External-Revised 65575757
External-Revised 65575757
Jan Dereziński
Dept. of Math. Methods in Phys.,
Faculty of Physics, University of Warsaw
Hoza 74, 00-682 Warszawa, Poland
email Jan.Derezinski@fuw.edu.pl
January 14, 2020
Abstract
The main purpose of these notes is a review of various models of Quan-
tum Field Theory involving quadratic Lagrangians. We discuss scalar and
vector bosons, spin 12 fermions, both neutral and charged. Beside free the-
ories, we study their interactions with classical perturbations, called, de-
pending on the context, an external linear source, mass-like term, current
or electromagnetic potential. The notes may serve as a first introduction
to QFT.
Contents
0 Introduction 7
1 Basic concepts 11
1.1 Minkowski space . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.1.1 Coordinates in Minkowski space . . . . . . . . . . . . . . 11
1.1.2 Causal structure . . . . . . . . . . . . . . . . . . . . . . . 12
1.1.3 Invariant measure . . . . . . . . . . . . . . . . . . . . . . 13
1.1.4 Fourier transform . . . . . . . . . . . . . . . . . . . . . . . 13
1.1.5 Lorentz and Poincaré groups . . . . . . . . . . . . . . . . 14
1.1.6 Double coverings of Lorentz and Poincaré groups . . . . . 15
1.1.7 Finite dimensional representations of the Lorentz group . 17
1.2 Symplectic dynamics and its quantization . . . . . . . . . . . . . 18
1.2.1 Symplectic form vs Poisson bracket in classical mechanics 18
1.3 Darboux coordinates and symplectic vector spaces . . . . . . . . 19
1.4 General concepts of quantum field theory . . . . . . . . . . . . . 19
1
1.4.1 Quantum mechanics . . . . . . . . . . . . . . . . . . . . . 19
1.4.2 Time reversal . . . . . . . . . . . . . . . . . . . . . . . . . 20
1.4.3 Relativistic quantum mechanics . . . . . . . . . . . . . . . 21
1.4.4 Haag-Kastler axioms for observable algebras . . . . . . . . 23
1.4.5 Haag-Kastler axioms for field algebras . . . . . . . . . . . 23
1.4.6 Global symmetries . . . . . . . . . . . . . . . . . . . . . . 24
1.4.7 Neutral quantum fields . . . . . . . . . . . . . . . . . . . . 25
1.4.8 Wightman axioms for neutral fields . . . . . . . . . . . . . 25
1.4.9 Relationship between Haag-Kastler and Wightman axioms 26
1.4.10 Global symmetries in the Wightman formalism . . . . . . 27
1.4.11 Charged fields . . . . . . . . . . . . . . . . . . . . . . . . 27
1.4.12 Wightman axioms for neutral and charged fields . . . . . 28
1.4.13 U (1) symmetry . . . . . . . . . . . . . . . . . . . . . . . . 29
1.4.14 Charge conjugation . . . . . . . . . . . . . . . . . . . . . . 30
1.4.15 Parity invariance . . . . . . . . . . . . . . . . . . . . . . . 30
1.4.16 Time reversal invariance . . . . . . . . . . . . . . . . . . . 31
1.4.17 The CPT Theorem . . . . . . . . . . . . . . . . . . . . . . 31
1.4.18 The CPT Theorem in a P and T -invariant theory . . . . 32
1.4.19 N -point Wightman and Green’s functions . . . . . . . . . 33
1.5 General scattering theory . . . . . . . . . . . . . . . . . . . . . . 34
1.5.1 Time ordered exponential . . . . . . . . . . . . . . . . . . 34
1.5.2 Schrödinger and Heisenberg picture . . . . . . . . . . . . 35
1.5.3 Schrödinger and Heisenberg picture for time-dependent
Hamiltonians . . . . . . . . . . . . . . . . . . . . . . . . . 35
1.5.4 Time-dependent perturbations . . . . . . . . . . . . . . . 37
1.5.5 Vacuum energy . . . . . . . . . . . . . . . . . . . . . . . . 38
1.5.6 Time-ordered Green’s functions . . . . . . . . . . . . . . . 38
1.5.7 Adiabatic switching and the energy shift . . . . . . . . . . 39
1.5.8 Adiabatic switching and Green’s functions . . . . . . . . . 42
1.5.9 Adiabatic scatttering theory . . . . . . . . . . . . . . . . . 43
2
2.2 Neutral scalar bosons with a linear source . . . . . . . . . . . . . 61
2.2.1 Classical fields . . . . . . . . . . . . . . . . . . . . . . . . 61
2.2.2 Lagrangian and Hamiltonian formalism . . . . . . . . . . 62
2.2.3 Quantization . . . . . . . . . . . . . . . . . . . . . . . . . 63
2.2.4 Operator valued source . . . . . . . . . . . . . . . . . . . 64
2.2.5 Scattering operator . . . . . . . . . . . . . . . . . . . . . . 64
2.2.6 Green’s functions . . . . . . . . . . . . . . . . . . . . . . . 66
2.2.7 Path integral formulation . . . . . . . . . . . . . . . . . . 67
2.2.8 Feynman rules . . . . . . . . . . . . . . . . . . . . . . . . 67
2.2.9 Vacuum energy . . . . . . . . . . . . . . . . . . . . . . . . 69
2.2.10 Problems with the scattering operator . . . . . . . . . . . 69
2.2.11 Energy shift and scattering theory for a stationary source 70
2.2.12 Travelling source . . . . . . . . . . . . . . . . . . . . . . . 71
2.2.13 Scattering cross-sections . . . . . . . . . . . . . . . . . . . 73
2.2.14 Inclusive cross-section . . . . . . . . . . . . . . . . . . . . 74
2.3 Neutral scalar bosons with a mass-like perturbation . . . . . . . 75
2.3.1 Classical fields . . . . . . . . . . . . . . . . . . . . . . . . 75
2.3.2 Lagrangian and Hamiltonian formalism . . . . . . . . . . 76
2.3.3 Dynamics in the interaction picture . . . . . . . . . . . . 77
2.3.4 Quantization . . . . . . . . . . . . . . . . . . . . . . . . . 78
2.3.5 Quantum Hamiltonian . . . . . . . . . . . . . . . . . . . . 79
2.3.6 Path integral formulation . . . . . . . . . . . . . . . . . . 79
2.3.7 Feynman rules . . . . . . . . . . . . . . . . . . . . . . . . 80
2.3.8 Vacuum energy . . . . . . . . . . . . . . . . . . . . . . . . 81
2.3.9 Pauli-Villars renormalization . . . . . . . . . . . . . . . . 82
2.3.10 Renormalization of the vacuum energy . . . . . . . . . . . 84
2.3.11 Method of dispersion relations . . . . . . . . . . . . . . . 85
2.3.12 Wick rotation . . . . . . . . . . . . . . . . . . . . . . . . . 85
2.3.13 Dimensional renormalization . . . . . . . . . . . . . . . . 87
2.3.14 Energy shift . . . . . . . . . . . . . . . . . . . . . . . . . . 88
3 Massive photons 89
3.1 Free massive photons . . . . . . . . . . . . . . . . . . . . . . . . . 90
3.1.1 Space of solutions . . . . . . . . . . . . . . . . . . . . . . 90
3.1.2 Equations of motion . . . . . . . . . . . . . . . . . . . . . 90
3.1.3 Symplectic structure on the space of solutions . . . . . . . 91
3.1.4 Smeared 4-potentials . . . . . . . . . . . . . . . . . . . . . 92
3.1.5 Lagrangian formalism and stress-energy tensor . . . . . . 92
3.1.6 Diagonalization of the equations of motion . . . . . . . . 94
3.1.7 Plane waves . . . . . . . . . . . . . . . . . . . . . . . . . . 96
3.1.8 Positive frequency space . . . . . . . . . . . . . . . . . . . 97
3.1.9 Spin averaging . . . . . . . . . . . . . . . . . . . . . . . . 97
3.1.10 Quantization . . . . . . . . . . . . . . . . . . . . . . . . . 98
3.2 Massive photons with an external 4-current . . . . . . . . . . . . 100
3.2.1 Classical 4-potentials . . . . . . . . . . . . . . . . . . . . . 100
3.2.2 Lagrangian and Hamiltonian formalism . . . . . . . . . . 101
3
3.2.3 Quantization . . . . . . . . . . . . . . . . . . . . . . . . . 102
3.2.4 Causal propagators . . . . . . . . . . . . . . . . . . . . . . 104
3.2.5 Feynman rules . . . . . . . . . . . . . . . . . . . . . . . . 105
3.2.6 Path integral formulation . . . . . . . . . . . . . . . . . . 106
3.2.7 Energy shift . . . . . . . . . . . . . . . . . . . . . . . . . . 108
3.3 Alternative approaches . . . . . . . . . . . . . . . . . . . . . . . . 108
3.3.1 Classical 4-potentials without the Lorentz condition . . . 108
3.3.2 The Lorentz condition . . . . . . . . . . . . . . . . . . . . 110
3.3.3 Diagonalization of the equations of motion . . . . . . . . 110
3.3.4 Positive frequency space . . . . . . . . . . . . . . . . . . . 112
3.3.5 “First quantize, then reduce” . . . . . . . . . . . . . . . . 113
3.3.6 Quantization without reduction on a positive definite Hilbert
space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
3.3.7 The Gupta-Bleuler approach . . . . . . . . . . . . . . . . 115
4
5 Charged scalar bosons 137
5.1 Free charged scalar bosons . . . . . . . . . . . . . . . . . . . . . . 138
5.1.1 Classical fields . . . . . . . . . . . . . . . . . . . . . . . . 138
5.1.2 Smeared fields . . . . . . . . . . . . . . . . . . . . . . . . 139
5.1.3 Lagrangian formalism . . . . . . . . . . . . . . . . . . . . 140
5.1.4 Classical 4-current . . . . . . . . . . . . . . . . . . . . . . 140
5.1.5 Stress-energy tensor . . . . . . . . . . . . . . . . . . . . . 141
5.1.6 Plane waves . . . . . . . . . . . . . . . . . . . . . . . . . . 142
5.1.7 Positive and negative frequency subspace . . . . . . . . . 143
5.1.8 Plane wave functionals . . . . . . . . . . . . . . . . . . . . 144
5.1.9 Quantization . . . . . . . . . . . . . . . . . . . . . . . . . 145
5.1.10 Quantum 4-current . . . . . . . . . . . . . . . . . . . . . . 147
5.1.11 Quantization in terms of smeared fields . . . . . . . . . . 149
5.2 Charged scalar bosons in an external 4-potential . . . . . . . . . 149
5.2.1 Classical fields . . . . . . . . . . . . . . . . . . . . . . . . 149
5.2.2 Lagrangian and Hamiltonian formalism . . . . . . . . . . 150
5.2.3 Classical discrete symmetries . . . . . . . . . . . . . . . . 152
5.2.4 Quantization . . . . . . . . . . . . . . . . . . . . . . . . . 153
5.2.5 Quantum Hamiltonian . . . . . . . . . . . . . . . . . . . . 154
5.2.6 Quantized discrete symmetries . . . . . . . . . . . . . . . 154
5.2.7 2N -point Green’s functions . . . . . . . . . . . . . . . . . 156
5.2.8 Path integral formulation . . . . . . . . . . . . . . . . . . 157
5.2.9 Feynman rules . . . . . . . . . . . . . . . . . . . . . . . . 157
5.2.10 Vacuum energy . . . . . . . . . . . . . . . . . . . . . . . . 159
5.2.11 Pauli-Villars renormalization . . . . . . . . . . . . . . . . 160
5.2.12 Renormalization of the vacuum energy . . . . . . . . . . . 162
5.2.13 Method of dispersion relations . . . . . . . . . . . . . . . 163
5.2.14 Dimensional renormalization . . . . . . . . . . . . . . . . 164
5.2.15 Abstract gauge covariance . . . . . . . . . . . . . . . . . . 165
5.2.16 Ward identities . . . . . . . . . . . . . . . . . . . . . . . . 165
5.2.17 Energy shift . . . . . . . . . . . . . . . . . . . . . . . . . . 167
5
6.1.13 Fermionic Hamiltonian formalism . . . . . . . . . . . . . . 179
6.1.14 Fermionic Lagrangian formalism . . . . . . . . . . . . . . 180
6.1.15 Classical 4-current . . . . . . . . . . . . . . . . . . . . . . 181
6.1.16 Quantum 4-current . . . . . . . . . . . . . . . . . . . . . . 183
6.2 Dirac fermions in an external 4-potential . . . . . . . . . . . . . . 184
6.2.1 Dirac equation in an external 4-potential . . . . . . . . . 184
6.2.2 Lagrangian and Hamiltonian formalism . . . . . . . . . . 185
6.2.3 Classical discrete symmetries . . . . . . . . . . . . . . . . 186
6.2.4 Quantization . . . . . . . . . . . . . . . . . . . . . . . . . 187
6.2.5 Quantum Hamiltonian . . . . . . . . . . . . . . . . . . . . 188
6.2.6 Quantized discrete symmetries . . . . . . . . . . . . . . . 188
6.2.7 2N -point Green’s functions . . . . . . . . . . . . . . . . . 189
6.2.8 Path integral formulation . . . . . . . . . . . . . . . . . . 191
6.2.9 Feynman rules . . . . . . . . . . . . . . . . . . . . . . . . 191
6.2.10 Vacuum energy . . . . . . . . . . . . . . . . . . . . . . . . 192
6.2.11 Pauli-Villars renormalization . . . . . . . . . . . . . . . . 193
6.2.12 Method of dispersion relations . . . . . . . . . . . . . . . 194
6.2.13 Dimensional renormalization . . . . . . . . . . . . . . . . 194
6.2.14 Energy shift . . . . . . . . . . . . . . . . . . . . . . . . . . 195
A Appendix 204
A.1 Second quantization . . . . . . . . . . . . . . . . . . . . . . . . . 204
A.1.1 Fock spaces . . . . . . . . . . . . . . . . . . . . . . . . . . 204
A.1.2 Creation/annihilation operators . . . . . . . . . . . . . . . 205
A.1.3 Weyl/antisymmetric and Wick quantization . . . . . . . . 206
A.1.4 Second quantization of operators . . . . . . . . . . . . . . 208
A.1.5 Implementability of Bogoliubov translations . . . . . . . . 208
A.1.6 Implementability of Bogoliubov rotations . . . . . . . . . 209
A.1.7 Infimum of a van Hove Hamiltonian . . . . . . . . . . . . 209
6
A.1.8 Infimum of a Bogoliubov Hamiltonian . . . . . . . . . . . 210
A.2 Miscellanea . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 211
A.2.1 Identities for Feynman integrals . . . . . . . . . . . . . . . 211
A.2.2 Identities for the dimensional regularization . . . . . . . . 212
A.2.3 Operator identities . . . . . . . . . . . . . . . . . . . . . . 213
A.2.4 Coulomb and Yukawa potential . . . . . . . . . . . . . . . 213
A.2.5 Vector fields . . . . . . . . . . . . . . . . . . . . . . . . . . 213
A.2.6 Dispersion relations . . . . . . . . . . . . . . . . . . . . . 214
0 Introduction
In these notes we discuss various models of Quantum Field Theory in 1+3
dimensions involving quadratic Lagrangians or, equivalently, quadratic Hamil-
tonians.
First of all, we describe basic types of free fields:
(1) neutral scalar bosons,
(2) neutral massive vector bosons (“massive photons”),
(3) neutral massless vector bosons (“massless photons”),
(4) charged scalar bosons,
(5) (charged) Dirac fermions,
(6) (neutral) Majorana fermions.
We also consider free fields perturbed by a linear or quadratic perturbation
involving a classical (c-number) function.
(1) neutral scalar bosons interacting with a linear source,
(2) neutral scalar bosons interacting with a mass-like perturbation,
(3) massive photons interacting with a classical 4-current,
(4) massless photons interacting with a classical 4-current,
(5) charged scalar bosons interacting with an electromagnetic 4-potential,
(6) Dirac fermions interacting with an electromagnetic 4-potential,
(7) Majorana fermions interacting with a mass-like perturbation.
All the above models are (or at least can be) well understood in the non-
perturbative sense. Perturbation theory is not necessary to compute their scat-
tering operators and Green’s functions, which is not the case (at least so far) of
truly interacting models.
Quantum fields interacting with classical perturbations is a topic with many
important applications to realistic physical systems. Therefore, the formalism
developed in our text is well motivated physically.
Clearly, many important issues of quantum field theory are outside of the
scope of free fields interacting with classical perturbations. However, surpris-
ingly many difficult topics can be discussed already on this level. Therefore, we
7
believe that our text has pedagogical value, as a kind of an introduction to full
quantum field theory.
In our text we stress the deductive character of quantum field theory. Models
that we discuss are quite rigid and built according to strict principles. Among
these principles let us mention the Poincaré covariance, the Einstein causality
and the boundedness of the Hamiltonian from below. Some of these principles
are encoded in the Haag-Kastler and Wightman axioms. Even if these axioms
are often too restrictive, they provide useful guidelines.
The only known models for Haag-Kastler or Wightman axioms in 1+3 dimen-
sions are free theories. Their scattering theory is trivial. To obtain interesting
physical information one needs interacting theories. Unfortunately, interacting
theories are known only perturbatively.
Free theories are the quantizations of covariant 2nd order linear hyperbolic
equations on the Minkowski space. These equations can be perturbed by 0th or
1st order terms involving an arbitrary space-time functions called, depending on
the context, a classical (=external) linear source, mass-like term, 4-current or
electromagnetic 4-potential. We can consider the quantization of the perturbed
equation. Such a theory is still essentially exactly solvable, since the Hamilto-
nian is quadratic. It has no Poincaré covariance. However, it still gives rise to
a net of observable algebras satisfying the Einstein causality.
In our discussion we always start from the study of a classical theory, which
we discuss from the Hamiltonian and Lagrangian point of view. Then we dis-
cuss its quantization. Even though in all the cases we consider the Hamiltonian
is quadratic, its quantization often has various subtle points. In some cases,
especially for vector fields, there are several natural approaches to quantization,
which in the end lead to the same physical results. We try to discuss vari-
ous possible approaches. In our opinion, the existence of seemingly different
formalisms for the same physical system constitutes one of the most confusing
aspects of quantum field theory.
Classical perturbations that we consider are usually described by smooth
space-time functions that decay fast both in space and time. In particular,
their dynamics is typically described by time-dependent Hamiltonians. This is
a certain minor difficulty, which is often ignored in the literature. We discuss
how to modify the usual formalism in order to deal with this problem.
The models that we discuss illustrate many problems of interacting theories,
such as the ultraviolet problem, the infrared problem and the gauge invariance.
The ultraviolet problem means that when we try to define a theory in a
naive way some integrals are divergent for large momenta. In the context of our
paper this is never due to classical perturbations, which we always assume to be
smooth – the source of ultraviolet divergences is the behavior of propagators.
The ultraviolet problem is already visible when we consider neutral fields
with a masslike perturbation or charged fields with a classical electromagnetic
4-potential. In these systems classical dynamics exists under rather weak as-
sumptions. However there are problems with the quantum dynamics.
In some cases the quantum dynamics cannot be implemented on a Hilbert
space. This is the case of charged particles (bosons or fermions) in the presence
8
of variable spatial components of the 4-potential. On the other hand, the the
scattering operator exists under rather weak assumptions for 4-potential going
to zero in the past and future.
Even if we are able to implement the classical dynamics or the classical scat-
tering operator, we encounter another unpleasant surprise. The only quantity
that is not fixed by the classical considerations is the phase factor of the scat-
tering operator, written as e−iE/~ , where E is usually called the vacuum energy.
Computed naively, it often turns out to be divergent. In order to make this
phase factor finite it is necessary to renormalize the naive expression. This di-
vergence appears in low order vacuum energy diagrams. It was first successfully
studied by Heisenberg and Euler in the 30’s. A quantity closely related to this
phase factor is the effective action, which for a constant field was computed
exactly by Schwinger.
The infrared problem means that in the naive theory some integrals are di-
vergent for small momenta. This problem appears already in non-relativistic
quantum mechanics – in scattering theory with Coulomb forces. These forces
are long-range, which makes the usual definition of the scattering operator im-
possible [14]. Its another manifestation is the appearance of inequivalent repre-
sentations of canonical commutation relations, when we consider scattering of
photons against a classical 4-current that has a different direction in the past
and in the future [13, 15]. Thus, even in these toy non-relativistic situations
the usual scattering operator is ill-defined. Therefore, it is not surprising that
(much bigger) problems are present eg. in the full QED. One can cope with the
infrared problem by approximating massless photons with massive ones and re-
stricting computations only to inclusive cross-sections justified by an imperfect
resolution of the measuring device [57, 27, 55].
The expression gauge invariance has in the context of quantum field theory
several meanings.
(1) The most common meaning, discussed already in the context of classi-
cal electrodynamics, is the fact that if a total derivative is added to a
4-potential solving the Maxwell equation, then it still solves the Maxwell
equations. Of course, this no longer holds for the Proca equations – the
massive generalization of the Maxwell equations. Therefore, it is often
stressed that gauge invariance implies that the photons are massless.
(2) There exists another meaning of gauge invariance: we can multiply charged
fields by a space-time dependent phase factor and compensate it by chang-
ing the external potentials.
1. and 2. go together in the full QED, which is invariant with respect to
these two gauge transformations applied simultaneously.
(3) One often uses the term “gauge invariance” in yet another meaning: To
compute the scattering operator we can use various (free) photon prop-
agators. Equivalently, we have the freedom of choosing a Lagrangian in
the path integral formalism. This meaning applies both to massive and
massless photons. Some of these propagators are distinguished, such as
9
the propagator in the Feynman or the Coulomb gauge. (Note, however,
that time-ordered N -point Green’s functions depend on the choice of the
propagator).
All these three meanings of gauge invariance can be illustrated with models
that we consider.
The paper is most of the time rigorous mathematically. In the places where
it is not, we believe that many readers can quite easily make it rigorous. We
try to make the presentation of various models parallel by applying, if possible,
coherent notation and formalism. This makes our text sometimes repetitious –
we believe that this helps the reader to understand small but often confusing
differences between distinct models.
Mathematical language that we use is most of the time elementary. Some-
times we use some mathematical concepts and facts that are, perhaps, less
commonly known, such as C ∗ -algebras, von Neumann algebras, the Schwartz
Kernel Theorem. The readers unfamiliar with them should not be discouraged
– their role in the article is minor.
Most of the material of this work has been considered in one way or another
in the literature. Let us give a brief and incomplete review of references.
On the formal level examples of quantum fields with classical perturbations
are discussed in most textbooks on quantum field theory, see eg. [26, 27, 46, 50,
55, 54, 5].
Linear hyperbolic equations is a well established domain of partial differential
equations, see eg [3].
Axioms of quantum field theory are discussed in [52, 23, 22].
A necessary and sufficient condition for the implementability of Bogoliubov
transformation was given by Shale for bosons [48] and by Shale and Stinespring
for fermions [49], see also [15]
Problems with implementability of the dynamics of charged particles in ex-
ternal potentials was apparently first noticed on a heuristic level in [45]. It was
studied rigorously by various authors. In particular, charged bosons were stud-
ied in [47, ?, 36, 37, 25, 1] and charged fermions in [40, 30, 29, 43, 12]. Rigorous
discussion of the smeared out local charge for charged fermions is contained in
[33].
The renormalization of the vacuum energy goes back to pioneering work of
[24]. In the mathematically rigorous literature it leads to the concept of a causal
phase discussed in the fermionic case in [44, 21].
The infrared problem goes back to [7, 28], see also [13].
The Gupta-Bleuler method of quantization of photon fields goes back to
[19, 6]. The C ∗ -algebraic formulation of the subsidiary condition method is
discussed in [53].
Rigorous study of vacuum energy for Dirac fermions in a stationary potential
is given in [18].
A topic that not included in these notes are anomalies in QFT, which to
a large extent can be treated in the context of external classical perturbations
[20, 32, 10]
10
The notes also treat only dimension 1+3. Note, however, that related prob-
lems can be considered in other dimensions. Of particular importance is the
case of 1+1 dimension with a large literature, eg. [11, 34]
1 Basic concepts
1.1 Minkowski space
1.1.1 Coordinates in Minkowski space
The coordinates of the Minkowski space R1,3 will be typically denoted by xµ ,
µ = 0, 1, 2, 3. By definition, the Minkowski space is the vector space R4 equipped
with the canonical pseudo-Euclidean form of signature (− + ++)
3
X
gµν xµ xν = −(x0 )2 + (xi )2 .
i=1
(Throughout these notes the velocity of light has the value 1 and we use the
Einstein summation convention). We use metric tensor [gµν ] to lower indices
and its inverse [g µν ] to raise indices:
xµ = gµν xν , xµ = g µν xν .
∂f (x)
= ∂xµ f (x) = ∂µ f (x) = f,µ (x).
∂xµ
11
Writing R3 we will typically denote the spatial part of the Minkowski space
obtained by setting x0 = 0. If x ∈ R1,3 , then ~x will denote the projection of
x onto R3 . Latin letters i, j, k will sometimes denote the spatial indices of a
vector. Note that xi = xi .
ijk denotes the 3-dimensional Levi-Civita tensor (the fully antisymmetric
tensor satisfying 123 = 1).
~ x) we define its divergence and rotation in
For a vector field R3 3 ~x 7→ A(~
the standard way:
~ = ∂i Ai , (rotA)
divA ~ i = ijk ∂j Ak .
We write ∂~ A
~ as the shorthand for the tensor ∂i Aj , moreover,
~ 2 :=
X 2
∂~ A
∂i Aj .
ij
1,3
On R we have the standard Lebesgue measure denoted dx. The notation
d~x will be used for the Lebesgue measure on R3 ⊂ R1,3 .
We will often write t for x0 = −x0 . The time derivative will be often denoted
by a dot:
∂f (t) ∂f (x0 )
f˙(t) = = ∂t f (t) = = ∂0 f (x0 ) = f,0 (x0 ).
∂t ∂x0
θ(t) will denote the Heaviside function. We set |t|+ := θ(t)|t|.
12
1.1.3 Invariant measure
Let f be a function on R. The following fact (under appropriate assumptions)
is easy:
Lemma 1.1 Let δ be an approximate delta function, that is
Z
lim δ (t)φ(t)dt = φ(0).
&0
Then Z X φ(si )
lim δ (f (s))φ(s)ds = . (1.1)
&0 |f 0 (si )|
f (si )=0
Thus we have
p p
δ k 0 − ~k 2 + m2 δ k 0 + ~k 2 + m2
δ(k 2 + m2 )dk = p d~k + p d~k. (1.3)
2 ~k 2 + m2 2 ~k 2 + m2
To derive (1.3) from (1.2) we fix ~k and use
d(k 2 + m2 )
= 2k 0 ,
dk 0
p
and k 2 + m2 = 0 iff k 0 = ± ~k 2 + m2 .
Now (1.3) is a measure on R1,3 invariant wrt Lorentz transformations. In
fact,
1 1
2πiδ(k 2 + m2 ) = lim + , (1.4)
&0 k 2 + m2 − i k 2 + m2 + i
where the rhs is obviously Lorentz invariant.
Often, we will drop F – the name of the variable will indicate whether we use
the position or momentum representation:
Z Z
~ −i~ 1 ~
f (k) = e k·~
x
f (~x)d~x, f (~x) = eik·~x f (~k)d~k.
(2π)3
For the time variable (typically t) we reverse the sign in the Fourier transform:
Z Z
1
iεt
f (ε) = e f (t)dt, f (t) = e−iεt f (ε)dε.
2π
13
1.1.5 Lorentz and Poincaré groups
The pseudo-Euclidean group O(1, 3) is called the full Lorentz group. Its con-
nected component of unity is denoted SO↑ (1, 3) and called the connected Lorentz
group.
The full Lorentz group contains special elements: the time reversal T and
the space inversion (the parity) P and the space-time inversion X := PT:
T(x0 , ~x) = (−x0 , ~x), P(x0 , ~x) = (x0 , −~x), Xx = −x.
It consists of four connected components
SO↑ (1, 3), T·SO↑ (1, 3), P·SO↑ (1, 3), X·SO↑ (1, 3).
