The quantized Maxwell field provided us already
with an example of a relativistic quantum field theory. On the
other hand, the description of relativistic charged particles
requires Klein-Gordon fields for scalar particles and Dirac fields
for fermions. Relativistic fields are apparently relevant for high
energy physics. However, relativistic effects are also important in
photon-matter interactions, spectroscopy, spin dynamics, and for
the generation of brilliant photon beams from ultra-relativistic
electrons in synchrotrons. Quasirelativistic effects from linear
dispersion relations in materials like
Graphene and in Dirac semimetals have also reinvigorated the need
to reconsider the role of Dirac and Weyl equations in materials
science. In applications to materials with quasirelativistic
dispersion relations c and
m become effective velocity
and mass parameters to describe cones or hyperboloids in regions of
space.
We start our discussion of relativistic matter
fields with the simpler Klein-Gordon equation and then move on to
the more widely applicable Dirac equation. We will also discuss
covariant quantization of photons, since this is more convenient
for the calculation of basic scattering events than quantization in
Coulomb gauge.
21.1 The Klein-Gordon equation
A limitation of the Schrödinger equation in the
framework of ordinary quantum mechanics is its lack of covariance
under Lorentz transformations1. On the other hand, we have
encountered an example of a relativistic wave equation in
Chapter 18, viz. the inhomogeneous Maxwell equation
This equation is manifestly covariant (or rather, form invariant)
under Lorentz transformations because it is composed of quantities
with simple tensorial transformation behavior under Lorentz
transformations, and it relates a 4-vector ∂ μ F μ ν to a 4-vector j ν , such that the equation holds
in this form in every inertial reference frame.
Another, simple reasoning to come up with a
relativistic wave equation goes as follows. We know that the
standard Schrödinger equation for a free massive particle arises
from the non-relativistic energy-momentum dispersion relation
upon
substitution of the classical energy-momentum vector through
differential operators, .
Following the same procedure in the relativistic dispersion
relation
yields the free Klein-Gordon equation2
(21.1)
Furthermore, the gauge principle or minimal
coupling prescription
yields the coupling of the charged Klein-Gordon field to
electromagnetic potentials,
(21.2)
Complex conjugation of
equation (21.2) leads to the Klein-Gordon equation for a
scalar field with charge − q. Therefore the charge conjugate Klein-Gordon field is
simply gotten by complex conjugation,
(21.3)
The Klein-Gordon field is relevant in particle
physics. E.g. π-mesons are
described by Klein-Gordon fields as soon as their kinetic energy
becomes comparable to their mass MeV, when relativistic effects
have to be taken into account. Another important application of the
Klein-Gordon field is the Higgs field for electroweak symmetry
breaking in the Standard Model of particle physics.
The Klein-Gordon field also provides a simple
introduction into the relativistic quantum mechanics of charged
particles. Therefore it is also useful as a preparation for the
study of the Dirac field. We will focus in particular on the
canonical quantization of freely evolving Klein-Gordon fields,
since this describes Klein-Gordon operators in the practically
relevant interaction picture representation. Conservation laws for
full scalar quantum electrodynamics are discussed in
Problems 21.6a and 21.7.
Mode expansion and quantization
of the Klein-Gordon field
Fourier transformation of
equation (21.1) yields the general solution of the free
Klein-Gordon equation in space,
where is just the k space expression for the relativistic
dispersion relation,
Frequency-time Fourier transformation (5.12) yields
and the general free Klein-Gordon wave function in space is
(21.4)
(21.5)
For the inversion of the Fourier transformation
in the sense of solving for and we need
equation (21.5) and
Inversion of both equations yields
Here the alternating derivative is defined as
(21.6)
(21.7)
Substituting (21.6, 21.7) back into (21.5) and formal exchange
of integrations yields
with the time evolution kernel for free scalar fields,
This distribution satisfies the initial value problem
(21.8)
(21.9)
For canonical quantization we need the Lagrange
density for the complex Klein-Gordon field
or the real Klein-Gordon field
In the following we will continue with the discussion of the
complex Klein-Gordon field.
(21.10)
(21.11)
Canonical quantization proceeds from
(21.10)
without any problems. The conjugate momenta
yield the canonical commutation relations in space,
and in space,
The Lagrangian for interacting Klein-Gordon and
Maxwell fields is
(21.12)
The charge operator of the Klein-Gordon
field
The Klein-Gordon Lagrangian (21.10) is invariant under
phase transformations
According to Section 16.2 this implies a local
conservation law (16.13) for a conserved charge
Q. After cancelling the
superfluous factor α, the
charge following from (16.14) is (after normal ordering of
the integrand in space, see the remarks following
equations (18.40, 18.41))
(21.13)
The charge density
is not positive definite, and therefore division of the charge
density by q does not yield
a probability density for the location of a particle, contrary to
the Schrödinger field. Lack of a single particle interpretation is
a generic property of relativistic fields which we had also
encountered for the Maxwell field.
Hamiltonian and momentum operators
for the Klein-Gordon field
The invariance of the Klein-Gordon Lagrangian
(21.10) under
constant translations
implies a local conservation law (16.15)
with corresponding conserved Hamilton and momentum operators
(16.17). This yields the following
expressions for energy and momentum of Klein-Gordon fields,
(21.14)
(21.15)
(21.16)
(21.17)
The commutation relations and the charge operator
(21.13), the
Hamilton operator (21.15), and the momentum operators
(21.17) imply
that the operator creates a particle of
momentum , energy and charge
q, while creates a particle of
momentum , energy and charge −
q.
