required to change the hadron’s internal structure.
328
Table 16: The Observed Energy Levels for the charmonium system (cc) in units
of MeV/c2 .
1 3 3 3 3
S0 S1 P0 P1 P2
which yields a ratio of 1.36. This ratio is far too small for the triplet of π mesons
since Mπ ∼ 139 MeV/c2 , and Mn ∼ 938 MeV/c2 . Although it is in adequate
for the pseudo-scalar mesons, the MIT bag model is more appropriate for the ω
vector meson which is composed of √12 (uu + dd) and has a mass of Mω ∼ 783
MeV/c2 , or the ρ vector meson √1 2 (uu + dd) with a mass Mρ ∼ 776 MeV/c2 .
Hence, at best, the MIT Bag model produces mixed results. The MIT Bag
model is also quite unappealing, since the basic assumptions of the bag model
do not follow from Quantum Chromodynamics, and the model is neither re-
normalizable nor is it Lorentz invariant.
329
Table 17: The Observed Energy Levels for the Upsilon system (bb) in units of
MeV/c2 .
1 3 3 3
S0 P0 P1 P2
that at small separations the quarks only interact weakly. This property is
called asymptotic freedom. It was the realization by ’t Hooft122 , Gross and
Wilczek123 and Politzer124 that non-Abelian gauge theories possessed the prop-
erties of asymptotic freedom that led to the acceptance of the theory of Quantum
Chromodynamics. The screening of the color force between the quarks at large
distances (due to virtual quark/anti-quark pairs) is more than compensated by
an anti-screening due to virtual gluon pairs. However, at small distances the
color force vanishes.
Exercise:
Show that the positive energy eigenvalues of the Dirac equation with the
mass m(r) given by
m(r) c = m0 c − i ω α . r (1894)
are determined as √
En,j,l = m0 c2 tA + 1 (1895)
122 G. t’ Hooft, unpublished (1972).
123 D. J. Gross and F. A. Wilczek, Phys. Rev. Lett. 30, 1343 (1973).
124 H. D. Politzer, Phys. Rev. Lett. 30, 1346 (1973).
125 D. Ito, K. Mori and E. Carriere, Nuovo Cimento, 51 A, 1119, (1967).
330
where the dimensionless parameter t corresponding to the string tension is given
by
h̄ ω
t = (1896)
m0 c2
and A is given in terms of the quantum numbers as
1
A = 2 (n − j) + 1 if j = l +
2
1
A = 2 (n + j) + 3 if j = l − (1897)
2
Hence, find the best fit to the excitation spectra of quarkonium.
The Dirac wave function ψ(r) can be expressed in terms of two two-component
spinors A
φ (r)
ψ(r) = (1898)
φB (r)
One only need specify the upper component φA (r), since once φA (r) has been
specified φB (r) is completely determined. For example, for the in and out
asymptotes, the Dirac equation reduces to
E p − m c2
A
− c p̂ . σ φ (r)
= 0 (1899)
− c p̂ . σ E p + m c2 φB (r)
331
In the scattering experiment, a plane-wave with momentum p parallel to the
ê3 -axis falls incident on the target. The in-asymptote can be described by a
state which is in a superposition of eigenstates of Ŝ (3) given by
in χ± pr
ψ± (r) = NEp c p
± Ep + m c2 χ± exp i cos θ (1901)
h̄
From the Rayleigh expansion, one observes that the in-asymptotes are not eigen-
states of (Ĵ)2 = (L̂ + Ŝ)2 since they are formed of linear superpositions of many
2 2
states with differing eigenvalues of L̂ but have a fixed eigenvalue of Ŝ . How-
ever, the in-asymptote are eigenstates of Jˆ(3) = L̂(3) + Ŝ (3) with eigenvalues
± h̄2 .
Ψin
(θ,ϕ)
p
Ψout
Figure 60: The geometry of the asymptotic final state of Mott scattering. At
large r, the beam separated into an unscattered beam ψin and a spherical out-
going wave ψout .
active in the vicinity of the target. In spherical polar coordinates, the orbital
332
angular momentum raising and lowering operators are given by
∂ ∂
L̂± = ± h̄ exp[±iϕ] ± i cot θ (1903)
∂θ ∂ϕ
In light of the comment about the upper two-component spinor, one sees that
the scattered wave is determined by
f (θ)
(1908)
g(θ) exp[+iϕ]
if the initial beam has a negative helicity. The quantities f (θ) and g(θ) are gen-
eralized scattering amplitudes that have the dimensions of length, and depend
on θ but do not depend on ϕ as both the in and out asymptotes are eigenstates
of Jˆ(3) with eigenvalues ± h̄2 . A partial wave analysis can be performed on the
Dirac equation to yield expressions for the scattering amplitudes f (θ) and g(θ)
in terms of phase shifts. A detailed knowledge of the scattering amplitudes is
not required for the following analysis.
333
If the in-asymptote has the spin quantized along the direction given by
(sin θs cos ϕs , sin θs sin ϕs , cos θs ), the upper component of the Dirac wave spinor
is determined by the two-component spinor
f (θ) cos θ2s exp[−i ϕ2s ] − g(θ) sin θ2s exp[+i ϕ2s ] exp[−iϕ]
A 1 pr
φ (r) = exp i
g(θ) cos θ2s exp[−i ϕ2s ] exp[+iϕ] + f (θ) sin θ2s exp[+i ϕ2s ] r h̄
(1911)
The probability for scattering is proportional to
2
θs ϕs θs ϕs
I(θ, ϕ) ∝ f (θ) cos exp[−i ] − g(θ) sin exp[+i ] exp[−iϕ]
2 2 2 2
2
θs ϕs θs ϕs
+ g(θ) cos
exp[−i ] exp[+iϕ] + f (θ) sin exp[+i ]
2 2 2 2
2 2 ∗ ∗
= | f (θ) | + | g(θ) | + sin θs sin(ϕ − ϕs ) i f (θ) g(θ) − f (θ) g (θ)
(1912)
If the initial beam is unpolarized, the direction of the initial spin (θs , ϕs )
should be averaged over by integrating over the solid angle dΩs = dϕs dθs sin θs .
This process yields the scattering probability for the unpolarized beam
Z
dΩs 2 2
I(θ, ϕ) = | f (θ) | + | g(θ) | (1913)
4π
which is independent of the azimuthal angle ϕ. It should be noted that the
unpolarized cross-section differs from the polarized cross-section.
Even if the initial beam is unpolarized, the final beam will be partially
polarized. The direction of the net polarization is determined by evaluating the
matrix elements of Ŝ and averaging over the direction of the initial spin, θs and
ϕs . The result is proportional to
∗
f (θ) g(θ) − f (θ) g ∗ (θ)
h̄
Ŝ = i (sin ϕ, − cos ϕ, 0) (1914)
2 | f (θ) |2 + | g(θ) |2
Hence, the polarization is perpendicular to the scattering plane. It should also
be noted that the net polarization of the scattered wave is determined by the
relative deviation of the scattering cross-section for polarized electrons from the
unpolarized scattering cross-section.
334
12.11.2 Partial Wave Analysis
The Dirac equation with a spherically symmetric potential V (r) has solutions
of the form
j± 12
!
f (r)
r Ω j,jz (θ, ϕ)
ψ(r) = j∓ 12 (1915)
i g(r)
r Ω j,jz (θ, ϕ)
j± 1
where the two-component spinor spherical harmonics Ωj,jz2 (θ, ϕ) are given by
q 1 1
j± 12
j+ 2 ± 2 ∓jz
j± 1 ∓ 2j+1±1 Y jz − 12
(θ, ϕ)
Ωj,jz2 (θ, ϕ) = q 1 1 j± 12
(1916)
j+ 2 ± 2 ±jz
2j+1±1 Y j + 1 (θ, ϕ)
z 2
c2 h̄2 k 2 = E 2 − m2 c4 (1918)
335
where δκ (k) are the phase shifts that characterize the potential. The phase
shifts depend directly on κ (and the energy) and only depend indirectly on j
and l through κ. The phase shifts are defined so that the asymptotic variation
of the radial functions is given by
fκ (r) cos(kr − (κ + 1) π2 + δκ (k))
∼ eiδκ (k) (1923)
r r
and only differs from the asymptotic variation of the free particle solutions
through the phase shifts. Furthermore, if this is decomposed in terms of incom-
ing and outgoing spherical waves,
exp i k r − (κ + 1) π2 + 2 δκ (k)
fκ (r)
∼
r 2r
π
exp − i k r − (κ + 1) 2
+ (1924)
2r
their fluxes are equal due to conservation of particles and, as written, the in-
coming spherical waves are not modified by the phase-shifts.
The general asymptotic r → ∞ form of the wave function for the scattering
is composed of the un-scattered wave and a spherical outgoing wave. The polar-
axis is chosen to be parallel to direction of the incident beam which is also chosen
to be the quantization axis for the spin. If the incident beam is polarized with
spin-up, the upper two-component spinor has the form
exp i k r
1 f (θ)
φA
↑ (r) = exp i k r cos θ +
0 g(θ) exp[ i ϕ ] r
(1925)
whereas for a down-spin polarized incident beam
exp i k r
0 − g(θ) exp[ − i ϕ ]
φA↓ (r) = exp i k r cos θ +
1 f (θ) r
(1926)
On recalling the Rayleigh expansion
X
exp i k r cos θ = il ( 2 l + 1 ) jl (kr) Pl (cos θ) (1927)
l
one can find the scattered spherical outgoing wave by subtracting the un-
scattered beam from the total wave function. On using the asymptotic large r
variation, one obtains the asymptotic form
cos(kr − (l + 1) π2 )
X
exp i k r cos θ → il ( 2 l + 1 ) Pl (cos θ) (1928)
kr
l
336
which has a similar form to the asymptotic form of the total wave function.