O(1, 3) has three subgroups of index two: the special Lorentz group (preserving
the spacetime orientation), the orthochronous Lorentz group (preserving the
forward light cone) and the chiral Lorentz group (preserving the parity):
SO(1, 3) = SO↑ (1, 3) ∪ X·SO↑ (1, 3), (1.5)
↑ ↑ ↑
O (1, 3) = SO (1, 3) ∪ P·SO (1, 3), (1.6)
Ochir (1, 3) = SO↑ (1, 3) ∪ T·SO↑ (1, 3). (1.7)
The affine extension of the full Lorentz group R1,3 o O(1, 3) is called the full
Poincaré group. Its elements will be typically written as (y, Λ). We will often
write y instead of (y, 1l) and Λ instead of (0, Λ). It is the full symmetry group
of the Minkowski space.
Quantum field theory models are often not invariant wrt the full Poincaré
group but one of its subgroups: the connected, special, ortochronous or chiral
Poincaré group, which have the obvious definitions.
Example 1.2 Let us determine O(1, 1). We set
1 1
x+ := x + t, x+ := x − t; x= (x+ + x− ), t= (x+ − x− ).
2 2
a b
Now, let A = .
c d
x2 − t2 = x+ x0 = (ax+ + bx− )(cx+ + dx− )
is solved by
ad + bc = 1, ac = 0, bd = 0.
This has 4 types of solutions:
a > 0, d > 0, b = c = 0, (1.8)
a < 0, d < 0, b = c = 0, (1.9)
b > 0, c > 0, a = d = 0, (1.10)
b > 0, c > 0, a = d = 0. (1.11)
Finally, we set
1 1 −1 a b 1 1
A= .
2 1 1 c d −1 1
14
1.1.6 Double coverings of Lorentz and Poincaré groups
The full Poincaré group or one of its subgroups discussed above is sufficient to
describe spacetime symmetries on the level of observables. On the level of the
Hilbert space one needs to replace it by one of its double coverings.
There exists a unique, up to an isomorphism, connected group Spin↑ (1, 3)
such that the following short exact sequence is true:
We say that Spin↑ (1, 3) is a connected double covering of SO↑ (1, 3). The group
Spin↑ (1, 3) happens to be isomorphic to SL(2, C). The kernel of the homomor-
phism Spin↑ (1, 3) → SO↑ (1, 3) consists of 1l and −1l (in the notation inherited
from SL(2, C)).
We would like to extend (1.12) to O(1, 3). There are two natural choices
defined by adjoining the elements P̃± , T̃± that cover P , T , and demanding that
they satisfy
P̃2± = ±1l, T̃2± = ±1l, P̃± T̃± = −T̃± P̃± . (1.13)
One obtains the groups P in+ (1, 3) and P in− (1, 3), which satisfy the following
diagram with exact rows and columns commutes:
1l 1l 1l
↓ ↓ ↓
1l → Z2 → Spin↑ (1, 3) → SO↑ (1, 3) → 1l
↓ ↓ ↓
1l → Z2 → P in± (1, 3) → O(1, 3) → 1l, (1.14)
↓ ↓ ↓
1l → Z2 × Z2 → Z2 × Z2 → 1l
↓ ↓
1l 1l
15
above construction). We obtain a diagram
1l 1l 1l
↓ ↓ ↓
1l → Z2 → Spin↑ (1, 3) → SO↑ (1, 3) → 1l
↓ ↓ ↓
1l → Z4 → P inext (1, 3) → O(1, 3) → 1l. (1.15)
↓ ↓ ↓
1l → Z2 → Z2 × Z2 × Z2 → Z2 × Z2 → 1l
↓ ↓ ↓
1l 1l 1l
Note that if we set X̃ := P̃+ T̃+ = −P̃− T̃− , then X̃2 = −1l and P̃± , T̃± , X̃
anticommute among themselves.
Remark 1.3 As noted in [51] Sect. 3.10, there exists 8 nonisomporphic groups
that are double coverings extending (1.12), that is groups G such that can be
put in the diagram (1.14) in the place of P in± (1, 3). Indeed, we can demand
independently that the elements P̃, T̃ that cover P, T satisfy
Each of the groups (1.5), (1.6) and (1.7) has two nonisomorphic double
coverings. We always prefer those contained in P inext (1, 3).
We have
1l 1l 1l
↓ ↓ ↓
1l → Z2 → Spin↑ (1, 3) → SO↑ (1, 3) → 1l
↓ ↓ ↓
1l → Z2 → Spin(1, 3) → SO(1, 3) → 1l, (1.17)
↓ ↓ ↓
1l → Z2 → Z2 → 1l
↓ ↓
1l 1l
The group Spin(1, 3) is contained in both P in+ (1, 3) and P in− (1, 3). It is
obtained from Spin↑ (1, 3) by adjoining X̃ satisfying X̃2 = −1l. (The other
double covering of SO(1, 3), obtained by adjoining X̃ satisfying X̃2 = 1l, is not
contained in P inext (1, 3)).
We also have two double coverings of O↑ (1, 3) extending (1.12), one contained
16
in P in+ (1, 3), the other in P in− (1, 3):
1l 1l 1l
↓ ↓ ↓
1l → Z2 → Spin↑ (1, 3) → SO↑ (1, 3) → 1l
↓ ↓ ↓
1l → Z2 → P in↑± (1, 3) → O↑ (1, 3) → 1l, (1.18)
↓ ↓ ↓
1l → Z2 → Z2 → 1l
↓ ↓
1l 1l
P inchir 2
± (1, 3) is obtained by adjoining T̃± satisfying T̃± = ±1l.
17
extension to Spin(1, 3) given by the formula in (1.20). Note in particular that
the representation of X̃ is
ij−k 1lCj+1 ⊗Ck+1 .
d
F (ζ) = dF (ζ)V (ζ). (1.22)
dt
The phase space Y is typically a symplectic manifold. This means, it is
equipped with a symplectic form ω, that is, a nondegenerate differential 2-form
ω satisfying dω = 0. If y ∈ Y and ζ1 , ζ2 ∈ Ty Y, then we can evaluate ω(ζ1 , ζ2 ).
Actually, it is convenient to write this as
where ω is treated as a linear map from Ty Y to its dual (Ty Y)# . Then we
can define the Poisson bracket, which is a bilinear antisymmetric map C ∞ (Y) ×
C ∞ (Y) → C ∞ (Y)
{F, G} := −dF · ω −1 dG = ω ω −1 (dF ), ω −1 (dG) , F, G ∈ C ∞ (Y). (1.24)
One usually assumes that the dynamics is Hamiltonian. This means that
the vector field V (ζ) is given by a function H ∈ C ∞ (Y), called a Hamiltonian,
by
V (ζ) = ω −1 dH. (1.26)
The evolution of observables is given by
d
F (ζ) = {F, H}(ζ). (1.27)
dt
By the Jacobi identity, a Hamiltonian dynamics preserves the Poisson bracket,
and hence also the symplectic form.
18
1.3 Darboux coordinates and symplectic vector spaces
The Darboux Theorem says that on any symplectic manifold locally we can
always choose coordinates, say φi , πj , i = 1, . . . , n, such that
To simplify, assume for the moment that we have one degree of freedom.
0 1
The symplectic form is given by the matrix ω = . Clearly, −ω −1 =
−1 0
0 1
. Noting that dF = (∂φ F, ∂π F ), dG = (∂φ G, ∂π G), and going back to
−1 0
an arbitrary number of degrees of freedom, we can write
The phase space has often the structure of a vector space and φi , πj , i =
1, . . . , n, can be chosen to be the coordinates in a basis. Then the tangent
space to Ty Y at any point y ∈ Y can be identified with Y itself and the form
ω is simply a nondegenerate antisymmetric bilinear form on Y. The Darboux
Theorem says that we can identify a symplectic manifold with a symplectic
vector space at least locally.
The space Y # of linear functionals on Y obviously is contained in C ∞ (Y). For
a linear functional on Y, its derivative is the original functional itself. Therefore,
(1.28) can be simplified and written as
ω = φi ∧ πi , (1.31)
19
Definition 1.4 We will say that H, H, Ω satisfy the standard requirements of
quantum mechanics (QM) if
(1) H is a Hilbert space;
(2) H is a positive self-adjoint operator on H (called the Hamiltonian);
(3) Ω is a normalized eigenvector of H with eigenvalue 0;
(4) Ω is nondegenerate as an eigenvector of H.
where w = [wa ] ∈ W. We will call [ρab ] the matrix of ρ. Note that it is natural
to denote the operator and its matrix by the same symbol. In particular, the
matrix of the product of linear operators is simply the product of their matrices.
Let κ be an antilinear operator on W. Then there exists a matrix [κab ] such
that X
(κw)a = κab wb , (1.33)
b
where, as usual, the bar denotes the complex conjugation. We will say that
[κab ] is the matrix of κ. Unfortunately, it is dangerous to use the same letter
for an antilinear operator and its matrix, even if we will sometimes do so, as in
(1.33). Note in particular that
X
(κσ)ac = κab σ bc . (1.34)
b
20
Let G be a group equipped with a homomorphism θ : G → Z2 = {1, −1}. It
yields an obvious partition of G:
G = G1 ∪ G−1 .
which together form a representation of G. One can write (1.35) and (1.36)
more compactly:
21
project onto the spaces of representations of integer, resp. half-integer spin. We
will write
I := U (0, −1l).
Obviously, U (y, Λ̃)I = IU (y, Λ̃). Anticipating the connection of spin and statis-
tics we will call I the fermionic parity.
Denote the ∗-automorphism defined by U (y, Λ̃) by U(y,Λ̃) :
Definition 1.5 The following conditions will be called the basic requirements
of relativistic quantum mechanics (RQM):
(1) Existence of a Poincaré invariant vacuum: There exists a (normalized)
vector Ω invariant with respect to R1,3 o Spin↑ (1, 3).
(2) Spectral condition: The joint spectrum of the energy-momentum operator
is contained in the forward light cone, that is, sp(P ) ⊂ J + .
(3) Uniqueness of the vacuum: The vector Ω is unique up to a phase factor.
(4) Integer and half-integer spin states live in separate superselection sectors:
Observables are contained in {I}0 .
22
1.4.4 Haag-Kastler axioms for observable algebras
We still need some postulates that express the idea of causality. In the mathe-
matical physics literature one can find two kinds of axioms that try to formalize
this concept: the Haag-Kastler and the Wightman axioms. Even though the
Wightman axioms were formulated earlier, it is more natural to start with the
Haag-Kastler axioms.
Definition 1.7 We keep the basic requirements of RQM.
In addition, to each open bounded set O ⊂ R1,3 we associate a von Neumann
algebra A(O) ⊂ {I}0 . We will say that the family A(O), O open in R1,3 , is a
net of observable algebras satisfying the Haag-Kastler axioms if the following
conditions hold:
(1) Isotony: O1 ⊂ O2 implies A(O1 ) ⊂ A(O2 ).
(2) Poincaré covariance: for (y, Λ̃) ∈ R1,3 o Spin↑ (1, 3), we have
U(y,Λ̃) A(O) = A (y, Λ)O .
23
Definition 1.9 We assume the basic requirements of RQM. We say that a fam-
ily of von Neumann algebras F(O) ⊂ B(H) associated to bounded open subsets
O of R1,3 is a net of field algebras in the sense of Haag-Kastler axioms if the
following conditions hold:
(1)’ Isotony: O1 ⊂ O2 implies F(O1 ) ⊂ F(O2 ).
(2)’ Poincaré covariance: for (y, Λ̃) ∈ R1,3 o Spin↑ (1, 3), we have
U(y,Λ̃) F(O) = F (y, Λ)O .
The main reason for introducing the twisted Einstein causality is the need
to accommodate anticommuting fermionic fields. Clearly, if the net F(O), O ⊂
R1,3 satisfies the Haag-Kastler axioms for field algebras, then the net of their
fermionic even subalgebras
F0 (O) := {B ∈ F(O) : IBI = B}, O ⊂ R1,3 ,
satisfies the Haag-Kastler axioms for observable algebras.
Note that in our formulation the decomposition H = H0 ⊕ H1 given by the
operator I plays a double role.
(1) It describes the decomposition of the Hilbert space into integer and half-
integer spin representations.
(2) In the Einstein causality axiom, block-diagonal operators have the bosonic
character and block-off-diagonal operators have the fermionic character.
A priori it is not obvious that these two properties should give the same decom-
position. However, one can show that it is natural to assume from the beginning
that this is the case. This is the content the theorem about the connection of
the spin and statistics, described eg. in [52].
Setting Λ̃ = −1l in Axiom (2)’ shows that the bosonic/fermionic superselec-
tion rule is local, ie., IF(O)I = F(O) for all O.
24
We define the gauge invariant subalgebras
Fgi (O) = {B ∈ F0 (O) : Rg (B) = B, g ∈ G}
or, equivalently,
Fgi (O) = F0 (O) ∩ {R(g) : g ∈ G}0 .
Then the net O 7→ Fgi (O) satisfies the Haag-Kastler axioms for observable
algebras.
25
(1) Continuity: For any Φ, Ψ ∈ D,
is continuous.
(2) Poincaré covariance: for (y, Λ̃) ∈ R1,3 o Spin↑ (1, 3) we have
h i
U(y,Λ̃) φ̂[f ] = φ̂ σ(Λ̃)f ◦ (y, Λ)−1 .
(3) Einstein causality: Let suppf1 × suppf2 , where fi have values in Vji , i =
1, 2. Then
φ̂[f1 ]φ̂[f2 ] = (−1)j1 j2 φ̂[f2 ]φ̂[f1 ].
(4) Cyclicity of the vacuum: Let Falg denote the algebra of polynomials gener-
ated by φ̂[f ]. Then Falg Ω is dense in H.
(5) Hermiticity: For any Φ, Ψ ∈ D,
In what follows a map (1.44) satisfying Axiom (1) will be called an operator
valued distribution. By saying that it is cyclic we will mean that it satisfies
Axiom (4).
Setting Λ̃ = −1l in Axiom (2), we see that f ∈ S(R1,3 , Vj ) implies
26
(1) A and B preserve D and are essentially self-adjoint on D.
(2) A and B commute on D.
(3) A and B do not commute strongly.
(4) D is dense.
More about what is known about the relationship between the Haag-Kastler
and Wightman axioms the reader can find in [2], Sect. 4.9.
27
after smearing with complex test functions
XZ
ψ̂[h] := ha (x)ψ̂a (x)dx,
a
XZ
ψ̂ ∗ [h] := ha (x)ψ̂a∗ (x)dx,
a
28
(3) Einstein causality: (...) Let supph1 ×supph2 , where hi have values in Wji ,
i = 1, 2. Then (...)
(4) Cyclicity of the vacuum: Let Falg denote the algebra of polynomials in φ̂[f ],
ψ̂[h] and ψ̂ ∗ [h]. Then Falg Ω is dense in H.
(5) Hermiticity: (...)
(Φ|ψ̂[h]Ψ) = (ψ̂ ∗ [h]Φ|Ψ).
Note that the cyclicity of vacuum implies that the sum of Hn is dense in H.
Assume that Hn are mutally orthogonal, so that we have the decomposition
H = ⊕ Hn . For θ ∈ U (1) we define R(θ) := ⊕ einθ . Clearly, R(θ)Ω = Ω and
n∈Z n∈Z
U (1) 3 θ 7→ R(θ) is a unitary representation commuting with U (y, Λ̃). Let Rθ
be the corresponding ∗-automorphism:
Rθ (A) = R(θ)AR(−θ).
We then have
Rθ φ̂a (x) = φ̂a (x),
e−iθ ψ̂a (x),
Rθ ψ̂a (x) =
Rθ ψ̂a∗ (x) eiθ ψ̂a∗ (x).
=
29
1.4.14 Charge conjugation
Let C be a unitary operator such that CΩ = Ω. Let C be the corresponding
∗-automorphism:
C(A) := CAC −1 .
We say that it is a charge conjugation if it satisfies
X
−1
C φ̂a (x) = αab φ̂b (x), (1.51)
b
X
κ−1 ∗
C ψ̂a (x) = ab ψ̂b (x), (1.52)
b
and hence
X
C ψ̂a∗ (x) κ−1
= ab ψ̂b (x), (1.53)
b
The resulting set of axioms will be called the Wightman axioms of a P -invariant
theory.
30
In particular, the space inversion (parity) P̃± ∈ P in↑± (1, 3) is represented in
the Hilbert space by the unitary operator P± := U (P̃± ). P± := UP̃± denotes
the corresponding automorphism. It acts on the fields as follows:
X
−1
P± φ̂a (x0 , ~x) = (P̃± )φ̂b (x0 , −~x),
σab
b
X
0 −1
(P̃± )ψ̂b (x0 , −~x).
P± ψ̂a (x , ~x) = τab
b
P inchir
± (1, 3) 3 Λ̃ 7→ σ(Λ̃) ∈ L(V), (1.61)
P inchir
± (1, 3) 3 Λ̃ 7→ τ (Λ̃) ∈ Lθ(Λ̃) (W), (1.62)
R1,3 o P inchir
± (1, 3) 3 (y, Λ̃) 7→ U (y, Λ̃) ∈ Uθ(Λ̃) (H). (1.63)
Note that we demand that (1.62) is θ-linear and (1.63) is θ-unitary. We denote
by [τab (Λ̃)] the matrix of τ . The resulting set of axioms will be called the
Wightman axioms of a T -invariant theory.
In particular, the time reversal is implemented by the anti-unitary operator
T± := U (T̃± ). T± := UT̃± denotes the corresponding automorphism. The time
reversal acts on the fields as follows
X
−1
T± φ̂a (x0 , ~x) = (T̃± )φ̂b (−x0 , ~x),
σab
b
X
0 −1
(T̃± )ψ̂b (−x0 , ~x).
T± ψ̂a (x , ~x) = τab
b
31
Let us stress that (1.65) is linear and not θ-linear! A deep theorem, called the
CP T Theorem, says that we can extend the representation (1.57) to a θ-unitary
representation
Note that we demand that (1.70) is θ-linear and (1.71) is θ-unitary. We assume
(1.49) and (1.50) for P inext (1, 3). In particular, we have the antilinear operators
σ0 (X̃), τ0 (X̃), and U0 (X̃).
By the CPT Theorem we also have the representations (1.64) and (1.65)
of Spin(1, 3) satisfying (1.67) and (1.68). In particular, we have the operators
σ(X̃), τ (X̃± ), and U (X̃). Note that we put the subscript 0 in (1.69), (1.70) and
(1.71) to distinguish the representations used in Wightman axioms from the
representations obtained by the CPT Theorem.
Define
α := σ(X̃)σ0 (X̃−1 ),
κ := τ (X̃)τ0 (X̃−1 )
C := U (X̃)U0−1 (X̃−1 ).
32
Then C is unitary, and the corresponding automorphism C(A) = CAC −1 satis-
fies (1.51) and (1.52).
If P = U0 (P̃+ ) and T = U0 (T̃+ ), then U0 (X̃) = U0 (P̃+ )U0 (T̃+ ), and hence
X = CP T . This explains the name of the CPT Theorem. (Let us stress,
however, that the theorem holds also if the theory is not P and T invariant, so
that we cannot write X = CP T , as described in the previous subsection).
S(R1,3 , V) × · · · × S(R1,3 , V)
3 (fN , . . . , f1 ) 7→ (Ω|φ̂[fN ] · · · φ̂[f1 ]Ω) ∈ C, (1.72)
From the point of view of the Wightman axioms, the collection of Wightman
functions WN , N = 0, 1, . . . , contains all the information about a given quantum
field theory. In particular,
φ̂[fN ] · · · φ̂[f1 ]Ω|φ̂[gM ] · · · φ̂[g1 ]Ω
Z
= W (y1 , . . . , yN , xM , . . . , x1 )
×f1 (x1 ) · · · fN (xN )gM (yM ) · · · g1 (y1 )dx1 · · · dxN dyM · · · dy1 .
where sgna (σ) is the sign of the permutation of the fermionic elements among
N, . . . , 1.
33
Note that we multiply a distribution with a discontinuous function in (1.73),
which strictly speaking is illegal. Disregarding this problem, Green’s functions
are covariant due to the commutativity/anticommutativity of fields at spacelike
separations.
where (σ1 , . . . , σn ) is the permutation such that tσn ≥ · · · ≥ tσ1 and sgna (σ) is
the sign of this permutation restricted to the odd elements among Bn , . . . , B1 .
Consider a family of self-adjoint operators
t 7→ H(t). (1.74)
We will assume that H(t) are even. For t+ > t− , we define the time-ordered
exponential
!
Z t+
Texp −i H(t)dt (1.75)
t−
∞
X Z Z
:= (−i)n ··· H(tn ) · · · H(t1 )dtn · · · dt1
n=0 t+ ≥tn ≥···≥t1 ≥t−
∞ Z t+ Z t+
X
n 1
= (−i) ··· T (H(tn ) · · · H(t1 )) dtn · · · dt1 .
n=0 t− t− n!
For brevity, we will write U (t+ , t− ) for (1.75) and call it the dynamics generated
by t 7→ H(t). (Of course, if H(t) are unbounded, the above definition should
be viewed only as a heuristic indication how to define the family of unitary
operators U (t+ , t− ). In general, in most of this subsection we are not very
precise about the boundedness of operators, limits, etc.)
We also set U (t− , t+ ) := U (t+ , t− )−1 .
Clearly, if H(t) = H, then U (t+ , t− ) = e−i(t+ −t− )H .
34
1.5.2 Schrödinger and Heisenberg picture
Suppose that H is a (time-independent) Hamiltonian. If we prepare a state ρ at
time 0 and measure an observable A at time t > 0, then the expectation value
of the measurement is
TrρeitH Ae−itH . (1.76)
In quantum physics two equivalent ways of expressing (1.76) are used:
(1) The Schrödinger picture: We let the state evolve ρt := e−itH ρeitH and keep
the observable constant. Then (1.76) equals Trρt A.
(2) The Heisenberg picture: We let the observable evolve At := eitH Ae−itH
and keep the state constant. Then (1.76) equals TrρAt .
(By the Schrödinger picture one also means the unitary evolution Ψt := e−itH Ψ
on H.)
where we treat t = 0 as the reference time. Equivalently, AHp (t) is the solution
of
d
AHp (t) = i [HHp (t), AHp (t)] , (1.82)
dt
A(0) = A, (1.83)
35
where the Hamiltonian in the Heisenberg picture is defined as
d
U (t, 0) = −U (t, 0)iHHp (t); (1.85)
dt
d
U (0, t) = iHHp (t)U (0, t). (1.86)
dt
Above we assumed that the observable A is time-independent in the Schrödinger
picture. It is sometimes useful to be more general and to consider an obsevable
t 7→ ASp (t), which is time-dependent in the Schrödinger picture. Then in the
Heisenberg picture it is defined as
Let us note that a similar distinction between two pictures exists in classical
dynamical systems. Consider a flow on Rd given by the equation
d
x(t) = v t, x(t) .
dt
For any initial condition x(0) = x0 ∈ Rd , we obtain a solution R 3 t 7→ x(t, x0 ).
Thus any time dependent observable F has two descriptions:
In fluid dynamics, (1.88) is sometimes called the Eulerian description, and (1.89)
the Lagrangian description.
In classical mechanics the phase space is described by coordinates (φ, π) ∈
Rm × Rm . The time evolution is described by the Hamilton equations
φ̇(t) = ∂π H t, φ(t), π(t) ,
π̇(t) = −∂φ H t, φ(t), π(t) .
For any initial condition φ0 , π0 ∈ Rm × Rm , we obtain a solution of the
Hamilton equations. Similarly as in the quantum case, we have two kinds of the
classical Hamiltonian:
36
(1.90) is the Hamiltonian in the Eulerian description and (1.91) is the Hamilto-
nian in the Lagrangian description. The former is the analog of the Schrödinger
picture and the latter of the Heisenberg picture. If they do not depend on time,
they coincide.
We will use the classical Hamiltonian in the Lagrangian description and the
quantum Hamiltonian in the Schrödinger picture as the standard ones. Note
that for the time being we are rather fussy about puttimg the subscripts such as
Sp and Hp. In practice they are usually omitted and determined by the context.
d
Afp (t) = i [Hfr , Afp (t)] ,
dt
Afp (0) = A,
We will also use the free Heisenberg picture for time-dependent observables
t 7→ ASp (t):
Note that
!
Z t+
UInt (t+ , t− ) = Texp −i Vfp (t)dt .
t−
serves as the Hamiltonian for the interaction picture, therefore, (1.94) has an
alternative notation HInt (t).
37
We define the scattering operator by
Clearly, S = S +∗ S − .
Then we compute
38
Suppose that there exist
Note that the generating function is the vacuum expectation value of a certain
scattering operator:
Z(f ) = (Φfr |S(f )Φfr ), (1.106)
where S(f ) is the scattering operator (1.95) with λV (t) replaced by
X
λV (t) + fi (t)Ai .
i
39
Let > 0. We define V (t) := e−|t| V . We will write
40
Proof. Applying (1.109) to Φfr yields (1.111). We scalar multiply it with
Φfr obtaining (1.112). Combining (1.111) with (1.112) gives (1.113). 2
In the following theorem we argue that the adiabatic switching often allows
us to compute an eigenvector of H and its eigenvalue.
Theorem 1.14 (1) Assume that
there exist a nonzero lim Φ±
. (1.114)
&0
Φ±
GL := lim |(Φfr |S± Φfr )|Φ±
.
&0
Then
HΦ± ± ±
GL = EGL ΦGL .
±
Proof. (1) The existence of EGL is immediate. Next we note that (Φ± ±
|Φ ) =
|(Φfr |S± Φfr )|−2 . Hence lim |(Φfr |S± Φfr )| exists. This implies the existence of
&0
Φ±
GL .
By (1.113), we have
±
(H − EGL ) lim Φ±
= 0.
&0
41
Proof. The right hand side of (1.117) equals
−
(S+ Φfr |BS− Φfr ) (Φ+ −
|BΦ ) (Φ+
GL |BΦGL )
lim = lim = .
&0 (S+ Φfr |S− Φfr ) + −
&0 (Φ |Φ ) + −
(ΦGL |ΦGL )
2
The following theorem describes the Sucher formula often used in practical
computations of the energy shift.
Theorem 1.16
iλ
E − Efr = lim ∂λ log(Φfr |S Φfr ). (1.118)
&0 2
Proof. We sandwich (1.110) with Φfr and divide with (Φfr |S Φfr ) obtaining
Suppose that H = Hfr +λV . The Gell-Mann and Low Theorem about Green’s
functions allows us to express interacting Green’s functions by the free ones:
Theorem 1.17 Suppose that (1.114), (1.115) and (1.116) are true, so that we
can apply the results of the previous subsubsection. Then
∞
1 X (−iλ)n
hAk (tk ) · · · A1 (t1 )i = lim (1.119)
&0 (Φfr |S Φfr ) n!
n=0
Z ∞ Z ∞
× dsn · · · ds1 hV (sn ) · · · V (s1 )Ak (tk ) · · · A1 (t1 )ifr ,
−∞ −∞
∞ n Z∞ Z ∞
X (−iλ)
(Φfr |S Φfr ) = lim dsn · · · ds1 (1.120)
&0
n=0
n! −∞ −∞
42
Proof. (1.120) follows from (1.95) applied to U .
Let us prove (1.119). Let tk ≥ · · · ≥ t1 . Let Ai, (t) denote the operator Ai
in the Heisenberg picture for the evolution U . The left-hand side of (1.119) is
Theorem 1.18 Suppose that (1.114), (1.115) and (1.116) are true.
43
(1) Assume also that there exist the adiabatic or Gell-Mann–Low Møller oper-
ators
± |(Φfr |S± Φfr )| ±
SGL := w− lim S , (1.124)
&0 (Φfr |S± Φfr )
and
|(Φfr |S± Φfr )| ±
w− lim λ∂λ S = 0. (1.125)
&0 (Φfr |S± Φfr )
Then
± ±
SGL (Hfr − Efr ) = (H − ReE)SGL , (1.126)
±
SGL Φfr = Φ±
GL . (1.127)
44
2 Neutral scalar bosons
In this section we consider the Klein-Gordon equation
and we quantize the space of its real solutions. We study two kinds of interac-
tions: an external linear source
can be written as
Z
dk
ζ(x) = eikx g(k)δ(k 2 + m2 )
(2π)3
√
XZ d~k q 0 ~2 2 ~
= p g ± ~k 2 + m2 , ~k e∓ix k +m +i~xk ,
±
3 ~
(2π) 2 k + m2 2
where H1± are the Hankel functions and K1 is the MacDonald function of
the 1st order.