The operators (21.5) and ,
are the field operators in the Dirac picture, or the free field
operators in the Heisenberg picture. They satisfy the Heisenberg
evolution equations
with the free Hamiltonian (21.15). The corresponding integrals follow in
the standard way,
etc. In the Schrödinger picture theory, this amounts to operators
, , and time evolution of the
states
with the free Hamiltonian (21.15) for free states or a corresponding
minimally coupled Hamiltonian which follows from (21.12) for interacting
states, see Problem 21.5. This is the statement that we have
Heisenberg and Schrödinger type evolution equations also in
relativistic quantum field theory.
The Klein-Gordon equation also follows from the
iterated Heisenberg equation,
cf. (18.46) for photons.
(21.18)
Non-relativistic limit of the Klein-Gordon
field
We have in the non-relativistic limit
and therefore in leading order also .
Suppose that the -space amplitudes and are negligibly small unless
. In this
case we can approximate equation (21.5) by
However, this expression automatically contains two fields
and
which satisfy the free Schrödinger equation, i.e. the complex
Klein-Gordon field will reduce to a Schrödinger field if the -space amplitudes also satisfy
.
Substitution of the remaining approximation
into the charge, current, energy and momentum densities of the
Klein-Gordon field yields the corresponding expressions for the
Schrödinger field,
Furthermore, the free Klein-Gordon equation (21.1) becomes with
the free Schrödinger equation
as it should, because we have already observed in the derivation of
(21.19) that
satisfies the free
Schrödinger equation.
(21.19)
(21.20)
For the non-relativistic limit of the real
Klein-Gordon field we find
but we have to include first order time derivatives of and in the evaluation of
and , and then use the
Schrödinger equation to find that remnant fast oscillation terms
proportional to reduce to
boundary terms.
21.2 Klein’s paradox
The commutation relations for the field operators
and imply that the operator
(21.5) describes both
particles and anti-particles simultaneously, and therefore the
Klein-Gordon equation cannot support a single particle
interpretation. This is also obvious from the charge operator
(21.13) and
the corresponding lack of a conserved probability density for
Klein-Gordon particles. Klein’s paradox provides a particularly
neat illustration of the failure of single particle interpretations
of relativistic wave equations.
Klein observed that using relativistic quantum
fields to describe a relativistic particle running against a
potential step yields results for the transmission and reflection
probabilities which are incompatible with a single particle
interpretation3. This
observation can be explained by pair creation in strong fields and
the fact that relativistic fields describe both particles and
anti-particles simultaneously. We will explain Klein’s paradox for
the Klein-Gordon field.
In the following we can neglect the y and z coordinates and deal only with the
x and t coordinates. We are interested in a
scalar particle of charge q
scattered off a potential step of height V > 0. The step is located at
x = 0, and can be
implemented through an electrostatic potential ,
(21.21)
A monochromatic solution without any apparent
left moving component for x > 0 is (after omission of an
irrelevant constant prefactor)
The frequency follows from the solution of the Klein-Gordon
equation in the two domains,
It has to be the same in both regions for continuity of the wave
function at x = 0.
(21.23)
(21.24)
The sign in the last equation of (21.24) depends on the
sign of . We apparently have to use the
minus sign if and only if . Note that in our solution
we always have .
Solving for κ yields
However, we have to be careful with the sign in (21.25). The group
velocity in x > 0 for
(i.e. for the negative
sign in (21.24)) is
i.e. we have to take the negative root for κ for to ensure positive
group velocity in the region x > 0. We can collect the results
for κ in the equations
(21.25)
(21.26)
The current density
is
Note that in x > 0 we
have j∕q < 0 if , in spite of the fact
of positive group velocity in the region. Since charges
q cannot move to the left
in x > 0, this means
that the negative value of j∕q in x > 0 for must correspond to
right moving charges − q.
We will see that this arises as a consequence of the generation of
anti-particles near the potential step for .
(21.27)
The junction conditions
yield
and the corresponding reflection and transmission coefficients are
(21.28)
(21.29)
(21.30)
The resulting behavior of the reflection
coefficient is summarized in Table 21.1.
Table 21.1
Reflection and transmission for different
relations between height V
of the potential step and energy of the incident particle
|
|
1 > R ≥ 0
|
0 < T ≤ 1
|
---|---|---|---|
|
k ≥ κ ≥ 0
|
0 ≤ R ≤ 1
|
1 ≥ T ≥ 0
|
|
|
R = 1
|
T = 0
|
|
0 ≥ κ ≥ −k
|
|
|
|
|
|
|
For an explanation of the unexpected result
R > 1 for , recall
that the solution for in x > 0 has κ < 0. If we write the solution as
and compare with the anti-particle contribution to the free
solution (21.5), we recognize the solution in the region
x > 0 as an
anti-particle solution with momentum and energy
This is acceptable, because the anti-particle has charge −
q and therefore experiences
a potential U = −V in the region x > 0. Further support for this
energy assignment for the anti-particles comes from the equality
for the kinetic+rest energy of the anti-particles,
We expect at least in
the non-relativistic limit for the anti-particles.
(21.31)
(21.32)
(21.33)
The anti-particles move to the right,
d(−ω)∕d(−κ) > 0, and yield a negative
particle current density due to
the opposite charge. We therefore get R > 1 and T < 0 for due to pair creation.
The generated particles move to the left because they are repelled
by the potential . They add to the
reflected particle in x < 0 to generate a formal
reflection coefficient R > 1. The anti-particles move to
the right because they can only move in the attractive potential −
V in x > 0. The movement of charges −
q to the right generates a
negative apparent transmission coefficient .
Please note that the last two lines in
Table 21.1 do not state that extremely large
potentials V ≫ 2mc
2 are less efficient for pair creation. They only state
that a potential V > 2mc 2 is particularly
efficient for generation of particle–anti-particle pairs with
energies .