In particular, the spin and orbital angular momentum eigenstates can be de-
composed in terms of the spinor spherical harmonics. Thus, for the up-spin
polarized incident beam one has the upper two-component spinor
r
4π
Pl (cos θ) χ+ = Y l (θ, ϕ) χ+
2l + 1 0
√
√ √
4π
= l + 1 Ωll+ 1 , 1 − l Ωll− 1 , 1 (1929)
2l + 1 2 2 2 2
337
and the down-spin component is given by
√ r
4π X l(l + 1)
exp[ 2 i δ−l−1 (k) ] − 1
2ik 2l + 1
l
exp i k r
− exp[ 2 i δl (k) ] − 1 Y1l (θ, ϕ) (1935)
r
In the above expressions, the index on the phase-shifts δκ (k) refer to the value
of κ. Hence, for a spin-up polarized incident beam, the scattering amplitudes
are given in terms of the phase-shifts via
√
4π X (l + 1)
f (θ) = √ exp[ 2 i δ−l−1 (k) ] − 1
2ik 2l + 1
l
l
+ √ exp[ 2 i δl (k) ] − 1 Y0l (θ, ϕ) (1936)
2l + 1
and
√ r
4π X l(l + 1)
g(θ) exp[ i ϕ ] = exp[ 2 i δ−l−1 (k) ] − 1
2ik 2l + 1
l
− exp[ 2 i δl (k) ] − 1 Y1l (θ, ϕ) (1937)
A = B x êy (1939)
338
In the standard representation, the energy eigenvalue equation is represented
by the set of coupled equations
q
( E − m c2 ) φA (r) = c σ . ( p̂ − A ) φB (r)
c
q
( E + m c2 ) φB (r) = c σ . ( p̂ − A ) φA (r)
c
(1941)
Substituting the expression for φB from the second equation into the first, one
obtains the second-order differential equation for φA
2
2 2 4 A 2 q
(E − m c )φ = c σ . ( p̂ − A) φA (r)
c
2 q 2 q h̄
= c ( p̂ − A) − σ . B φA (r)
c c
= p̂ c + q B x − 2 q p̂y c B x − q c h̄ σ B φA (r)
2 2 2 2 2 (z)
(1942)
Since p̂y and p̂z commute with x, one can find simultaneous eigenstates of Ĥ,
p̂y and p̂z . Hence, the two-component spinor φA can be expressed as
A
φ (r) = exp i ky y + i kz z ΦA (x) (1943)
where
σ (z) χσ = σ χσ (1946)
in which the eigenvalues of σ (z) are denoted by σ. Therefore, the eigenvalue
equation can be reduced to
2
∂2
c h̄ ky
− h̄2 c2 + ( q B )2 x− f (x) = ( E 2 − m2 c4 − c2 h̄2 kz2 + q c h̄ B σ ) f (x)
∂x2 qB
(1947)
339
which (apart from an overall scale factor) is formally equivalent126 to the (non-
relativistic) energy eigenvalue equation for a shifted harmonic oscillator, with
frequency 2 c | q | B. The modulus sign was inserted to ensure that the frequency
ωHO is positive. The energy eigenvalues are determined from
1
( E 2 − m2 c4 − c2 h̄2 kz2 + q c h̄ B σ ) = 2 | q | c h̄ B ( n + ) (1948)
2
Hence, for an electron with negative charge q = − e one finds that the positive-
energy eigenvalue is given by the solution
r
| e | h̄
E = c m2 c2 + h̄2 kz2 + ( 2 n + 1 + σ ) B (1949)
c
This expression has an infinite degeneracy as it is independent of the continuous
variable ky . It also has a discrete (two-fold) degeneracy between the levels with
quantum numbers (n, σ = 1) and (n + 1, σ = −1). The two-fold degeneracy
can be understood as a consequence of the generalized helicity σ . ( p̂ − qc A )
commuting with the Hamiltonian Ĥ. This results in the spin’s alignment with
the electron’s velocity being preserved, as the spin’s precession is precisely bal-
anced by the electron’s orbital precession. It should be noted that if the g factor
deviates from 2, and such an anomaly in the g factor is expected from Quantum
Electrodynamics and has been found in experiment, then this degeneracy will
be lifted. The calculated ( g − 2 ) anomaly for an electron is given by
2 3 4
g − 2 1 α α α α
= − 0.3284986 + 1.17611 − 1.434 + ...
2 Theor 2 π π π π
(1950)
where 2
e
α = (1951)
h̄ c
is the fine structure constant. The experimentally determined value of the g
anomaly is found as
g − 2
= 0.0011659208 (1952)
2 Expt
126 The explicit (but dimensionally incorrect) analogy is obtained by setting the Harmonic
340
and differs from the theoretical value in the last two decimal places127 . In
the non-relativistic limit, the expression for the relativistic energy eigenvalue
reproduces the expression for energies of the well-known Landau levels
h̄2 kz2
1 + σ | e | h̄ B
E ≈ m c2 + + (n + ) (1953)
2m 2 mc
which are doubly-degenerate.
Aµ = Aµ (φ) (1955)
∂µ Aµ = kµ Aµ (φ)0 = 0 (1956)
where the prime indicates differentiation with respect to φ. The classical vector
potential must satisfy the source-free wave equation
∂ν ∂ ν Aµ = kν k ν Aµ (φ)00 = 0 (1957)
kν k ν = 0 (1958)
The Dirac equation for a spin one-half particle with charge q can be used to
obtain the second-order differential equation
q2
2 µ q µ µ 2 2 q µ ν 0
− h̄ ∂µ ∂ − 2 i h̄ A ∂µ + 2 Aµ A − m c − i h̄ γ kµ γ Aν (φ) ψ = 0
c c c
(1959)
127 This discrepancy could indicate the importance of virtual processes in which heavy par-
ticle/antiparticle pairs are created. The (g − 2) anomalies for the muon and its anti-particle
have also been measured [G. W. Bennett et al., Phys. Rev. Lett. 92, 1618102 (2004).].
These experiments show that particles and anti-particles precess at the same rate. However,
the value of the (g − 2) anomaly is inconsistent with the theoretical prediction based on the
standard model of particle physics.
341
where ψ is the four-component Dirac spinor. In deriving this, the Lorenz gauge
condition has been used to re-write
µ ν µ ν µ,ν
γ γ ∂µ Aν ψ = γ γ ∂µ Aν ψ − g ∂µ Aν ψ
(1960)
in the diagonal terms.
does commutes with the Hamiltonian and, therefore, is conserved. The con-
servation of this quantity can be interpreted in terms of the energy absorbed
or emitted by the electron due to interaction with the classical electromagnetic
field being accompanied by the absorption or emission of similar amount of mo-
mentum129 . Despite the different interpretation of pµ in the presence of the
classical field, the four-vector pµ shall be chosen to satisfy the condition
pµ pµ = m2 c2 (1964)
which is the dispersion relation for a free electron130 .
342
The form of the wave function of eqn(1961) is to be substituted into the
second-order differential eqn(1959). It shall be noted that
Aµ ∂µ F (φ) = kµ Aµ F (φ)0 = 0
∂ µ ∂µ F (φ) = k µ kµ F (φ)00 = 0 (1965)
µ µ
since A satisfies the Lorenz gauge condition and k satisfies the dispersion
relation for electromagnetic waves in vacuum. On substituting the ansatz into
the second-order equation, using the above two equations and the choice of pµ
satisfying the free-electron dispersion relation, one finds that the second-order
equation reduces to a first-order differential equation for the spinor F (φ)
q2
q q
2 i h̄ pµ k µ F (φ)0 = 2 Aµ pµ − 2 Aµ Aµ + i h̄ γ µ kµ γ ν Aν (φ)0 F (φ)
c c c
(1966)
which only depends on φ since the exponential phase-factor which depends on
pµ xµ has been factored out. The first-order equation can be integrated w.r.t.
φ to yield
Z φ
q γ µ kµ γ ν Aν
iq µ 0 1 q µ 0 0 0
F (φ) = exp − p Aµ (φ ) − A (φ ) A µ (φ ) dφ + F (0)
h̄ c pλ k λ 0 2 c c 2 pλ k λ
(1967)
where F (0) is an arbitrary constant four-component spinor. The exponential of
the matrix is defined in terms of its series expansion.
Z φ
iq µ 0 1 q µ 0 0 0
F (φ) = exp − p A µ (φ ) − A (φ ) A µ (φ ) dφ
h̄ c pλ k λ 0 2 c
q γ µ kµ γ ν Aν
× exp F (0) (1968)
c 2 pλ k λ
The above form can be simplified by expanding the last exponential factor due
to the identity n
µ ν
γ kµ γ Aν = 0 (1969)
for all integers n such that n > 1. The identity can be proved by
γ µ kµ γ ν Aν γ τ kτ γ ρ Aρ = − γ µ kµ γ τ kτ γ ν Aν γ ρ Aρ + 2 g ν,τ Aν kτ γ µ kµ γ ρ Aρ
= − γ µ kµ γ τ kτ γ ν Aν γ ρ Aρ (1970)
where the first line follows by using the anti-commutation relations for the γ
matrices and the second line follows from applying the Lorenz gauge condition.
The expression can be further simplified by noting that on anticommuting the
first pair of γ matrices, one has
= − γ µ kµ γ τ kτ γ ν Aν γ ρ Aρ
= γ τ kτ γ µ kµ γ ν Aν γ ρ Aρ + 2 g µ,τ kµ kτ
= γ τ kτ γ µ kµ γ ν Aν γ ρ Aρ
= γ µ kµ γ τ kτ γ ν Aν γ ρ Aρ (1971)
343
the third line follows from the condition k µ kµ = 0 and the last line follows
from interchanging the first two pairs of summation indices. On comparing the
first and last lines, one notes that the right-hand side is zero. Therefore, one
has proved the identity
γ µ kµ γ ν Aν γ τ kτ γ ρ Aρ = 0 (1972)
Therefore, one demands that F (0) satisfies the above supplementary condition
which is the same as for a free particle. Hence, one can set
χσ
F (0) = NF (1978)
p . σ
p(0) + m c
χσ
344
The spectrum of eigenvalues of the electron’s energy can be found by Fourier
transforming the above solution with respect to time, which shows that the elec-
tron absorbs and emits radiation in multiples of h̄ ω. The Volkov solutions have
been used to describe the Compton scattering of electrons by intense coherent
laser beams, and is also the basis of the strong-field approximation sometimes
found useful in atomic physics131 .
q γ ν Aν γ µ kµ
† † S
ψ = F (0) Iˆ + exp − i (1981)
c 2 pλ k λ h̄
q pν Aν q 2 Aν Aν
c q µ
j µ = (0) pµ − A + kµ − (1982)
p V c c k λ pλ c2 2 k λ pλ
γ µ p̂µ ψ = m c ψ (1984)
where the γ matrices are any set of matrices which satisfy the anti-commutation
relations
γ µ γ ν + γ ν γ µ = 2 g µ,ν Iˆ (1985)
131 L. V. Keldysh, Zh. Eksp. Teor. Fiz. 47, 1945 (1964). [Sov. Phys. J.E.T.P. 20, 1307
(1965).]
F. H. M. Faisal, J. Phys. B 6, L89 (1973).
H. R. Reiss, Phys. Rev. A 22, 1786 (1980).
345
The Dirac equation is independent of the specific representation of the γ matri-
ces. We have chosen the representation
I 0
γ (0) = (1986)
0 −I
and
σ (i)
(i) 0
γ = (1987)
−σ (i) 0
where σ (i) are the Pauli-matrices. This is the standard representation.
γ µ0 = Û γ µ Û † (1989)
invariant.
346
The components φL and φR are related to the components of ψ in the standard
representation via
L0 A
φ − φB
φ 1
= √ (1996)
φR0 2 φA + φB
The chiral representation is particularly useful for the description of massless
spin one-half particles, such as might be the case for the neutrino. The neutrino
masses are extremely small. The masses have evaded direct experimental mea-
surement. However, direct measurements have set upper limits on the masses
which decrease with time132 . In this case, with the limit m → 0, the Dirac
equation takes the form
∂
0 ∂t + c σ . ∇ L0
φ
= 0 (1997)
φR0
∂
∂t − c σ . ∇ 0
Hence, the Dirac equation for a massless free particle reduces to two uncoupled
equations, each of which are equations proposed by Weyl133
∂
+ c σ . ∇ φR0 = 0 (1998)
∂t
and
∂
− c σ . ∇ φL0 = 0 (1999)
∂t
The Weyl equation describes a spin one-half massless particle by a two com-
ponent spinor wave function. The Weyl equation violates parity invariance.