45
(2) The Pauli-Jordan or the commutator function:
Z
dk
D(x) = i eikx sgn(k 0 )δ(k 2 + m2 )
(2π)3
d~k
Z q
x~
i~ k 0 ~2 2
= p e sin x k + m
(2π)3 ~k 2 + m2
1 msgnx0 θ(−x2 ) p
= sgnx0 δ(x2 ) − √ J1 (m −x2 ),
2π 4π −x2
where J1 is the Bessel function of the 1st order. D(x) is the unique solution
of the Klein-Gordon equation satisfying
Solutions of
(−2 + m2 )ζ(x) = δ(x), (2.5)
are called Green’s functions or fundamental solutions of the Klein-Gordon equa-
tion. Formally, they can be written as
eikx
Z
dk
(k + m2 ) (2π)4
2
eikx
Z
dk
D± (x) =
(k + m ∓ i0sgnk ) (2π)4
2 2 0
1 mθ(−x2 )θ(±x0 ) p
= θ(±x0 )δ(x2 ) − √ J1 (m −x2 ).
2π 4π −x 2
In the literature, D+ (x) is usually denoted Dret (x) and D− (x) is usually
denoted Dadv (x).
(2) The Feynman(-Stueckelberg) Green’s function:
eikx
Z
dk
DF (x) =
(k + m2 − i0) (2π)4
2
1 mθ(−x2 ) − p 2 miθ(x2 ) √
= δ(x2 ) − √ H1 (m −x ) + √ K1 (m x2 ).
4π 8π −x2 4π 2 x2
The special solutions and Green’s functions introduced above are often called
propagators.
Remark 2.1 Both the Feynman propagator is called causal Green’s function by
Bogoliubov-Shirkov. The Pauli-Jordan bisolution is called the causal propagator
by e.g. Fredenhagen. We will avoid these names.
46
Proposition 2.2 We have suppD± ⊂ J ± and suppD ⊂ J. The propagators
satisfy the following relations
eikx
Z
dk
D+ (x) =
(k + m − i0sgnk 0 ) (2π)4
2 2
0 0 ~
e−ik x +ik~x dk 0 d~k
Z
= .
~k 2 + m2 − (k 0 + i0)2 (2π)4
eikx
Z
dk
DF (x) =
~k 2 + m2 − (|k 0 | + i0)2 (2π)4
0 0 ~
e−ik x +ik~x
Z
1
= dk (2.14)
(2π)4
p p
2 ~k 2 + m2 ~k 2 + m2 − |k 0 | − i0
0 0 ~
e−ik x +ik~x
Z
dk
+ p . (2.15)
~k 2 + m2 + |k 0 | + i0 (2π)4
p
2 ~k 2 + m2
In (2.15) we can replace i0 with −i0. Then the parts of (2.14) and (2.15) with
47
k 0 < 0 are swapped:
0 0 ~
e−ik x +ik~x
Z
dk
= p p 4
2 ~k 2 + m2 ~k 2 + m2 − k 0 − i0 (2π)
0 0 ~
e−ik x +ik~x
Z
dk
+ p p 4
,
~ 2
2 k +m 2 ~ 2 2 0
k + m + k − i0 (2π)
Z −i√~k2 +m2 x0 +i~k~x ~
0 e dk
= iθ(x ) p 3
~ 2
2 k +m 2 (2π)
√
Z i ~k2 +m2 x0 +i~k~x ~
0 e dk
+iθ(−x ) p 3
2 ~k 2 + m2 (2π)
= θ(x0 )D(+) (x) − θ(−x0 )D(−) (x).
0
dk 0 = 2πie∓ix ε θ(±x0 ).
ε ∓ k − i0
Finally, let us prove (2.12). (2.6) and (2.7) imply D(+) = −D(−) +D+ −D− .
Inserting this into (2.11) we obtain
= D(+) + D− .
Theorem 2.3 Let ς, ϑ ∈ Cc∞ (R3 ). Then there exists a unique ζ ∈ Csc
∞
(R1,3 )
that solves
(−2 + m2 )ζ = 0 (2.16)
with initial conditions ζ(0, ~x) = ς(~x), ζ̇(0, ~x) = ϑ(~x). It satisfies suppζ ⊂
J(suppς ∪ suppϑ) and is given by
Z Z
ζ(t, ~x) = Ḋ(t, ~x − ~y )ς(~y )d~y + D(t, ~x − ~y )ϑ(~y )d~y . (2.17)
R3 R3
48
∞
For ζ1 , ζ2 ∈ Csc (R1,3 ) we define
Let YKG , resp. CYKG denote the space of real, resp. complex, space-compact
solutions of the Klein-Gordon equation. If ζ1 , ζ2 ∈ CYKG , then
∂µ j µ (x) = 0.
Clearly, the form (2.21) is well defined also if only ζ2 ∈ YKG , and ζ1 is a
distributional solution of the Klein-Gordon equation.
The Poincaré group R1,3 o O(1, 3) acts on YKG and CYKG by
r(y,Λ) are symplectic (preserve the symplectic form) for Λ ∈ O↑ (1, 3), otherwise
they are antisymplectic (change the sign in front of the symplectic form).
The Pauli-Jordan function D can be used to construct solutions of the Klein-
Gordon equation parametrized by space-time functions, which are especially
useful in the axiomatic formulation of QFT.
Proposition 2.4 (1) For any f ∈ Cc∞ (R1,3 , R), D ∗ f ∈ YKG , where
Z
D ∗ f (x) := D(x − y)f (y)dy.
49
(4) If suppf1 × suppf2 , then
(D ∗ f1 )ω(D ∗ f2 ) = 0.
Proof. Let us prove (2.22). Choose time t later than suppfi , i = 1, 2. Then
we have D ∗ fi = D+ ∗ fi . Now, using Green’s identity (2.20), we obtain
−(D ∗ f1 )ω(D ∗ f2 )
Z
= (Ḋ+ ∗ f1 )(t, ~x)(D+ ∗ f2 )(t, ~x) − (D+ ∗ f1 )(t, ~x)(Ḋ+ ∗ f2 )(t, ~x) d~x
Z
= − (2 − m2 )(D+ ∗ f1 )(x)(D+ ∗ f2 )(x)
x0 <t
+(D+ ∗ f1 )(x)(2 − m2 )(D+ ∗ f2 )(x) dx
Z
f1 (x)(D+ ∗ f2 )(x) − (D+ ∗ f1 )(x)f2 (x) dx
=
Z Z
f1 (x)(D+ ∗ f2 )(x) − f1 (x)(D− ∗ f2 )(x) dx =
= f1 (x)(D ∗ f2 )(x)dx.
2
The right hand side of (2.22) is sometimes called the Peierls bracket of f1
and f2 .
Note that in this context the star does not denote the Hermitian conjugation
(which in our text is the standard meaning of the star).
Let us stress that the space YKG is real, which reflects the fact that in this
section we consider neutral fields. In the section devoted to charged fields the
main role will be played by the complexification of YKG , that is WKG := CYKG .
For x ∈ R1,3 , φ(x), π(x) will denote the functionals on YKG given by
They are called classical fields. Clearly, for any ζ ∈ YKG we have
(−2 + m2 )hφ(x)|ζi = 0.
50
Thus the equation
(−2 + m2 )φ(x) = 0 (2.23)
is a tautology.
# #−1
On YKG we have the action of the Poincaré group (y, Λ) 7→ r(y,Λ) . Note that
#−1
r(y,Λ) φ(x) = φ(Λx + y).
or more simply, Z
ω= φ(t, ~x) ∧ π(t, ~x)d~x. (2.25)
By (2.25), the symplectic structure on the space YKG leads to the Poisson
bracket
The relations (2.26) can be viewed as mnemotechnic identities that yield the
correct Poisson bracket for more regular functions, eg. the smeared out fields in
(??) or (2.29) described below. Note that formally φ(t, ~x) and π(t, ~x) generate
the algebra of all functions on YKG .
Using (2.24) we obtain
−(ωζ)ρ = ζωρ.
51
#
Therefore, each element of YKG defines an element of YKG by the pairing given
by the symplectic form. It is convenient to allow complex smearing functions
paired antilinearly. More precisely, for ζ ∈ CYKG we introduce the functional
on YKG given by
hφ((ζ))|ρi := ζωρ, ρ ∈ YKG .
A possible alternative notation for φ((ζ)) is −ωζ or ζω.
Clearly,
Z
ζω = φ((ζ)) = −ζ̇(t, ~x)φ(t, ~x) + ζ(t, ~x)π(t, ~x) d~x. (2.27)
Note that
{φ((ζ1 )), φ((ζ2 ))} = ζ 1 ωζ 2 . (2.28)
We can also smear fields with space-time functions. For f ∈ Cc∞ (R1,3 , R),
we set Z
φ[f ] := f (x)φ(x)dx.
We have
Z Z
{φ[f1 ], φ[f2 ]} = f1 (x)D(x − y)f2 (y)dxdy. (2.29)
Proposition 2.5 Here is the relationship between the two kinds of smearing:
52
In the Lagrangian formalism one also uses a classical field, which we will
denote by φ(x), as before. But now, this field is off-shell. This means, we do
not enforce any equation on φ(x). One can interpret φ(x) as the functional on,
say, C ∞ (R1,3 ) or Csc
∞
(R1,3 ) such that hφ(x)|f i := f (x).
In QFT one uses the local Lagrangian density L(x), which is a function of
the field φ(x), ∂µ φ(x) =: φ,µ (x) and of x ∈ R1,3 . The Euler-Lagrange equation
reads then
∂L(x)
∂φ(x) L(x) − ∂µ =0 (2.31)
∂φ,µ (x)
To obtain the Klein-Gordon equation, using φ(x) in the off-shell formalism,
introduce the Lagrangian density
L(x) = − 21 ∂µ φ(x)∂ µ φ(x) − 21 m2 φ(x)2 . (2.32)
The Euler-Lagrange equation yields the Klein-Gordon equation (2.1).
When we go from the Lagrangian to Hamiltonian formalism, we enforce the
on-shell condition, that is, the Euler-Lagrange equation, and we introduce the
variable conjugate to φ(x):
∂L(x)
π(x) := = φ,0 (x) = φ̇(x).
∂φ,0 (x)
Then we express everything in terms of φ(x) and π(x).
53
We introduce the (total) Hamiltonian and momentum:
Z Z
µ0
H := T (x)dsµ (x) = H(t, ~x)d~x, (2.34)
ZS Z
P i := T µi (x)dsµ (x) = P i (t, ~x)d~x. (2.35)
S
where
Z
1 ~ 0 ~
ζ (±) = ζ (±) (k) p q e±i(−ε(k)x +k~x) d~k. (2.37)
(2π)3 2ε(~k)
We have
Z Z
(+) (+) (−) (−)
iζ1 ωζ2 = ζ1 (k)ζ2 (k)d~k − ζ1 (k)ζ2 (k)d~k (2.38)
(+)
Note that (3.67) restricted to WKG is positive definite. For g1 , g2 ∈ W (+)
we will write
(g1 |g2 ) := ig 1 ωg2 .
The Hilbert space of positive energy solutions is denoted ZKG , and is the
(+)
completion of WKG in this scalar product.
1,3 ↑
R o O (1, 3) leaves ZKG invariant.
(+)
We have a natural identification of YKG with WKG . Indeed, ζ ∈ YKG can
(+)
be projected onto ζ (+) ∈ WKG , as in (2.36)-(2.53). This identification allows us
to define a real scalar product on YKG :
(+) (+)
hζ1 |ζ2 iY := Re(ζ1 |ζ2 ).
54
We can compute explicitly this scalar product:
Z Z
hζ1 |ζ2 iY = ζ̇1 (0, ~x)D(+) (0, ~x − ~y )ζ̇2 (0, ~y )d~xd~y (2.39)
Z Z
+ ζ1 (0, ~x)(−∆~x + m2 )D(+) (0, ~x − ~y )ζ2 (0, ~y )d~xd~y .
d~k
Z
eikx a(k) + e−ikx a∗ (k) ,
φ(x) = q
(2π)3 2ε(~k)
p
q
Z d~k ε(~k)
√ eikx a(k) − e−ikx a∗ (k) .
π(x) = p
3
i (2π) 2
We have
{a(k), H} = −iε(~k)a(k), (2.43)
∗
{a (k), H} = iε(~k)a∗ (k); (2.44)
0 ∗ ∗ 0
{a(k), a(k )} = {a (k), a (k )} = 0, (2.45)
∗
{a(k), a (k )} 0
= −iδ(~k − ~k 0 ). (2.46)
a(k), a∗ (k) diagonalize simultaneously the Hamiltonian, momentum and
symplectic form:
Z
H = d~kε(~k)a∗ (k)a(k), (2.47)
Z
P~ = d~k~ka∗ (k)a(k), (2.48)
Z
iω = d~ka∗ (k) ∧ a(k). (2.49)
55
With ζ1 , ζ2 ∈ YKG , the last identity is the shorthand for
Z
iζ1 ωζ2 = ha(k)|ζ1 iha(k)|ζ2 i − ha(k)|ζ1 iha(k)|ζ2 i d~k.
Note that negative frequency plane waves in the neutral case play secondary
role.
Following Dirac, we denote plane waves using the “ket notation” |k) when
they appear on the right of a bilinear form. We also write (x|k) for the evaluation
of |k) at the point x ∈ R1,3 .
If a plane wave appears on the left, we employ the “bra notation”, which
implies an additional complex conjugation:
1 1
(k|x) = p q e−ikx , (k|x) = p q eikx .
(2π)3 2ε(~k) (2π)3 2ε(~k)
#
Recall that if ζ ∈ YKG , then it determines ζω, an element of YKG . We can
∗
interpret the plane wave functionals a(k), a (k) as linear functionals on the real
space YKG
a(k)ζ := i(k|ωζ, (2.55)
∗
a (k)ζ := −i(k|ωζ. (2.56)
56
In other words,
a(k) = iφ((|k)))
Z
= i ∂t (k|0, ~x)φ(0, ~x) − (k|0, ~x)π(0, ~x) d~x,
a∗ (k) = −iφ((|k)))
Z
= −i ∂t (k|0, ~x)φ(0, ~x) − (k|0, ~x)π(0, ~x) d~x.
Z
φ(x) = (x|k)a(k) + (x|k)a∗ (k) d~k.
2.1.10 Quantization
Let us describe the quantization of the Klein-Gordon equation, following the
formalism of quantization of neutral bosonic systems [15]. We will use the
“hat” to denote the quantized objects.
We want to construct H, Ĥ, Ω satisfying the standard requirements of QM
(1)-(3) and a self-adjoint operator valued distribution
˙
such that, with π̂(x) := φ̂(x),
(1) (−2 + m2 )φ̂(x) = 0,
(2) [φ̂(0, ~x), φ̂(0, ~y )] = [π̂(0, ~x), π̂(0, ~y )] = 0,
[φ̂(0, ~x), π̂(0, ~y )] = iδ(~x − ~y ).
(3) eitĤ φ̂(x0 , ~x)e−itĤ = φ̂(x0 + t, ~x).
(4) Ω is cyclic for φ̂(x).
The above problem has a solution, which is essentially unique. Indeed, let
H, Ĥ, Ω, R1,3 3 x 7→ φ̂(x), π̂(x) solve the above problem. Decorating (2.41) and
(2.42) with hats leads to the definitions of two operator valued distributions
Hermitian conjugate to one another:
s !
Z
d~x ~ ε( ~k) i
â(k) : = p e−ik~x φ̂(0, ~x) + q π̂(0, ~x) , (2.58)
(2π)3 2
2ε(~k)
s !
ε(~k)
Z
∗ d~x i~
k~x i
â (k) = p e φ̂(0, ~x) − q π̂(0, ~x) . (2.59)
(2π)3 2
2ε(~k)
57
Using (2) and (3) we obtain the quantized versions of (2.43)-(2.46):
Using again (2.60), (2.63) and (2.64) we see that the scalar product of two
vectors Ψ, Ψ0 is zero if n 6= n0 , and otherwise it is
Z
(Ψ|Ψ0 ) = n! Ψ(~k1 , . . . , ~kn )Ψ0 (~k1 , . . . , ~kn )d~k1 · · · d~kn .
Therefore, H can be identified with Γs L2 (R3 ) , Ω with the Fock vacuum, â∗ (k)
with the creation operators in the “physicist’s notation”, the quantum field is
d~k
Z
eikx â(k) + e−ikx â∗ (k) .
φ̂(x) := q
(2π)3 2ε(~k)
p
By (2.54) we can identify L2 (R3 ) with the positive frequency Hilbert space
ZKG . Using the “mathematician’s notation” on the right we can write
â∗ (k) = â∗ |k) .
(2.65)
58
This is true even though we only required that time translations are imple-
mented.
We have
[φ̂(x), φ̂(y)] = −iD(x − y).
For f ∈ Cc∞ (R1,3 , R) set
Z
φ̂[f ] := f (x)φ̂(x)dx. (2.66)
YKG 3 ζ 7→ φ̂((ζ))
(2)
φ̂((r(t,~0) ζ)) = eitĤ φ̂((ζ))e−itĤ .
59
R1,3 o O↑ (1, 3) acts on CCR(YKG ) by ∗-automorphisms defined by
r̂(y,Λ) (W (ζ)) := W r(y,Λ) (ζ) .
We are looking for a cyclic representation of this algebra with the time evolution
generated by a positive Hamiltonian.
The solution is provided by the state on CCR(YKG ) defined by
1
ψ W (ζ) = exp − hζ|ζiY .
2
Let (Hψ , πψ , Ωψ ) be the GNS representation generated by the state ψ. Then this
representation has the required properties. Hψ can be identified with Γs (ZKG )
and the fields are related to the Weyl operators by
πψ (W (ζ)) = eiφ̂((ζ)) .
60
This yields the identities for spatially smeared fields and Weyl operators, where
the scalar product h·|·iY on YKG was introduced in (2.39):
Z Z
2
(Ω|φ̂((ζ)) Ω) = ζ̇(0, ~x)D(+) (0, ~x − ~y )ζ̇(0, ~y )d~xd~y
Z Z
ζ(0, ~x)(−∆~x + m2 )D(+) (0, ~x − ~y )ζ(0, ~y )d~xd~y
= hζ|ζiY , (2.76)
1
(Ω|eiφ̂((ζ)) Ω)
= exp − hζ|ζiY . (2.77)
2
Fix a function
R1,3 3 x 7→ j(x) ∈ R, (2.79)
which will be called the (external) linear source. In most of this subsection we
will assume that (2.79) is Schwartz. The interacting fields satisfy the equation
We also require that the interacting fields have the same equal-time Poisson
brackets as the free fields:
61
Let us mention some alternative ways to define the interacting fields φ(x).
First of all, there is nothing special about the time t = 0 in (2.83) – we can
replace it with any t = t0 . Alternatively, we can demand
lim (φfr (t, ~x) − φ(t, ~x)) = 0, lim (πfr (t, ~x) − π(t, ~x)) = 0,
t→∞ t→∞
or lim (φfr (t, ~x) − φ(t, ~x)) = 0, lim (πfr (t, ~x) − π(t, ~x)) = 0.
t→−∞ t→−∞
Another possibility is to introduce YKG (j), the space of smooth real space-
compact solutions of
(−2 + m2 )ζ(x) = −j(x), (2.85)
and define φ(x) by
hφ(x)|ζi := ζ(x), ζ ∈ YKG (j).
62
2.2.3 Quantization
We will use the notation φ̂fr (x) for the free quantum fields studied in the previous
subsection. We are now looking for interacting quantum fields φ̂(x) satisfying
(−2 + m2 )φ̂(x) = −j(x). (2.89)
We also set
˙
π̂(x) := φ̂(x) (2.90)
and require the equal time commutation relations
[φ̂(t, ~x), φ̂(t, ~y )] = [π̂(t, ~x), π̂(t, ~y )] = 0,
[φ̂(t, ~x), π̂(t, ~y )] = iδ(~x − ~y ). (2.91)
We would like to solve (2.89) and (2.91) in terms of free fields. That means,
we are looking for φ̂(x) on the Hilbert space of the free Klein-Gordon fields,
Γs (ZKG ). We will in addition demand that the interacting and free fields at
time t = 0 coincide:
φ̂(~x) : = φ̂(0, ~x) = φ̂fr (0, ~x),
π̂(~x) : = π̂(0, ~x) = π̂fr (0, ~x). (2.92)
Clearly, the unique solution is obtained by decorating (2.84) with hats:
φ̂(x) := φ̂fr (x)
Z
+ D+ (x − y)θ(y 0 ) + D− (x − y)θ(−y 0 ) j(y)dy. (2.93)
It can be written as
Z 0 Z t
φ̂(t, ~x) = Texp −i Ĥ(s)ds φ̂(0, ~x)Texp −i Ĥ(s)ds , (2.94)
t 0
63
2.2.4 Operator valued source
So far we assumed that j(x) is a c-number. Most of the formalism works, at
least formally, for operator valued sources. The main additional difficulty is the
need to distinguish between the source in various pictures.
Let us start with the Schrödinger picture. Let R1,3 3 x 7→ ĵ(x) be an
operator-valued function (or distribution) that commutes with time zero fields:
which is obtained from (2.97) by replacing j(t, ~x) with ĵInt (t, ~x).
d~k
Z
−itε(~
k) ~k)â(k) + eitε(~k) j(t, ~k)â∗ (k) .
ĤInt (t) = q e j(t,
(2π)3 2ε(~k)
p
64
The scattering operator (1.95) can be computed exactly. On the level of creation
and annihilation operators it acts as
j ε(~k), ~k
∗ ∗ ∗
Ŝâ (k)Ŝ = â (k) + i p q , (2.99)
(2π)3 2ε(~k)
j ε(~k), ~k
∗
Ŝâ(k)Ŝ = â(k) − i p q . (2.100)
(2π)3 2ε(~k)
|j(k)|2
Z Z
i 1 1 dk
= +
2 2ε(~k) (ε(~k) − k 0 − i0) (ε(~k) + k 0 − i0) (2π)4
|j(k)|2
Z Z
i dk
= .
2 ε(~k)2 − (k 0 )2 − i0 (2π)4
65
2.2.6 Green’s functions
Recall that the N -point Green’s function is defined for xN , . . . , x1 as follows:
hφ̂(xN ) · · · φ̂(x1 )i
:= Ω+ |T φ̂(xN ) · · · · · · φ̂(x1 ) Ω− , (2.103)
where
Z 0
Ω± := lim Texp −i Ĥ(s)ds Ω
t→±∞ t
Z 0
= Texp −i ĤInt (s)ds Ω.
±∞
∂N
hφ̂(xN ) · · · φ̂(x1 )i = iN Z(f ) . (2.106)
∂f (xN ) · · · ∂f (x1 ) f =0
hφ̂(kN ) · · · φ̂(k1 )i
Z Z
:= dxn · · · dx1 e−ixn kn −···−ix1 k1 hφ̂(xN ) · · · φ̂(x1 )i.
66
Amputated Green’s functions can be used to compute scattering amplitudes:
kn++ , . . . , k1+ | Ŝ |kn−− , . . . , k1− (2.108)
Note that the Feynman rules for scattering amplitudes follow now from
the rules for the vacuum expectation value of the scattering amplitude, if we
Radd additional insertion vertices—one-legged vertices corresponding to the term
dxf (x)φ̂fr (x).
Consider heuristically the space of all (off-shell) configurations with the Lebesgue
measure Π dφ(x). Physicists like to rewrite (2.109) as
x
R R
Π dφ(x) exp i L(x) − f (x)φ(x) dx
x
Z(f ) = R R , (2.110)
Π dφ(x) exp i Lfr (x)dx
x
which follows by basic rules of Gaussian integrals. Note that strictly speaking
(2.110) is ambiguous, since DF , the causal propagator, is only one of many
inverses (Green’s functions) of −2 + m2 . The choice of the causal propagator
is an additional convention that is not explicitly contained in the expression
(2.110).
67
We have 1 kind of lines and 1 kind of vertices. At each vertex just one
line ends. Vertices are denoted by solid dots. Lines have no distinguished
orientation. However, when we fix the orientation of a line, we can associate to
it a momentum k.
Diagrams for Green’s functions, in addition to internal lines have external
lines ending with insertion vertices, which will be denoted by small circles. To
compute Green’s functions we do as follows:
(1) We draw all possible Feynman diagrams. More precisely, we put N dots for
insertion vertices, labelled 1, . . . , N . We put n dots, labelled 1, . . . , n, for
interaction vertices. Then we connect them in all possible allowed ways.
The expression for the diagram is then divided by n!.
(2) To each vertex we associate the factor −ij(k), where k is the momentum
flowing towards this vertex.
(3) To each line we associate the propagator
c −i
−iDfr (k) = .
k 2 + m2 − i0
d4 k
(4) For internal lines we integrate over the variables with the measure (2π)4 .
68
Figure 2: Diagram for a scattering amplitude.
∞
X 1 (2m)! m
(Ω|ŜΩ) = D = exp(D/2).
m=0
(2m)! 2m m!
|j(k)|2 |j(k)|2
Z Z
1 dk 1 dk
Re = (2.111)
2 (k 2 + m2 − i0) (2π)4 2 (k 2 + m2 ) (2π)4
69
can be infinite. This is not a very serious problem. (2.111) is responsible only
for the phase of the scattering amplitude and does not influence scattering cross-
sections.
We can try to remedy the problem by an apropriate renormalization of the
phase. In particular, in the case of a stationary source or, more generally, a
source travelling with a constant velocity, we can use the adiabatic switching
and the Gell-Mann and Low construction to obtain a meaningful scattering
operator. We will describe this construction in the next subsubsections.
The problem with Ŝ is more serious if
is infinite. Then no unitary operator Ŝ satisfies the relations (2.99) and (2.100),
see Thm A.1. The scattering operator is ill defined. However, as we describe in
Subsubsect. 2.2.14, also in this situation there is a pragmatic solution – we can
define inclusive cross-sections.
Note that if k 7→ j(k) is Schwartz, then Ŝ is well-defined, even if m = 0.
2 m2
Z
1 1
Ĥ = : π̂(~x)2 + ∂~ φ̂(~x) + φ̂(~x)2 + j(~x)φ̂(~x) :d~x (2.112)
2 2 2
d~k
Z Z
∗ ~ ~ j(~k)a(k) + j(~k)â∗ (k) p
= â (k)â(k)ε(k)dk + q . (2.113)
(2π)3 2ε(~k)
By the method of completing the square (A.14) we compute the infinum of Ĥ:
e−m|~x−~y|
Z
1
E=− j(~x) j(~y )d~xd~y (2.114)
2 4π|~x − ~y |
|j(~k)|2 d~k
Z
1
=− . (2.115)
2 (2π)3 (~k 2 + m2 )
Obviously, the standard Møller operators Ŝ ± (1.123) are ill defined and we
need to use the Gell-Mann–Low construction. Let us replace j with j± (t, ~x) :=
θ(±x)j(~x)e−|t| . Its Fourier transform is j± (k 0 , ~k) = ∓i(k 0 ∓i)−1 j(~k), Inserting
70
this into (2.101) we obtain
±
Z j(~k) d~k
Ŝ = exp − q a∗ (k) p
2ε(~k)(ε(~k) ∓ i) (2π)3
Z j(~k) d~k
× exp q a(k) p
2ε(~k)(ε(~k) ± i) (2π)3
1Z |j(~k)|2 d~k |j(~k)|2 d~k
Z
i
× exp − + .
2 2ε(~k)(ε(~k)2 + 2 ) (2π)3 2 2(ε(~k)2 + 2 ) (2π)3
±
Note that the phase of Ŝ± behaves as O(−1 ). In the definition of SGL we take
this phase away and put & 0, see (1.124). We obtain
±
Z j(~k) d~k
ŜGL = exp − q a∗ (k) p
2ε(~k)3 (2π)3
Z j(~k) d~k
× exp q a(k) p
2ε(~k)3 (2π)3
1 Z |j(~k)|2 d~k
× exp − .
2 2ε(~k)3 (2π)3
R
If m > 0 or if j(~x)d~x = 0, then Ĥ has a ground state and the operators
±
ŜGL are well defined. We have
± ±
ŜGL Ĥfr = (Ĥ − E)ŜGL .
+ −
Note a somewhat disappointing feature: ŜGL = ŜGL , and hence the scatter-
+∗ −
ing operator ŜGL := ŜGL ŜGL = 1l is trivial.