The conclusion in a nutshell is that if we wish
to calculate scattering in the potential V > 2mc 2 for incident particles
with energies in the pair creation region , then the
ongoing pair creation will yield the seemingly paradoxical results
R > 1 and T = 1 − R < 0, see Figure 21.1.
Fig. 21.1
Particles of charge q experience the potential V for x > 0, while anti-particles with
charge − q experience the
potential − V. If the
potential satisfies V > 2mc 2, it can produce
particles with energy E
p , , in the
region x < 0 and
anti-particles with energy , , in
the region x > 0. This
corresponds to a kinetic+rest energy , , see
equation (21.33). Pair creation is most efficient for
Please note that a more satisfactory discussion
of energetics of the problem would also have to take into account
the dynamics of the electromagnetic field ,
and then use the Hamiltonian density (21.132) of quantum
electrodynamics with scalar matter. This would also imply an
additional energy cost for separating the oppositely charged
particles and anti-particles. The potential V would therefore decay due to pair
creation until it satisfies the condition V ≤ 2mc 2, when pair creation
would seize and the standard single particle results
0 ≤ T = 1 − R ≤ 1 apply for incident particles with
any energy, or the potential would have to be maintained through an
external energy source.
21.3 The Dirac equation
We have seen in equation (21.13) that the conserved
charge of the complex Klein-Gordon field does not yield a conserved
probability, and therefore has no single particle interpretation.
This had motivated Paul Dirac in 1928 to propose a relativistic
wave equation which is linear in the derivatives5,
Since the relativistic dispersion relation implies that the field
should
still satisfy the Klein-Gordon equation,
equation (21.34) should imply the Klein-Gordon equation.
Applying the operator
yields
This is the Klein-Gordon equation if the coefficients γ μ can be chosen to satisfy
In four dimensions, equation (21.35) has an up to
equivalence transformations unique solution in terms of (4 ×
4)-matrices (see Appendix G for the relevant proofs and for
the construction of γ
matrices in d spacetime
dimensions).
(21.34)
(21.35)
The Dirac basis for γ matrices is
where the (4 × 4)-matrices are expressed in terms of (2 ×
2)-matrices. Another often used basis is the Weyl basis:
The two bases are related by the orthogonal transformation
(21.36)
(21.37)
The Dirac equation with minimal photon coupling
follows from the Lagrange density of quantum electrodynamics,
The conserved current density for the phase invariance
is
Variation of (21.39) with respect to the vector potential
shows that j
μ appears as the
source term in Maxwell’s equations,
(21.38)
(21.39)
(21.40)
Solutions of the free Dirac equation
We temporarily set and c = 1 for the construction of the
general solution of the free Dirac equation.
We can use any representation of the γ matrices to find
i.e. the solutions of (21.41) must have the form
with
and
The normalization factors in (21.42) are included for later convenience when
we quantize the Dirac field.
(21.42)
(21.43)
(21.44)
To find the eigenspinors , , we observe
i.e. the columns of the matrix
solve equation (21.43) while the columns of the matrix
solve equation (21.44). However, only two columns of each of
the two matrices are linearly
independent.
We initially use a Dirac basis (21.36) for the
γ matrices. A suitable
basis for the general solution of the free Dirac equation is then
given by the spin basis in the Dirac representation,
where was used. The
spin labels indicate that describes spin
up or down particles, while describes spin
up or down anti-particles.
(21.45)
(21.46)
(21.47)
(21.48)
The general solution of the free Dirac equation
then has the form
where is understood:
.
(21.49)
Calculations involving 4-spinors are often
conveniently carried out with and c = 1, and restoration of the constants
is usually only done in the final results from the requirement of
correct units. For completeness I would also like to give the
general solution of the free Dirac equation with the constants
and
c restored. We can choose
the basic spinors (21.45–21.48) to have units of square roots of energy,
e.g.
and the solution (21.49) is
with .
In these conventions the Dirac field has the same dimensions
length−3∕2 as the Schrödinger field. The free field
also describes the freely evolving field operator in the interaction picture.
(21.50)
(21.51)
Some useful algebraic properties of the spinors
(21.45–21.48) are frequently used in the calculations
of cross sections and other observables,
(21.52)
(21.53)
(21.54)
(21.55)
(21.56)
The following equations contain 4 × 4 unit
matrices 1 on the right
hand sides,
It is actually clumsy to write down unit matrices when their
presence is clear from the context, and the action e.g. of the
scalar mc 2 on a
4-spinor has
the same effect as the matrix mc 2 1. Therefore we will usually adopt
the practice of not writing down 4 × 4 unit matrices
explicitly.
(21.57)
(21.58)
(21.59)
(21.60)
(21.61)
Equations (21.52) and (21.53) are used e.g. in
the inversion of the Fourier representation (21.51),
Substituting these equations back into (21.51) yields
with the time evolution kernel
This satisfies the initial value problem
It is related to the time evolution kernel (21.9) of the Klein-Gordon
field through
(21.62)
(21.63)
(21.64)
(21.65)
(21.66)
(21.67)
It is sometimes useful to express
equation (21.49) and the corresponding equation in
space in bra-ket notation, similar
to equations (18.24, 18.25) for the Maxwell field. With
the definitions
we can write the free Dirac field in the forms
and
where a ∈ { 1, …4} is a Dirac spinor index,
labels particles (+) or
anti-particles (−), and s
is the spin label. The equations (21.52) and the first
equation in (21.53) are
Equation (21.54) is
and equation (21.57) is
(21.68)
(21.69)
(21.70)
(21.71)
Charge operators and quantization of the Dirac
field
We can apply the results from
Section 16.2 to calculate the energy and
momentum operator for the Dirac field. The free Dirac Lagrangian
yields the positive definite normal ordered Hamiltonian
but only if we assume anti-commutation properties of the
d s and d s + operators.