The Weyl equation was considered to be un-physical until the discovery of the
(anti-)neutrino134 and the associated violation of parity invariance135 . After the
parity violation of the weak interaction was established, the Weyl equation was
adopted to describe the neutrino136 .
Inexplicably nature seems to have selected the Weyl equation for φL , but not
φR to describing neutrinos. The solutions of the Weyl equation for free particles
∂
− c σ . ∇ φL = 0 (2000)
∂t
132 L. Langer and R. Moffat, Phys. Rev. 88, 689 (1952).
1413 (1957).
136 T. D. Lee and C. N. Yang, Phys. Rev. 105, 1671 (1957).
347
can be written as
(0)
u 1 i
φL = √ exp − ( E t − p . r ) (2001)
u(1) V h̄
Since helicity is conserved, one can choose the direction of p as the axis of
quantization. The positive-energy solution is given by
L 0 1 i
φ− = √ exp − (Et − pz) (2002)
1 V h̄
which has negative helicity and has energy given by
E− = c p (2003)
E+ = − c p (2005)
This negative-energy solution will describe anti-particles. The Weyl equation for
φR has a positive-energy solution with positive helicity, and a negative-energy
solution with negative helicity. Since only neutrinos with negative helicity are
observed in nature, only φL is needed. The anti-neutrinos have positive helicity
and are represented by φR .
E Λ=+1 Λ=−1
R L Λ=−1
φ φ ν∗ ν
Elementary Excitations
Λ=+1
Figure 61: The dispersion relations for φL and φR . The elementary excitations
are the negative-helicity neutrino ν and a positive-helicity anti-neutrino ν.
The Neutrino
348
decay products included a proton and an electron. However, it was observed
that the emitted electron had a continuous range of kinetic energies. Therefore,
another neutral particle must have been emitted in the decay. This particle was
termed the anti-neutrino, and the reaction can be written as
n → p + e− + ν e (2006)
Conservation of angular momentum requires that the neutrino has a spin of h̄2 .
Furthermore, since an energy of 1.2934 MeV is released in the transformation of
a neutron to a proton, and since sometime the decay processes produce electrons
which seem to take up all the released energy, the neutrino was suggested as
having zero mass. An upper limit on the neutrino’s mass of a few eV follows
from the Fermi-Kurie plot137 . The Fermi-Kurie plot of the electron energy
200
2 1/2
150
[N(p)/Fp ]
100
50
0
0 5 10 15 20
Energy [keV]
Figure 62: The Fermi-Kurie plot of the energy distribution of the electrons
emitted in the beta decay of tritium, 3 H → 3 He + e− + ν e . The decay releases
18.1 keV. It is seen that the electrons produced in the decay process have a
non-zero probability for carrying off most of the released energy. Hence, one
concludes that the anti-neutrinos are almost massless. The dashed blue curve
is the curve expected if the neutrino had a mass of 3 keV.
distribution is based on the phase space available for the emission of the electron
and anti-neutrino138 . The joint phase-space available for the electron of four-
momentum (Ee /c, p) and the anti-neutrino of four-momentum is (Eν /c, q) is
proportional to the factor
Z
dΓ = dp p2 dq q 2 δ(E − Ee (p) − Eν (q))
Z
p Ee (p) q Eν (q)
= dEe (p) 2
dEν (q) δ(E − Ee (p) − Eν (q))
c c2
137 L. Langer and R. Moffat, Phys. Rev. 88, 689 (1952).
349
1 p
2 − m2 c4 E (q)
= 5
dE e (p) p E e (p) E ν (q) ν ν
c
Eν (q)=E−Ee (p)
1 p
= dEe (p) p Ee (p) ( E − Ee (p) )2 − m2ν c4 ( E − Ee (p) )
c5
(2007)
The process of beta decay does not conserve parity. The non-conservation
of parity was discovered in the experiments of C. S. Wu et al.139 . In these
experiments, the spin of a 60 Co nucleus was aligned with a magnetic field. The
spin S = 5h̄ 60 Co nucleus decayed into a spin S = 4h̄ 60 N i nucleus by emitting
an electron and an anti-neutrino.
60
Co → 60
N i + e− + ν e (2008)
Since angular momentum is conserved, the spin of the electron and the anti-
neutrino initially must both be aligned with the field. In the experiment, the
angular distribution of the emitted electrons was observed. Because the helicity
of the electrons is conserved, the angular distribution of the electrons can be
used to prove that the electrons all have negative helicity, and hence it is inferred
that the anti-neutrinos should have positive helicity. Since helicity should be
reversed under the parity operation, and since only negative helicity electrons
are observed, the process is not invariant under parity. Hence, parity is not
conserved.
The electrons that are emitted in beta decay have negative helicities. If
the momentum of an emitted electron is given by (p, θp , ϕp ), then its helicity
operator is
cos θp sin θp exp[−iϕp ]
Λp = (2009)
sin θp exp[+iϕp ] − cos θp
The helicity operator has eigenstates χ given by
Λp χ± ±
θp = ± χθp (2010)
139 C. S. Wu, E. Ambler, R. W. Hayward, D. D. Hoppes and R. F. Hudson, Phys. Rev. 108,
1413 (1957).
350
which are determined as
!
θ ϕ
cos 2p exp[−i 2p ]
χ+
θp = θ ϕ
sin 2p exp[+i 2p ]
!
θ ϕ
− sin 2p exp[−i 2p ]
χ−
θp = θ ϕ (2011)
cos 2p exp[+i 2p ]
Since angular momentum is conserved and the emitted electrons only have neg-
ative helicity, the angular distribution of the emitted electrons is proportional to
the square of the overlap of the initial electron spin-up spinor with the negative
helicity spinors
† − 2 θp
| χ+
θ=0 χθp | = sin2
2
1
= ( 1 − cos θp ) (2012)
2
which is in exact agreement with the experimentally observed distribution. From
the distribution of emitted electrons one is led to expect that the anti-neutrino
has positive helicity.
νe
S=h/2
e-
S=h/2
Co
S=5h Ni
S=4h
Figure 63: The spin S = 5h̄ of the Co nucleus is aligned with the magnetic field.
The Co undergoes beta decay to N i which has S = 4h̄ by emitting an electron
e− and an anti-neutrino ν e . The spin of the electron and the anti-neutrino
produced by the decay must initially be aligned with the magnetic field, due to
conservation of angular momentum.
351
1.2
0.8
I(θ)
0.6
0.4
0.2
0
0 0.2 0.4 0.6 0.8 1
θ/π
Figure 64: The angular distribution of the emitted electron in the beta decay
experiment of Wu et al.
captures an electron from the K-shell and decays to the excited state of a 152 Sm
nucleus with angular momentum J = h̄ and emits a neutrino.
152
Eu + e− → 152
Sm∗ + νe (2013)
Goldhaber et al. measured the photons with the full Doppler shift, from which
they were able to infer the direction of the recoil of the nucleus. The photons
e-
Eu
Sm*
νe
J=1
γ Sm νe
Λγ=−1 J=0
352
oriented along the direction of motion of the emitted neutrino. Since the sum of
the angular momentum of the excited state (J = h̄) and the emitted neutrino
must equal the spin of the captured electron h̄2 , the neutrino must have its spin
oriented anti-parallel to the angular momentum of the Sm∗ nucleus. Hence, the
neutrino has negative helicity.
The Lagrangian equation of motion is found from the variational principle which
states that the action is extremal with respect to ψ and ψ † . The condition that
the action is extremal with respect to variations in ψ † leads to the Dirac equation
q
i h̄ γ µ ( ∂µ + i Aµ ) ψ = m c ψ (2018)
c h̄
after the resulting equation has been multiplied by a factor of γ (0) . On making
a variation of the action with respect to ψ, one finds the Hermitean conjugate
equation
q † †
− i h̄ c ( ∂µ − i Aµ ) ψ γ µ − m c2 ψ = 0 (2019)
c h̄
That this is the Hermitean conjugate of the Dirac equation can be shown by
taking its Hermitean conjugate, which results in
q
i h̄ c γ µ† γ (0) ( ∂µ + i Aµ ) ψ − m c2 γ (0) ψ = 0 (2020)
c h̄
The above equation can be reduced to the conventional form by multiplying by
γ (0) and by using the identities
γ (0) γ (0) = Iˆ
γ (0) γ µ† γ (0) = γµ (2021)
353
Hence, the equation found by varying ψ is just the Hermitean conjugate of the
Dirac equation
q
i h̄ γ µ ( ∂µ + i Aµ ) ψ = m c ψ (2022)
c h̄
Furthermore, it is surmised that the starting Lagrangian is appropriate to de-
scribe the Dirac field theory.
354
Likewise, (c times) the momentum density T 0 j is found from
†
T 0j = i h̄ c ψ γ (0) ∂j ψ
= i h̄ c ψ † ∂j ψ
= c ψ † p̂j ψ (2028)
where the partial derivative has been identified with the covariant momentum
operator. Hence, the contravariant component of the momentum is given by
∂
T 0,j = − i h̄ c ψ † ψ
∂xj
= c ψ † p̂(j) ψ (2029)
One can also determine the conserved Noether charges by noting that the
Lagrangian is invariant under a global gauge transformation
0
ψ → ψ = exp + i ϕ ψ
∗
∗ 0∗
ψ →ψ = exp − i ϕ ψ (2032)
δψ = + i δϕ ψ
δψ ∗ = − i δϕ ψ ∗ (2033)
δL = 0 (2034)
so we have
0 = δ L
∂L ∂L ∗ ∂L ∂L
= δψ + δψ + δ(∂µ ψ) + δ(∂µ ψ ∗ )
∂ψ ∂ψ ∗ ∂(∂µ ψ) ∂(∂µ ψ ∗ )
(2035)
355
After substituting the Euler-Lagrange equations for the derivatives w.r.t. the
fields ψ and ψ ∗ , the variation is expressed as
∂L ∂L ∗
0 = ∂µ δψ + δψ (2036)
∂(∂µ ψ) ∂(∂µ ψ ∗ )
For an arbitrary gauge transformation through the fixed infinitesimal angle δϕ,
this condition becomes
∂L ∂L ∗
0 = i δϕ ∂µ ψ − ψ (2037)
∂(∂µ ψ) ∂(∂µ ψ ∗ )
Hence, one finds that there is a current j µ which satisfies the continuity equation
∂µ j µ = 0 (2038)
where (apart from the infinitesimal constant of proportionality) the current is
given by
∂L ∂L ∗
j µ ∝ i δϕ ψ − ψ (2039)
∂(∂µ ψ) ∂(∂µ ψ ∗ )
For the Dirac Lagrangian, the second term is identically zero and the first term
is non-zero. Hence, on adopting a conventional normalization, the conserved
current is identified as
†
jµ = c ψ γµ ψ (2040)
This is the same expression for the conserved current that was previously derived
for the one-electron Dirac equation. Hence, the one-particle Dirac equation
yields the same expectation values and obeys the same conservation laws as the
(classical) Dirac field theory.
where φL and φR are two-component Dirac spinors and the two sets of quantities
σ µ and σ̃ µ are expressed in terms of the Pauli matrices as
µ
σL = ( σ0 , − σ )
µ
σR = ( σ0 , σ ) (2044)
356
µ µ
The difference between σL and σR reflect the different chirality of φL and φR . In
the absence of the mass term, the Dirac Lagrangian possesses two independent
scalar gauge transformations. These transformations corresponds to the global
gauge transformations
L L0 L
φ → φ = φ exp i θL
R R0 R
φ → φ = φ exp i θR (2045)
where θL and θR are independent angles. The Lagrangian has a U (1) × U (1)
gauge symmetry. The presence of a mass term would couple the two fields and
reduce the gauge transformation to one in which θR = θL .