R
If m = 0 and j(~x)d~x 6= 0, then Ĥ has no ground state (even though it is
±
bounded from below) and the operators ŜGL are ill defined.
71
This is the interaction picture Hamiltonian for a time-independent perturbation
where in the dynamics we replace the 1-particle energy ε(~k) by ε(~k) − ~v~k.
We use the Gell-Mann–Low type adiabatic switching, so that we replace j
by
j (t, ~x) := e−|t| j(t, ~x).
We slightly generalize the Gell-Mann–Low Møller operators:
± |(Ω|Ŝ± Ω)|
ŜGL = lim Ŝ±
&0 (Ω|Ŝ± Ω)
Z q(~k) d~k
= exp − q a∗ (k) p
2ε(~k)(ε(~k) − ~v~k) (2π)3
Z q(~k) d~k
× exp q a(k) p
2ε(~k)(ε(~k) − ~v~k) (2π)3
1Z |q(~k)|2 d~k
× exp − .
2 2ε(~k)(ε(~k) − ~v~k)2 (2π)3
Note that if |v| < 1 (if the source is slower than light) and m > 0, then
+ −
ŜGL = ŜGL are well defined unitary operators.
±
R
If m = 0 and q(~x)d~x 6= 0, then the infrared problem shows up: ŜGL are ill
defined.
It is interesting to assume that the source has a different asymptotics in the
future and in the past. For simplicity, suppose that the change occurs sharply
at time t = 0 and consider
q− (~x − t~v− ), t < 0,
j(t, ~x) =
q+ (~x − t~v+ ), t > 0.
We can introduce first the scattering operator Ŝ with the adiabatically
switched interaction. Then we can define a Gell-Mann–Low type scattering
operator by taking & 0 and renormalizing the phase:
|(Ω|Ŝ Ω)|
lim Ŝ (2.117)
&0 (Ω|Ŝ Ω)
!
q+ (~k) q− (~k) d~k
Z
1
∗
= exp − â (k) p
~ ~ ~ ~
q
2ε(~k) (ε(k) − ~v+ k) (ε(k) − ~v− k) (2π)3
!
q + (~k) q − (~k) d~k
Z
1
× exp − + â(k) p
(ε(~k) − ~v+~k) (ε(~k) − ~v−~k)
q
2ε(~k) (2π)3
!
q+ (~k) q− (~k) 2 d~
Z
1 1 k
× exp − − .
2 2ε(~k) (ε(~k) − ~v+~k) (ε(~k) − ~v−~k) (2π)3
Let m = 0. Then (2.117) is ill defined if
72
R R
(1) q+ (~x)d~x 6= q− (~x)d~x,
or
R R
(2) q+ (~x)d~x = q− (~x)d~x 6= 0 and v+ 6= v− .
(2.117) is given by (2.101) where we replace
|j(k)|2
Z
dk
(k + m − i0) (2π)4
2 2
with
If we do not like the adiabatic switching approach we can directly define the
Møller operators by removing the (possibly infinite) phase shift from (2.101).
Alternatively, we can multiply the two Møller operators:
!
~k)q− (~k) d~k
2Im q (
Z
+∗ − i 1 +
ŜGL ŜGL = ŜGL exp − .
2 2ε(~k) (ε(~k) − ~v+~k)(ε(~k) − ~v−~k) (2π)3
Note that the crosssections are sometimes zero. In the next subsubsection
we describe how to cope with this problem.
73
2.2.14 Inclusive cross-section
Let δ > 0. The 1-particle Hilbert space can be split as Z = Z<δ ⊕ Z>δ corre-
sponding to the soft momenta |~k| < δ and hard momenta |~k| > δ. Clearly,
Assume first that m > 0 and the scattering operator is computed as above.
Clearly, the scattering operator and scattering cross-sections factorize:
|~q1+ |, . . . , |~qm
+
q1− |, . . . , |~qm
+ |, |~
−
− | < δ. (2.118)
Likewise, let
|~k1+ |, . . . , |~kn++ |, |~k1− |, . . . , |~kn−− | > δ. (2.119)
The corresponding hard scattering cross-section are
We have
74
σ>δ describes the experiment that does not measure outgoing particles of mo-
mentum less than δ and in the incoming state there are no particles of mo-
mentum less than δ. Actually, we would have obtained the same scattering
cross-section if the part of the incoming state below the momentum δ was ar-
bitrary. This is an example of an inclusive cross-section – a cross-section which
involves summing over many unobserved final states.
If m & 0, the soft scattering operator Ŝ<δ has no limit. All σ<δ go to zero.
In fact, they are proportional to
Z |j(ε(~q), ~q)|2 d~q
σ<δ = exp − .
|~
q |<δ 2ε(~q) (2π)3
The hard scattering operator Ŝ>δ and σ>δ have well defined limits and can have
a physical meaning.
One can imagine various experimental scenarios that lead to different inclu-
sive cross-sections. For example, imagine that our apparatus does not detect
the details of an outgoing state if the total energy of soft particles is less than
δ. This leads to the following inclusive cross-section:
app
k1+ , . . . , kn++ ; kn−− , · · · , k1− := σ k1+ , . . . , kn++ ; kn−− , · · · , k1−
σ>δ
X∞ Z
σ k1+ , . . . , kn++ , q1 , . . . , qj ; kn−− , · · · , k1− d~q1 · · · d~qj .
+
j=1 ε(~
q1 )+···+ε(~
qj )<δ
app
Note that both σ>δ and σ>δ are proportional to one another:
app
k1+ , . . . , kn++ ; kn−− , · · · , k1−
σ>δ
σ>δ k1+ , . . . , kn++ ; kn−− , · · · , k1−
∗
:= (Ω<δ |Ŝ<δ 1l[0,δ] (Ĥfr )Ŝ<δ Ω<δ ) = σ<δ (; ) (2.121)
∞
X Z
+ σ<δ (q1 , . . . , qj ; ) d~q1 · · · d~qj .
j=1 ε(~
q1 )+···+ε(~
qj )<δ
75
Let us define the corresponding retarded and advanced propagators as the
unique distributional solutions of
satisfying
suppD± ⊂ {x, y : x ∈ J ± (y)}.
We also generalize the Pauli-Jordan function:
Note that
suppD ⊂ {x, y : x ∈ J(y)}.
The function D can be used to solve the initial value problem of (2.122):
Z
φ(t, ~x) = − ∂s D(t, ~x, s, ~y ) φ(0, ~y )d~y
s=0
Z
+ D(t, ~x, 0, ~y )π(0, ~y )d~y . (2.124)
76
2.3.3 Dynamics in the interaction picture
The classical interaction picture Hamiltonian can be expressed in terms of plane
wave functionals:
Z
1
HInt (t) = κ(t, ~x)φ2fr (t, ~x)d~x (2.125)
2
d~k1 d~k2 κ(t, ~k1 + ~k2 ) −itε(~k1 )−itε(~k2 )
Z
1
= q q e a(−k1 )a(−k2 )
2
(2π)3 2ε(~k1 ) 2ε(~k2 )
~ ~ ~ ~
+2eitε(k1 )−itε(k2 ) a∗ (k1 )a(−k2 ) + eitε(k1 )+itε(k2 ) a∗ (k1 )a∗ (k2 ) .
The solution of these equations at two times are related by a matrix of the form
" #
pt+ ,t− qt+ ,t−
(2.126)
qt+ ,t− pt+ ,t−
(2.126) has a limit as t+ , −t− → ∞, which can be called the classical scattering
operator.
One can try to solve the equations of motion by iterations. The first iteration
is often (at least in the quantum context) called the Born approximation, and
it gives the following formula for the elements of (2.126):
t+
κ(s, −~k + ~k1 )
Z
~ ~
pBorn
t+ ,t− (k, k1 ) = δ(~k − ~k1 ) + i ds q q e−isε(k)+isε(k1 ) ,
t− (2π)3 2ε(~k) 2ε(~k1 )
t+
κ(s, −~k + ~k1 )
Z
~ ~
qtBorn
+ ,t−
(k, k1 ) = i ds q q e−isε(k)−isε(k1 ) .
t− (2π)3 2ε(~k) 2ε(~k1 )
77
2.3.4 Quantization
Again, we are looking for quantum fields R1,3 7→ φ̂(x) satisfying
(−2 + m2 )φ̂(x) = −κ(x)φ̂(x), (2.127)
˙
with the conjugate field π̂(x) := φ̂(x) having the equal time commutators (2.91),
and coinciding with the free field at time 0, as in (2.92). The solution is given
by putting “hats” onto (2.124).
We would like to check whether the classical scattering operator and the clas-
sical dynamics are implementable in the Fock space for nonzero κ. By Thm A.2,
we need to check the Shale condition, that is, whether the off-diagonal elements
of (2.126) are square integrable. For simplicity, we will restrict ourselves to the
Born approximation; the higher order terms do not change the conclusion.
The verification of the Shale condition is easier for the scattering operator.
Consider
Z ∞
κ(s, −~k + ~k1 ) ~ ~
Born
q∞,−∞ (k, k1 ) = i ds q q e−isε(k)−isε(k1 ) . (2.128)
−∞ (2π)3 2ε(~k) 2ε(~k1 )
Recall that κ is a Schwartz function. Therefore, we can integrate by parts as
many times as we want:
Z ∞ ~ ~
Born n+1 ∂sn κ(s, −~k + ~k1 ) e−isε(k)−isε(k1 )
q∞,−∞ (k, k1 ) = i ds . (2.129)
~ ~ n
q q
−∞ (2π)3 2ε(~k) 2ε(~k1 ) ε(k) + ε(k1 )
This decays in ~k and ~k1 as any inverse power, and hence is square integrable on
R3 × R3 . Therefore the classical scattering operator is implementable.
Next let us check the implementability of the dynamics, believing again that
it is sufficient to check the Born approximation. We integrate by parts once:
qtBorn
+ ,t−
(k, k1 )
~ ~ ~ ~
−κ(t+ , −~k + ~k1 )e−it+ ε(k)−it+ ε(k1 ) + κ(t− , −~k + ~k1 )e−it− ε(k)−it− ε(k1 )
= q q
(2π)3 2ε(~k) 2ε(~k1 ) ε(~k) + ε(~k1 )
~ ~
∂s κ(s, −~k + ~k1 )e−isε(k)−isε(k1 )
Z t+
+ ds q q . (2.130)
t− (2π)3 2ε(~k) 2ε(~k1 ) ε(~k) + ε(~k1 )
Using that κ(s, ~k + ~k1 ) decays fast in the second variable, we see that (2.130)
can be estimated by
C
,
(ε(k) + ε(~k1 ))2
~
which is square integrable. Therefore, the dynamics is implementable for any
t− , t+ .
By a similar computation we check that if we freeze t0 ∈ R, the dynamics
generated by the momentary Hamiltonian HInt (t0 ) is implementable.
78
2.3.5 Quantum Hamiltonian
We may try to write the quantum Hamiltonian as
Z
1 1 2 1
Ĥ(t) := : π̂ 2 (~x) + ∂~ φ̂(~x) + (m2 + κ(t, ~x))φ̂2 (x) :d~x. (2.131)
2 2 2
We will see later on that the Wick-ordered expression (2.131) does not define
an operator. However we will use it to derive the Feynman rules, which unfor-
tunately will lead to divergent diagrams.
Formally (2.94) remains true if we add a time dependent constant C(t) to
(2.131). We will see that in order to define correct Hamiltonians Ĥ(t) this
constant has to be infinite. We will obtain bounded from below Hamiltonians
Ĥren (t), however the vacuum will not be contained in their form domain. There-
fore, the condition (Ω|Ĥren (t)Ω) = 0 for all t, which is equivalent to the Wick
ordering, cannot be imposed.
The interaction picture Hamiltonian is
Z
1
ĤInt (t) = κ(t, ~x):φ̂2fr (t, ~x):d~x (2.132)
2
d~k1 d~k2 κ(t, ~k1 + ~k2 ) −itε(~k1 )−itε(~k2 )
Z
1
= q q e â(−k1 )â(−k2 )
2
(2π)3 2ε(~k1 ) 2ε(~k2 )
~ ~ ~ ~
+2eitε(k1 )−itε(k2 ) â∗ (k1 )â(−k2 ) + eitε(k1 )+itε(k2 ) â∗ (k1 )â∗ (k2 ) .
Here, the determinant is understood (at least formally) as the Fredholm deter-
1
minant on the space L2 (R1,3 ). The term exp κDfr
c 2
is responsible for the Wick
ordering.
79
Similarly as in the case of (2.110), (2.133) is often expressed in terms of path
integrals as
Z Z
C Π dφ(x) exp i L(x) − f (x)φ(x) dx . (2.134)
x
80
Feynman rules are similar as in the case of a linear source. The difference
is that now vertices have 2-legs. The rule (2) for calculating Green’s functions
changes: for each vertex with incoming momenta k1 , k2 we insert the number
−iκ(k1 + k2 ), where k1 and k2 are the momenta of lines entering the vertex.
Another difference is that we do not allow a line to begin and end at the same
vertex – this is because we use the Wick ordered Ĥ(t).
Diagrams can be decomposed into connected components of two kinds:
1. lines ending at insertion vertices (for Green’s functions) or on-shell parti-
cles (for scattering amplitudes) with 0, 1, 2, . . . interaction vertices;
2. loops with 2, 3, . . . interaction vertices.
Note that loops with 1 interaction vertex do not appear because of the Wick
ordering.
Diagrams without loops (both for Green’s functions and scattering ampli-
tudes) are finite, because the external momenta are fixed and on interaction
vertices we have the fast decaying function κ.
Consider a loop with 4-momenta k1 , . . . , kn flowing around it. On vertices we
have the function κ, which essentially identifies ki with ki+1 . The propagators
give the power |ki |−2 . Thus we are left with 4 degrees of freedom and the
integrand that behaves as |k|−2n . This is integrable if n > 2, but divergent for
the 2-vertex loop. We will see that only the imaginary part of this diagram is
divergent.
81
Figure 6: Vacuum energy
n vertices. This is a special case of a more general rule saying that to compute
log(Ω|ŜΩ) we need to sum over all connected diagrams with no external lines
divided by the symmetry factor (the order of the group of the symmetries of
the diagram). In the case of a loop with n vertices its group of symmetries is
the nth dihedral group, hence the symmetry factor is 2n.
where the right hand side defines the vacuum energy function π(k 2 ). Unfortu-
nately, computed naively, π(k 2 ) is logarithmically divergent.
The renormalization of a mass-like perturbation is not very difficult and can
be done in many ways. We will describe 3 methods of renormalization. All of
them will lead to the same renormalized vacuum energy function π ren (k 2 ).
We start with the Pauli-Villars method. In the context of a mass-like per-
turbation, the Pauli-Villars regularization consists in introducing an additional
fictitious field that has a (large) mass M and appears only in loops. (Thus we
ignore diagrams involving external lines of the fictitious particle). In addition,
each loop of the fictitious field has a (nonphysical) coefficient −1. We organize
our computations by setting m0 = m, C0 = 1, m1 = M , and C1 = −1. The
Pauli-Villars regularized vacuum energy function is the sum of the loop of the
physical particle and of the fictitious one:
82
d4 q X
Z
2 1
4πM (k ) = i Ci
(2π)4 i ((q + 21 k)2 + m2i − i0)((q − 12 k)2 + m2i − i0)
Z ∞ Z ∞
d4 q
Z
X
2 1 2 2
= −i dα1 dα2 Ci exp −i(α1 + α2 ) q + k + mi − i(α1 − α2 )qk
(2π)4 0 0 i
4
Z ∞ Z ∞
1 X 1 2 α1 α2 2
= − dα 1 dα 2 C i exp −i(α 1 + α2 )m i − i k
(4π)2 0 0 i
(α1 + α2 )2 α1 + α2
Z 1 Z ∞
(1 − v 2 )k 2
1 dρ X
= − 2
dv Ci exp −iρ m2i +
(4π) 0 0 ρ i 4
Z 1 X
1 k 2 (1 − v 2 )
= 2
dv Ci log m2i + − i0
(4π) 0 i
4
Z 1 X
(1 − v 2 )k 2
1 2
= dv Ci log 1 + − i0 + log mi .
(4π)2 0 i
4m2i
83
Here is the calculation. First we assume that k 2 > 0. Then we can drop i0 and
Z 1 k 2 (1 − v 2 )
log 1 + dv (2.138)
0 4m2
√
k2 r !
k 2 + 4m2
Z 2m
2m
=√ √ log − w dw. (2.139)
k 2 − 2m k2 4m2
Using log(w)dw = w log(w) − w, this gives the value of π ren (k 2 ) in the first
R
84
The counterterm has an infinite coefficient π∞ (0). Otherwise, it is quite well
behaved – it depends locally on the interaction, and therefore the renormaliza-
tion preserves the Einstein causality. This manifests itself in the identity
Using log(t − i0) = log |t| − iπθ(−t), we see that the imaginary part of the
logarithm is very simple. Hence
Z 1
ren 2 π (1 − v 2 )k 2
Imπ (k ) = − θ − 1 − dv
4(4π)2 0 4m2
r
π
= − √ − k 2 − 4m2 .
4(4π) −k 2
2 +
We can obtain the real part by using the fact that π ren (0) = 0 and the once
subtracted dispersion relations for the variable s = −k 2 , see Thm A.4:
Z ∞
ren 2 1 ren 1 1
Reπ (k ) = P dsImπ (−s) − . (2.143)
π 4m2 s + k2 s
85
The physical region R1,3 of (2.144) lies at the boundary–on ]0, ∞[×R3 from
above and on ] − ∞, 0[×R3 from below:
hp|qiE := p0 q 0 + p~~q,
Clearly, we can express the causal propagator in terms of the Euclidean propa-
gator with help of the Wick rotation:
d4 q G(p2 , pq, q 2 )
Z
R1,3 3 p 7→ F (p) := , (2.146)
(2π)4 ap2 + 2bpq + cq 2 + m2 − i0 n
a b
where G is holomorphic and the matrix is positive definite. Then
b c
instead of F we can consider the holomorphic function
C\ ] − ∞, −m] ∪ [m, ∞[ × R3 3 (p0 , p~) (2.147)
d4 q G(p2 , pq, q 2 )
Z
7→ F (p) := ,
(2π)4 ap2 + 2bpq + cq 2 + m2 n
where in the integral we substituted (iq 0 , ~q) for (q 0 , ~q). This substitution can be
reached from the original variables inside the holomorphy domain by the Wick
86
rotation, hence it does not affect the integral. F E is holomorphic on the domain
of (2.145). We can retrieve the physical values of F from F E by
d4 q
Z
1
4π E (k 2 ) = −
(2π)4 ((q + 21 k)2 + m2 )((q − 12 k)2 + m2
1 1 d4 q
Z Z
1
= − dv 4 2
2 −1 (2π) q + 2 k 2
2
4 + m + vqk
Z 1 Z
d4 q 1
= − dv , (2.148)
0 (2π) q 2 + (1 − v 2 ) + m2 2
4 k 2
4
dq 4 µ4−d Ωd ∞ d−1
Z Z
is replaced by |q| d|q|, (2.149)
(2π)4 (2π)d 0
where Ωd is the “area of the unit sphere in d dimension”, see (A.30). Thus
instead of (2.148) we consider its dimensionally regularized version:
Z ∞
µ4−d Ωd 1 |q|d−1
Z
4π E,d (k 2 ) = − d
dv 2 d|q|
(2π) 0 0
k 2
q 2 + 4 (1 − v 2 ) + m2
Z 1 k2
1 2 2 2
' − dv − γ + log(µ 4π) − log (1 − v ) + m
(4π)2 0 4
1
− . (2.150)
(4π)2 (2 − d/2)
87
which coincides with the Wick rotated result obtained by the Pauli-Villars
method. Thus the renormalization of (2.150) amounts to choosing
µ2
log = γ − log 4π, (2.151)
m2
dropping the pole term and setting d = 4.
Above, we rewrote the square root by using the identities (A.35) and (A.36),
expanded the denominator in the Neumann series and at the end we used the
identity (A.37). Note that the nth term of the above expansion corresponds to
the loop with n vertices. They are all well defined except for n = 2, which needs
renormalization. We can guess that the renormalized energy shift is
d~k
Z Z
1
E ren = π ren (~k 2 )|κ(~k)|2
3
+ Tr κ (2.152)
(2π) (−∆ + m2 + τ 2 )
1 1 1 dτ
× 2 2
κ 2 2
κ 2 2
τ2 ,
(−∆ + m + τ ) (−∆ + m + κ + τ ) (−∆ + m + τ ) 2π
where we rewrote the sum of terms with n ≥ 3 in a compact form, and π ren was
introduced in (2.137).
88
Another way to derive the expression for E ren is to use Sucher’s formula.
We introduce the adiabatically switched perturbation e−|t| κ(~x) multiplied by a
coupling constant λ, which will be put to 1 at the end. The Fourier transform
of the switching factor e−|t| is 2 2
+τ 2 Therefore,
Eren = i log(Ω|Ŝren Ω)
42 dτ d~k
Z
= λ2 π ren (−τ 2 + ~k 2 ) 2 2 2
|κ(~k)|2 + O(λ3 ).
( + τ ) (2π)4
By Sucher’s formula,
iλ
E ren = lim ∂λ log(Ω|Ŝren Ω)
&0 2
Z
43 ~
= lim λ 2
π ren (−τ 2 + ~k 2 ) |κ(~k)|2 dτ dk + O(λ3 )
&0 (2 + τ 2 )2 (2π)4
d~k
Z
= λ2 π ren (~k 2 )|κ(~k)|2 + O(λ3 ),
(2π)3
43
R
where we used (2 +τ 2 )2 dτ = 2π. Eventually, we put λ = 1 and we obtain
(2.152).
3 Massive photons
Let m > 0. In this section we discuss the quantization of the Proca equation
where
F µν := ∂ µ Aν − ∂ ν Aµ . (3.2)
Beside the free equation, we will also consider the Proca equation interacting
with a given vector function J µ , called an external 4-current:
∂ν J ν (x) = 0. (3.4)
There are several possible approaches to the Proca equation on the classical
and, especially, quantum level. In particular, one can use from the beginning the
reduced phase space, both for the classical description and quantization. This
is the approach that we will treat as the standard one. Alternative approaches
will be discussed later.
89
3.1 Free massive photons
3.1.1 Space of solutions
Let YPr , resp. CYPr denote the set of real, resp. complex smooth space-compact
solutions of the Proca equation
r(y,Λ) are symplectic for Λ ∈ O↑ (1, 3), otherwise they are antisymplectic.
We introduce the functionals Aµ (x) called 4-potentials. They act on ζ ∈ YPr
giving
We also introduce the field tensor and the electric field vector:
~˙ = 0,
(−∆ + m2 )A0 + divA (3.8)
(−2 + m2 )A~ = 0. (3.9)
~˙ − ∂A
~ =A
Taking the divergence of the definition of the electric field E ~ 0 , then
~
using (3.8), we can express A0 in terms of E:
m2 A0 = ~
−divE. (3.10)
Ȧ0 = ~
divA. (3.11)
90
Finally, we have the following version of the evolution equations in terms of
~ A
E, ~ with only first order derivatives:
~˙ =
A ~ − 1 ∂div
E ~ E, ~ (3.12)
m2
~˙
E = −(−∆ + m2 )A ~ A.
~ − ∂div ~ (3.13)
The flux of (3.17) over a Cauchy surface is zero. Hence (3.16) gives the same
symplectic form as (3.14).
The symplectic form on YPr (3.15) can be written as
Z
~ ~x) ∧ E(t,
ωPr = A(t, ~ ~x)d~x.
We have
∂µ ∂ν
{Aµ (x), Aν (y)} = gµν − D(x − y),
m2
where D(x − y) is the Pauli-Jordan function.
91
Indeed, this follows after we insert (3.12), (3.10) and (3.11) into
Z
~˙ ~y ) + Ḋ(t, ~x − ~y )A(0,
~ ~x) =
A(t, D(t, ~x − ~y )A(0, ~ ~y ) d~y ,
Z
A0 (t, ~x) = D(t, ~x − ~y )Ȧ0 (0, ~y ) + Ḋ(t, ~x − ~y )A0 (0, ~y ) d~y ,
~ ~x).
and then we commute them with A0 (0, ~x) and A(0,
hA((ζ))|ρi := ζωρ.
Note that
{A((ζ1 )), A((ζ2 ))} = ζ 1 ωζ 2 .
Z
˙
~ 0 ~ ~ ~
A((ζ)) = −(ζ(t, ~x) − ∂ζ (t, ~x))A(t, ~x) + ζ(t, ~x)E(t, ~x) d~x. (3.19)
Another way of smearing the 4-potentials is also useful. For a space-time vec-
tor valued functions f ∈ Cc∞ (R1,3 , R1,3 ) the corresponding space-time smeared
4-potential is Z
A[f ] := fµ (x)Aµ (x)dx. (3.20)
∂µ ∂ ν
ζµ = −D ∗ fµ + D ∗ fν .
m2
Adding to f µ a derivative ∂ µ χ for χ ∈ Cc∞ (R1,3 ) does not change (3.20).
1 m2
L := − Fµν F µν − Aµ Aµ .
4 2
The resulting Euler-Lagrange equations
∂L ∂L
= ∂µ
∂Aα ∂Aα,µ
92
The canonical stress-energy tensor, which follows directly from the Noether
Theorem, equals
µν ∂L
Tcan = g µν L − A ,ν
∂Aα,µ α
1 m2
= −g µν Fαβ F αβ + Aα Aα + F µα Aα,ν .
4 2
One usually prefers to replace it with the Belifante-Rosenfeld stress-energy ten-
sor. It is defined as
T µν µν
= Tcan − ∂α Σµνα
1 m2
= −g µν Fαβ F αβ + Aα Aα + ∂α F αµ Aν + F µα F να ,
4 2
where
Σµνα = −Σανµ := F µα Aν . (3.21)
On solutions of the Euler-Lagrange equations we have
1 m2
T µν = −g µν Fαβ F αβ + Aα Aα + m2 Aµ Aν + F µα F να ,
4 2
µν
∂µ Tcan = ∂µ T µν = 0.
Using (3.12) and (3.13) we check that H generates the equations of motion and
P~ the translations.
The observables H, P~ are in involution.
93
3.1.6 Diagonalization of the equations of motion
For ~k ∈ R3 , ~k 6= ~0 fix two spatial vectors ~e1 (~k), ~e2 (~k) that form an oriented
orthonormal basis of the plane orthogonal to ~k. Define
1
~e(~k, ±1) := √ ~e1 (~k) ± i~e2 (~k) .
2
Note that
~k × ~e(~k, ±1) = ±i|~k|~e(~k, ±1),
~e(~k, σ) · ~k = 0,
ei (~k, σ)ei (~k, σ 0 ) = δσ,σ0 ,
X ki kj
ei (~k, σ)ej (~k, σ) = δij − .
~k 2
σ=±1
p
Let k ∈ R1,3 with k 0 = ε(~k) = ~k 2 + m2 . Introduce
|~k| ε(~k)~k
u(k, 0) := , , (3.22)
m m|~k|
u(k, ±1) := 0, ~e(~k, ±1) . (3.23)
Note that
Set
Z
~ t (~k)
A ~ ~x)e−i~k~x p d~x ,
=
A(t,
(2π)3
Z
~ t (~k) =
E ~ ~x)e−i~k~x p d~x .
E(t,
(2π)3
~
~˙ t (~k)
A ~ t (~k) + k ~k·E
= E ~ t (~k),
m2
~˙ t (~k)
E = −(~k 2 + m2 )A~ t (~k) + ~k ~k·A
~ t (~k),
the relations
A∗i (~k) = Ai (−~k), Ei∗ (~k) = Ei (−~k),
94
and the Poisson brackets
Set
the relations
We set
s
ε(~k) i
at (k, σ) := At (~k, σ) + q Et (~k, σ),
2
2ε(~k)
s
ε(~k) ∗ ~ i
a∗t (k, σ) := At (k, σ) − q Et∗ (~k, σ).
2 ~
2ε(k)
We will usually write a(k, σ), a∗ (k, σ) for a0 (k, σ), a∗0 (k, σ), so that
~
at (k, σ) = e−itε(k) a(k, σ),
~
a∗t (k, σ) = eitε(k) a∗ (k, σ).
95
The direct definitions of a(k, σ), a∗ (k, σ) are
s !