(21.74)
(21.75)
The normal ordered momentum operator is then
(21.76)
The electromagnetic current density (21.40) yields the charge
operator
(21.77)
The normalization in equation (21.51) has been chosen
such that the quantization condition
for the components of yields
with the other anti-commutators vanishing. The
equations (21.75–21.77) then imply that the operator
creates a fermion of mass
m, momentum and charge q, while creates a particle with the
same mass and momentum, but opposite charge − q.
For an explanation of the spin labels of the
spinors , we notice that
the spin operators corresponding to the rotation generators
are both in the Dirac and in the Weyl representation given by
see Appendix H for an explanation of generators of Lorentz
boosts and rotations for Dirac spinors.
(21.78)
Equation (21.78) implies that the rest frame spinors
transform under
rotations around the z axis
as spinors with z-component
of spin .
For an explanation of the spin labels of the
spinors , we have to look
at charge conjugation. Both in the Dirac and the Weyl
representation of γ
matrices we have
Therefore complex conjugation of the Dirac equation
followed by multiplication with iγ 2 from the left yields
with the charge conjugate field
In particular, we have
and
i.e. the negative energy spinors for charge q, momentum and spin projection
correspond to positive energy spinors for charge − q, momentum and spin projection
.
(21.79)
21.4 The energy-momentum tensor for quantum electrodynamics
We use the symmetrized form of the QED Lagrangian
(21.39),
This yields according to (16.16) a conserved energy-momentum
tensor
(21.80)
According to the results of
Section 16.2, this yields on-shell conserved
charges, i.e. we can use the equations of motion to simplify this
expression. The Dirac equation then implies
We can also add the identically conserved
improvement term
where Maxwell’s equations
have been used. This yields the gauge invariant tensor
(21.81)
However, we can go one step further and replace
t μ ν with a symmetric
energy-momentum tensor. The divergence of the spinor term in
t μ ν is
where again the Dirac equation was used.
(21.82)
The symmetrization of t μ ν also involves the commutators
of γ matrices,
Since we can write a product always as a sum of an anti-commutator
and a commutator, we have
and the commutators also satisfy6
Equations (21.83–21.85) together with
imply also
Therefore the local conservation law also holds for
the symmetrized energy-momentum tensor
This yields in particular the Hamiltonian density
and the momentum density with components ,
Elimination of the time derivatives using the Dirac equation yields
(21.83)
(21.84)
(21.85)
(21.86)
(21.87)
(21.88)
(21.89)
(21.90)
The spin contribution
with the vector of 4 × 4 spin matrices
(21.78)
appears here as an additional contribution compared to the orbital
momentum density
that follows directly from the tensor (21.81). The spin term in
the momentum density (21.90) generates the spin contribution in the
total angular momentum density
from
if the symmetric energy-momentum tensor is used in the calculation
of angular momentum. This is explained in Problem 21.16c, see in particular
equations (21.147–21.150).
Energy and momentum in QED in Coulomb
gauge
In materials science it is convenient to
explicitly disentangle the contributions from Coulomb and photon
terms in Coulomb gauge .
We split the electric field components in Coulomb gauge according
to
The equation for the electrostatic potential decouples from the
vector potential in Coulomb gauge,
and is solved by
Furthermore, the two components (21.91) of the electric field are orthogonal in
Coulomb gauge,
and the contribution from to the Hamiltonian is
where the summation is over 4-spinor indices. The presentation of
the ordering of the field operators was conventionally chosen as
the correct ordering in the non-relativistic limit, cf.
(18.65), but (21.93) must actually be
normal ordered such that the particle and anti-particle creation
operators and appear leftmost in the
Coulomb term in the forms , , etc. Substituting the mode expansions
and normal ordering therefore
leads to the attractive Coulomb terms between particles and their
anti-particles.
(21.91)
(21.92)
(21.93)
The resulting Hamiltonian in Coulomb gauge
therefore has the form
This Hamiltonian yields the corresponding Dirac equation in the
Heisenberg form
if canonical anti-commutation relations are used for the spinor
field. The Coulomb gauge wave equation (18.10) with the relativistic current
density (21.40) follows in the form
if the commutation relations (18.36, 18.39) are used. This
confirms the canonical relations between Heisenberg, Schrödinger
and Dirac pictures, and the consistency of Coulomb gauge
quantization with the transverse δ function (18.27) also in the fully
relativistic theory. It also implies appearance of the Dirac
picture time evolution operator in the scattering matrix in the now
familiar form.
(21.94)
(21.95)
The momentum operator in Coulomb gauge follows
from (21.90)
and
as
where boundary terms at infinity were discarded.
(21.96)
21.5 The non-relativistic limit of the Dirac equation
The Dirac basis (21.36) for the
γ-matrices is convenient
for the non-relativistic limit. Splitting off the time dependence
due to the rest mass term
in the Dirac equation (21.38) yields the equations
This yields in the non-relativistic regime
and substitution into the equation for ψ yields Pauli’s equation7
The spin matrices for spin-1/2 Schrödinger fields are the upper
block matrices in the spin matrices (21.78) for the full Dirac
fields, ,
see also Section 8.1 and in particular
equation (8.12).