φL
ψ = (2049)
φR
The first factor represents the usual global gauge transformation for the Dirac
Lagrangian with finite mass. This transformation yields the usual conserved
four-vector current jVµ defined by
†
jVµ = c ψ γ µ ψ (2050)
The second factor is specific to the Dirac Lagrangian with zero mass. It is
called the chiral transformation or axial U (1) transformation. Using the anti-
commutation relation
{ γ (4) , γ µ }+ = 0 (2051)
one can show that the exponential factor in the chiral gauge transformation has
the property that
µ θR − θL (4) θR − θL (4)
γ exp i γ = exp − i γ γ µ (2052)
2 2
357
This property can be used to show that the Lagrangian is invariant under the
chiral transformation because
† θR − θL θR − θL
ψ 0 γ µ ∂µ ψ 0 = ψ † exp − i γ (4) γ (0) γ µ ∂µ exp i γ (4) ψ
2 2
= ψ † γ (0) γ µ ∂µ ψ
†
= ψ γ µ ∂µ ψ (2053)
which involves two commutations. Since the massless Dirac Lagrangian is invari-
ant under the chiral transformation, Noether’s theorem shows that the current
µ †
jA = c ψ γ µ γ (4) ψ (2054)
which is the difference between the number of particles with positive helicity
and the number of particles with negative helicity.
141 Schwinger has noted that the condition of gauge invariance does not necessarily result
in the photon being massless. He argued that if the electromagnetic coupling strength were
larger, the photon could have finite mass. [J. Schwinger, “Gauge Invariance and Mass”,
Physical Review, 125, 397 (1962).]
358
Exercise:
Dirac noted that if the negative-energy states were all filled, then the Pauli
exclusion principle would prevent the decay of positive-energy particles into
the negative-energy states. Furthermore, in the absence of any positive-energy
particles, the Pauli exclusion principle would cause the set of particles in the
negative-energy state to be completely inert. In this picture, the filled sea
of negative-energy states would represent the physical vacuum, and would be
unobservable in experiments. For example, if charge is measured, it is the
non-uniform part of the charge distribution that is measured, but the infinite
number of particles in the negative-energy states do produce a uniform charge
density. Likewise, when energies are measured, the energy is usually measured
with respect to some reference level. For the case of a vacuum in which all
the negative-energy states are filled with electrons, the measured energies cor-
respond to energy differences and so the infinite negative energy of the vacuum
359
2
Unoccupied Positive Energy States
2 1
E/mc
-1
Figure 66: A cartoon depicting the vacuum for Dirac’s Hole Theory, in which
the negative-energy states are filled and the positive-energy states are empty.
should cancel. Therefore, Dirac postulated that the vacuum consists of the
state in which all the negative-energy states are all filled with electrons142 . Fur-
thermore, physical states correspond to the states were a relatively few of the
positive-energy states are filled with electrons and a few negative states are un-
occupied. In this case, the electrons in the positive-energy states are identified
with observable electrons, and the unfilled states or holes in the distribution of
negative-energy states are also observable. These holes are known as positrons
and are the anti-particles of the electrons. The properties of a positron are
found by computing the difference between the property for a state with an
absent negative-energy electron and the property of the vacuum state.
q p = ( N − 1 ) q e − N qe (2059)
Therefore, one finds that the positron has the opposite charge to that of an
electron
qp = − qe (2060)
142 P. A. M. Dirac, Proc. Roy. Soc. A 126, 360 (1930).
360
Hence, the positron has a positive charge. Likewise, the energy of the vacuum
in which all the electrons occupy all the negative-energy states is denoted by
E0 . The positron energy will be denoted as Ep (pe ). The positron corresponds
all states with negative energy being filled except for the state with the energy
p
Ee (pe ) = − m2 c4 + p2e c4 (2061)
= − Ee (pe )
p
= m2 c4 + p2e c4 (2062)
= − pe (2063)
Hence, the momentum of the positron is the negative of the momentum of the
missing electron
p p = − pe (2064)
Likewise, the spin of the positron is opposite to the spin of the missing electron,
etc. The velocity of an electron is defined as the group velocity of a wave packet
of momentum pe . Hence, one finds the velocity of the negative energy-electron
from
∂
ve = Ee (pe )
∂pe
pe c2
= − q (2065)
m2 c4 + p2e c2
361
Table 18: The relation between properties of Negative Energy Electron and
Positron States.
h̄
Electron −|e| −|E| +p + 2 σ σ.p v
h̄
Positron +|e| +|E| −p − 2 σ σ.p v
pe c2
= − q
m2 c4 + p2e c2
= ve (2066)
Therefore, the positron and the negative-energy electron states have the same
velocities.
e + e → 2γ (2067)
In this process, it is necessary that the excess energy be carried off by two
photons if the energy-momentum conservation laws are to be satisfied. Likewise,
by supplying an energy greater than a threshold energy of 2 m c2 , it should be
possible to promote an electron from a negative-energy state, thereby creating
an electron-positron pair. Since it is unlikely that more than one photon can
be absorbed simultaneously, electron-positron pair creation only occurs in the
vicinity of a charged nucleus which can carry off any excess momentum.
γ → e + e (2068)
362
2
Unoccupied Positive Energy States
1
2
E/mc
(k,α)
-1
of the charged particles in a cloud chamber in the presence of a magnetic field. Anderson
inferred the charge of the particles from their direction of motion. The insertion of a lead
plate in a cloud chamber caused the particles to lose energy on one side of the plate which
was observed as a change in the radius of curvature of the particle’s track. Therefore, the
examination of the radius of curvature of the track on both sides of the plate allowed the
direction of motion to be established.
145 P. M. S. Blackett and G. P. S. Occhialini, Proc. Roy. Soc. A 139, 688 (1933). These
authors were the first who correctly identified the positively charged particle as the anti-
particle of the electron, in full accord with the predictions of Dirac’s hole theory.
146 J. Thibaud, Phys. Rev, 35, 78 (1934).
147 P. A. M. Dirac, Proc. Roy. Soc. A126, 360 (1930).
363
to the radiation field is produced by γ (0) γ . A. The interaction operator can
be expressed as
0 σ
ĤInt = − q α . A = − q .A (2069)
σ 0
which only connects the upper and lower two-component spinors of the initial
and final states ψn and ψn0 . Hence, as light scattering processes are at least of
second-order in A, the intermediate state ψn00 must involve a negative-energy
electron state. Since the Pauli exclusion principle forbids the occupation of
the filled negative-energy states, hole theory ascribes the intermediate states
as involving virtual electron-positron creation and annihilation processes. This
shows that, even for processes which appear to involve a single electron in the
initial and final states, one must abandon single-particle quantum mechanics
and adopt a multi-particle description. Therefore, a purely single-particle de-
scription is inadequate and one must consider a many-particle description such
as quantum field theory.
and where q indicates all the quantum numbers of a positive-energy free electron
state. The sum over q 00 represents a sum over all possible intermediate states
of the electron, no matter whether they are positive or negative-energy states.
The matrix element M is composed of a coherent superposition of matrix el-
ements for virtual processes which represent the absorption of a photon (k, α)
followed by the subsequent emission of a photon (k 0 , α0 ) and the process where
the emission of light precedes the absorption process.
364
(k,α) (k',α')
(k,α) (k',α')
q''
q''
q' q q'
q
k + q = q 00
q 00 = k 0 + q 0 (2072)
which leads to the identification of the momentum of the intermediate and final
states as
q 00 = q + k
q0 = q + k − k0 (2073)
In the second process, where the emission process precedes the absorption, con-
servation of momentum yields
k + q = k + k 0 + q 00
k + k 0 + q 00 = k 0 + q 0 (2074)
which yields
q 00 = q − k 0
q0 = q + k − k0 (2075)
The limit in which the initial electron is at rest q = 0 shall be considered. The
momenta of the incident and scattered photon will be assumed sufficiently low
so that the momentum of the electron in the intermediate state can be neglected
since q 00 ≈ 0. That is, the Compton scattering process will be consider in the
limit k → 0 and k 0 → 0.
365
If the initial (positive-energy) electron is stationary and has spin σ, its wave
function can be represented by the Dirac spinor
1 χσ
ψσ,q (r) = √ (2076)
V 0
Because the interaction Hamiltonian has the form of an off-diagonal 2 × 2 block
matrix
q 0 σ
ĤInt = − .A (2077)
c σ 0
the only non-zero matrix elements are those which connect the upper two-
component spinor to the lower two-component spinor of the virtual state. Also,
momentum conservation requires that the virtual state also be one of almost
zero momentum. Hence, the electron in the virtual state must have the form of
a negative-energy eigenstate
1 0
ψσ00 ,q00 (r) ≈ √ (2078)
V χσ00
since the contribution from a positive-energy state with small momentum is
negligibly small. Therefore, the electronic part of the matrix elements involving
the initial electron simply reduce to the expression
< ψσ00 ,q00 | ĤInt | ψσ,q > = | e | χ†σ00 σ χσ . A (2079)
Likewise, the matrix elements which involve the final (positive energy) electron
are evaluated as
< ψσ0 ,q0 | ĤInt | ψσ00 ,q00 > = | e | χ†σ0 σ χσ00 . A (2080)
From these one finds that, to second-order, the matrix elements that appear in
the transition rate are given by
χσ00 χ†σ00 = I
X
(2083)
σ 00
366
the matrix elements are evaluated as
e2 2 π h̄ c2
† 0 0
M ≈ √ χσ0 ( σ .
ˆα (k) ) ( σ .
ˆα0 (k ) ) + ( σ .
ˆα0 (k ) ) ( σ .