ε(~k) ~
Z
d~x ~
−ik~ x ~ i ~ ~
a(k, σ) = p e ~e(k, σ)A(0, ~x) + q ~e(k, σ)E(0, ~x) ,
(2π)3 2
2ε(~k)
s !
ε(~k) ~
Z
d~x i~ i
∗
a (k, σ) = p e k~x ~ ~x) − q
~e(k, σ)A(0, ~e(~k, σ)E(0,
~ ~x) .
(2π)3 2
2ε(~k)
X Z
H = d~kε(~k)a∗ (k, σ)a(k, σ),
σ=0,±1
X Z
P~ = d~k~ka∗ (k, σ)a(k, σ),
σ=0,±1
X Z
iω = a∗ (k, σ) ∧ a(k, σ)d~k.
σ=0,±1
96
a(k, σ) can be called plane wave functionals:
Note that
X Z d~k
gµ (x) = q eikx uµ (k, σ)ha(k, σ)|gi.
~
p
σ=0,±1 3
(2π) 2ε(k)
(+)
For g1 , g2 ∈ WPr we define the scalar product
X Z
(g1 |g2 ) := ig 1 ωg2 = ha(k, σ)|g1 iha(k, σ)|g2 id~k. (3.31)
σ=0,±1
(+)
We set ZPr to be the completion of WPr in this scalar product. R1,3 o
O↑ (1, 3) leaves ZPr invariant.
We have
ha(k, σ)|gi = (k, σ|g).
We can identify ZPr with L2 (R3 , C3 ) and rewrite (3.31) as
X Z
(g1 |g2 ) = (k, σ|g1 )(k, σ|g2 )d~k.
σ=0,±1
M µ kµ = N ν kν = 0.
97
The following identity allows us to average over spin and is useful in computa-
tions of scattering cross-sections:
X
M µ uµ (k, σ)uν (k, σ)N ν = M µ Nµ . (3.32)
σ=0,±1
In fact,
X kµ kν
uµ (k, σ)uν (k, σ) = gµν + .
σ=0,±1
m2
(M · k)(N · k)
M µ gµν N ν + .
m2
But
k · M = k · N = 0.
3.1.10 Quantization
We want to construct (H, Ĥ, Ω) satisfying the standard requirements of QM and
a self-adjoint operator-valued distribution R1,3 3 x 7→ µ (x) such that, setting
~ ~˙
Ê = Â − ∂~ Â0 , we have
(1) −∂ µ (∂µ Âν − ∂ν µ ) + m2 Âν (x) = 0;
(2) [Âi (0, ~x), Âj (0, ~y )] = [Êi (0, ~x), Êj (0, ~y )] = 0,
[Âi (0, ~x), Êj (0, ~y )] = iδij δ(~x − ~y );
(3) eitĤ µ (x0 , ~x)e−itĤ = µ (x0 + t, ~x);
(4) Ω is cyclic for µ (x).
The above problem has an essentially unique solution, which we describe
below.
For the Hilbert space we should take the bosonic Fock space H = Γs (ZPr )
and for Ω the Fock vacuum. With ZPr ' L2 (R3 , C3 ) and k on shell we have
creation operators
â∗ (k, σ) = â∗ |k, σ) ,
98
satisfy the required commutation relations. The quantum Hamiltonian and mo-
mentum are
X Z
Ĥ = ε(~k)â∗ (k, σ)â(k, σ)d~k,
σ=0,±1
~ X Z
P̂ = ~kâ∗ (k, σ)â(k, σ)d~k.
σ=0,±1
1,3 ↑
R o O (1, 3) is unitarily implemented on H by U (y, Λ) :=
The group
Γ r(y,Λ) . We have
ZPr
U (y, Λ)µ (x)U (y, Λ)∗ = (Λ−1 )µν Âν (y, Λ)x .
Moreover,
∂µ ∂ν
[µ (x), Âν (y)] = −i gµν − D(x − y).
m2
d~k ik(x−y) X
Z
=(2π)−3 e uµ (k, σ)uµ (k 0 , σ 0 )
2ε(~k) σ=0,±1
Z ~
dk ik(x−y) kµ kν
∂µ ∂ν
−3
=(2π) e gµν + = −i gµν − D(+) (x − y).
2ε(~k) m2 m2
99
Here is a proof of (3.34):
∂ν J ν (x) = 0. (3.36)
In most of this subsection we will assume that (3.35) is Schwartz. In its presence
the Proca equation takes the form
−∂ µ (∂µ Aν − ∂ ν Aµ ) + m2 Aν = −J ν . (3.37)
∂ν Aν = 0. (3.38)
We have therefore
(−2 + m2 )Aµ = −J µ . (3.39)
100
The only dynamical variables are the spatial components, satisfying the
equation
~ = −J.
(∂02 − ∆ + m2 )A ~ (3.40)
The temporal component of (3.37) has no time derivative:
~ + m2 A0 = −J0 .
−∆A0 + ∂0 divA (3.41)
We can introduce
~
E(x) ~˙
= A(x) ~ 0 (x).
− ∂A (3.43)
~
We can compute A0 in terms of E:
1 ~
A0 = − (J0 + divE). (3.44)
m2
The current (3.14) is conserved also in the presence of interaction and defines
the symplectic form (3.15). Consequently, the equal time Poisson brackets are
the same as in the free case:
1 m2
L := − Fµν F µν − Aµ Aµ − Jµ Aµ
4 2
1 1 m2
= − ∂µ Aν ∂ µ Aν + ∂µ Aν ∂ ν Aµ − Aµ Aµ − Jµ Aµ
2 2 2
2 2
1 ~ 2 + 1 (∂A ~ 0 )2 + 1 A
~˙ 2 − A
~ 0 + m A2 − m A
~˙ ∂A ~ 2 − J~A
~ + J 0 A0 .
= − (rotA)
2 2 2 2 0 2
~
As noted before, only spatial components A(x) are dynamical. The conjugate
~
variable is E(x).
The definition of the canonical and Belifante-Rosen energy-momentum ten-
sor are defined analogously as in the free case. In particular, the canonical
101
Hamiltonian density is
∂L(x)
Hcan (x) = −L(x) + Ȧi (x)
∂ Ȧi (x)
1 ~ 2 (x) − 1 (∂A~ 0 )2 (x) + 1 A ~˙ 2 (x)
= (rotA)
2 2 2
m2 2 m2 ~ 2 ~ A(x)~
− A (x) + A (x) + J(x) − J0 (x)A0 (x)
2 0 2
1 ~ 2 (x) + 1 (E)~ 2 (x) + E(x)
~ ∂A ~ 0 (x)
= (rotA)
2 2
m2 2 m2 ~ 2 ~ A(x)~
− A0 (x) + A (x) + J(x) − J0 (x)A0 (x).
2 2
~ 0
We add to it a spatial divergence div E(x)A (x) and use (3.44) to eliminate
A0 , obtaining the usual (Belifante-Rosen) Hamiltonian density
1 ~2 1 ~ 2 (x)
H(x) := E (x) + (rotA)
2 2
2
1 ~ 2 (x) + m A~ 2 (x) + J(x)
~ A(x).
~
+ 2 J 0 − divE)
2m 2
The Hamiltonian
Z Z
H(t) = H(t, ~x)d~x = Hcan (t, ~x)d~x (3.46)
generates the equations of motion. Using the splitting of A ~ and E ~ into the
transversal and longitudinal part as in (A.41) and (A.42), we can rewrite H(t)
as
Z 1
~ tr
2 1~ ~ tr (t, ~x) + J(x)
~ A ~ tr (x)
H(t) = d~x E (t, ~x) + A x)(−∆ + m2 )A
tr (t, ~
2 2
Z 1 1 1 2
+ d~x Elg (t, ~x)2 + 2
J 0 (t, ~x) − (−∆) 2 Elg (t, ~x)
2 2m
m2
+ Alg (t, ~x)2 . (3.47)
2
We can interpret interacting fields as functionals on YPr satisfying
~ ~x) = A
A(0, ~ fr (0, ~x), ~ ~x) = E
E(0, ~ fr (0, ~x).
3.2.3 Quantization
We are looking for operator valued distributions R1,3 3 x 7→ µ (x) satisfying
−∂µ (∂ µ Âν (x) − ∂ ν µ (x)) + m2 Âν (x) = −J ν (x),
˙
having the standard equal time commutation relations with Ê i := Âi − ∂i Â0
[Âi (0, ~x), Âj (0, ~y )] = [Êi (0, ~x), Êj (0, ~y )] = 0,
[Âi (0, ~x), Êj (0, ~y )] = iδij δ(~x − ~y ).
102
~ ~
We will assume that Â, Ê coincide with free fields at t = 0:
We have
Z 0 Z t
µ (t, ~x) := Texp −i Ĥ(s)ds µ (~x)Texp −i Ĥ(s)ds ,
t 0
where the Hamiltonian Ĥ(t), and the corresponding Hamiltonian in the inter-
action picture are
Z 1
~ 1 ~
Ĥ(t) = d~x : Ê 2 (~x) + (J 0 (t, ~x) − divÊ(~x))2
2 2m2
1 ~ m2 ~ 2 ~
~ ~x)Â(~
+ (rotÂ)2 (~x) + Â (~x) + J(t, x) :, (3.48)
Z 2 2
1 ~ ~ ~x)Â ~ 1 0
ĤInt (t) = d~x − 2 J 0 (t, ~x)divÊfr (t, ~x) + J(t, fr (t, ~
x) + J (t, ~
x ) 2
.
m 2m2
Recall that by (3.10)
~
m2 Â0fr = divÊfr . (3.49)
Using this we express ĤInt (t) in terms of creation/annihilation operators:
Z 1 0
ĤInt (t) = d~x Jµ (t, ~x)µfr (t, ~x) + J (t, ~
x ) 2
2m2
d~k
Z
~
= q eitε(k) Jµ (t, ~k)uµ (k, σ)â∗ (k, σ)
(2π)3 2ε(~k)
p
Z Z
i dk µ 0 ν i 0 2
Ŝ = exp − J (k)Dµν (k)J (k) − J (x) dx
2 (2π)4 2m2
X Z d~k u µ (k, σ)
× exp −i p a∗ (k, σ) q J µ (ε(~k), ~k)
(2π) 3
σ=0,±1 ~
2ε(k)
X Z ~
dk uµ (k, σ) µ ~ ~
× exp −i p a(k, σ) q J (ε(k), k) , (3.50)
(2π) 3
σ=0,±1 2ε(~k)
where
0 1 kµ kν
Dµν (k) = 2 gµν + . (3.51)
m + k 2 − i0 m2
103
0
(The superscript 0 over Dµν will be explained later on).
For xN , . . . , x1 , the N -point Green’s function is defined as follows:
hµN (xN ) . . . µ1 (x1 )i
:= Ω+ |T µN (xN ) · · · µ1 (x1 ) Ω− .
+ + − −
uµ1 (k1+ , σ1+ ) · · · uµn+ (kn++ , σn++ )uµn− (kn−− , σn−− ) · · · uµ1 (k1− , σ1− )
= q q q q
n+ +n−
(2π) 2 2ε(k1+ ) · · · 2ε(kn++ ) 2ε(kn−− ) · · · 2ε(k1− )
×hµ+ (k1+ ) · · · µ+ (kn++ )µ− (−kn−− ) · · · µ− (−k1− )iamp .
1 n+ n− 1
104
For any α ∈ R, we can pass to the following propagators
α 1 kµ kν
Dµν = gµν + (1 − α) 2 .
m2 + k 2 − i0 αk + m2
The above propagator for α = 0 was obtained in the Hamiltonian approach, see
(3.34). For α = 1 we obtain the so-called propagator in the Feynman gauge
Feyn 1
Dµν (k) = .
m2 + k 2 − i0
α = ∞ corresponds to the propagator in the Landau or Lorentz gauge:
Lan 1 kµ kν
Dµν = gµν − .
m2 + k 2 − i0 k2
k0 ki
f0Yuk (k) = , fiYuk (k) = − .
(k 2 + m2 − i0)2(m2 + ~k 2 ) (k 2 + m2 − i0)2(m2 + ~k 2 )
(The propagator in the Yukawa gauge is the massive analog of the propagator
in the Coulomb gauge.)
The propagator in the temporal gauge is
tem tem tem 1 ki kj
D00 = 0, D0j = 0, Dij = 2 δij − .
k + m2 − i0 k02
tem Feyn
We have Dµν = Dµν + kµ fνtem (k) + fµtem (k)kν , where
1 ki
f0tem (k) = , fitem (k) = − 2 .
(m2 + k 2 − i0)2k0 (m + k 2 − i0)2k02
105
d4 k
(4) For internal lines we integrate over the variables with the measure (2π)4 .
To compute scattering amplitudes with N − incoming and N + outgoing par-
ticles we draw the same diagrams as for N − + N + -point Green’s functions. The
rules are changed only concerning the external lines.
(i) With each incoming external line we associate √ 1 u(k, σ).
(2π)3 2ε(~
k)
1
(ii) With each outgoing external line we associate √ u(k, σ).
(2π)3 2ε(~
k)
0
If we prefer, we can use a different causal propagator instead of Dµν . Green’s
functions change, because of external lines, however scattering amplitudes do
not.
Z(f ) (3.52)
−2
Z
i (gµν + m k k
µ ν )
= exp J µ (k) + f µ (k) (J ν (k) + f ν (k))dk .
2 (k 2 + m2 − i0)
0 g +m−2 k k 0
Note that Dµν (k) = µνk2 +m2 −i0
µ ν
, or in the position representation Dµν =
−2 F µν 2 µ ν
(gµν − m ∂µ ∂ν )D is one of the inverses of g (−2 + m ) + ∂ ∂ . Therefore,
(3.52) is often formally rewritten as
Π Π dAµ (x) exp i L(x) − (J µ (x) + f µ (x))Aµ (x) dx
R R
µ x
Z(f ) = R R .
Π Π dAµ (x) exp i Lfr (x)dx
µ x
•
Let Dµν be one of the propagators considered in Subsubect. 3.2.4. Let B•µν
be its inverse. We have the corresponding “free action”
Z
1
T•fr = − Aµ (x)B•µν (x − y)Aν (y)dxdy.
2
106
We define the corresponding generating function as
Z• (f ) (3.53)
Z
i
:= exp (J µ (k) + f µ (k)D•µν (k)(J ν (k) + f ν (k))dk
2
Z
i µ µ ν ν
= exp (J (x) + f (x))D•µν (x − y)(J (y) + f (y))dxdy
2
Π Π dAµ (x) exp iT•fr + i (J µ (x) + f µ (x))Aµ (x)dx
R R
µ x
= R .
Π Π dAµ (x) exp (iT•fr )
µ x
•
In general, Z• (f ) differs for various propagators Dµν , unless f satisfies the
Lorentz condition. However, all Z• (f ) can be used to compute the same scat-
tering operator.
Likewise, the Euler-Lagrange equations obtained from those various action
integrals differ from the Proca equation. However, YPr belong always to their
solutions.
If we take the Lagrangian
1
− ∂µ Aν (x)∂ µ Aν (x) + m2 Aν (x)Aν (x)
2
+(α − 1)∂µ Aµ (x)∂ν Aν (x) , (3.54)
α
then we obtain the propagator Dµν . Indeed,
g µν (k 2 + m2 ) + (α − 1)k µ k ν
α
is the inverse of Dµν (k).
If we restrict the integration by the Lorentz condition
∂µ Aµ (x) = 0. (3.55)
and take the free Lagrangian (3.54) (they now coincide for all α), then we obtain
the propagator in the Landau/Lorentz gauge.
If we take the free Lagrangian
1
− ∂µ Ai (x)∂ µ Ai (x) + m2 Ai (x)Ai (x)
2
1
+ 2 ∂µ ∂i Ai (x)∂ µ ∂j Aj (x) + ∂i Ai (x)∂j Aj (x)
m
−∂i A0 (x)∂i A0 (x) − m2 A0 (x)2 ,
Yuk
we obtain Dµν . Indeed,
ki kj
2 2
(k + m ) δij + 2 − δµ0 δ0ν (~k 2 + m2 )
m
107
Yuk
is the inverse of Dµν (k).
If we take the free action
Z
1
∂µ Ai (x)∂ µ Ai (x) + m2 Ai (x)Ai (x) dx
−
2
Z
1
− ∂µ ∂i Ai (x)(−2)−1 (x − y)∂ µ ∂j Aj (y)
2
+∂i Ai (x)(−2)−1 (x − y)∂j Aj (y) dxdy,
tem
(which is nonlocal and does not involve A0 ), we obtain Dµν . Indeed,
ki kj
(k 2 + m2 ) δij − 2
k
tem
is the inverse of Dij (k).
108
for vector fields (3.6) together with the Lorentz condition (3.7). This suggests
an alternative approach to the massive photons.
In this approach one considers first the Klein-Gordon equation on functions
with values in R1,3 :
(−2 + m2 )ζµ (x) = 0. (3.61)
The space of smooth real space-compact solutions of (3.61) will be denoted by
Yvec . The following 4-current
µ
jvec (ζ1 , ζ2 , x) := ∂ µ ζ1,ν (x)ζ2ν (x) − ζ1,ν (x)∂ µ ζ2ν (x)
is conserved, that is
µ
∂µ jvec (x) = 0.
It defines in the usual way a symplectic form on Yvec
Z
µ
ζ1 ωvec ζ2 = jvec (ζ1 , ζ2 , x)dsµ (x)
S
Z
= −ζ̇1ν (t, ~x)ζ2ν (t, ~x) + ζ1ν (t, ~x)ζ̇2ν (t, ~x) d~x,
We clearly have
(−2 + m2 )Aµ (x) = 0. (3.62)
We can use the Lagrangian
1 m2
L(x) := − Aµ,ν (x)Aµ,ν (x) − Aµ (x)Aµ (x).
2 2
The conjugate variables are
∂
Πµ (x) := L(x) = Ȧµ (x).
∂ Ȧµ (x)
109
The Hamiltonian and momentum density are
1 1 m2
H(x) = T 00 (x) = Πµ (x)Πµ (x) + Aµ,i (x)Aµ,i (x) + Aµ (x)Aµ (x),
2 2 2
P i (x) = T 0i (x) = −Πµ (x)Aµ,i (x).
110
Set
Z
~ t (~k)
A = ~ ~x)e−i~k~x p d~x ,
A(t,
(2π)3
Z
~ t (~k) =
Π ~ ~x)e−i~k~x p d~x .
Π(t,
(2π)3
We have the equations of motion
Ȧt (~k) = Πt (~k),
Π̇t (~k) = −ε(~k)2 At (~k),
the relations
A∗t (~k) = At (−~k), Π∗t (~k) = Πt (−~k),
and the Poisson brackets
{A∗tµ (~k), Atν (~k 0 )} = {Π∗tµ (~k), Πtν (~k 0 )} = 0,
{A∗ (~k 0 ), Πtµ (~k)}
tν = gµν δ(~k − ~k 0 ). (3.65)
Set
At (~k, σ) := uµ (~k, σ)Aµt (~k),
Πt (~k, σ) := uµ (~k, σ)Πµt (~k).
the relations
A∗t (~k, σ) = At (−~k, −σ), Π∗t (~k, σ) = Πt (−~k, −σ),
and the Poisson brackets
{A∗t (~k, σ), At (~k 0 , σ 0 )} = {Π∗t (~k, σ), Πt (~k 0 , σ 0 )} = 0,
{A∗ (~k, σ), Πt (~k 0 , σ 0 )}
t = κσσ0 δ(~k − ~k 0 ), (3.66)
where κσ,σ0 = 1 for σ = σ 0 = ±1, 0 and κsc,sc = −1. We set
s
ε(~k) i
at (k, σ) := At (~k, σ) − q Πt (~k, σ),
2 ~
2ε(k)
s
ε(~k) ∗ ~ i
a∗t (k, σ) := At (k, σ) + q Π∗t (~k, σ).
2 ~
2ε(k)
111
We have the equations of motion
σ=0,±1
X Z Z
P~ = d~k~ka∗ (k, σ)a(k, σ) − d~k~ka∗ (k, sc)a(k, sc).
σ=0,±1
d~k
X Z
gµ (x) = q eikx uµ (k, σ)ha(k, σ)|gi.
~
p
σ=0,±1,sc 3
(2π) 2ε(k)
(+)
For g1 , g2 ∈ Wvec we have a natural scalar product
X Z
(g1 |g2 ) := ig 1 ωvec g2 = ha(k, σ)|g1 iha(k, σ)|g2 id~k
σ=0,±1
Z
− ha(k, sc)|g1 iha(k, sc)|g2 id~k
Z
= g µν haµ (k)|g1 ihaν (k)|g2 id~k. (3.67)
112
Unfortunately, the above definition gives an indefinite scalar product. We can
also introduce a positive definite scalar product, which unfortunately is not
covariant:
XZ
(g1 |g2 )+ := haµ (k)|g1 ihaµ (k)|g2 id~k.
µ
(+)
The positive frequency space Wvec equipped with the scalar product (3.67) can
be completed in the norm given by (·|·)+ . It will be called Zvec . It is an example
of the so-called Krein space, which is a space with an indefinite scalar product
and has a topology given by a positive scalar product.
(+)
Using the projection (3.64), Wvec can be decomposed into the direct sum
(+) (+) (+)
of orthogonal subspaces WLor and Wsc . On WLor the scalar product (3.67)
(+)
is positive definite, on Wsc it is negative definite. Their completions will be
denoted ZLor and Zsc .
Every ζ ∈ Yvec can be uniquely written as ζ = ζ (+) + ζ (+) , where ζ (+) ∈
(+)
Wvec . This allows us to define a real scalar product on Yvec :
(+) (+)
hζ1 |ζ2 iY := Re(ζ |ζ2 ) (3.68)
Z Z1
= ζ̇1µ (0, ~x)(−i)D(+) (0, ~x − ~y )ζ̇2µ (0, ~y )d~xd~y
Z Z
+ ζ1µ (0, ~x)(−∆~x + m2 )(−i)D(+) (0, ~x − ~y )ζ2µ (0, ~y )d~xd~y .
Γs (ZLor ⊕ Z sc ) (3.69)
113
satisfy the standard commutation relations
d~k
Z
+ p q uµ (k, sc)eikx b̂∗ (k, sc) + uµ (k, sc)e−ikx b̂(k, sc) ,
(2π)3 2ε(~k)
X Z Z
Ĥ = d~kε(~k)â∗ (k, σ)â(k, σ) − d~kε(~k)b̂∗ (k, sc)b̂(k, sc),
σ=0,±1
X Z Z
P̂ = d~k~kâ∗ (k, σ)â(k, σ) − d~k~k b̂∗ (k, sc)b̂(k, sc).
σ=0,±1
114
It will be convenient to describe this method in the C ∗ -algebraic language.
Let CCR(Yvec ) denote the (Weyl) C ∗ -algebra of the CCR over Yvec , that is, the
C ∗ -algebra generated by W (ζ), ζ ∈ Yvec , such that
ζ1 ωvec ζ2
W (ζ1 )W (ζ2 ) = e−i 2 W (ζ1 + ζ2 ), W (ζ)∗ = W (−ζ). (3.71)
πψ (W (ζ)) = eiÂ((ζ)) .
â(k, σ)Ω = 0.
The expressions for the Hamiltonian, momentum and 4-potentials are the same
as in the classical case:
X Z Z
Ĥ = d~kε(~k)â∗ (k, σ)â(k, σ) − d~kε(~k)â∗ (k, sc)â(k, sc),
σ=0,±1
~ X Z Z
P̂ = d~k~kâ∗ (k, σ)â(k, σ) − d~k~kâ∗ (k, sc)â(k, sc),
σ=0,±1
XZ d~k
ikx −ikx ∗
µ (x) = q uµ (k, σ)e â(k, σ) + uµ (k, σ)e â (k, σ) .
(2π)3 2ε(~k)
p
σ
115
Note that all eigenvalues of Ĥ are positive, however its expectation values (wrt
the indefinite scalar product) can be negative. We have
4 Massless photons
In this section we discuss the quantization of the Maxwell equation
F µν := ∂ µ Aν − ∂ ν Aµ .
Similarly as in the massive case, there are several possible approaches to the
Maxwell equation on the classical and, especially, quantum level. The approach
based from the beginning on the reduced phase space, both for the classical
description and quantization, will be treated as the standard one. The situation
is however somewhat more complicated than in the massive case, since the
Lorentz condition is not enough to fully reduce the phase space. Alternative
approaches will be discussed later.
We try to make the discussion of massive and massless photons as parallel as
possible. This is not entirely straightforward. In particular, the massless limit is
quite subtle – to describe it one needs to fix the time coordinate. The covariant
4-potential converges as m & 0 in an appropriate sense to the noncovariant
4-potential in the Coulomb gauge.
116
4.1 Free massless photons
4.1.1 Space of solutions and the gauge invariance
It is well known that the Maxwell equation
g /{ζ ∈ YMax
YMax := YMax g : ζ = ∂χ}.
−∆A0 + divA ~˙ = 0,
~ − ∂~ Ȧ0 + ∂div
∂02 − ∆ A ~ A ~
= 0.
2A
~ tr = 0, (4.7)
117
where
~ tr
A := A ~
~ + ∂(−∆)−1 ~
divA.
ζ0Coul := 0, (4.9)
ζ~Coul := ζ~tr = ζ~ + ∂(−∆)
~ −1 ~
divζ. (4.10)
Note that
ζµCoul := ζµ + ∂µ χ, (4.11)
where χ(t, ~x) := (−∆) −1 ~ ~x).
divζ(t, (4.12)
Clearly, ζ Coul is the unique solution of the Maxwell equation gauge equivalent
to ζ satisfying
ζ0Coul = 0, divζ~Coul = 0.
Neither χ nor ζ Coul have to be space-compact. They however
R decay in space
~ ~x)d~x = 0,
directions quite fast. In fact, the Stokes theorem yields that divζ(t,
therefore χ = O(|~x|−2 ) because of (A.39). Hence
ACoul
0 (x) := 0, (4.14)
~ Coul (x) := A
A ~ tr (x) = A(x)
~ ~
+ ∂(−∆) −1 ~
divA(x). (4.15)
Note that
hACoul
µ (x)|ζi = hAµ (x)|ζ Coul i = ζµCoul (x).
ACoul (x) does not depend on the gauge, hence can be interpreted as a functional
on YMax . It is not, however, Lorentz covariant.
Thus the classical 4-potential in the Coulomb gauge ACoul (x) satisfies
ACoul
0 = 0, 2A
~ Coul = 0, ~ Coul = 0.
divA
Clearly,
~ = ∂t A
E ~ Coul , ~
divE(x) = 0. (4.16)
118
4.1.2 Symplectic structure on the space of solutions
We introduce a current defined by the same formula as in the case of the Proca
equation:
j µg (ζ1 , ζ2 , x) (4.17)
Max
µ
∂ν ζ1µ (x) ζ2ν (x) − ζ1ν (x) ∂ µ ζ2ν (x) − ∂ ν ζ2µ (x) .
:= ∂ ζ1ν (x) −
ζ1 ωMax
g ζ2 (4.18)
Z
˙ ~ 10 (t, ~x) ζ~2 (t, ~x) + ζ~1 (t, ~x) ζ~˙2 (t, ~x) − ∂ζ
= − ζ~1 (t, ~x) − ∂ζ ~ 20 (t, ~x) d~x.
However, unlike in the massive case, this form is only presymplectic not sym-
plectic (it is degenerate).
(4.18) does not depend on the gauge. To see this it is enough to note that
for ζ ∈ YMax
g
j µ (∂χ, ζ) = −∂ν χ(∂ µ ζ ν − ∂ ν ζ µ ) .
(4.19)
The presymplectic form can be written as
ζ1 ωMax
g ζ2 = ζ1Coul ω g ζ2Coul (4.20)
Z Max
˙ ˙
= −ζ~1Coul (t, ~x)ζ~2Coul (t, ~x) + ζ~1Coul (t, ~x)ζ~2Coul (t, ~x) d~x.
Note that by (4.13) the integrand of (4.20) behaves as O(|~x|−6 ), hence is inte-
grable.
Proposition 4.1 Let ζ ∈ YMax
g . We have the following equivalence:
(1) ζ ∈ KerωMax
g.
(2) ζ Coul = 0.
(3) ζ = ∂χ.
∞
There exists a unique ξ ∈ Csc (R4 , R4 ) such that
˙ ~x) = 0, ~u(~x) , ξ(0, ~x) = 0, ~v (~x) , 2ξ = 0.