(21.97)
(21.98)
(21.99)
(21.100)
(21.101)
If the external magnetic field is approximately constant over the
extension of the wave function we can use
Substitution of the vector potential in equation (21.101) then yields the
following linear terms in in the Hamiltonian on the right hand
side,
Here is the Bohr magneton, and we
used the short hand notation
for the representation of the angular
momentum operator. Recall that this operator is actually given by
(21.102)
Equation (21.102) shows that the Dirac equation explains
the double strength magnetic coupling of spin as compared to
orbital angular momentum (often denoted as the magneto-mechanical anomaly of the
electron or the anomalous
magnetic moment of the electron). The corresponding
electromagnetic currents in the non-relativistic regime are
where
is the three-dimensional tensor with the (2 × 2)-matrix entries
(we can think of it as a (3 × 3)-matrix containing (2 × 2)-matrices
as entries). Substitution of
yields
with a spin term
However, this term does not accumulate or diminish charges in any
volume, ,
and can therefore be neglected in the calculation of electric
currents.
(21.103)
The non-relativistic approximations for the
Lagrange density , the energy density and the momentum density are
(21.104)
(21.105)
(21.106)
The Hamiltonian and momentum operators in Coulomb
gauge are
and (cf. equation (21.96))
(21.107)
(21.108)
It is interesting to note that if we write the
current density (21.103) as
we can write Ampère’s law with Maxwell’s correction term as
i.e. the “spin density”
adds a spin magnetic field to the magnetic field which is generated by
orbital currents and time-dependent electric fields
,
Higher order terms and spin-orbit
coupling
We will discuss higher order terms in the
framework of relativistic quantum mechanics, i.e. our basic quantum
operators are x and
p etc., but not quantum
fields. This also entails a semi-classical approximation for the
electromagnetic fields and potentials.
For the discussion of higher order terms, we
write the Dirac equation in Schrödinger form,
with the Hamilton operator
The operator is
and
is the a-th component of
the 4-spinor
(21.97) in
representation.
(21.109)
We continue to use the Dirac basis (21.36) of γ matrices in this section, such that
as a matrix valued vector is given by
The part of the Hamiltonian (21.109) which mixes the
upper and lower components of the 4-spinor
is
Operators which mix upper and lower 2-spinors in 4-spinors are also
denoted as odd terms in the
Hamiltonian.
We can remove the odd contribution K(t) by using the anti-hermitian operator
which implies subtraction of K(t) from the new transformed Hamiltonian
. However, we also have
to take into account that the transformed state
satisfies the equation
(21.110)
Therefore the transformed Hamiltonian is actually
We also wish to expand the Hamiltonian up to
terms of order , where contains contributions from the kinetic
energy of the particle and from its interactions with the
electromagnetic fields.
Equation (21.110) implies
The relevant commutators are
and
We don’t need to evaluate the final higher order
commutator
because this is an odd term of order , which is eliminated in
the next step through a unitary transformation, to which it
contributes in order . We only need to observe
that C odd (3)(t) contains one term proportional to
, and other terms proportional
to
such that . This will
become relevant for the elimination of C odd (3)(t) in the next step.
However, for now our transformed Hamiltonian is
The last line contains three odd contributions
which we can eliminate exactly as in the previous
step by using a unitary transformation
with
This yields a new Hamiltonian
which is in the required order
This contains again an odd piece
which is eliminated by another unitary transformation of the form
with
The resulting Hamiltonian after this transformation still contains
an odd piece
which is eliminated in a final transformation
Therefore up to terms of order , we finally
find an equation which is diagonal in upper and lower 2-spinors
with
(21.111)
(21.112)
(21.113)
The Hamiltonian acting on the upper 2-spinor is
The first three terms are again the Pauli Hamiltonian from
(21.101).
(21.114)
It is of interest to write some of the higher
order terms in the Hamiltonian (21.114) also in terms of the charge density
which generates the
electromagnetic fields.
The term
amounts to a contact interaction between the particles described by
equation (21.112) (e.g. electrons) and the particles
which generate the electromagnetic fields. This term is known as
the Darwin term. The contact interaction has the counter-intuitive
property to lower the interaction energy between like charges, but
recall that it emerged from eliminating the anti-particle
components up to terms of order . It should not
surprise us that a positronic component in electron wave functions
contributes an attractive term to the electron-electron
interaction. The Hamiltonian (21.114) is in excellent agreement with
spectroscopy if radiative corrections are also taken into account,
see e.g. [18].
(21.115)
The term
is apparently a coupling between spin
and induced potentials from time-dependent charge-current
distributions.
(21.116)
In the static case we can write the term
in (21.114)
in the form
Here
is the orbital angular momentum operator with respect to the point
, and the term
apparently contains a charge weighted sum over angular momentum
operators. The term
is therefore the origin of spin-orbit coupling. In particular, for
a radially symmetric charge distribution
one finds
This implies equation (8.20) for spin-orbit coupling in
hydrogen atoms.
(21.117)
So far we have emphasized the emergence of
terms from
the term,
and historically the coupling of spin and orbital angular momentum
had provided the initial motivation for the designation as
spin-orbit coupling term. However, the direct coupling of orbital
momentum and spin provides just as good a
reason for the name spin-orbit coupling, and another important
special case of the term
arises for a uni-directional electric field e.g. in z direction. In this case the term
takes the form
For homogeneous electric field this yields a spin-orbit coupling
term of the form
with constant α
R . This
particular form of a spin-orbit coupling term is known as a Rashba
term9. Spin-orbit
coupling was always relevant not only for atomic and molecular
spectroscopy, but also for electronic energy band structure in
materials where they are often significantly enhanced e.g. due to
low effective masses. In recent years spin-orbit coupling terms in
low-dimensional systems, and Rashba terms in particular, have also
attracted a lot of interest because of their relevance for
spintronics10.