ˆα (k) ) χσ
2 m c2 V ωk ωk 0
(2084)
The products in the above expression can be evaluated with the aid of the Pauli
identity. The result is
( σ . ˆα (k) ) ( σ . ˆα0 (k 0 ) ) = ( ˆα (k) . ˆα0 (k 0 ) ) + i σ . ( ˆα (k) ∧ ˆα0 (k 0 ) ) (2085)
Therefore, after combining both terms and noting that the pair of vector product
terms cancel since the vector product is antisymmetric under the interchange
k ↔ k 0 , one finds that the matrix elements reduce to
e2 2 π h̄ c2
† 0
M ≈ √ χσ0 2 ˆα (k) . ˆα (k ) χσ
0
2 m c2 V ωk ωk 0
e2 2 π h̄ c2
M ≈ √ δσ,σ0 2 ˆα (k) . ˆα0 (k 0 ) (2086)
2 m c2 V ωk ωk 0
Hence the matrix elements are diagonal in the spin indices. The above matrix
elements are identical to the matrix elements that occur in the non-relativistic
quantum theory of Thomson scattering. On substituting this result into eqn(2070),
one recovers the non-relativistic expression for the differential scattering cross-
section
2 2
ωk 0
dσ e
≈ δσ,σ 0 | ˆα (k) . ˆα0 (k 0 ) |2
dΩk0 ωk m c2
2 2
ωk 0
e
≈ δσ,σ0 cos2 Θ (2087)
ωk m c2
where
cos Θ = ˆα (k) . ˆα0 (k 0 ) (2088)
is the angle subtended by the initial and final polarization vectors. Hence, one
concludes that the negative-energy states do play an important role in light
scattering processes which involve low-energy electrons. The result, although
correct, does need re-interpretation, since the states of negative energy are as-
sumed to be filled with electrons in the vacuum and, therefore, the electron is
forbidden to occupy these levels in the intermediate states.
Electron-Positron Interpretation
The first contribution to the matrix elements, which was described above,
has to be re-interpreted as representing a process in which an electron that
initially occupies the negative-energy state q 00 makes a transition to the positive-
energy state q 0 while emitting the photon (k 0 , α0 ). This transition is subsequently
followed by the positive-energy electron q absorbing the photon (k, α) and falling
367
q q
e- e-
(k',α') (k,α)
q''
e+ e+ q''
q' (k',α')
(k,α)
e-
-
e
q'
into the empty negative-energy state. In this process, the negative-energy states
are completely occupied in the initial and final state, and the energy of the initial
and final states are conserved. By re-ordering the factors in the matrix elements
and noting that since
the contribution to the matrix element of these two descriptions are identical
(apart from an over all negative sign).
one finds an identical expression (and the multiplicative factor of minus one).
Hence, Dirac hole-theory does lead to the correct classical result.
The above description is quite cumbersome, but can be made more concise by
adopting an anti-particle description of the unoccupied negative-energy states.
The first contribution to M first involves the creation of a virtual electron-
positron pair with the emission of the photon (k 0 , α0 ). The electron which has
just been created in the momentum eigenstate (q 0 , σ 0 ) remains unchanged in
the final state. Subsequently, the positron annihilates with the initial electron
(q, σ) while absorbing the photon (k, α). Since the intermediate state is a vir-
tual state, energy does not have to be conserved. The second contribution to M
involves the creation of a virtual electron-positron pair with the absorption of
the photon (k, α). The created electron (q 0 , σ 0 ) remains in the final state while
368
the positron subsequently annihilates with the initial electron (q, σ) and emits
the photon (k 0 , α0 ). This process is also a virtual process if the energy of the
incident light h̄ ωk is less than 2 m c2 .
We shall multiply the complex conjugate of the Dirac equation by γ (2) and anti-
commute γ (2) with the real γ µ∗ and commute γ (2) with the γ (2)∗ matrix. This
procedure changes the sign in front of the term originating from the differential
momentum operator w.r.t. the sign of the mass term. This procedure yields
q
γ (2) µ∗
γ ( − i h̄ ∂µ − Aµ ) − m c ψ ∗ = 0
c
q
µ
γ ( i h̄ ∂µ + Aµ ) − m c γ (2) ψ ∗ = 0 (2095)
c
148 O. Klein and Y. Nishina, Zeit. für Physik, 52, 843 (1928).
369
Hence, one sees that γ (2) ψ ∗ describes a Dirac particle with mass m and a charge
of − q moving in the presence of a vector potential Aµ . The fact that the opera-
tion of charge conjugation (in any representation) involves complex conjugation
is related to gauge invariance. Charge conjugation is a new type of symmetry for
particles that have complex wave functions which relates particles to particles
with opposite charges. The charge conjugate field ψ c is defined as
ψ c = Ĉ ψ ∗ (2096)
which is the result of the complex conjugation followed by the action of a linear
operator Ĉ. The joint operation can be represented as an anti-unitary operator.
The charge conjugation operator Ĉ is defined as the unitary and Hermitean
operator
Ĉ = − i γ (2) (2097)
The charge conjugation operator is Hermitean as
Ĉ † = + i γ (2)† = − i γ (2) = Ĉ (2098)
and it is unitary since
Ĉ † Ĉ = − γ (2) γ (2) = Iˆ (2099)
where the anti-commutation relations of the γ matrices have been used. It was
through this type of logic that Kramers149 discovered the form of the charge
conjugation transformation which turns a particle into an anti-particle.
where we have used the identity z = (z ∗ )∗ in the second line. However, since Ĉ
is real, one finds
Z Z ∗
3 c† c 3 † ∗
d r ψ (r) Â ψ (r) = d r ψ (r) Ĉ Â Ĉ ψ(r)
Z ∗
= − d3 r ψ † (r) γ (2) Â∗ γ (2) ψ(r)
(2102)
149 H. A. Kramers, Proc. Amst. Akad. Sci. 40, 814 (1937).
370
This shows the relation between expectation values of a general operator  in
a state ψ(r) and its charge conjugated state ψ c (r).
We shall examine the effect of charge conjugation on the plane wave solutions
of the Dirac equation. The plane-wave solutions can be written as
r
( E + m c2 )
χσ µ
ψσ,k (x) = c h̄ σ . k exp − i k xµ (2103)
2EV E + m c2 χσ
where
Ĉ = − i γ (2)
−iσ (2)
0
=
iσ (2) 0
0 0 0 −1
0 0 1 0
=
0 1 0 0
(2105)
−1 0 0 0
cos θ2 exp[+i ϕ2 ]
χ+σ (θ, ϕ)∗ = (2107)
sin θ2 exp[−i ϕ2 ]
150 Note that the helicity is invariant under the joint transformation
σ → −σ
k → −k
371
it turns it into the negative-eigenvalue eigenstate
− sin θ2 exp[−i ϕ2 ]
χ−σ (θ, ϕ) = (2108)
cos θ2 exp[+i ϕ2 ]
The end result is that the charge conjugated single-particle wave function has
the form
r
c h̄ ( σ . (−k) )
( E + m c2 )
c − 2 χ−σ µ
ψσ,k (x) = E + m c exp + i k xµ
2EV χ−σ
(2111)
The properties described above are the properties of a state of a relativistic free
particle with a negative energy eigenvalue − E, momentum − h̄ k and spin − σ.
The absence of an electron in the charge conjugated state describes a positron,
with positive energy E, momentum h̄ k and spin σ.
Exercise:
Exercise:
372
Prove the completeness relation for the set of solutions for the Dirac equation
for a free particle
X
φ†α (r)λ φα (r0 )ρ + φcα † (r)λ φcα (r0 )ρ = δ 3 (r − r0 ) δλ,ρ (2112)
α
The role of the source term for the electromagnetic field is played by the term
in the Dirac Lagrangian which represents the coupling to the vector potential.
The expression for the QED action
Z
SQED = dx4 LQED (2114)
Aµ → A0µ = Aµ + ∂µ Λ
iq
ψ → ψ0 = ψ − Λψ
h̄ c
iq
ψ† → ψ †0 = ψ † + Λ ψ† (2115)
h̄ c
151 Frequently, the relativistic free electron states are given a manifestly covariant normaliza-
tion, in order to facilitate covariant perturbation theory. The use of different normalization
conventions results in changes the form of the completeness relation.
373
Applying Noether’s theorem to the action yields an infinite number of continuous
currents, which, up to a multiplicative factor, are given by
ν ∂LQED ∂LQED iq
jΛ = ∂µ Λ − Λψ (2116)
∂ (∂ν Aµ ) ∂ (∂ν ψ) h̄ c
which satisfy the continuity conditions
∂ν jΛν = 0 (2117)
The currents are identified as
1 †
jΛν = F µ,ν ∂µ Λ + q ψ γ ν Λ ψ (2118)
4π
On substituting the Euler-Lagrange equation for the electromagnetic field
4π †
∂µ F µ,ν = q c ψ γν ψ (2119)
c
into the second term of the four-vector current density, one finds that the con-
tinuous current has the form
1
jΛν = ∂µ F µ,ν Λ (2120)
4π
The conserved charge Q can be defined as a volume integral of the temporal
component of the four-vector current density
Z
1 (0)
QΛ = d3 r jΛ (r)
c
Z
1
= d3 r ∂µ F µ,0 Λ
4πc
Z
1
= d3 r ∇ . ( E Λ ) (2121)
4πc
For a global transformation for which Λ is constant over all space-time, one can
write Z
Λ
QΛ=const. = d3 r ( ∇ . E ) (2122)
4πc
On using Gauss’s law
∇.E = 4πρ (2123)
one finds that global gauge invariance ensures conservation of electrical charge
Z
Λ
QΛ=const. = d3 r ρ(r) (2124)
c
However, for any of the infinite number of spatially varying Λ which vanish on
the boundaries
Z
1
QΛ = d3 r ∇ . ( E Λ )
4πc
Z
1
= d2 S . E Λ
4πc
= 0 (2125)
374
Hence, although local transformations lead to an infinite number of symmetries
and an infinite number of continuous currents, all the conserved charges are
zero. Thus, local gauge invariance does not leads to new conserved quantities.
It does, however, constrain the form of the interactions.
{ Â , B̂ }+ ≡ Â B̂ + B̂ Â (2126)
The fermion creation and annihilation operators, ĉ†α and ĉα , satisfy the anti-
commutation relations
{ ĉ†α , ĉ†β }+ = 0
{ ĉα , ĉβ }+ = 0 (2127)
and
{ ĉ†α , ĉβ }+ = δα,β (2128)
where the quantum numbers α and β describe a complete set of single-particle
states.