ξ(0,
119
ξ clearly belongs to YMax
g and is in the Coulomb gauge. We have
Coul
ξωMax
gζ = ξωMax
gζ
Z
~˙ ~x)ζ~Coul (0, ~x) + ξ(0,
~ ~x)ζ~˙ Coul (0, ~x) d~x,
= −ξ(0,
g. 2
which equals (4.21) and is nonzero. Hence ζ 6∈ KerωMax
Thus we can write
YMax := YMax g,
g /KerωMax
which means that YMax is obtained by the symplectic reduction of the presym-
plectic space YMax
g . Clearly, YMax is equipped with a natural symplectic form
ωMax . R1,3 o O↑ (1, 3) acts on YMax by symplectic transformations.
In what follows we will usually drop the subscript Coul from ACoul (x). This
introduces a possible ambiguity with A(x) defined in (4.5). However, when we
speak about YMax , then (4.5) is ill defined, only ACoul (x) is well defined, so we
think that the risk of confusion is small.
The symplectic structure on the space YMax can be written as
Z
ωMax = Ai (t, ~x) ∧ Ei (t, ~x)d~x, (4.22)
Note that
{A((ζ1 )), A((ζ2 ))} = ζ 1 ωMax ζ 2 ,
120
Z
A((ζ)) = −ζ̇µ (t, ~x)Aµ (t, ~x) + ζµ (t, ~x)E µ (t, ~x) d~x. (4.25)
Let us stress that A((ζ)) depends on ζ only modulo gauge transformations and
is Lorentz covariant.
We can also introduce space-time smeared 4-potentials in the Coulomb gauge,
which are the functionals on YMax , for f ∈ Cc∞ (R1,3 , R1,3 ) given by
Z
A[f ] := f µ (x)Aµ (x)dx. (4.26)
∂i ∂ j
ζi = −D ∗ fi − fj , ζ0 = 0.
∆
~
(4.26) is not Lorentz covariant. Adding to f µ any function of the form (ρ, ∂χ)
∞ 1,3
for ρ, χ ∈ Cc (R ) does not change (4.26), because of the Coulomb gauge.
µν ∂L
Tcan = g µν L − A ,ν
∂Aα,µ α
1
= −g µν Fαβ F αβ − F µα A,να .
4
One usually replaces it with the Belifante-Rosenfeld stress-energy tensor. It is
defined as
T µν = µν
Tcan + ∂α Σµνα
1
= −g µν Fαβ F αβ + F µα F να ,
4
where
Σµνα = −Σανµ := Fµα Aν . (4.27)
On solutions of the Euler-Lagrange equations we have
∂ µ Tµν
can
= ∂ µ Tµν = 0.
121
~ = 0. We express T can
conjugate to Ai is ∂Ȧi L = E i , which also satisfies divE µν
and Tµν in terms of A~ and E.
~ We introduce the Hamiltonian, momentum
1 ~2
~ 2 (x) ,
H(x) := T 00 (x) = E (x) + (rotA)
2
P j (x) := T 0j (x) = E i (x)F ji (x)..
a(k, σ)
s
ε(~k) −i~x~k
Z
i ~
d~x
= e uj (k, σ)Aj (0, ~x) − q e−ik~x uj (k, σ)E j (0, ~x) p .
2 (2π)3
2ε(~k)
Plane waves are defined as in the massive case, with σ = ±1. We have
and
X Z
Aµ (x) = (x|k, σ)a(k, σ) + (x|k, σ)a∗ (k, σ) d~k.
σ=±1
122
4.1.6 Positive frequency space
(±)
WMax will denote the subspace of CYMax consisting of classes of solutions that
in the Coulomb gauge have positive, resp. negative frequencies.
(+)
Every g ∈ WMax can be written as
X Z d~k
g(x) = q eikx u(k, σ)ha(k, σ)|gi.
~
p
σ=±1 3
(2π) 2ε(k)
(+)
For g1 , g2 ∈ WMax we define the scalar product
X Z
(g1 |g2 ) := ig 1 ωMax g2 = ha(k, σ)|g1 iha(k, σ)|g2 id~k. (4.28)
σ=±1
(+)
The definition of WMax depends on the choice of coordinates. It is however
(+)
easy to see that the space WMax is invariant w.r.t. R1,3 o O↑ (1, 3).
(+)
We set ZMax to be the completion of WMax in this scalar product.
We have
ha(k, σ)|gi = (k, σ|g).
We can identify ZMax with L2 (R3 , C2 ) and rewrite (4.28) as
X Z
(g1 |g2 ) = (k, σ|g1 )(k, σ|g2 )d~k.
σ=±1
(+)
We can identify YMax with WMax and transport the scalar product onto
YMax , which for ζ1 , ζ2 is given by
(+) (+)
hζ1 |ζ2 iY := Re(ζ1 |ζ2 ) (4.29)
Z Z
Coul
= ζ̇1i (0, ~x)(−i)D(+) (0, ~x − ~y )ζ̇2i
Coul
(0, ~y )d~xd~y
Z Z
Coul
+ ζ1i (0, ~x)(−∆~x )(−i)D(+) (0, ~x − ~y )ζ2i Coul
(0, ~y )d~xd~y .
M µ kµ = N ν kν = 0.
Then we have X
M µ uµ (k, σ)uν (k, σ)N ν = M µ Nµ . (4.30)
σ=±1
123
To see (4.30), note that
X ~kµ~kν
uµ (k, σ)uν (k, σ) = gµν + δµ0 δν0 − .
σ=±1 |~k|2
~ ~k)(N
(M ~ ~k)
M µ gµν N ν + M 0 N 0 − .
|~k|2
But
~k M ~ ~k N~
M0 = , N0 = .
|~k| |~k|
4.1.8 Quantization
We would like to quantize the Maxwell equation starting from the symplectic
space YMax . We will use the 4-potentials in the Coulomb gauge (where, as usual,
we drop the superscript Coul). The quantization is similar to the Proca equation
based on YPr described in Subsubsect. 3.2.3, with Condition (1) replaced by
d~k
Z X
Âi (x) := q ui (k, σ)eikx â(k, σ) + ui (k, σ)e−ikx â∗ (k, σ) .
(2π)3 2ε(~k) σ=±1
p
~ X Z
P̂ := â∗ (k, σ)â(k, σ)~kd~k.
σ=±1
124
1,3 ↑
The wholegroup R oO (1, 3) is unitarily implemented on H by U (y, Λ) :=
Γ r(y,Λ) . We have
ZMax
0 0
U (y, Λ)F̂µν (x)U (y, Λ)∗ = Λµµ Λνν F̂µ0 ν 0 (y, Λ)x .
Moreover,
∂i ∂j
[Âj (x), Âi (y)] = −i δij − D(x − y).
∆
Note the identities
∂i ∂j
(Ω|Âi (x)Âj (y)Ω) = −i δij − D(+) (x − y),
∆
∂i ∂j
(Ω|T(Âi (x)Âj (y))Ω) = −i δij − Dc (x − y).
∆
The family
Z
Cc∞ (R1,3 , R1,3 ) 3 f 7→ Â[f ] := f µ (x)µ (x)dx
with D := Γfin
s (ZMax ) does not satisfy the Wightman axioms because of two
problems: the noncausality of the commutator and the absence of the Poincaré
covariance.
If we replace µ with F̂µν , we restore the causality and the Poincaré covari-
ance.
For an open set O ⊂ R1,3 we set
ζ1 ω g ζ2
Max
W (ζ1 )W (ζ2 ) = e−i 2 W (ζ1 + ζ2 ), W (ζ)∗ = W (−ζ).
We are looking for a cyclic representation of this algebra with the time evolution
generated by a positive Hamiltonian.
Consider the state on CCR(YMax g ) defined for ζ ∈ YMaxg by
1
ψ W (ζ) = exp − hζ|ζiY ,
2
125
where hζ|ζiY is defined in (4.29).
Note that the state is gauge and Poincare invariant. Let (Hψ , πψ , Ωψ ) be
the GNS representation. Hψ is naturally isomorphic to Γs (ZMax ). Ωψ can
be identified with the vector Ω. πψ (W (ζ)) can be identified with eiÂ((ζ)) . In
particular, if ζ1 and ζ2 are gauge equivalent, then Â((ζ1 )) = Â((ζ2 )). However,
Â(x) in the sense of (4.5) is not well defined.
∂ν J ν (x) = 0. (4.31)
where
ζ~tr := ζ~ − ∂∆
~ −1 divζ,
~
J~tr ~ ~ −1
:= J − ∂∆ divJ. ~
ζ0 = ∆−1 J0 ,
2ζ~ = J~tr ,
divζ~ = 0. (4.36)
126
The Coulomb gauge seems to be the most natural gauge for the Hamiltonian
approach.
Let ζ be a space compact solution of (4.33). Setting
ζµCoul := ζµ + ∂µ χ,
~ ~x), we obtain a solution of (4.36). ζ Coul is the
where χ(t, ~x) := (−∆)−1 divζ(t,
unique solution of (4.36) gauge equivalent to ζ. It does not have to be space
compact.
Thus the classical 4-potential in the Coulomb gauge ACoul (x) satisfies
ACoul
0 = −(−∆)−1 J0 ,
2A
~ Coul = J~tr ,
~ Coul
divA = 0.
~˙ ∂A
~ = A−
The electric field is E ~ 0 . It is easy to see that if we use the Coulomb
~ ~˙ Coul
gauge, then Etr = A .
Similarly as in the previous subsection, we will drop the superscript Coul in
what follows.
127
generates the equations of motion.
The longitudinal part of A ~ does not enter in the Hamiltonian because of
(A.45). The longitudinal part of E~ is fixed by the constraint (4.37). Therefore,
we can impose the Coulomb gauge, so that only the transversal part of A ~=A ~ tr .
The corresponding conjugate variable is E ~ tr . We can rewrite the Hamiltonian
as
Z
1 ~2 1 ~ 2 1
H(t) = Etr (x) + (∂~ A) ~ A(x)
(x) + J(x) ~ + J0 (−∆)−1 J0 (x) d~x.
2 2 2
4.2.3 Quantization
To quantize the Maxwell equation in the presence of an external 4-current we
will use the Coulomb gauge, dropping as usual the subscript Coul.
We are looking for quantum 4-potentials R1,3 3 x 7→ µ (x) satisfying
ÂCoul
0 = −(−∆)−1 J0 ,
~ Coul
2Â = J~tr ,
~
divÂCoul = 0,
~˙
~
having the following commutation relations with E(x) = Â(x) − ∂~ Â0 (x).
~ ~
The above conditions determine Â0 . To fix  and Ê we assume that they
coincide with their free quantum counterparts at t = 0:
~ ~ ~
Â(0, ~x) = Âfr (0, ~x) =: Â(~x),
~ ~ ~
Ê(0, ~x) = Êfr (0, ~x) =: Ê(~x).
128
The scattering operator can be computed exactly:
Z
i dk µ Coul ν
Ŝ = exp J (k)Dµν (k)J (k)
2 (2π)4
X Z d~k uµ (k, σ)
× exp −i p â∗ (k, σ) q J µ (ε(~k), ~k)
(2π)3
σ=±1 ~
2ε(k)
X Z ~
dk uµ (k, σ) µ ~ ~
× exp −i p â(k, σ) q J (ε(k), k) , (4.39)
(2π)3
σ=±1 ~
2ε(k)
Coul Feyn
We have Dµν = Dµν + kµ fνCoul (k) + fµCoul (k)kν , where
k0 ki
f0Coul (k) = , fiCoul (k) = − .
(k 2 − i0)2~k 2 (k 2 − i0)2~k 2
The propagator in the temporal gauge
tem tem tem 1 ki kj
D00 = 0, D0j = 0, Dij = δij − .
k 2 − i0 k02
tem Feyn
We have Dµν = Dµν + kµ fνtem (k) + fµtem (k)kν , where
1 ki
f0tem (k) = , fitem (k) = − 2 .
(k 2 − i0)2k0 (k − i0)2k02
129
4.2.5 Path integral formulation
Let Dµν•
be one of the propagators considered in Sect. 3.2.4. Let B•µν be its
inverse. Then we can use the corresponding action to express the generating
function by path integrals, as described in Sect. 3.2.6, where this approach for
massive vector fields was considered.
α tem
The discussion of the propagators Dµν and Dµν is an obvious generalization
of the massive case.
Coul
To obtain the propagator in the Coulomb gauge Dµν , we take the La-
grangian
1
− ∂µ Ai (x)∂ µ Ai (x) − ∂i A0 (x)∂i A0 (x) ,
2
and restrict the integration by the condition
~
divA(x) = 0.
where Dc is one of the propagators for the Proca equation and in the expressions
where we use the 3-dimensional integration d~k, the 4-momenta are on shell, that
is, k = (ε(~k), ~k). Using the propagator in the Yukawa gauge and multiplying it
130
by an obvious phase factor we can write it as
Z
i 0 2
Ŝ exp J (x) dx
2m2
Z
i dk i 1 ki kj
= exp J (k) 2 gij − J i (k)
2 (2π)4 m + k 2 − i0 m2 + ~k 2
Z !
i dk 1 0 2
− |J (k)|
2 (2π)4 ~k 2 + m2
X Z ~
dk uµ (k, σ) µ
× exp −i p â∗ (k, σ) q J (k)
(2π) 3
σ=0,±1 ~
2ε(k)
X Z
d~k u µ (k, σ)
× exp −i p â(k, σ) q J µ (k)
(2π) 3
σ=0,±1 ~
2ε(k)
= Ŝtr ⊗ Ŝlg ,
131
!!
k02
Z
i dk 0 2
1 1 1
Ŝlg = exp |J (k)| − −
2 (2π)4 (m2 + k 2 − i0) ~k 2 (m2 + ~k 2 ) (~k 2 + m2 )
d~k
Z
uµ (k, 0) µ
× exp −i p â∗ (k, 0) q J (k)
(2π)3 2ε(~k)
Z
d ~k u µ (k, 0)
× exp −i p â(k, 0) q J µ (k)
(2π)3 2ε(k)~
!
dk m2 |J0 (k)|2 |J0 (k)|2
Z
i
= exp −
2 (2π)4 (m2 + k 2 − i0)~k 2 ~k 2
d~k mJ 0 (k)
Z
× exp i p
â∗ (k, 0) q
(2π)3 |~k| 2ε(~k)
Z
d~k mJ 0 (k)
× exp i p â(k, 0) q
(2π)3 ~
|k| 2ε(k) ~
0 2
dk |J (k)|
Z
i
→ exp − .
2 (2π)4 ~k 2
Thus the renormalized massive scattering operator converges to the massless
scattering operator in the Coulomb gauge as m & 0.
132
in (4.39) with Z
dk µ
Im J (k)Dµν (k)J ν (k). (4.41)
(2π)4
R
(4.41) is infrared divergent if m = 0, q(~x)d~x 6= 0 and ~v+ 6= ~v− .
We could try to justify the use of ŜGL similarly as in Subsubsect. 2.2.12,
by introducing the Gell-Mann–Low adiabatic switching. This justification is
adopted by many physicists, eg. [31]. One could criticize this approach, since
after multiplying by the switching function e−|t| the 4-current is no longer
conserved. Therefore, as indicated above, we prefer to define the scattering
operator ŜGL simply by removing the (typically infinite) phase shift.
133
4.3.2 The Lorentz condition
Recall that the Proca equation is equivalent to the Klein-Gordon equation for
vector fields together with the Lorentz condition. Therefore, one can first de-
velop its theory on the symplectic space Yvec , and then reduce it to the subpace
YLor , as described before.
One can follow a similar route for the Maxwell equation. However, there is
a difference: the reduction by the Lorentz condition is insufficient, one has to
make an additional reduction.
Anyway, let us start, as described in Subsubsect. 3.3.1, by introducing the
space Yvec , the form ωvec , the subspace YLor , the 4-potentials Aµ (x), Πµ (x) :=
Ȧµ (x), where now m = 0.
In the massive case YLor was symplectic (that means, the form ωvec restricted
to YLor was nondegenerate). This is no longer true in the massless case. Instead,
the following is true.
Proof. Using −2∂µ Aµ (x) = 0 we see that, for any fixed t, we can replace
∂µ Aµ (x) = 0 (4.44)
with
It is clear that (4.47) and (4.48 are true. To see (4.49) we compute:
2
YLor is a subspace of YMax
g and on YLor the forms ωMax
g and ωvec coincide.
∂µ ζ µ = ξ+ + ξ− ,
134
ξ− is past space compact and ξ+ is future space compact. By using the advanced
and retarded Green’s functions we can solve
−2χ− = ξ− , −2χ+ = ξ+ ,
Note that the definition (4.50) does not depend on the choice of coordinates and
is invariant wrt. the group R1,3 o O↑ (1, 3).
The scalar product is positive semidefinite, but not strictly positive definite.
(+) (+)
Let WLor,0 be the subspace of elements WLor with a zero norm. Using Prop.
(+) (+) (+)
4.1 we see that WLor,0 consists of pure gauges. The factor space WLor /WLor,0
has a nondegenerate scalar product. Its completion is naturally isomorphic to
the space ZMax , which we constructed in Subsubsect. 4.1.6.
(+)
We have a natural identification of YLor with WLor given by the obvious
projection. For ζ ∈ YLor we will denote by ζ (+) the corresponding element
(+)
of WLor . This identification allows us to define a positive semidefinite scalar
product on YLor :
(+) (+)
hζ1 |ζ2 iY := Re(ζ |ζ2 )
Z Z1
Coul
= ζ̇1i (0, ~x)(−i)D(+) (0, ~x − ~y )ζ̇2i
Coul
(0, ~y )d~xd~y
Z Z
Coul
+ ζ1i (0, ~x)(−∆~x )(−i)D(+) (0, ~x − ~y )ζ2i
Coul
(0, ~y )d~xd~y .
135
this approach there are problems with the 4-potential µ (x). Besides, the full
Hilbert space turns out to be non-separable.
In the Gupta-Bleuler approach the 4-potentials µ (x) are well defined and
covariant. Unfortunately it uses indefinite scalar product spaces.
136
Note that Hψ is non-separable – it is an uncountable direct sum of superselection
sectors corresponding to various values of the Lorentz condition. All these
superselection sectors are separable.
Special role is played by the (separable) subspace (superselection sector)
corresponding to the Lorentz condition Ξ = 0. We can choose ζΞ=0 = 0 and
thus this subspace is naturally isomorphic to Γs (ZMax ) with the fields obtained
by the usual quantization obtained by the method “first reduce, then quantize”.
Note that πψ (W (ζ)) maps between various sectors of (4.52) if ζ 6∈ YLor . The
unitary group R 3 t 7→ πψ (W (tζ)) is strongly continuous if and only if ζ ∈ YLor .
If this is the case, we can write πψ (W (ζ)) = eiÂ((ζ)) . We have Â((ζ1 )) = Â((ζ2 ))
if in addition ζ1 differs from ζ2 by a pure gauge. Â((ζ)) is ill defined if ζ 6∈ YLor .
To my knowledge, the approach that we described above, restricted to the
0th sector, was essentially one of the first approaches to the quantization of
Maxwell equation. It is typical for older presentations, eg. [27]. However,
without the language of C ∗ -algebras it is somewhat awkward to describe. One
usually says that the Lorentz condition ∂µ µ (x) = 0 is enforced on the Hilbert
space of states and constitutes a subsidiary condition.
137
5.1 Free charged scalar bosons
5.1.1 Classical fields
WKG will denote the space of smooth space-compact complex solutions of the
Klein-Gordon equation
(−2 + m2 )ζ = 0. (5.2)
(In the context of neutral fields, it was denoted CYKG , because it was an aux-
iliary object, the complexification of the phase space YKG . Now it is the basic
object, the phase space itself).
Clearly, the space WKG is equipped with a complex conjugation ζ 7→ ζ and
a U (1) symmetry ζ 7→ eiθ ζ, θ ∈ R/2πZ = U (1).
If T is a real linear functional on W, then we have two kinds of natural
complex conjugations of T :
If we decompose elements of WKG into their real and imaginary part ζ = ζR +iζI ,
then the real part of the 4-current splits into a part depending on ζR and on ζI :
Rej µ (ζ 1 , ζ2 , x)
= ∂ µ ζR,1 (x)ζR,2 (x) − ζR,1 (x)∂ µ ζR,2 (x)
+∂ µ ζI,1 (x)ζI,2 (x) − ζI,1 (x)∂ µ ζI,2 (x).
Thus WKG can be viewed as the direct sum of two symplectic spaces with the
form
2Reζ 1 ωζ2 = 2ζR,1 ωζR,2 + 2ζI,1 ωζI,2 .
For x ∈ R1,3 , one can introduce the fields φR (x), φI (x), πR (x), πI (x) as the
real linear functionals on WKG given by
√ √
hφR (x)|ζi := 2Reζ(x), hφI (x)|ζi := 2Imζ(x), (5.5)
√ √
hπR (x)|ζi := 2Reζ̇(x), hπI (x)|ζi := 2Imζ̇(x). (5.6)
Clearly, we have the usual equal time Poisson brackets (we write only the non-
vanishing ones):
{φR (t, ~x), πR (t, ~y )} = {φI (t, ~x), πI (t, ~y )} = δ(~x − ~y ). (5.7)
138
In practice instead of (5.5) and (5.6) one prefers to use complex fields ψ(x), η(x) ∈
W # defined by
Clearly,
1 1
ψ(x) = √ φR (x) + iφI (x) , ψ ∗ (x) = √ φR (x) − iφI (x) ,
2 2
1 1
η(x) = √ πR (x) + iπI (x) , η ∗ (x) = √ πR (x) − iπI (x) .
2 2
Note that
Z Z
ψ(t, ~x) = Ḋ(t, ~x − ~y )ψ(0, ~y )d~y + D(t, ~x − ~y )η(0, ~y )d~y . (5.8)
hψ((ζ))|ρi := ζωρ,
hψ ∗ ((ζ))|ρi := ζωρ, ρ ∈ WKG .
Equivalently,
Z
ψ((ζ)) = −ζ̇(t, ~x)ψ(t, ~x) + ζ(t, ~x)η(t, ~x) d~x,
Z
ψ ∗ ((ζ)) = −ζ̇(t, ~x)ψ ∗ (t, ~x) + ζ(t, ~x)η ∗ (t, ~x) d~x.
Note that
139
We can also introduce space-time smeared fields. To a space-time function
f ∈ Cc∞ (R1,3 , C) we associate
Z
ψ[f ] := f (x)ψ(x)dx,
Z
ψ ∗ [f ] := f (x)ψ ∗ (x)dx.
Clearly,
∂L
η ∗ (x) := = ∂0 ψ ∗ (x),
∂ψ,0 (x)
∂L
η(x) := ∗ (x) = ∂0 ψ(x).
∂ψ,0
∂µ J µ (x) = 0,
µ ∗
J (x) = J µ (x).
140
Up to a coefficient, it coincides with (5.4) viewed as a quadratic form:
hJ µ (x)|ζi = ij µ (ζ, ζ, x)
i ∂ µ ζ(x)ζ(x) − ζ(x)∂ µ ζ(x) .
=
The 0th component of the 4-current is called the charge density
Q(x) := J 0 (x) = i −η ∗ (x)ψ(x) + ψ ∗ (x)η(x) .
141
It is conserved on shell
∂µ T µν (x) = 0.
The components of the stress-energy tensor with the first temporal coordinate
are called the Hamiltonian density and momentum density. We express them
on-shell in terms of ψ(x), ψ ∗ (x), η(x) and η ∗ (x):
H and P~ are the generators of the time and space translations. The observables
H, P1 , P2 , P3 and Q are in involution.
142
5.1.7 Positive and negative frequency subspace
When we discussed neutral scalar fields we introduced positive/negative fre-
quency spaces, which in the notation used in the charged case can be defined
by
(+)
WKG := {g ∈ CYKG : (p|ωg = 0, p0 < 0},
(−) (+)
WKG := WKG = {g ∈ CYKG : (p|ωg = 0, p0 > 0}.
(−)
Instead of WKG for quantization we will use the corresponding complex
(−)
conjugate space denoted WKG and equipped with the scalar product
(−) (−) (−) (−)
(ζ 1 |ζ 2 ) := iζ1 ωζ2 id~
p. (5.18)
(−) (+)
Note that WKG = WKG , where we use the usual (internal) complex conju-
gation in WKG . In paticular,
| − p) = |p). (5.19)
(−) (+)
Therefore in principle we could identify ZKG and ZKG . This identification will
be important for the definition of the charge conjugation. Normally, however,
(−) (+)
we treat ZKG and ZKG as two separate Hilbert spaces.
1,3 ↑ (+) (−)
R o O (1, 3) acts on ZKG and ZKG in a natural way.
143
5.1.8 Plane wave functionals
Plane wave functionals are defined as linear or antilinear functionals on the
complex space WKG , for any ζ ∈ WKG given by
ha(p)|ζi = i(p|ωζ = (p|ζ (+) ), (5.20)
∗
ha (p)|ζi = −i(p|ωζ = (p|ζ (+) ), (5.21)
hb(p)|ζi = i(−p|ωζ = (−p|ζ (−) ), (5.22)
∗
hb (p)|ζi = −i(−p|ωζ = (−p|ζ (−) ). (5.23)
Thus
Z r
E(~
p) i d~x
a(p) = ψ(0, ~x) + p η(0, ~x) e−i~p~x p ,
2 2E(~ p) (2π)3
Z r
∗ E(~
p) ∗ i d~x
a (p) = ψ (0, ~x) − p η ∗ (0, ~x) ei~p~x p ,
2 2E(~ p) (2π)3
Z r
E(~
p) ∗ i d~x
b(p) = ψ (0, ~x) + p η ∗ (0, ~x) e−i~p~x p ,
2 2E(~ p) (2π)3
Z r
∗ E(~
p) i d~x
b (p) = ψ(0, ~x) − p η(0, ~x) ei~p~x p ,
2 2E(~ p) (2π)3
The only non-vanishing Poisson bracket are
{a(p), a∗ (p0 )} = {b(p), b∗ (p0 )} = −iδ(~
p − p~0 ).
We have the following expressions for the fields:
Z
d~p
eipx a(p) + e−ipx b∗ (p) ,
ψ(x) = p p
3
(2π) 2E(~ p)
Z p
p E(~
d~ p)
√ eipx a(p) − e−ipx b∗ (p) .
η(x) = p
i (2π)3 2
We have accomplished the diagonalization of the basic observables:
Z
p) a∗ (p)a(p) + b∗ (p)b(p) ,
H = d~
pE(~
Z
~ pp~ a∗ (p)a(p) + b∗ (p)b(p) ,
P = d~
Z
p a∗ (p)a(p) − b∗ (p)b(p) .
Q = d~
144
5.1.9 Quantization
In principle, we could quantize the complex Klein-Gordon equation as a pair of
real Klein-Gordon fields. However, we will use the formalism of quantization of
charged bosonic systems [15].
We want to construct (H, Ĥ, Ω) satisfying the usual requirements of QM
(1)-(3) and an operator valued distribution
[ψ̂(0, ~x), η̂ ∗ (0, ~y )] = iδ(~x − ~y ), [ψ̂ ∗ (0, ~x), η̂(0, ~y )] = iδ(~x − ~y ); (5.25)
145
Equivalently, for any t
Z
Ĥ = : η̂ ∗ (t, ~x)η̂(t, ~x) + ∂~ ψ̂ ∗ (t, ~x)∂~ ψ̂(t, ~x) + m2 ψ̂ ∗ (t, ~x)ψ̂(t, ~x) :d~x,
Z
~
P̂ = : −η̂ ∗ (t, ~x)∂~ ψ̂(t, ~x) − ∂~ ψ̂ ∗ (t, ~x)η̂(t, ~x) :d~x,
Z
Q̂ = i : −η̂ ∗ (t, ~x)ψ̂(t, ~x) + ψ̂ ∗ (t, ~x)η̂(t, ~x) :d~x.
Thus all these operators are expressed in terms of the Wick quantization of their
classical expressions.
Note thatthe whole
R1,3 o O↑ (1, 3) acts unitarily on H by U (y, Λ) :=
group
Γ r(y,Λ) (+)
⊗ Γ r(y,Λ) (−)
, with
ZKG ZKG
Moreover,
[ψ̂(x), ψ̂ ∗ (y)] = −iD(x − y), [ψ̂(x), ψ̂(y)] = 0.