(21.118)
21.6 Covariant quantization of the Maxwell field
We have seen in Section 18.2 how to quantize the Maxwell
field and describe photons in Coulomb gauge. This is useful if our
problem contains non-relativistic charged particles, since the
Hamiltonian in Coulomb gauge conveniently describes the
electromagnetic interaction between the charged particles through
Coulomb terms. The free interaction picture photon operators
or the
corresponding Schrödinger picture operators are then only needed
for the calculation of absorption, emission or scattering of
external photons. Exchange of virtual photons provides only small
corrections to Coulomb interactions for non-relativistic charged
particles. The relevant Hamiltonian is (21.107) with Schrödinger
fields and Coulomb terms for all the different kinds of charged
particles.
Coulomb gauge can also be used for problems
involving relativistic fermions. These can be described by the
Hamiltonian (21.94) including Dirac fields and Coulomb
interaction terms for all the different kinds of spin-1∕2 particles
in the problem. Indeed, we will calculate basic scattering
processes involving relativistic charged particles in
Sections 22.2 and 22.4 in Coulomb gauge, and the
calculations will explicitly show how the Coulomb interaction terms
between charged particles dominate over photon exchange terms if
the kinetic energies of the charged particles are small compared to
their rest energies, see in particular equation (22.29).
However, if the problem indeed contains
relativistic charged particles, then the interaction of those
particles with other charged particles is more conveniently
described through a covariant quantization of photons in Lorentz
gauge,
Suppose the potential does not satisfy the Lorentz
gauge condition. We can construct the Lorentz gauge vector
potential A
μ (x) by performing a gauge transformation
with
Here
is the retarded massless scalar Green’s function, cf.
(J.37, J.60). This also helps us to solve Maxwell’s equations
in Lorentz gauge,
in the form
where the Liénard-Wiechert potentials
solve the inhomogeneous Maxwell equations (21.122) and satisfy the
Lorentz gauge condition due to charge conservation. The remainder
A D μ (x) must therefore satisfy
(21.119)
(21.120)
(21.121)
(21.122)
(21.123)
(21.124)
(21.125)
To quantize this, we observe that Maxwell’s
equations in Lorentz gauge follow from the Lagrange density of
electromagnetic fields (18.1) if we take into account the
Lorentz gauge condition,
This yields canonically conjugate momentum fields for all
components of the vector potential,
The principles of canonical quantization and the Lorentz gauge
condition then motivate the following quantization condition for
electromagnetic potentials in Lorentz gauge (with ),
(21.126)
The general solution of (21.125)
satisfies the quantization condition if
and the other commutators vanish.
(21.127)
(21.128)
A possible choice for the polarization vectors
is e.g.
to choose and
as spatial
orthonormal vectors without time-like components and perpendicular
to such that
and choose
This formalism can be motivated as a limiting case of the
quantization of massive vector fields, and it has the advantage of
faster and easier calculation of scattering amplitudes involving
electromagnetic interactions of relativistic charged particles,
because there are no separate amplitudes for photon exchange and
Coulomb interactions, which need to be added to give the full
scattering amplitude. The spatially longitudinal photons generated
by , ,
incorporate the contributions from the Coulomb interactions11. Why then don’t we see
photon states ?
These photon states are actually spurious gauge degrees of freedom.
We could perform another gauge transformation
with
which takes us right back to Coulomb gauge,
without any freely oscillating time-like component. Since
was the only photon state with a time-like component, (21.129) has removed that
photon state. We can think of the photons with longitudinal spatial
components and corresponding time-like components as virtual place
holders for the Coulomb interaction.
(21.129)
(21.130)
21.7 Problems
21.1. Show that for an appropriate class
of integration contours in the complex k 0 plane the scalar
propagator (21.9) can be written in the form
21.2. We have discussed the
non-relativistic limit of the Klein-Gordon field in the case
.
However, there must also exist a non-relativistic limit for the
anti-particles. How does the non-relativistic limit work in the
case of negligible particle amplitude ?
21.3. Derive the energy density
and the momentum density for the real Klein-Gordon
field.
21.4. Calculate the non-relativistic
limits for the Hamilton operator H and the momentum operator
of the real Klein-Gordon
field.
21.5. Derive the energy-momentum tensor
for QED with scalar matter (21.2),
The corresponding densities of energy, momentum, and energy current
are
(21.131)
(21.132)
(21.133)
Solution.
The Lagrange density for quantum electrodynamics
with scalar matter is
This yields according to (16.16) a conserved energy-momentum
tensor
To find a gauge invariant energy-momentum tensor
we add the identically conserved improvement term
where Maxwell’s equations
were used. This yields the gauge invariant tensor (21.131) from
21.6. If we write the solution
(21.5) of the
free Klein-Gordon equation as the sum of the positive and negative
energy components,
the charge densities
are separately conserved, and therefore we can also identify
conserved particle and anti-particle numbers
21.6b. Why is it not possible to derive
separately conserved (anti-)particle numbers N ± for the scalar particles
in the interacting theory (21.2)?
21.7. Show that contrary to the case of
spinor electrodynamics, it is not possible in relativistic scalar
electrodynamics to derive a Coulomb gauge Hamiltonian (although
Coulomb gauge can still be imposed on the Maxwell field, of
course).
Why can we nevertheless find a Coulomb gauge
Hamiltonian in the non-relativistic limit for the scalar
fields?
Hint: The Gauss law of scalar electrodynamics in
Coulomb gauge takes the form
21.8. Show that the junction conditions
(21.28) are
necessary and sufficient to ensure that the Klein-Gordon equation
holds at the step of the potential.
21.9. Generalize the reasoning from
Section 21.2 to the case of oblique incidence against
the potential step, e.g. by considering a scalar boson running
against the potential (21.21) with initial momentum components
and .
Remarks on the Solution.