375
The choice of anti-commutation relations results in the eigenvalues of the num-
ber operator to be restricted to either nα = 1 or nα = 0. This can be seen by
examining the identity
n̂α n̂α = n̂α (2132)
which follows from
where we have used the anti-commutation relation for the creation and annihila-
tion operator to obtain the second line and used the anti-commutation relation
for two annihilation operators to obtain the last line. On comparing the second
and third lines, one recognizes that
Hence, we have
376
The state | 0 > is also an eigenstate of the number operator with eigenvalue
zero since
n̂α | 0 > = ĉ†α ĉα | 0 >
= 0 (2140)
A general eigenstate of the number operator with eigenvalue nα can be expressed
as
( ĉ† )nα
| nα > = √α |0 > (2141)
nα !
as can be seen by using the commutation relation
[ n̂α , ( ĉ†α )nα ] = nα ( ĉ†α )nα (2142)
If this operator equation acts on the state where the quantum state α is unoc-
cupied | 0 >, and using the condition
n̂α | 0 > = 0 (2143)
one finds the state of equation(2140) satisfies the eigenvalue equation
n̂α | nα > = nα | nα > (2144)
with eigenvalue of either unity or zero.
where the allowed values of the set of occupation numbers nα are either unity
or zero. The sequencing or ordering of the creation operators in this expression
is crucial, since the interchange the positions of the operators may result in a
change in sign of the state. For example, the action of a creation operators on
a general number eigenstate has the effect
Pβ
ĉ†β | n1 n2 . . . nβ . . . > = ( − 1 )( i=1 ni ) | n1 n2 . . . nβ + 1 . . . > (2146)
where the sign occurs since this involves anti-commutating ĉ†β with
Pβ
i=1 ni
other creation operators to bring it into the β-th position.
377
with a different notation from the positive-energy states. The change of no-
tation is to reflect the intent of describing the (quasi-particle) excitations of
the system and not to describe the many-particle ground state which is unob-
servable. The wave functions φα (r) describing the positive-energy states of the
non-interacting electrons are indexed by the set of quantum numbers α ≡ (k, σ).
The negative-energy states are described as the charge conjugates of the positive-
energy states. Therefore, the negative-energy states are described by the same
set of indices α and the corresponding wave functions are denoted by φcα (r).
The annihilation operator for electrons in the positive-energy state α is denoted
by ĉα . However, the operator which removes an electron from the (negative-
energy) charge conjugated state φcα (r) is denoted by a creation operator b̂†α . The
change from annihilation operator to creation operator merely represents that
creating a positron with quantum numbers α is equivalent to creating a hole in
the negative-energy state154 . The effect of the annihilation operators on Dirac’s
vacuum | 0 >, in which all the negative-energy states are fully occupied are
ĉα | 0 > = 0
b̂α | 0 > = 0 (2147)
where the first expression follows from the assumed absence of electrons in the
positive-energy states, and the second expression follows from the assumption
that all the negative-energy states are completely filled, so adding an extra elec-
tron to the state φcα is forbidden by the Pauli-exclusion principle. More concisely,
the above relations state that the vacuum contains neither (positive-energy)
electrons nor positrons. It is seen that the form of the anti-commutation rela-
tions are unchanged by this simple change of notation. The anti-commutation
relations become
The mixed electron/positron anti-commutation relations are all zero, since the
operators describe electrons in different single-particle energy eigenstates. In
154 W. H. Furry and J. R. Oppenheimer, Phys. Rev. 45, 245 (1934).
378
this notation, the field operators are expressed as155
X
c †
ψ̂(r) = φα (r) ĉα + φα (r) b̂α (2151)
α
and
X
ψ̂ † (r) = φ∗α (r) ĉ†α + φcα ∗ (r) b̂α (2152)
α
The field operators ψ̂(r) and ψ̂ † (r) are expected to be canonically conjugate, as
we shall show below.
Π̂(r) =
1 δL ˆ † (r) γ (0) = i h̄ ψ̂ † (r)
= i h̄ ψ (2154)
c δ(∂0 ψ̂)
Hence, one expects that the field operators ψ̂ † (r) and ψ̂(r) are canonically con-
jugate and, therefore, satisfy the equal-time anti-commutation relations
where λ and ρ label the components of the Dirac spinor. The anti-commutation
relations for the field operators can be verified by noting that
X
† 0
{ ψ̂ (r) , ψ̂(r ) }+ = { ĉ†α , ĉβ }+ φ∗α (r) φβ (r0 ) + { ĉ†α , b̂†β }+ φ∗α (r) φcβ (r0 )
α,β
+ { b̂α , ĉβ }+ φcα ∗ (r) φβ (r0 ) + { b̂α , b̂†β }+ φcα ∗ (r) φcβ (r0 )
X
= δα,β φ∗α (r) φβ (r0 ) + δα,β φcα ∗ (r) φcβ (r0 )
α,β
X
= φ∗α (r) φα (r0 ) + φcα ∗ (r) φcα (r0 )
α
= δ (r − r0 )
3
(2156)
where the fermion anti-commutation relations have been used in arriving at the
second line. The positive-energy states and their charge conjugated states form
a complete set of basis states for the single-particle Dirac equation, so their
155 W. Heisenberg and W. Pauli, Zeit. für Physik, 56, 1 (1929).
379
completeness condition has been used in going from the third to the fourth line.
The equal-time field anti-commutation relations can be generalized to field anti-
commutators at space-time points with a general type of separation. In the case
where the two field points x and x0 have a space-like separation
µ
( xµ − x0 ) ( xµ − x0 µ ) < 0
causality dictates that the anti-commutators are zero
{ ψ̂ † (x) , ψ̂(x0 ) }+ = 0
That is, for space-like separations, there is no causal connection156 so a mea-
surement of a local field at x0 cannot affect a measurement at x. N. Bohr and
∆x2 > 0
ct
r
2
∆x < 0
Figure 70: Due to causality, the anti-commutator of the field operator should
vanish for space-like separations. The anti-commutators can be non-zero inside
or on the light cone.
L. Rosenfeld157 have put forward general arguments that the commutation rela-
tions also place limitations on the measurement of fields at time-like separations.
The Hamiltonian density for the (non-interacting) quantized Dirac field the-
ory can be expressed as the operator
Ĥ = ψ̂ † γ (0) c − i h̄ γ . ∇ + m c ψ̂ (2157)
When the expansion of the quantized field in terms of single-particle wave func-
tions is substituted into the Hamiltonian, one finds
X
† c †
Ĥ = Eα ĉα ĉα + Eα b̂α b̂α
α
156 Outside the light-cone there is no way to distinguish between future and past.
157 N. Bohr and L. Rosenfeld, Kon. Dansk. Vid. Selskab., Mat.-Fys. Medd. XII, 8 (1933).
380
X
= Eα ĉ†α ĉα − Eα b̂α b̂†α (2159)
α
where the expression for the energy of the charge conjugated state
Eαc = − Eα (2160)
has been used. On anti-commuting the positron and annihilation operators, one
finds
X
Ĥ = Eα ĉ†α ĉα + b̂†α b̂α − 1 (2161)
α
The last term, when summed over α, yields the infinitely negative energy of
Dirac’s vacuum in which all the negative-energy states are filled. The vacuum
energy shall be used as the reference energy, so the Hamiltonian becomes
X
† †
Ĥ = Eα ĉα ĉα + b̂α b̂α (2162)
α
which describes the energy of the excited state as the sum of the energies of the
excited electrons and the excited positrons. The energies of the positrons and
electrons are given by positive numbers.
which is just the sum of the momenta of the (positive-energy) electrons and the
positrons. The spin operator is defined as
Z
h̄
Ŝ = d3 r ψ̂ † σ̂ ψ̂ (2164)
2
This is evaluated by substituting the expression for the field operators in terms
of the single-particle wave functions and the particle creation and annihilation
operators. The expectation value of the spin operator in the charge conjugated
state φcα is given by
Z Z ∗
d3 r φcα † (r) σ̂ φcα (r) = − d3 r φ†α (r) γ (2) σ ∗ γ (2) φα (r)
Z ∗
= d3 r φ†α (r) σ (2) σ̂ ∗ σ (2) φα (r)
Z ∗
3 †
= − d r φα (r) σ̂ φα (r)
Z
3 †
= − d r φα (r) σ̂ φα (r) (2165)
381
The third line follows from the identity
The last line follows since σ is Hermitean. Hence, the spin operator is evaluated
as
h̄ X † † †
Ŝ = χσ00 σ χσ0 ĉk,σ00 ĉk,σ0 + b̂k,σ00 b̂k,σ0 (2167)
2 0 00
k;σ ,σ
which is just the sums of the spins of the electrons and positrons.
The last term in the parenthesis, when summed over all states α, yields the total
charge of the vacuum which is to be discarded. Hence, the observable charge is
defined as
X
Q̂ = ĉ†α ĉα − b̂†α b̂α (2169)
α
which shows that the total electrical charge defined as the difference between
the number of electrons and the number of positrons is conserved.
14.3.1 Parity
The parity eigenvalue equation for a multi-particle state with parity ηψ can be
expressed as
P̂ | ψ > = ηψ | ψ > (2170)
Since the action of the parity operator on states is described by a unitary opera-
tor, operators transform under parity according to the general form of a unitary
382
transformation. In particular, the effect of the parity transformation on the
field operator is determined as
one has
X
P̂ ψ̂(r) P̂ = ĉα P̂ φα (r) P̂ + b̂†α P̂ φcα (r) P̂ (2173)
α
where ηαP is a phase factor which represents the intrinsic parity of the state.
Furthermore, since P̂ 2 = I,ˆ then the intrinsic parities ηαP and ηαP c have to
satisfy the conditions
( ηαP )2 = 1
( ηαP c )2 = 1 (2175)
So the intrinsic parities are ±1. The intrinsic parity of a state φα (r) and its
charge conjugated state φcα (r) are related by
This follows since charge conjugation flips the upper and lower two-component
spinors and these two-component spinors have opposite intrinsic parity. There-
fore, the state φα (r) and the charge conjugates state φcα (r) have opposite pari-
ties. Therefore, it follows that the field operator transforms as
X
P̂ ψ̂(r) P̂ = ηαP ĉα φPα (r) − ηαP b̂†α φcPα (r) (2177)
α
383
The relations between parity reversed states and parity reversed charge con-
jugated states can be verified by examining the free particle solutions of the
Dirac equation and noting that the parity operator consists of the product of
γ (0) and spatial inversion r → − r. This spatial inversion acting on a wave
function with momentum k and spin σ becomes a wave function with momentum
−k and spin σ, up to a constant of proportionality. A free particle momentum
eigenstate is given by
χσ (0)
φσ,k (x) = N c h̄ k . σ exp − i ( k0 x − k.r) (2178)
E + m c2 χσ
The application of the parity operator to the above wave function yields
(0) χσ (0)
P̂ φσ,k (x) = N γ c h̄ k . σ exp − i ( k0 x + k.r)
E + m c2 χσ
χσ
= N c h̄ k . σ exp − i ( k0 x(0) + k . r )
− E + m c2 χσ
= φσ,−k (x) (2179)
where
Ĉ = − i γ (2)
0 0 0 −1
0 0 1 0
=
0 1 0 0
(2181)
−1 0 0 0
384
where in the first line the parity operator has sent r → − r and the factor of
γ (0) has flipped the sign of the lower components. In the second line we have
re-written k as −(−k) in the two two-component spinor, in anticipation of the
comparison with eqn(2180) which allows us to identify the factor of φcσ,−k (x).
This example shows that a state and its charge conjugate have opposite intrinsic
parities.