Note the identities
The observable algebra A(O) is the subalgebra of F(O) fixed by the automor-
phism
B 7→ eiθQ̂ Be−iθQ̂ .
The algebras F(O) and A(O) satisfy the Haag-Kastler axioms.
146
5.1.10 Quantum 4-current
Let us try to introduce the (quantum) 4-current density by
i µ ∗
Jˆµ (x) = ∂ ψ̂ (x)ψ̂(x) + ψ̂(x)∂ µ ψ̂ ∗ (x)
2
−ψ̂ ∗ (x)∂ µ ψ̂(x) − ∂ µ ψ̂(x)ψ̂ ∗ (x) . (5.29)
(Ω|Jˆµ (x)Ω) = 0.
Thus Jˆµ (x) can be defined both as the Weyl quantization (5.29) and the Wick
quantization (5.30) of the corresponding quadratic classical expression.
Formally, we can check the relations
∂ µ Jˆµ (x) = 0,
Jˆµ (x)∗ = Jˆµ (x).
Similarly, as in the classical case, for χ ∈ Cc∞ (R3 , R), let αχ denote the cor-
responding gauge transformation at time t = 0 defined as the ∗-automorphism
of the algebra generated by the fields operators satisfying
Obviously,
147
Assume that χ 6= 0. Let us check whether αχ is unitarily implementable.
On the level of annihilation operators we have
Z Z s s !
E(~
p1 ) E(~p) d~xd~
p1 i(~p1 −~p)~x−ieχ(~x)
αχ (â(p)) = + e â(p1 )
E(~p) p1 ) 2(2π)3
E(~
Z Z s s !
E(~
p1 ) E(~p) d~xd~
p1 −i(~p1 +~p)~x−ieχ(~x) ∗
+ − e b̂ (p1 ).
E(~p) p1 ) 2(2π)3
E(~
Let qχ (~
p, p~1 ) denote the integral kernel on the second line above. By the Shale
criterion (Thm A.2), we need to check whether it is square integrable. Now
s s !
E(~p1 ) E(~p)
− (5.34)
E(~
p) E(~
p1 )
p1 | − |~
(|~ p|)(|~
p1 | + |~
p|)
= p .
E(~
p) + E(~ p1 ) E(~ p)E(~ p1 )
But we know that αχ is not implementable. Thus for nonzero χ (5.35) cannot
be defined as a closable operator.
However, the (quantum) charge
Z
Q̂ = Q̂(t, ~x)d~x,
148
of creation and annihilation operators:
Z Z s s !
d~p1 d~
p2 E(~ p1 ) E(~p2 )
Q̂(x) = +
2(2π)3 E(~ p2 ) E(~p1 )
× e−ixp1 +ixp2 â∗ (p1 )â(p2 ) − eixp1 −ixp2 b̂∗ (p2 )b̂(p1 )
Z Z s s !
d~p1 d~
p2 E(~ p1 ) E(~ p2 )
+ −
2(2π)3 E(~ p2 ) E(~ p1 )
× −e−ixp1 −ixp2 â∗ (p1 )b̂∗ (p2 ) + eixp1 +ixp2 b̂(p1 )â(p2 ) ,
Z Z
~ d~
p d~ p
Ĵ (x) = p1 2 (~
p1 + p~2 )
2(2π)3 E(~ p1 )E(~p2 )
× −e−ixp1 +ixp2 â∗ (p1 )â(p2 ) + eixp1 −ixp2 b̂∗ (p2 )b̂(p1 )
Z Z
d~
p d~p
+ p1 2 p1 − p~2 )
(~
3
2(2π) E(~ p1 )E(~p2 )
× −e−ixp1 −ixp2 â∗ (p1 )b̂∗ (p2 ) + eixp1 +ixp2 b̂(p1 )â(p2 ) .
149
If ψ satisfies (5.38) and R1,3 3 x 7→ χ(x) ∈ R is smooth, then e−ieχ ψ satisfies
(5.38) with A replaced with A + ∂χ.
In this subsection, the field satisfying the Klein-Gordon equation with A = 0
will be denoted ψfr .
The retarded/advanced Green’s function is defined as the unique solution of
satisfying
suppD± ⊂ {x, y : x ∈ J ± (y)}.
We generalize the Pauli-Jordan function:
Clearly,
suppD ⊂ {x, y : x ∈ J(y)}.
The Cauchy problem of (5.38) can be expressed with help of the function D:
Z
ψ(t, ~x) = − ∂s D(t, ~x; s, ~y ) s=0 ψ(0, ~y )d~y (5.40)
R3
Z
+ D(t, ~x; 0, ~y )ψ̇(0, ~y )d~y .
R3
For definiteness, we will assume that ψ(x), η(x) act on WKG and at time t = 0
coincide with free fields:
150
Let us introduce the variable conjugate to ψ ∗ (x) and ψ(x):
∂L
η(x) := ∗ (x) = ∂0 ψ(x) + ieA0 (x)ψ(x),
∂ψ,0
∂L
η ∗ (x) = = ∂0 ψ ∗ (x) − ieA0 (x)ψ ∗ (x).
∂ψ,0 (x)
∂L(x) ∂L(x) ∗
H(x) = ψ̇(x) + ψ̇ (x) − L(x)
∂ ψ̇(x) ∂ ψ̇ ∗ (x)
= η ∗ (x)η(x) + ieA0 (x) (ψ ∗ (x)η(x) − η ∗ (x)ψ(x))
+(∂i − ieAi (x))ψ ∗ (x)(∂i + ieAi (x))ψ(x) + m2 ψ ∗ (x)ψ(x)
= η ∗ (x)η(x) + ∂i ψ ∗ (x)∂i ψ(x)
+ieA0 (x) (ψ ∗ (x)η(x) − η ∗ (x)ψ(x)) − ieAi (x) (ψ ∗ (x)∂i ψ(x) − ∂i ψ ∗ (x)ψ(x))
~ 2 ψ ∗ (x)ψ(x) + m2 ψ ∗ (x)ψ(x).
+e2 A(x)
The Hamiltonian Z
H(t) = H(t, ~x)d~x
ψ̇(t, ~x) = {ψ(t, ~x), H(t)}, η̇(t, ~x) = {η(t, ~x), H(t)}.
151
Z Z s s !
e d~p1 d~
p2 E(~
p1 ) E(~
p2 )
= +
2 (2π)3 E(~
p2 ) E(~
p1 )
× A0 (t, p~1 − p~2 )eitE(~p1 )−itE(~p2 ) a∗ (p1 )a(p2 ) − A0 (t, −~ p1 + p~2 )e−itE(~p1 )+itE(~p2 ) b(p1 )b∗ (p2 )
Z Z s s !
e d~p1 d~
p2 E(~ p1 ) E(~ p2 )
+ −
2 (2π)3 E(~ p2 ) E(~ p1 )
× A0 (t, p~1 + p~2 )eitE(~p1 )+itE(~p2 ) a∗ (p1 )b∗ (p2 ) − A0 (t, −~ p1 − p~2 )e−itE(~p1 )−itE(~p2 ) b(p1 )a(p2 )
Z Z
e d~
p d~ p
+ p 1 2 (~
p1 + p~2 )
2 3
(2π) E(~ p1 )E(~ p2 )
× −A(t,~ p~1 − p~2 )eitE(~p1 )−itE(~p2 ) a∗ (p1 )a(p2 ) + A(t, ~ −~ p1 + p~2 )e−itE(~p1 )+itE(~p2 ) b(p1 )b∗ (p2 )
Z Z
e d~
p d~ p
+ p 1 2 p1 − p~2 )
(~
2 3
(2π) E(~ p1 )E(~ p2 )
× −A(t,~ p~1 + p~2 )eitE(~p1 )+itE(~p2 ) a∗ (p1 )b∗ (p2 ) + A(t, ~ −~ p1 − p~2 )e−itE(~p1 )−itE(~p2 ) b(p1 )a(p2 )
e2
Z Z
d~p d~ p
+ p 1 2p
2 3
(2π) E(~ p1 ) E(~ p2 )
× A ~ (t, p~1 − p~2 )e
2 itE(~p 1 )−itE(~p2) ∗
a (p1 )a(p2 ) + A ~ 2 (t, −~
p1 + p~2 )e−itE(~p1 )+itE(~p2 ) b(p1 )b∗ (p2 )
+A~ 2 (t, p~1 + p~2 )eitE(~p1 )+itE(~p2 ) a∗ (p1 )b∗ (p2 ) + A
~ 2 (t, −~p1 − p~2 )e−itE(~p1 )−itE(~p2 ) b(p1 )a(p2 ) .
~
with ξP ψ(Px), ξP ψ ∗ (Px), A0 (Px), −A(Px)
~
with ξT ψ(Tx), ξT ψ ∗ (Tx), A0 (Tx), −A(Tx)
152
is a symmetry of (5.38) called time reversal and denoted T .
Let ξX := ξP ξT . The composition of C, P and T consists in replacing
ψ(x), ψ ∗ (x), A(x)
with ξX ψ ∗ (−x), ξX ψ(−x), −A(−x).
It is called the CPT symetry and is denoted X .
C, P, T and X commute with one another and we have the relations
C 2 = P 2 = T 2 = X 2 = id.
5.2.4 Quantization
We are looking for a quantum field satisfying
−(∂µ + ieAµ (x))(∂ µ + ieAµ (x)) + m2 ψ̂(x) = 0.
(5.43)
We set
η̂(x) := ∂0 ψ̂(x) + ieA0 (x)ψ̂(x).
We will assume that ψ̂, η̂ act on the Hilbert space of free fields
(+) (−)
Γs (ZKG ⊕ ZKG ),
and at time t = 0 we have
ψ̂(~x) := ψ̂(0, ~x) = ψ̂fr (0, ~x),
η̂(~x) := η̂(0, ~x) = η̂fr (0, ~x).
The solution is unique and is obtained by decorating (5.41) with “hats”.
We would like to ask whether the quantum fields are implemented by a
unitary dynamics. Equivalently, we want to check if the classical dynamics
generated by HInt (t) satisfies the Shale criterion (Thm A.2).
By following the discussion of Subsubsect. 2.3.4 we check that the classical
scattering operator is unitarily implementable.
The Shale criterion is satisfied for the dynamics from t− to t+ iff the spatial
part of the 4-potential is the same at the initial and final time:
~ + , ~x) = A(t
A(t ~ − , ~x), ~x ∈ R3 . (5.44)
To see this note that HInt (t) consists of three terms described in (5.42).
~ ~x)2 ψ ∗ (t, ~x)ψfr (t, ~x) is very similar to the mass-like pertur-
The term e2 A(t, fr
bation considered already in Subsubsect. 2.3.4, which did not cause problems
with the Shale criterion for the dynamics for any t+ , t− .
The same is true for the term eA0 (t, ~x)Qfr (t, ~x). Indeed, a similar term
was discussed before in the context of gauge transformations, see in particular
(5.34). Then there was a problem with the square integrability. But now we
can integrate by parts, which improves the decay.
~ ~x)J~fr (t, ~x) is problematic – it has worse behavior for large mo-
The term eA(t,
menta than the previous two terms. The integration by parts creates a boundary
term that is not square integrable unless (5.44) holds, when it vanishes.
153
5.2.5 Quantum Hamiltonian
Formally, the fields undergo a unitary dynamics given by
Z 0 Z t
ψ̂(t, ~x) := Texp −i Ĥ(s)ds ψ̂(0, ~x)Texp −i Ĥ(s)ds ,
t 0
Note that the above Hamiltonian is formally the Weyl quantization of its corre-
sponding classical expressions. This is perhaps not obvious the way it is written.
To see this we should note that equal time ψ̂ and ψ̂ ∗ commute, the same is true
for equal time η̂ and η̂ ∗ , finally the mixed term can be expressed by the 4-current
where the Wick and Weyl quantizations coincide, see Subsubsect. 5.1.10.
In any case, the analysis of the previous subsubsection shows that the above
Hamiltonian is often ill defined and should be understood as a formal expression,
even when we try renormalize by adding a constant C(t). We will need it to
develop perturbation expansion for the quantum scattering operator and to
compute the energy shift.
(5.45) can be compared with the free Hamiltonian without the Wick order-
ing, which differs from (5.28) by an (infinite) constant:
Z
Ĥfr = d~x η̂ ∗ (~x)η̂(~x) + ∂i ψ̂ ∗ (~x)∂i ψ̂(~x) + m2 ψ̂ ∗ (~x)ψ̂(~x) . (5.46)
154
First consider the charge conjugation. As we have already pointed out in
(+) (−)
Subsubsect. 5.1.7, the spaces ZKG and ZKG can be naturally identified. There-
(+) (−)
fore, we can define a unitary operator on ZKG ⊕ ZKG
χ(g1 , g 2 ) := (g 2 , g1 ).
Clearly,
χ|p) = | − p), χ| − p) = |p).
We set C := Γ(χ). We have C 2 = 1l, CΩ = Ω,
~ ~
C Q̂fr (x)C −1 = −Q̂fr (x), C Ĵfr (x)C −1 = −Ĵfr (x),
C Ŝ(A)C −1 = Ŝ(−A).
(+) (−)
Define a unitary operator on ZKG ⊕ ZKG
π(g1 , g 2 ) := ξP g1 ◦ P, ξP g2 ◦ P .
Clearly,
π|E, p~) = ξP |E, −~
p), π|−E, −~
p) = ξP |−E, p~).
We have a natural implementation of the parity P := Γ(π). It satisfies P 2 = 1l,
P Ω = Ω,
~ ~
P Q̂fr (x)P −1 = Q̂fr (Px), P Ĵfr (x)P −1 = −Ĵfr (Px),
~ −1 = Ŝ(A0 ◦ P, −A
P Ŝ(A0 , A)P ~ ◦ P).
(+) (−)
Define the following antiunitary operator on ZKG ⊕ ZKG :
τ (g1 , g 2 ) := ξT g1 ◦ T, ξT g2 ◦ T .
Clearly,
τ |E, p~) = ξT |E, −~
p), τ |−E, −~
p) = ξT |−E, p~).
Set T := Γ(τ ). We have T 2 = 1l, T Ω = Ω,
~ ~
T Q̂fr (x)T −1 = Q̂fr (Tx), T Ĵfr (x)T −1 = −Ĵfr (Tx),
~ −1 = Ŝ(A0 ◦ T, −A
T Ŝ(A0 , A)T ~ ◦ T).
155
5.2.7 2N -point Green’s functions
For yN , . . . y1 , xN , . . . , x1 , the 2N point Green’s function are defined as follows:
Z(g, g)
∞ Z
(−1)N ∗
X Z
:= ··· 2
ψ̂ (y1 ) · · · ψ̂ ∗ (yN )ψ̂(xN ) · · · ψ̂(x1 )
n=0
(N !)
×g(y1 ) · · · g(yN )g(xN ) · · · g(x1 )dy1 · · · dyN dxN · · · dx1
Z ∞ Z Z
∗
= Ω Texp −i ĤInt (t)dt − i g(x)ψ̂fr (x)dx − i g(x)ψ̂fr (x)dx Ω .
−∞
Set
|−p0n0 , . . . , −p01 , pn , . . . , p1 ) := b̂∗ (p0n0 ) · · · b̂∗ (p01 )â∗ (pn ) · · · â∗ (p1 )Ω.
One can compute scattering amplitudes from the amputated Green’s functions:
−0 −
−p+0
n +0 , . . . , p +
n + , . . . | Ŝ |−p n −0 , · · · , pn − , · · ·
−0 −
· · · ψ̂(p+ ∗ +0 ∗
n+ ) · · · ψ̂ (pn+0 )ψ̂(−pn−0 ) · · · ψ̂ (−pn− ) · · · amp
= q q q q ,
−0 −
p
(2π)3(n+ +n+0 +n−0 +n− ) · · · 2E(p+ n+ ) · · · 2E(p +0
n+0 ) 2E(p n−0 ) · · · 2E(p n− ) · · ·
where all p± ±0
i , pi are on shell.
156
5.2.8 Path integral formulation
Since the Hamiltonian that we consider is quadratic, we can compute exactly
the generating function in terms of the Fredholm determinant on L2 (R1,3 ):
Z(g, g) (5.48)
−1
det − 2 + m 2
− (∂µ + ieAµ (x))(∂ µ + ieAµ (x)) + m2 − i0
=
−1
× exp ig (∂µ + ieAµ (x))(∂ µ + ieAµ (x)) + m2 − i0 g
−1
= det 1l + −ieAµ (x)∂ µ − ie∂ µ Aµ (x) + e2 Aµ (x)Aµ (x) Dfr
c
c
µ µ 2 µ
c −1
× exp igDfr 1l + −ieAµ (x)∂ − ie∂ Aµ (x) + e Aµ (x)A (x) Dfr g .
Let us stress that the above formulas are based on the formal expression for
the Hamiltonian (5.47) where we used the Weyl quantization, in contrast to
the analogous formula (2.135) for the mass-like perturbation, which were Wick
ordered. The expression is to a large degree ill-defined.
Formally, (5.48) can be rewritten in terms of path integrals as
R
Π dψ ∗ (x) Π0 dψ(y) exp i L(x) − g(x)ψ ∗ (x) − g(x)ψ(x) dx
R
y y
R R .
Π dψ ∗ (y) Π dψ(y 0 ) exp i Lfr (x)dx
y y0
157
d4 p
(4) We integrate over the variables of internal lines with the measure (2π)4 .
It is immediate to derive the Feynman rules for charged scalar bosons from
the path integral formula (5.48).
The derivation of the Feynman rules within the Hamiltonian formalism using
the Dyson expansion of the scattering operator is relatively complicated, since
one has to use not only the two-point functions of “configuration space fields”
ψ, ψ ∗ , but also of conjugate fields η, η ∗ [26]:
∗ c
(Ω|T(ψ̂fr (x)ψ̂fr (y))Ω) = −iDfr (x − y),
∗ c
(Ω|T(η̂fr (x)ψ̂fr (y))Ω) = −i∂x0 Dfr (x − y),
∗ c
(Ω|T(ψ̂fr (x)η̂fr (y))Ω) = −i∂y0 Dfr (x − y),
∗ c
(Ω|T(η̂fr (x)η̂fr (y))Ω) = −i∂x0 ∂y0 Dfr (x − y) − iδ(x − y).
158
Figure 8: Diagram for scattering amplitudes.
Here D` is the value of the loop ` and n` is its symmetry factor. Any such a
loop is described by a cyclic sequence (α1 , . . . , αn ), where αj = 1, 2 correspond
to 1− and 2−photon vertices. The symmery factor n` is the order of the group
of the authomorphisms of this loop. The loop is oriented, hence this group is
always a subgroup of rotations. In particular, if the loop has n identical vertices,
the group is Zn and n` = n.
Actually, it is better to organize (5.49) not in terms of the number of vertices
on a loop but in terms of the order wrt e. Using the unitary charge conjugation
operator C and CΩ = Ω we obtain
(Ω|Ŝ(A)Ω) = (Ω|C Ŝ(A)C −1 Ω) = (Ω|Ŝ(−A)Ω).
Therefore, diagrams of an odd order in e vanish. This is the content of Furry’s
theorem for charged bosons. Hence (5.49) can be written as
∞
X
E= e2n En .
n=1
159
Figure 9: Divergent diagrams for vacuum energy.
The expressions for En obtained from the Feynman rules are convergent for
n ≥ 3. E2 is logarithmically divergent, but its physically relevant gauge invariant
part is convergent. E1 is quadratically divergent and its gauge-invariant part
is logarithmically divergent. It needs an infinite renormalization, which will be
described below.
m20 := m2 , C0 := 1,
m21 := m2 + 2Λ2 , C1 := 1,
m22 2
:= m + Λ , 2
C2 := −2.
Using
2
X 2
X
Ci = Ci m2i = 0 (5.51)
i=0 i=0
we can check that with this choice the sums used in the following computations
are integrable.
In the following formula we have a contribution of the loop with 2 single-
photon vertices and twice the contribution of the loop with a single 2-photon
160
vertex. It is convenient to write the latter as the sum of two terms, equal to
one another.
d4 q X
Z
2 4qµ qν
2ΠµνΛ (p) = ie Ci
(q + 12 p)2 + m2i − i0 (q − 21 p)2 + m2i − i0
(2π)4 i
gµν gµν
− −
(q + 21 p)2 + m2i − i0 (q − 12 p)2 + m2i − i0
∞ ∞
e2 (α1 − α2 )2
Z Z X
= − dα1 dα2 (g p2 − pµ pν )
Ci
4 µν
(4π)2
0 0 i
(α 1 + α2 )
2
!
α1 α2 i m i
+2gµν p2 − +
(α1 + α2 )4 (α1 + α2 )3 (α1 + α2 )2
α1 α2 2
× exp −i(α1 + α2 )m2i − i p
α1 + α2
=: (−gµν p2 + pµ pν )2Πgi 2 gd 2
Λ (p ) + 2ΠµνΛ (p ).
−Πgd 2
µνΛ (p )
Z ∞ Z ∞
e2
X
2 α1 α2 2
= Ci dα1 dα2 exp −i(α1 + α2 )mi − i p
i
(4π)2 0 0 α1 + α2
α1 α2 p2 m2i
i
×gµν − +
(α1 + α2 )4 (α1 + α2 )3 (α1 + α2 )2
2 Z ∞ Z ∞
X e 2 α1 α2 2
= Ci ρ∂ρ dα1 dα2 exp −iρ (α1 + α2 )mi + p
i
(4π)2 0 0 α1 + α2
igµν
×
ρ(α1 + α2 )3 ρ=1
Z ∞ Z ∞
e2
X
2 α1 α2 2
= Ci ρ∂ ρ dα1 dα 2 exp −i (α1 + α2 )m i + p
i
(4π)2 0 0 α1 + α2
igµν
× = 0.
(α1 + α2 )3
161
see (2.136), and we obtain
Z ∞ Z ∞
e2 X (α1 − α2 )2
Πgi
Λ (p2
) = − 2
dα 1 dα2 Ci
2(4π) 0 0 i
(α1 + α2 )4
α1 α2 2
× exp −i(α1 + α2 )m2i − i p
α1 + α2
2 Z 1 Z ∞
(1 − v 2 )p2
e dρ X
2 2
= − dv C i v exp −iρ m i +
2(4π)2 0 0 ρ i 4
1
e2 (1 − v 2 )p2
Z X
= 2
dv Ci v 2 log m2i + − i0 .
2(4π) 0 i
4
We define
Πren (p2 )
! p
e2 1 1+θ 2 2 p2
= log − − , θ= p , 0 < p2 ;
2 · 3(4π)2 θ3 1 − θ 3 θ2 p + 4m2
2
! p
e2 2 2 2 −p2
= arctan θ − − 2 , θ= p , −4m2 < p2 < 0;
2 · 3(4π)2 θ3 3 θ 2
p + 4m 2
! p
e2 1 θ+1 2 2 −p2
= log − iπ − − 2 , θ= p , p2 < −4m2 .
2 · 3(4π)2 θ3 θ−1 3 θ −p − 4m2
2
Hence
1
− Fµν (p)F µν (p) = −p2 |A(p)|2 + |pA(p)|2 . (5.54)
2
Thus the renormalized 1st order contribution to the vacuum energy is
Z
dp
E1ren = − Πren (p2 )Fµν (p)F µν (p). (5.55)
2(2π)4
162
gi
We can formally write Πgi
∞ (k) := lim ΠΛ (k) (which is typically infinite).
Λ→∞
Note that the renormalized scattering operator Ŝren is a well defined unitary
operator:
gi
i
Fµν (x)F µν (x)dx
R
Ŝren = e− 2 Π∞ (0) Ŝ. (5.56)
1 ∞
Z
ren 2 ren 1 1
ReΠ (p ) = dsImΠ (−s) − . (5.58)
π 4m2 s + p2 s
Note that (5.57) is nonzero only for p2 < −4m2 , and then it is negative. For
such p we can find a coordinate system with p = (p0 , ~0). Then
and
~ 0 , ~0)|2 .
−Fµν (p0 , ~0)F µν (p0 , ~0) = p20 |A(p (5.59)
Thus the imaginary part of (5.55) is negative (and is responsible for the decay).
163
5.2.14 Dimensional renormalization
We present an alternative computation of Πren µν based on the dimensional regu-
larization. We use the Euclidean formalism.
d4 q
Z
4qµ qν
2ΠEµν (p) = −e 2
4 1
(2π) ((q + 2 p) + m2 )((q − 12 p)2 + m2 )
2
2gµν
− 2
q + m2
d4 q 4qµ qν − 2gµν (q 2 + 41 p2 + m2 )
Z
= −e2
(2π)4 ((q + 21 p)2 + m2 )((q − 12 p)2 + m2 )
e2 1 d4 q 4qµ qν − 2gµν (q 2 + 14 p2 + m2 )
Z Z
= − dv 2
2 −1 (2π)4 (q 2 + p4 + m2 + vqp)2
p2
d4 q 4qµ qν − 2gµν (q 2 + 41 p2 + m2 ) + v 2 (pµ pν − gµν 2 )
Z 1 Z
= −e2 dv 2 , (5.60)
0 (2π)4 (q 2 + p4 (1 − v 2 ) + m2 )2
where we used the Feynman identity (A.28), replaced q + vp 2 withR q, used the
R1 R1 1
symmetry v → −v to remove −1 dvv and replace 12 −1 dv with 0 dv. After
this preparation, we use the dimensional regularization:
d4 q µ4−d Ωd ∞ d−1
Z Z
is replaced by |q| d|q|, (5.61)
(2π)4 (2π)d 0
Z ∞
qµ qν d4 q µ4−d Ωd
Z
is replaced by gµν |q|d+1 d|q|, (5.62)
(2π)4 d(2π)d 0
164
We can now renormalize (5.63):
This coincides with the Wick rotated result obtained by the Pauli-Villars method.
Then
Z t+
Ŵ (t+ )Û (t+ , t− )Ŵ ∗ (t− ) = Texp − i ĤR (s)ds ,
t−
∂
∂yµ Ŝ(A) = 0.
∂Aµ (y)
165
In the momentum representation these identities read
∂
pµ Ŝ(A) = 0.
∂Aµ (p)
We will write hψ̂ ∗ (x01 ) · · · ψ̂ ∗ (x0N )ψ̂(xN ) · · · ψ̂(x1 )iA to express the depen-
dence of Green’s functions on the external 4-potential A. We have
hψ̂ ∗ (x01 ) · · · ψ̂ ∗ (x0N )ψ̂(xN ) · · · ψ̂(x1 )iA+∂χ (5.66)
0 0
∗
= hψ̂ (x01 ) · · · ψ̂ ∗ (x0N )ψ̂(xN ) · · · ψ̂(x1 )iA eieχ(x1 )+···+ieχ(xN )−ieχ(xN )−···−ieχ(x1 ) .
By differentiating with respect to χ(y) and setting χ = 0 we obtain the Ward(-
Takahashi) identities for Green’s functions in the position representation:
∂
∂yµ hψ̂ ∗ (x01 ) · · · ψ̂ ∗ (x0N )ψ̂(xN ) · · · ψ̂(x1 )iA
∂Aµ (y)
N N
!
X X
0
= e i δ(y − xj ) − i δ(y − xj ) hψ̂ ∗ (x01 ) · · · ψ̂ ∗ (x0N )ψ̂(xN ) · · · ψ̂(x1 )iA .
j=1 j=1
(5.65) and (5.66) are essentially obvious if we use the path integral expres-
sions. It is instructive to derive these statements also in the Hamiltonian for-
malism. This derivation is not fully rigorous, since transformations cannot be
implemented, and in general the dynamics does not have a well defined Hamil-
tonian.
Formally, we define the gauge transformation as a unitary operator
Z
Ŵ (χ, t) := exp −ie d~x χ(t, ~x)Q̂(~x)
Z t Z
= exp −ie ds d~x χ̇(s, ~x)Q̂(~x) (5.67)
−∞
Z t Z
= Texp −ie ds d~x χ̇(s, ~x)Q̂(~x) .
−∞
To see the second identity it is enough to note that [Q̂(~x), Q̂(~y )] = 0, hence
we can replace Texp with exp in (5.67). Clearly,
Ŵ (χ, t)ψ̂(~x)Ŵ (χ, t)∗ = eieχ(t,~x) ψ̂(~x),
Ŵ (χ, t)η̂(~x)Ŵ (χ, t)∗ = eieχ(t,~x) η̂(~x).
166
Let Ĥ(A, t) denote (5.45), that is the Schrödinger picture Hamiltonian. Let
Û (A, t+ , t− ) be the corresponding dynamics.