The ansatz for the Klein-Gordon wave
function which complies with the boundary conditions on the
incoming particle and the requirement of smoothness for all times
t and values of
y along the interface
x = 0 is
The frequency follows again from the solution of the Klein-Gordon
equation in the two domains,
All other pertinent results follow also exactly as in
Section 21.2 if we make the substitutions k → k x , κ → κ x and .
This applies in particular also to Table 21.1 and
Figure 21.1. In particular, we have generation of pairs
of particles and anti-particles in the energy range
if the height of the potential step satisfies
The wave number κ
x in this energy
range is
and writing the solution for x > 0 as
shows that it is an anti-particle solution with energy and momentum
components , . The kinetic+rest energy of the
generated anti-particles in the region x > 0 is
and the energy of the anti-particles is just the sum of their
kinetic+rest energy and their potential energy, .
(21.134)
(21.135)
21.10. Calculate the boson number
operator
for the free Klein-Gordon field in representation.
21.11. Show that scattering of a
Klein-Gordon field off a hard sphere yields the same result
(11.36) as the non-relativistic
Schrödinger theory, except that the definition (where E is the kinetic energy of the
scattered particle) is replaced by
The hard sphere is taken into account through a boundary condition
of vanishing Klein-Gordon field on the surface of the sphere, like
the condition on the Schrödinger wave function in
Section 11.3, i.e. we do not model it as a
potential. We could think of the hard sphere in this case as
arising from a hypothetical interaction which repels particles and
anti-particles alike (just like gravity is equally attractive for
particles and anti-particles).
21.12. You could also model an
impenetrable wall for a Klein-Gordon field in the manner of the
hard sphere of Problem 11. Which wave function for the Klein-Gordon
field do you get if the impenetrable wall prevents the field from
entering the region x > 0? Why does this result not
contradict the Klein paradox?
21.13. Calculate the fermion number
operator
for the free Dirac field in representation.
21.14. Calculate the reflection and
transmission coefficients for a Dirac field of charge q in the presence of a potential step
. How do your
results compare with the results for the Klein-Gordon field in
Section 21.2?
21.15a. What are the non-relativistic
limits of the spinor plane waves (21.72)?
21.15b. Verify the relations
(21.73) for
the relativistic spinor plane wave states.
21.16. Angular momentum in relativistic
field theory
21.16a. Show that if T μ ν is a symmetric conserved
energy momentum tensor, then the currents
are also conserved:
(21.136)
(21.137)
21.16b. The quantities have the
properties
and are therefore associated with angular momentum conservation and
conservation of the center of energy motion (18.128, 18.129)
in relativistic field theories. Show that invariance of the
relativistic field theory
under rotations and Lorentz boosts
yields the conservations laws (21.137) from the results of
Section 16.2 if proper improvement terms are
added. The Lorentz generators S α β in the spinor representation
are defined in equation (H.12).
(21.138)
(21.139)
(21.140)
(21.141)
21.16c. We have seen in the previous
problem that invariance of the relativistic theory (21.138) under the
rotations
yields densities of conserved charges which we can express in vector
form through ,
,
i.e. .
On the other hand, we have seen in Problem 16.6 that the total
angular momentum density of non-relativistic
fermions contains a spin term which is not proportional to any space-time
coordinates x
α , and yet we
have also seen in Problem 16.7 that only the combination of both terms in
(16.26) yields the density of a
conserved quantity in the presence of spin-orbit coupling. How can
that be?
Replace time derivatives on spinor fields in the
momentum density using the Dirac equation. This yields spin
contributions to the momentum density. Show that partial
integration of the resulting spin contributions to the angular
momentum density yields spin terms which reduce to the spin term in
equation (16.26) in the non-relativistic
limit.
Solution for 16b.
The electric current density for the Lagrange
density (21.138) is
Addition of the identically conserved improvement term
to the conserved current (16.13) for the transformation
(21.139–21.141) yields the gauge invariant conserved
current
(21.142)
(21.143)
The divergence of the spinor contributions to
this current density is
The relation (21.85) implies that on-shell
and comparison of (21.82) and (21.86) implies that we can write the remaining
part of in the form
The conserved current (21.144) is therefore equivalent to the
conserved current
with the symmetric stress-energy tensor for the Lagrange density
(21.138)
(cf. (21.87, 21.131))
(21.144)
(21.145)
(21.146)
Solution for 16c.
We discuss the angular momentum densities in
vector form, ,
with the momentum densities (cf. (21.89, 21.133)),
The γ matrices satisfy
with the vector of 4 × 4 spin matrices ,
cf. (21.78).
The Dirac equation then implies
and we can write the momentum density in the form
The spin term
in the momentum density contributes a term to the angular momentum
of the form
such that we can write the total angular momentum density also in
the form
with a spin contribution ,
and an orbital angular momentum
with the orbital momentum density:
(21.147)
(21.148)
(21.149)
(21.150)
(21.151)
21.17.
New charges from local phase invariance?
We have derived expressions for charge and
current densities from phase invariance
of Lagrange densities, see e.g. (16.31, 16.32) for the charge
and current densities of non-relativistic charged matter fields,
and (21.142)
for relativistic charged matter fields. In the final expressions we
always divided out the irrelevant constant parameter .
However, introduction of the electromagnetic potentials rendered
the Lagrange densities invariant under local phase transformations
In this case we cannot discard the phase parameter from the current densities for the
local symmetry. Does this provide us with additional useful notions
of conserved charges for quantum electronics and quantum
electrodynamics?
21.17a. Show that application of the
result (16.13) to local phase
transformations yields current densities
where j μ are the current densities which
were derived for constant phase parameter ,
e.g. (16.31, 16.32) or (21.142).