From the general form of the parity transformation on Dirac spinors, one
infers that the parity transform of the field operator is given by
X
P̂ ψ̂(r) P̂ = ĉα ηαP φPα (r) − b̂†α ηαP φcPα (r) (2184)
α
Thus, the parity operation can also be interpreted as only affecting the particle
creation and annihilation operators, and not the wave functions. Quantum
mechanically, this interpretation corresponds to viewing that the particles as
being transferred into their parity reversed states
X
† c
P̂ ψ̂(r) P̂ = P̂ ĉα P̂ φα (r) + P̂ b̂α P̂ φα (r) (2187)
α
In this new interpretation, the effects of parity on the fermion operators are
found by identifying the operators multiplying the single-particle wave functions
in the previous two equations. The resulting operator equations are
P
P̂ ĉα P̂ = ηPα ĉPα (2188)
and
P̂ b̂†α P̂ = − ηPα
P
b̂†Pα (2189)
which shows that fermion particles and anti-particles have opposite intrinsic
parities. Therefore, we conclude that, irrespective of which interpretation is
used, the field operator transforms as
X
P P † c
P̂ ψ̂(r) P̂ = ηα ĉα φPα (r) − ηα b̂α φPα (r) (2190)
α
which shows that the quantum field operators transforms in a similar fashion
to the classical field.
385
14.3.2 Charge Conjugation
Under charge conjugation, the classical Dirac field transforms as
ψ → ψ c = − i γ (2) ψ ∗ (2191)
(up to an arbitrary phase) since this is how the single-particle wave functions
transform. Classically, the (anti-linear) charge conjugation operator Cˆ is the
product of complex conjugation and the unitary matrix operator Ĉ = − i γ (2) .
If the classical field is expressed as a linear superposition of energy eigenfunc-
tions, the amplitudes of the eigenfunctions are represented by complex numbers.
In the charge conjugated state, these amplitudes are replaced by the complex
conjugates. In the quantum field, the amplitudes must be replaced by parti-
cle creation and annihilation operators. If an amplitude is associated with an
annihilation operator, then the complex conjugate of the amplitude is usually
associated with a creation operator. Hence, we should expect that charge con-
jugation will result in the creation and annihilation operators being switched.
where, in accord with the earlier comment about the relation between the quan-
tum and classical fields, the single-particle operators have been replaced by their
Hermitean conjugates. However, under charge conjugation general Dirac spinors
satisfy
therefore,
X
ψ̂ (r) = Cˆ ψ̂(r) Cˆ =
c
η c
ĉ†α φcα (r) + b̂α φα (r) (2195)
α
386
For consistency, the two expressions for ψ̂ c (r) must be equivalent. Hence, the
operator coefficients of φα (r) and φcα (r) in the two expressions should be iden-
tical. Therefore, one requires that
Cˆ ĉα Cˆ = η c b̂α
Cˆ b̂†α Cˆ = η c ĉ†α (2197)
ψ̂ c = Cˆ ψ̂ Cˆ = − i η c γ (2) ψ̂ † (2198)
where ψ̂ † is the Hermitean conjugate (column) field operator. Apart from the
replacement of the complex amplitudes with the Hermitean conjugates of the
creation and annihilation operators, the above expression is identical to the ex-
pression for charge conjugation on the classical field.
The charge conjugation operator has the effect of reversing the current den-
sity operator
ˆ † γ µ ψ̂ Cˆ = − ψ
Cˆ ψ ˆ † γ µ ψ̂ (2199)
which is understood as the result in the change of the charge’s sign.
Thus, under time reversal, the time and spatial components of the position
four-vector have different transformational properties. Furthermore, the energy-
momentum four-vector transforms as
Hence, the position four-vector and momentum four-vector have different trans-
formational properties. Due to the above properties, angular momentum (in-
cluding spin) transforms as
T̂ J = − J (2202)
Therefore, we find that time reversal reverses momenta and flips spins.
387
T̂ interchanges the initial and final states, then
satisfies the Dirac equation with t → − t. For example, the plane wave
solutions of the Dirac equation can be shown to transform as
which flips the momentum and the spin angular momentum. It should be noted
that the matrix operator γ (1) γ (3) does not couple the upper and lower two-
component spinors, but nevertheless is closely related to the operator − i γ (2)
which occurs in the charge conjugation operator.
T̂ cα T̂ = cT α
T̂ bα T̂ = bT α (2210)
388
Table 19: Discrete Symmetries of Particles.
The charge conjugated of a state is a negative energy state with momentum −p
and spin −σ, that is interpreted as the state of antiparticle with momentum p
and spin σ.
Q p σ Λ
Charge Conjugation − + + +
Parity + − + −
Time Reversal + − − +
CPT − + − −
ψ 0 (x0 ) = Cˆ P̂ T̂ ψ(x)
∗
(2)
= −iγ P̂ T̂ ψ(x)
∗
= + i γ (2) γ (0) γ (1) γ (3) ψ ∗ (−x)
389
14.3.4 The CPT Theorem
The CPT theorem states that any local161 quantum field theory with a Her-
mitean Lorentz invariant Lagrangian which satisfies the spin-statistics theorem,
is invariant under the compound operation Cˆ P̂ T̂ , where the operators can be
placed in any order.
The proof of the theorem relies on the fact that any Lorentz invariant quan-
tity must be created out of contracting the indices of bi-linear covariants (quan-
tities such as the current density jµ which involve products of the γµ ) with the
indices of contravariant derivatives ∂ µ . Since the joint operation P̂ T̂ results in
each of the contravariant derivatives ∂ µ in the product changing sign, the theo-
rem ensures that the corresponding bi-linear covariants with which the deriva-
tives are contracted with must undergo an equivalent number of sign changes
under the compound operation Cˆ P̂ T̂ . The theorem only assumes invariance
under proper orthochronous Lorentz transformations and makes no assumptions
about reflection. The improper transformations are treated as analytic continu-
ation of the Lorentz transformation into complex space-time. The theorem was
first discussed by Lüders162 and Pauli163 , and then by Lee, Oehme and Yang164 .
The theorem has several consequences, such as the equality of the masses of
particles and their anti-particles. This follows since the mass mc is an eigenvalue
of p̂(0) in the particle’s rest frame and since one can find simultaneous eigenstates
of the commuting operators p̂µ and the product Cˆ P̂ T̂ . If one denotes the
compound operator as
Θ̂ = Cˆ P̂ T̂ (2212)
then
since the CPT theorem ensures that Θ̂ commutes with the Hamiltonian
Θ̂ Ĥ Θ̂−1 = Ĥ (2214)
interactions can be expressed in terms of products of fields at the same point in space-time.
It would be truly remarkable if this concept were to continue to work at arbitrarily small
distances!
162 G. Lüders, Dan. Mat. Fys. Medd. 28, 5 (1954).
390
then the state Θ̂ | Ψ > describes an anti-particle with flipped angular momen-
tum. This follows since the vacuum satisfies
one finds that the energy of a particle is equal to the energy of an anti-particle
with a reversed spin. However, as the rest mass cannot depend on the angular
momentum, the mass of a particle is equal to the mass of its anti-particle. For
unstable particles, the equality of the mass of the particle and anti-particle is
ensured by the invariance of the S-matrix under Cˆ P̂ T̂ .
Likewise, one can use the CPT theorem to show that the total decay rate
of a particle into products is equal to the total decay rate of the anti-particle
into its products165 . It should be noted that the partial decay rates into specific
final states are not equivalent, only the sums over all final states are equal.
which only has positive excitation energies. Hence, if the wave function changes
sign under the interchange of a pair of spin one-half particles the energy is
bounded from below. If the field operators had been chosen to obey commu-
tation relation, then the wave function would have been symmetric under the
165 T. D. Lee, R. Oehme and C. N. Yang, Phys. Rev. 106, 340 (1957).
166 W. Pauli, Phys. Rev. 58, 716 (1940).
391
interchange of particles. If this were the case, there would be a negative sign in
front of the positron energies so that the energy would have been unbounded
from below. This would have implied that the vacuum would not be stable, and
the theory is erroneous. This can be taken as implying that spin one-half par-
ticles must obey Fermi-Dirac Statistics. The other part of the theorem compels
integer spin particles to be bosons. Therefore, since photons have spin one, the
expression for the energy of the electromagnetic field is considered to be given
by
X h̄ ωk †
ĤPhoton = âk,α âk,α + âk,α â†k,α (2220)
2
k,α
which is the sum of the vacuum energy (the zero-point energies) and the ener-
gies of each excited photon. The excitation energies are positive. If it had been
assumed that the photon wave functions were anti-symmetric under the inter-
change of particles, then one would have found that the photon energies would
have been identically equal to zero. Furthermore, the excited photons would
have carried zero momentum and, therefore, be completely void of any physi-
cal consequence. Hence, one concludes that spin-one photons must obey Bose-
Einstein Statistics. The generalized theorem167 is an assertion that a non-trivial
integer spin field cannot have a anti-commutator that vanishes for space-like sep-
arations and a non-trivial odd half-integer spin field cannot have a commutator
that vanishes for space-like separations.
392
This is equivalent to assuming four independent real fields. The inner product
is defined as
Φ† Φ = Φ∗1 Φ1 + Φ∗2 Φ2 (2224)
where α(0) is an arbitrary scalar. The invariance of the Lagrangian under mul-
tiplication of the wave function by the phase factor, is equivalent to the usual
U (1) gauge invariance which has been discussed in the context of the electro-
magnetic field. The operator Û must be a unitary operator, if the norm of Φ is
conserved by the generalized gauge transformation
Φ†0 Φ0 = Φ† Û † Û Φ
= Φ† Φ (2226)
Therefore, one requires
Û † Û = Iˆ (2227)
and so Û must be a unitary operator. The operator Û is assumed to be an
arbitrary unitary matrix that acts on isospin states, that is, it acts on the two
components of Φ. Furthermore, it shall be assumed that the unitary matrix
has determinant + 1. Hence, the Lagrangian is assumed to be invariant under
a set of SU (2) gauge transformations. A general transformation of SU (2) is
generated by the three operators
(1) 0 1
τ =
1 0
(2) 0 −i
τ =
i 0
(3) 1 0
τ = (2228)
0 −1
where these matrices generate a Lie algebra. That is, the algebra of the com-
mutation relations is closed, since
[ τ (i) , τ (j) ] = 2 i ξ i,j,k τ (k) (2229)
where ξ i,j,k is the antisymmetric Levi-Civita symbol. An arbitrary unitary
transformation can be expressed as
X
k (k)
Û = exp − i α τ (2230)
k
393
where the αk are three real quantities. This represents an arbitrary rotation in
isospin space169 . The U (1) gauge transformation can also be represented in the
same way. Namely, the U (1) transformation can be expressed as
Û0 = exp − i α(0) τ (0) (2231)
We shall alter the Lagrangian, such that it is invariant under a gauge trans-
formation which varies from point to point in space. These are local gauge
transformations, in which the αk (x) depend on x. If the Lagrangian is to be
invariant under local gauge transformations, then one must introduce a coupling
to gauge fields Aµ . This coupling compensates for the change of the derivatives
under the gauge transformation, so that
†
µ µ
∂µ − i g Aµ Φ ∂ − igA Φ
†
= ∂µ − i g A0µ Φ0 ∂ µ − i g Aµ0 Φ0 (2235)
169 We shall not stop and contemplate the question of what restricts our measurements have to
be quantized along the isospin z-direction, and shall not ponder why there is a super-selection
rule at work.