(21.154)
21.17b. Show in particular that the
charge density can be written in
the form
and that the current density is
Apply these results to a static charge distribution .
The charge
is only conserved if charges do not escape or enter at
:
Show that this implies
Show also that differs from the standard electric
charge
only by a constant factor
where
is the angular average of the phase parameter at .
(21.156)
(21.157)
(21.158)
21.18. Show for a free electron that
equation (21.100) implies that a positron component
ϕ in the wave function is
not negligible any more relative to the electron wave function
ψ at a distance of order
This implies that we cannot use the wave packet for a strongly
localized free electron with beyond a distance of
about 0. 1 μm from the center. However, for a free electron wave
packet with mm the limit (21.159) is much larger
than the confines of any physics or chemistry lab and therefore of
no concern.
(21.159)
Show also that at time t, the estimate for the usable range of
the wave packet is
(21.160)
21.20. We have derived the
equations (12.21, 12.23) for particles which satisfy the
relativistic dispersion equation. In the meantime, we have seen
that at the quantum level these particles are described by scalar
fields ϕ, spinors
ψ, or vector fields
A μ . The factor g counts spin and internal symmetry
degrees of freedom and has the form , where
is the dimension of the
representation of the internal symmetry group G under which the fields
transform.
21.20a. Scalar fields have g s = 1. Show that in d + 1 space-time dimensions,
g s = 2⌊(d−1)∕2⌋ for Dirac fields and
g s = d − 1 for vector fields.
Hint: A Dirac spinor in d + 1 space-time dimensions has
2⌊(d+1)∕2⌋
components, see Appendix G (note that there d denotes the number of space-time
dimensions).
21.20b. We have seen that scalar fields
can be either real or complex (and similar remarks apply to spinor
and vector fields if we go beyond quantum electrodynamics into the
standard model of particle physics). However, a complex field has
twice as many degrees of freedom as a real field. Should
g therefore not include an
additional factor g
c with
g c = 2 for complex fields and
g c = 1 for real fields?
21.21. Formulate the basic relations for
basis kets , for the potentials
in Lorentz gauge in analogy to
the corresponding relations (18.24–18.27) in Coulomb gauge.
21.22. Show that the two representations
given in equation (21.130) for the gauge transformation function
are indeed equivalent (hint:
use the fact that A
μ (x) satisfies the Gauss law ).
Bibliography
18.
C. Itzykson, J.-B. Zuber,
Quantum Field Theory
(McGraw-Hill, New York, 1980)
Footnotes
1
However, we will see that in the second quantized
formalism in the Heisenberg and Dirac pictures, the time evolution
of the field operators is given by Heisenberg equations of motion,
and the corresponding time evolution of states in the Schrödinger
and Dirac pictures is given by corresponding Schrödinger equations
with relativistic Hamiltonians.
2
E. Schrödinger, Annalen Phys. 386, 109 (1926); W.
Gordon, Z. Phys. 40, 117 (1926); O. Klein, Z. Phys. 41, 407
(1927).
3
O. Klein, Z. Phys. 53, 157 (1929). Klein actually
discussed reflection and transmission of relativistic spin 1/2
fermions which are described by the Dirac
equation (21.38).
4
We cannot try to discuss motion of particles of
mass m in the presence of a
potential by simply including a scalar potential term in the form
in the Klein-Gordon equation. This would correspond to a local mass
rather than to a local potential, and yield tachyons in
x > 0 for .
5
P.A.M. Dirac, Proc. Roy. Soc. London A 117, 610
(1928). Dirac’s relativistic wave equation was a great success, but
like every relativistic wave equation, it also does not yield a
single particle interpretation. It immediately proved itself by
explaining the anomalous magnetic moment of the electron and the
fine structure of spectral lines, and by predicting
positrons.
6
The commutators S μ ν provide the spinor
representation of the generators of Lorentz transformations.
Furthermore, equation (21.85) is the invariance of the γ matrices under Lorentz
transformations, see Appendix H.
7
W. Pauli, Z. Phys. 43, 601 (1927). Pauli actually
only studied the time-independent Schrödinger equation with the
Pauli term in the Hamiltonian, and although he mentions Schrödinger
in the beginning, he seems to be more comfortable with Heisenberg’s
matrix mechanics in the paper.
9
E.I. Rashba, Sov. Phys. Solid State 2, 1109
(1960); Yu.A. Bychkov, E.I. Rashba, JETP Lett. 39, 78 (1984); J.
Phys. C 17, 6039 (1984).
10
See e.g. J. Nitta, T. Akazaki, H. Takayanagi, T.
Enoki, Phys. Rev. Lett. 78, 1335 (1997); D. Grundler, Phys. Rev.
Lett. 84, 6074 (2000); J. Sinova et al., Phys. Rev. Lett. 92, 126603
(2004); E.Y. Sherman, D.J. Lockwood, Phys. Rev. B 72, 125340
(2005); K.C. Hall et al.,
Appl. Phys. Lett. 86, 202114 (2005); P. Pietiläinen, T.
Chakraborty, Phys. Rev. B 73, 155315 (2006); E. Cappelluti, C.
Grimaldi, F. Marsiglio, Phys. Rev. Lett. 98, 167002 (2007).
11
Of course, this implies that one cannot naively
invoke Hamiltonians with Coulomb interaction terms if we describe
photons in Lorentz gauge. Otherwise we would overcount
interactions. Remember that the Coulomb interaction terms came from
the contributions to Hamiltonians from electromagnetic fields
in Coulomb gauge, see
Section 21.4.
12
A current J μ is sometimes denoted as
strongly conserved if the
local conservation law ∂
μ J μ = 0 is an identity.