394
Since
Φ = Û † Φ0 (2236)
we require that
∂ µ − i g Aµ0 Φ0 = Û ∂ µ − i g Aµ Φ (2237)
where the derivative only acts on the unitary transformation. Since the Û are
generated by τ (k) , there must be four components of Aµ , i.e. the fields have
four components Aµ,k . The matrix form of Aµ is given by
3
X
Aµ = Aµ,k τ (k)
k=1
Aµ,(3) Aµ,(1) − i Aµ,(2)
= µ,(1) (2239)
A + i Aµ,(2) − Aµ,(3)
We shall identify the contravariant derivative for the massive scalar particles
as170 as
Dµ = ∂ µ − i g Aµ − i g 0 Aµ0 (2240)
and one recognizes that this has the same form as the coupling of charged
particles to the EM field. In that case, the coupling occurs solely via τ (0) , the
coupling constant is given by g 0 = h̄q c and the field Aµ(0) = Aµ is the
four-vector potential. Since τ (0) commutes with all isospin operators, it is not
necessary to consider g 0 to be identical with the g value for the SU (2) gauge
fields.
relativity, if one follows the logic adopted by Weyl and considers GR as a gauge field theory.
395
where D is the covariant derivative only involving the SU (2) triplet of gauge
fields. It should be noted that since the gauge fields do not commute, this
involves terms which are second-order in the field amplitudes. That is
F µ,ν = ∂ µ Aν − ∂ ν Aµ − i g Aµ Aν − Aν Aµ (2242)
(2243)
where the indices i and j are summed over and ξ i,j,k is the Levi-Civita symbol.
In arriving at the above expression, we have used the identity
3
X
τ (i) τ (j) = δ i,j τ (0) + i ξ i,j,k τ (k) (2244)
k=1
as expected for an electromagnetic field. Since the SU (2) gauge fields don’t
commute, the field theory is a non-Abelian gauge field theory. Under an SU (2)
transformation, the field tensors transform according to
F µ,ν → F µ,ν 0 = Û F µ,ν Û † (2246)
which is just a local unitary transform in isospin space. The Lagrangian density
for all the free gauge fields can be expressed as
1
Lgauge = − Trace F µ,ν Fµ,ν (2247)
32 π
where the Trace is evaluated in isospin space and takes into account that there
are a total of four fields. The Lagrangian density can be expressed directly in
terms of the contributions from four components of the field. The result can be
expressed as
3
1 X k
Lgauge = − Fµ,ν F µ,ν,k (2248)
16 π
k=0
396
where we have decomposed the fields as
X
k
Fµ,ν = Fµ,ν τ (k) (2249)
k
evaluated the product of the Pauli spin matrices and used the fact that the
Pauli spin matrices τ (k) for k 6= 0 are traceless.
One can consider the k-components of the vector potential Aµ (i.e. the three
real components Aµk for fixed µ) as forming three-vectors Aµ in isospin space.
These quantities transform as three-vectors under transformations in isospin
space, and also the Aµ transform as four-vectors under Lorentz transformations
in Minkowsky space-time. The three-vector fields are spin-one bosons with
isospin one. Hence, we might expect that the isospin triplet should contain two
oppositely charged particles and one uncharged particle. These particles are
supplemented by the particle corresponding to the single uncharged field Aµ(0) .
In terms of this set of isospin vectors, the free gauge field Lagrangian density
can be written in the form of a sum of a scalar product in isospin space and an
isospin scalar
1 µ ν ν µ µ ν
Lgauge = − (∂ A − ∂ A ) + 2g A ∧ A . (∂µ Aν − ∂ν Aµ ) + 2g Aµ ∧ Aν
16 π
1 µ ν ν µ
− ∂ A(0) − ∂ A(0) ∂µ Aν,(0) − ∂ν Aµ,(0) (2250)
16 π
It should be noted that the Lagrangian reduces to the sum of four non-interacting
electromagnetic Lagrangians in the limit g → 0. However, at finite values of
g, the Lagrangian density contains cubic and quartic interactions with coupling
strengths that are fixed by gauge invariance in terms of the single gauge param-
eter g.
Figure 71: The interaction vertices representing the interaction of three and
four isospin triplet gauge field bosons.
Exercise:
397
Determine the equations of motion for the vector gauge fields, in the presence
of a source term
1
Lint = − Trace ( Aµ . j µ ) (2251)
c
where the current source j µ has also been decomposed in terms of Pauli spin
matrices.
In terms of these new combinations, the free Lagrangian for the gauge fields
become
1 1 1
Lgauge = − (0)
Fµ,ν F µ,ν,(0) − (3)
Fµ,ν F µ,ν,(3) − F − F µ,ν,+ (2256)
16 π 16 π 8 π µ,ν
where the first two terms are recognized as being similar to the Lagrangian den-
sity for the electromagnetic field. It was first hypothesized by Sheldon Glashow
that the electro-weak interaction is produced by the massless vector bosons de-
scribed by the above Lagrangian171 . Masses for the gauge bosons should not
be added by hand, since the resulting theory would not be renormalizable. To
retain renormalizability of the theory, and to have massive vector bosons, we
need to break the symmetry.
398
be modified, as will be the excitations of the gauge fields. Due to the symmetry-
breaking of the scalar field, the U (1) vector gauge field will become coupled to
the triplet of SU (2) gauge fields. When the symmetry is broken, the elementary
excitations of the coupled system of fields change and these new excitations will
represent the observable particles.
The symmetry is broken by assuming that the physical ground state corre-
sponds to one specific choice of the uniform field Φ. Given the specific ground
state which the system chooses spontaneously, one can make use of the global
gauge invariance to describe the ground state Φ0 as a field which has one non-
zero component which is real. That is, αk can be chosen so that
<e Φ1
Φ0 =
0
φ0
= (2259)
0
The excited states can be expressed as
φ0 + χ1
Φ = (2260)
0
where the local gauge degrees of freedom have been used to make χ1 real. This
excited field is invariant under the transformation
Φ → Φ0 = ÛEM Φ (2261)
where ÛEM is restricted to have the form
1 0
ÛEM = (2262)
0 exp − i Λ
399
and will turn out to represent the residual U (1) gauge invariance of the electro-
magnetic field.
The Lagrangian density for the isospin doublet of scalar fields and their
couplings can be evaluated for the excited state as
2
mc
Lscalar = (Dµ Φ)† Dµ Φ − χ21 (2264)
h̄
Aµ,(3) ( φ0 + χ1 )
µ µ,(0)
µ ∂ χ1 A ( φ0 + χ1 ) √
D Φ = − i g0 −ig
0 0 2 Aµ,− ( φ0 + χ1 )
(2265)
A new interaction strength λ can be defined as
q
λ = g02 + g 2 (2266)
g0 = λ cos θ
g = λ sin θ (2268)
Thus, the covariant derivative has the connection with the field
The field AµZ will turn out to be the field that describes the neutral Z particle.
The field orthogonal to the Z field is defined as
When expressed in terms of the transformed fields and constants, the covariant
derivative terms become
µ µ
Dµ Φ =
∂ χ1
− i ( φ0 + χ1 ) √λ AZµ,− (2271)
0 g 2A
The lowest-order terms in the Lagrangian density of the non-uniform scalar field
and all its couplings to the gauge fields are expressed as
2
mc
Lscalar = ∂µ χ1 ∂ µ χ1 − χ21 + λ2 φ20 AµZ Aµ,Z + 2 g 2 φ20 Aµ+ A− µ
h̄
(2272)
400
The higher-order terms, which have been neglected, describe the self-interactions
between the scalar field and the residual interactions between the scalar field
and the gauge fields.
µ,ν
F(3) = sin θ FZµ,ν + cos θ FEM
µ,ν
+ 2 i g ( Aµ− Aν+ − Aν− Aµ+ )
(2276)
The Lagrangian density describing the small amplitude excitations of the scalar
field and the gauge fields can be written as
2
µ mc
LF ree = ∂µ χ1 ∂ χ1 − χ21
h̄
1
− Fµ,ν,Z FZµ,ν + λ2 φ20 AµZ Aµ,Z
16 π
1 µ,ν
− Fµ,ν,EM FEM
16 π
1
− F − F µ,ν,+ + 2 g 2 φ20 Aµ+ A− (2277)
8 π µ,ν µ
In electro-weak theory, the first term represents the free uncharged scalar
boson. The second term describes an uncharged vector particle with mass MZ
proportional λ φ0 . The third term describes the uncharged massless vector
particle known as the photon. From the equations of motion for Aµ,± , the
remaining term can be shown to describe a pair of charged particles with masses
MW proportional to g φ0 = sin θ λ φ0 . These particles are known as the W +
401
and W − particles. The W + and W − particles are charged and the observed
charges are ± e. The interaction mediated by the massive vector bosons is found
to have a finite range (≈ 10−18 m), and is responsible for the weak interaction.
The experimentally determined masses173 are MW c2 ≈ 80.33 GeV and MZ c2 ≈
91.187 GeV. Nearly all the parameters of this theory have been determined
through experiment, the only exception is the mass m of the scalar particle
which remains to be discovered. The ratio of the masses determines the angle
θ via174
MW
= sin θ (2278)
MZ
which yields sin θ ≈ 0.8810. The W ± particles carry electrical charges ±e since
they couple to the electromagnetic field. This can be seen by examining the
W ± field tensor
F±µ,ν = ( ∂ µ ∓ 2 i g Aµ,(3) ) Aν± − ( ∂ ν ∓ 2 i g Aν,(3) ) Aµ± (2279)
where
Aµ,(3) = sin θ AµZ + cos θ AµEM (2280)
Therefore, the covariant derivatives of the fields Dµ Aν± couple them to the elec-
tromagnetic field AµEM with either a positive or negative coupling constant of
magnitude 2 g cos θ. Since only electrically charged particles couple to the
electromagnetic field, one can make the identification
e
= 2 g cos θ = λ sin 2θ (2281)
h̄ c
which determines the coupling strengths. Furthermore, because the coupling
strengths have been completely determined, the observed masses can be used
to determine φ0 . This leads to the identification
sin2 2θ ( MZ c2 )2
φ20 = e2
(2282)
8 π h̄ c ( h̄c )
√
which leads to φ0 ≈ 178 GeV / h̄c, where h̄c ≈ 197 MeV fm. Hence, the
only undetermined parameter is the mass of the Higgs particle m. Recent ex-
periments175 have found a narrow resonance with an energy of approximately
126 GeV. The resonance has properties consistent with those expected of the
Higgs particle. The resonance decays either into two photons or two vector
bosons. Yang Mills theories, even if symmetry is spontaneously broken as in the
Weinberg-Salam theory, were shown to be renormalizable by G. t’ Hooft176 .
173 G. Arnison, A. Astbury, B. Aubert, et al., Phys. Lett. B, 122 103-116 (1983).
402