Z
∗
Ŵ (χ, t)Ĥ(t, A)Ŵ (χ, t) + e χ̇(t, ~x)Q̂(~x)d~x
Z
d~x η̂ ∗ (~x)η̂(~x) − ie A0 (t, ~x) + χ̇(t, ~x) : ψ̂ ∗ (~x)η̂(~x) − η̂ ∗ (~x)ψ̂(~x) :
=
Therefore, by (5.64), we have the following identity, which expresses the gauge
covariance:
= lim eit+ Ĥ0 Ŵ (χ, t+ )Û (A, t+ , t− )Ŵ (χ, t− )∗ e−it− Ĥ0
t+ ,−t− →∞
= Ŝ(A),
It can be compared with the Weyl ordered free Hamiltonian (5.46). We can ap-
ply the formula (A.17) to compute the naive energy shift (the difference between
the ground state energies of Ĥ and Ĥfr ):
q q
Tr − ∂~ + ieA)
~ 2 + m2 − e2 A2 − −∂~ 2 + m2
0
∞
X
=: e2n En .
n=1
167
In the above sum all the terms with n ≥ 2 are well defined. The term with
n = 1 needs renormalization. The renormalized energy shift is
Z ∞
ren 2 ren 2 µν d~p X
E = −e Π (~
p )Fµν (~
p)F (~
p) + e2n En ,
(2π)3 n=2
6 Dirac fermions
In this section we study the Dirac equation
(−iγ µ ∂µ + m)ψ(x) = 0
[γ µ , γ ν ]+ = −2g µν ,
γ 0∗ = γ 0 , γ i∗ = −γ i , i = 1, 2, 3.
γ 5 := −iγ 0 γ 1 γ 2 γ 3 .
It satisfies
[γ 5 , γ µ ]+ = 0, (γ 5 )2 = 1l, γ 5∗ = γ 5 .
All irreducible representations of Dirac matrices are equivalent and act on
the space C4 . One of the most common is the so-called Dirac representation
0 1 0 0 ~σ
γ = , ~γ = ,
0 −1 −~σ 0
0 1
γ5 = .
1 0
168
Here is the Majorana representation:
0 0 −1 1 0 σ1 2 −1 0 3 0 σ3
γ =i , γ =i , γ =i , γ =i ,
1 0 σ1 0 0 1 σ3 0
0 σ2
γ5 = − ,
σ2 0
Tr1l = 4,
Tr(aγ)(bγ) = −4ab,
Tr(aγ)(bγ)(cγ)(dγ) = 4(ab)(cd) − 4(ac)(bd) + 4(ad)(bc).
0 iσ i
0i
σ = ,
iσ i 0
σk 0
σ ij = ijk
. (6.1)
0 σk
The operators σ µν form a representation of the Lie algebra so(1, 3) = spin(1, 3).
It is the infinitesimal version of the representation
−(−iγ∂ + m)(−iγ∂ − m) = −2 + m2 .
169
Therefore, if
(−2 + m2 )ζ(x) = 0,
then (iγ µ ∂µ + m)ζ(x) is a solution of the homogeneous Dirac equation:
(−iγ µ ∂µ + m)(iγ µ ∂µ + m)ζ(x) = 0.
In particular, we have special solutions of the homogeneous Dirac equation
S (±) (x) = (iγ∂ + m)D(±) (x),
S(x) = (iγ∂ + m)D(x),
where D(±) and D are the special solutions of the Klein-Gordon equation intro-
duced in Subsubsect. 2.1.1. We have suppS ⊂ J.
If
(−2 + m2 )ζ(x) = δ(x),
then (iγ µ ∂µ + m)ζ(x) is a Green’s function of the Dirac equation, that is
(−iγ∂ + m)(iγ∂ + m)ζ(x) = δ(x).
In particular, a special role is played by the Green functions
S ± (x) = (iγ∂ + m)D± (x),
c
S (x) = (iγ∂ + m)DF (x),
where D± and DF are the Green’s functions of the Klein-Gordon equation
introduced in Subsubsect. 2.1.1. We have suppS ± ⊂ J ± .
The Dirac propagators satisfy the identities
S(x) = −S(−x) = S (+) (x) + S (−) (x)
= S + (x) − S − (x),
S (+) (x) = S (−) (−x),
+ −
S (x) = S (−x) = θ(x0 )S(x),
S − (x) = θ(−x0 )S(x),
S c (x) = S c (−x) = θ(x0 )S (−) (x) − θ(−x0 )S (+) (x).
Recall that the bosonic causal Green’s function in the momentum represen-
tation can be written as
1
DF (p) = .
p2 + m2 − i0
The Dirac causal Green’s function can be written in a similar way:
−γp + m
S c (p) =
p2+ m2 − i0
1
= , (6.2)
γp + m − i
where i is the shorthand for i0 sgnpγ.
170
6.1.3 Space of solutions
We set αi = γ 0 γ i , i = 1, 2, 3, and β := γ 0 . We obtain matrices satisfying
∂µ j µ (x) = 0.
171
The group R1,3 o Spin↑ (1, 3), acts unitarily on WD by
S ∗ f1 · S ∗ f2 = 0.
By (6.3), Z
ψ(t, ~x) = −i S(t, ~x − ~y )βψ(0, ~y )d~y .
ψ̃(x) := βψ ∗ (x).
#−1
r(y,Λ̃)
ψ(x) = τ (Λ̃−1 )ψ(Λx + y).
hψ((ζ))|ρi := ζ · ρ,
hψ ∗ ((ζ))|ρi := ζ · ρ, ρ ∈ WD .
172
Clearly, for any t
Z
ψ((ζ)) = ζ(t, ~x)ψ(t, ~x)d~x,
Z
ψ ∗ ((ζ)) = ζ(t, ~x)ψ ∗ (t, ~x)d~x.
For f ∈ Cc∞ (R1,3 , C4 ), the corresponding space-time smeared fields are given
by
Z
ψ[f ] := f (x)ψ(x)dx = ψ((S ∗ f )),
Z
ψ ∗ [f ] := f (x)ψ ∗ (x)dx = ψ ∗ ((S ∗ f )).
−(p0 )2 + p~2 + m2 = 0.
p
Set E(~
p) := p~2 + m2 , so that p = (±E(~
p), p~). Define
1 0
χ+ := , χ− := .
0 1
173
Note that
(u(p, s)|u(p, s0 )) = δs,s0 ,
0
(u(p, s)|u(−p, s )) = 0.
The basic plane waves are defined as
1
|p, s) = p u(p, s)eipx .
(2π)3
By writing (p, s|, as usual, we will imply the complex conjugation. We have
0
(p, s|p0 , s0 ) = p − p~0 )δs,s0 ,
δ(~ sgn(p0 p 0 ) > 0,
0
(p, s|p0 , s0 ) = 0, sgn(p0 p 0 ) < 0.
Note that plane waves diagonalize simultaneously the Dirac Hamiltonian D,
the momentum p~ = −i∂~ and the scalar product:
D|p, s) = p0 |p, s),
~ s) = p~|p, s),
−i∂|p,
XZ
ζ 1 · ζ2 = (ζ1 |p, s)(p, s|ζ2 ) + (ζ1 | − p, −s)(−p, −s|ζ2 ) d~
p.
s
In addition, positive frequency plane waves diagonalize the “upper spin in the
3rd direction” and negative frequency plane waves diagonalize the “lower spin
operator in the 3rd direction”:
1 σ3 0
|p, s) = s|p, s), sgnp0 > 0,
2 0 0
1 0 0
|p, s) = s|p, s), sgnp0 < 0.
2 0 σ3
174
We have
XZ d~
p
u(p, s)eipx a(p, s) + u(−p, −s)e−ipx b∗ (p, s)
ψ(x) = p
s (2π)3
XZ
p |p, s)a(p, s) + | − p, −s)b∗ (p, s) .
= d~
s
(+) (+)
X Z (+) (+)
(ζ1 |ζ2 ) = (ζ1 |p, s)(p, s|ζ2 )d~ p.
s
(+) (+)
We set ZD to be the completion of WD in this scalar product.
(−)
Instead of WD for quantization we will use the corresponding complex
(−)
conjugate space denoted W D and equipped with the scalar product
(−) (−)
We set ZD to be the completion of W D in this scalar product.
(+) (−)
The action of R1,3 o P in↑ (1, 3) leaves ZD and ZD invariant.
175
In the following spin averaging identities due to H.B.C.Casimir, which are
useful in computations of scattering cross-sections, the trace involves only the
spin degrees of freedom:
X 2 TrB̃(−p+ γ + m)B(−p− γ + m)
ũ(p+ , s+ )Bu(p− , s− ) = ,
4E + E −
s+ ,s−
X 2 TrB̃(−p+ γ − m)B(−p− γ − m)
ũ(−p+ , −s+ )Bu(−p− , −s− ) = ,
4E + E −
s+ ,s−
X 2 TrB̃(−p+ γ − m)B(−p− γ + m)
ũ(−p+ , −s+ )Bu(p− , s− ) = ,
4E + E −
s+ ,s−
X 2 TrB̃(−p+ γ + m)B(−p− γ − m)
ũ(p+ , s+ )Bu(−p− , −s− ) = ,
4E + E −
s+ ,s−
6.1.10 Quantization
We would like to describe the quantization of the Dirac equation. As usual, we
will use the “hat” to denote quantized objects.
We will use the formalism of quantization of charged fermionic systems [15].
We want to construct (H, Ĥ, Ω) satisfying the standard requirements of QM
(1)-(3) and a distribution
R1,3 3 x 7→ ψ̂(x) (6.8)
with values in C4 ⊗ B(H) such that the following conditions are true:
(1) (−iγ∂ + m)ψ̂(x) = 0;
(2) [ψ̂a (0, ~x), ψ̂b∗ (0, ~y )]+ = δab δ(~x − ~y ), [ψ̂a (0, ~x), ψ̂b (0, ~y )]+ = 0;
(3) eitĤ ψ̂(x0 , ~x)e−itĤ = ψ̂(x0 + t, ~x);
(4) Ω is cyclic for ψ̂(x), ψ̂ ∗ (x).
176
The above problem has a solution unique up to a unitary equivalence, which
we describe below.
We set
(+) (−)
H := Γa (ZD ⊕ ZD ).
(+)
Creation/annihilation operators for the particle space ZD ' L2 (R3 , C2 ) are
(−)
denoted with the letter a and for the antiparticle space ZD ' L2 (R3 , C2 )
with the letter b. Thus, for p on the mass shell and s = ± 21 , using physicist’s
notation on the left and mathematician’s on the right, creation operators for
particles/antiparticles are written as
The whole group R1,3 o Spin↑ (1, 3) acts unitarily on H. Moreover, if we set
˜
ψ̂(x) := β ψ̂ ∗ (x), then
˜
[ψ̂a (x), ψ̂b (y)]+ = Sab (x − y), [ψ̂a (x), ψ̂b (y)]+ = 0. (6.14)
We have
˜ (+)
(Ω|ψ̂a (x)ψ̂b (y)Ω) = Sab (x − y),
˜ c
(Ω|T(ψ̂a (x)ψ̂b (y))Ω) = Sab (x − y).
177
We obtain an operator valued distribution satisfying the Wightman axioms with
(+) (−)
D := Γfin
a (ZD ⊕ ZD ).
For an open set O ⊂ R1,3 the field algebra is defined as
F(O) := {ψ̂ ∗ [f ], ψ̂[f ] : f ∈ Cc∞ (O, C4 )}00 .
The observable algebra A(O) is the subalgebra of F(O) fixed by the automor-
phism
B 7→ eiθQ̂ Be−iθQ̂ ,
where Q̂ will be defined in (6.13). The nets of algebras F(O) and A(O), O ⊂
R1,3 , satisfy the Haag-Kastler axioms.
178
The number operator will be rebaptized as the charge and denoted
Q = dΓ(1l).
[ψa (x), ψ̃b (y)]+ = Sab (x − y), [ψa (x), ψb (y)]+ = 0. (6.18)
The plane wave functionals a(p, s), a∗ (p, s), b∗ (p, s), b(p, s) defined as in (6.6)
and (6.7) in terms of ψ(x), ψ ∗ (x), can be used to diagonalize the Hamiltonian,
momentum and charge
Z X
H = (a∗ (p, s)a(p, s) − b(p, s)b∗ (p, s)) E(~
p)d~
p, (6.19)
s
Z X
P~ = (a∗ (p, s)a(p, s) − b(p, s)b∗ (p, s)) p~d~
p, (6.20)
s
Z X
Q = (a∗ (p, s)a(p, s) + b(p, s)b∗ (p, s)) p~d~
p. (6.21)
s
cpl
The vacuum of Γa (WD ) is annihilated by ψ(x), hence also by a(p, s) and
b∗ (p, s). It is the state of the lowest charge possible. Therefore, it will be called
the bottom of the Dirac sea. We will call the above described procedure the
Dirac sea quantization.
The reader should compare the formulas for H (6.19), P~ (6.20) and Q (6.21)
~
with Ĥ (6.11), P̂ (6.12) and Q̂ (6.13). They differ only by the order of a part of
field operators. So formally they coincide modulo an (infinite) additive constant.
The usual quantization, called the positive energy quantization and the Dirac
sea quantization are just two inequivalent representations of canonical anticom-
mutation relations. If WD had a finite dimension (which can be accomplished
by applying both an infrared and ultraviolet cutoff), then the Dirac sea quan-
tization would be unitarily equivalent with the positive energy quantization by
the procedure invented by Dirac and called often filling the Dirac sea. The
Hamiltonians H and Ĥ, and as we see later, the charges Q and Q̂ would differ
~
by a finite constant. The momenta P~ and P̂ would coincide.
179
examples of such spaces. Symmetries are described by symplectic transforma-
tions. The dynamics is generated by a (classical) Hamiltonian – a function on
the symplectic space.
An important element of the Hamiltonian formalism is the “algebra of clas-
sical observables” – the commutative algebra of functions on the symplectic
space equipped with the Poisson bracket. One can ask whether there exists an
analogous structure behind fermionic quantum fields.
Clearly, the space WD , which is equipped with a scalar product, is the ob-
vious fermionic analog of a (complex) symplectic space from the bosonic case.
The fermionic analog of the “algebra of classical observables” considered in the
cpl
literature, eg. [50], is the Z2 -graded algebra of operators on Γa (WD ) equipped
with the graded commutator.
cpl
The space Γa (WD ) is equipped with the fermionic parity operator, which
we denote by I := (−1l)Q . An operator A satisfying IAI = ±A will be called
even/odd. Operators that are either even or odd will be called homogeneous. If
A is homogeneous we will write |A| = 0 if A is even and |A| = 1 if A is odd.
The analog of the Poisson bracket is the graded commutator:
{A, B} := AB − (−1)|A| |B| BA. (6.22)
∗
Note that ψ(x), ψ (x) are odd operators and for such operators {·, ·} coin-
cides with the anticommutator. Thus, to make (6.18) look “classical”, we can
replace [·, ·]+ with {·, ·} in this identity.
Note that the “classical” version of the Dirac theory has a quantum charac-
ter. In particular, the “classical fermionic algebra” is an algebra of operators on
a Hilbert space and symmetries are unitary. Nevertheless, one has a far reaching
analogy with the usual commutative classical mechanics.
180
yield the Dirac equation.
One can define the stress-energy tensor
∂L(x) ν ∂L(x) ν
T µν (x) := − ∂ ψ(x) − ∂ ψ̃(x) + g µν L(x)
∂ψ,µ (x) ∂ ψ̃,µ (x)
1
= ψ̃(x)γ µ (−i∂ ν )ψ(x) + i∂ ν ψ̃(x)γ µ ψ(x)
2
1
−g µν
ψ̃(x)γ(−i∂)ψ(x) + i∂ ψ̃(x)γψ(x) + mψ̃(x)ψ(x) .
2
It is conserved on shell
∂ µ Tµν (x) = 0.
The components of the stress-energy tensor with the first temporal coordinate
are called the Hamiltonian density and momentum density.
H(x) := T 00 (x)
1 ∗ ~ ~ ∗ (x)~
= ψ (x)~α(−i∂)ψ(x) + i∂ψ αψ(x) + mψ ∗ (x)βψ(x),
2
P i (x) := T 0i (x)
1
= − ψ ∗ (x)(−i∂ i )ψ(x) + i∂ i ψ ∗ (x)ψ(x) .
2
Note that in (6.24) and (6.24) we put ψ ∗ on the left and ψ on the right.
This is the Wick ordering for the Dirac sea quantization, which can be called
the charge Wick ordering. The Hamiltonian and momentum defined from these
densities coincide with the operators defined by the Dirac sea second quantiza-
tion (6.16), (6.17):
Z
H = H(t, ~x)d~x,
Z
P~ = ~ ~x)d~x.
P(t,
∂µ J µ (x) = 0, (6.24)
µ ∗ µ
J (x) = J (x). (6.25)
181
The sesquilinear form given by J coincides with (6.4):
ζ 1 J µ (x)ζ2 = j µ (ζ 1 , ζ2 , x)
= ζ1 (x)βγ µ ζ2 (x), ζ1 , ζ2 ∈ WD .
The current or the spatial part of 4-current can be expressed in terms of the
α matrices:
J~ (x) = ψ ∗ (x)~
αψ(x).
The 0th component of the 4-current is called the charge density
The charge is
Z
Q := Q(t, ~x)d~x
XZ
a∗ (~ p, s)b∗ (~
= p, s)a(~
p, s) + b(~ p, s) d~
p.
s
where the bracket coincides now with the commutator, since Q is even.
For χ ∈ Cc∞ (R3 , R), let αχ denote the ∗-automorphism of the algebra of
operators on WD defined by
Obviously,
182
6.1.16 Quantum 4-current
Let us try to introduce the quantum 4-current density as an operator valued dis-
(+) (−)
tribution on Γa (ZD ⊕ ZD ) by the antisymmetric quantization of the classical
expression
1 ∗
J µ (x) ψ̂ (x)βγ µ ψ̂(x) − ψ(x)βγ µ ψ ∗ (x) .
:= (6.28)
2
(See Subsubsect. A.1.3 for the definition of antisymmetric quantization. Note
that (βγ µ )∗ = βγ µ , and hence βγ µ is the transpose of βγ µ ). The charge con-
jugation C, which we introduce later on in Subsubsect. 6.2.6, satisfies CΩ = Ω
and C Jˆµ (x)C ∗ = −J µ (x). Therefore, (Ω|J µ (x)Ω) = 0. Hence
˜
Jˆµ (x) = :ψ̂(x)γ µ ψ̂(x):.
Formally, we can check the quantum versions of the relations (6.24) the (6.25).
We have
~
Ĵ (x) = :ψ̂ ∗ (x)~
αψ̂(x):,
and the 0th component of the 4-current is called the charge density
Let qχ (~
p, s; p~1 , s1 ) denote the integral kernel on the second line above. We need
to check whether it is square integrable. Now
2
X
2 p + p~1 |2 + E(~
|~ p) − E(~
p1 )
|u(p, s)u(−p1 , −s1 )| = . (6.30)
s,s1
2E(~
p)E(~p1 )
183
After integrating in ~x we obtain fast decay in p~ + p~1 , which allows us to control
the numerator of (6.30). We obtain
Z
C
p, p~1 )|2 d~
|qχ (~ p∼ ,
p1 )2
E(~
which is not integrable. Therefore, αχ is not implementable by the Shale-
Stinespring criterion, see Thm A.2.
Formally, with Z
Q̂(χ) := χ(~x)Q̂(0, ~x)d~x, (6.31)
184
If ψ satisfies (6.34) and R1,3 3 x 7→ χ(x) ∈ R is an arbitrary smooth function,
then eieχ ψ satisfies (6.34) with A replaced with A + ∂χ.
Note the identity
is the retarded/advanced Green’s function of (6.34), that is, the unique solution
of
γ µ (−i∂µ + eAµ (x)) + m S ± (x, y) = δ(x − y)
(6.36)
satisfying
suppS ± ⊂ {x, y : x ∈ J ± (y)}.
We set
S(x, y) := S + (x, y) − S − (x, y).
Clearly,
suppS ⊂ {x, y : x ∈ J(y)}.
We would like to introduce a field R1,3 3 x 7→ ψ(x) satisfying (6.34). If we
assume that it acts on WD and coincides with the free field ψfr (x) at x0 = 0,
such a field is given by
Z
ψ(t, ~x) = −i S(t, ~x; 0, ~y )βψfr (0, ~y )d~y . (6.37)
R3
∂L(x) ∂L(x)
H(x) = ψ̇(x) + ψ̇ ∗ (x) − L(x)
∂ ψ̇(x) ∂ ψ̇ ∗ (x)
1 ∗ ~ ~ ∗ (x)~
= ψ (x)~ α(−i∂)ψ(x) + i∂ψ αψ(x)
2
~
+ψ ∗ (x) e~γ A(x)
+ mβ + eA0 (x) ψ(x).
185
The Hamiltonian Z
H(t) = H(t, ~x)d~x
cpl
can be interpreted as a self-adjoint operator on Γa (WD ) that generates the
“classical” dynamics
κκ = 1l, κγ µ κ−1 = −γ µ ,
where the bar denotes the complex conjugation. In particular, κβκ−1 = −β.
Note also that
κκu = u, u ∈ C4 .
If ζ solves the Dirac equation with the 4-potential A, then so does κζ with
the 4-potential −A. Thus replacing
κ := iγ 2 .
Choose ξP ∈ {1, −1}. Recall that P denotes the space inversion. Replacing
~
ψ(x), ψ ∗ (x), A0 (x), A(x)
ξP γ 0 ψ(Px), ξP γ 0 ψ ∗ (Px), ~
with A0 (Px), −A(Px)
186
Choose ξT ∈ {1, −1}. Recall that T denotes the time reflection. Replacing
(in the Dirac representation)
~
ψ(x), ψ ∗ (x), A0 (x), A(x)
~
with ξT γ 1 γ 3 ψ(Tx), ξT γ 1 γ 3 ψ ∗ (Tx), A0 (Tx), −A(Tx)
X = CPT
C 2 = P 2 = −T 2 = −X 2 = id,
CP + PC = CT + T C = 0,
X P + PX = X T + T X = 0,
CX − X C = PT − T P = 0.
P 2 = (CT )2 = −X 2 = id,
where id denotes the identity. Thus together with Spin↑ (1, 3) they represent
the group P in+ (1, 3), see Subsubsect. 1.1.6.
Besides,
(PT )2 = −id
and PT commutes with P, CT , X . Thus it behaves as i · id. Thus the group
generated by Spin↑ (1, 3), C, P and T is P inext (1, 3), see Subsubsect. 1.1.6.
6.2.4 Quantization
We are looking for a quantum field satisfying
such that
ψ̂(~x) := ψ̂(0, ~x) = ψ̂fr (0, ~x).
Clearly the solution is obtained by decorating (6.37) with hats.
187
As in the bosonic case, we ask whether the fields are implemented by a
a unitary dynamics. Equivalently, we want to check if the classical dynamics
generated by HInt (t) satisfies the Shale-Stinespring criterion.
Arguments parallel to those of Subsubsect. 2.3.4 show that the classical
scattering operator is unitarily implementable.
An analysis similar to that of Subsect. 5.2.4 shows that the dynamics from
t− to t+ is implementable on the Fock space iff the spatial part of the 4-potential
is the same at the initial and final time:
~ + , ~x) = A(t
A(t ~ − , ~x), ~x ∈ R3 . (6.40)
where the Schrödinger picture Hamiltonian Ĥ(t) and the corresponding inter-
action picture Hamiltonian are
Z
α(−i∂~ + eA(t,
d~x: ψ̂ ∗ (~x)(~ ~ ~x)) + mβ + eA0 (t, ~x))ψ̂(~x) :,
Ĥ(t) =
Z
ĤInt (t) = d~xeAµ (t, ~x)Jˆfrµ (t, ~x).
Note that unlike in the case of charged bosons we use the Wick ordering.
This is because Ĥ(t) differs from Ĥfr by a term involving the 4-current Jˆfrµ (t, ~x),
which is automatically Wick ordered. Therefore, we can assume that both Ĥ(t)
and Ĥfr are Wick ordered, which was impossible for charged bosons.
188
We set C := Γ(χ). We have C 2 = 1l,
~ ~
C Q̂fr (x)C −1 = −Q̂fr (x), C Ĵfr (x)C −1 = −Ĵfr (x),
C Ŝ(A)C −1 = Ŝ(−A).
(+) (−)
Define the following unitary operator on ZD ⊕ ZD :
π g1 , g 2 := ξP γ 0 g1 ◦ P, ξP γ 0 g 2 ◦ P .
We check that
~ ~
P Q̂fr (x)P −1 = Q̂fr (Px), P Ĵfr (x)P −1 = −Ĵfr (Px),
~ −1 = Ŝ(A0 ◦ P, −A
P Ŝ(A0 , A)P ~ ◦ P).
We check that
~ ~
T Q̂fr (x)T −1 = Q̂fr (Tx), T Ĵfr (x)T −1 = −Ĵfr (Tx),
~ −1 = Ŝ(A0 ◦ T, −A
T Ŝ(A0 , A)T ~ ◦ T).
189
One can organize Green’s functions in terms of the generating function:
Z(g, g̃)
∞ Z
(−1)N ˜
Z
X ˜
:= ··· 2
ψ̂(y1 ) · · · ψ̂(yN )ψ̂(xN ) · · · ψ̂(x1 )
n=0
(N !)
×g(y1 ) · · · g(yN )g̃(xN ) · · · g̃(x1 )dy1 · · · dyN dxN · · · dx1
Z ∞ Z Z
˜
= Ω Texp −i ĤInt (t)dt − i g(x)ψ̂fr (x)dx − i g̃(x)ψ̂fr (x)dx Ω ,
−∞
˜ ˜
× ψ̂(p01 ) · · · ψ̂(p0N )ψ̂(pN ) · · · ψ̂(p1 ) .
where all pi , p0i are on shell. Scattering amplitudes are the matrix elements of
the scattering operator Ŝ between plane waves. One can compute scattering
amplitudes from the amputated Green’s functions:
−0 −0 − −
−p+0n +0 , −s+0
n +0 ; . . . ; p+
n + , s+
n + ; . . . | Ŝ |−pn −0 , −sn −0 ; . . . ; p n − , sn − ; . . .
−0 −0 − −
· · · ũ(p+ + +0 +0
n+ , sn+ ) · · · u(−pn+0 , −sn+0 )ũ(−pn−0 , −sn−0 ) · · · u(pn− , sn− ) · · ·
= p
(2π)3(n+ +n+0 +n−0 +n− )
˜ +0 −0 ˜ −
× · · · ψ̂(p+
n+ ) · · · ψ̂(pn+0 )ψ(−pn−0 ) · · · ψ̂(−pn− ) · · · amp
.
The scattering operator and Green’s functions satisfy the Ward identities
analogous to those satisfied by charged bosons.
190
6.2.8 Path integral formulation
We have the following formula for the generating function:
(5) If two diagrams differ only by an exchange of two fermionic lines, there is
an additional factor (−1) for one of them. This implies, in particular, that
loops have an additional factor (−1).
To compute scattering amplitudes with N − incoming and N + outgoing par-
ticles we draw the same diagrams as for N − + N + -point Green’s functions. The
rules are changed only concerning the external lines.
(i) With each incoming external line we associate
1
• fermion: √ u(p, s).
(2π)3
1
• anti-fermion: √ ũ(−p, −s).
(2π)3
191
1
• fermion: √ ũ(p, s).
(2π)3
− log iγ µ ∂µ + m − i . (6.43)
i
= Tr log 1l + ie∂µ Aµ (x) + ieAµ (x)∂µ + e2 Aµ (x)Aµ (x)
2
e
+ σ µν Fµν (x) Dfr c
. (6.44)
2
We can compare (6.44) with a similar expression in the bosonic case (5.49).
192
6.2.11 Pauli-Villars renormalization
A single electron loop with two vertices coming from a 4-potential Aµ leads to
a contribution of the form
Z
dp
E1 = Aµ (−p)Aν (p)Πµν (p).
(2π)4
Unfortunately, computed naively, Πµν (p) is divergent.
We will compute it using the Pauli-Villars regularization, similarly as for
charged bosons, see Subsubsect. 5.2.11:
∞ ∞
e2
Z Z
X 8α1 α2
= Ci dα1 dα2 (pµ pν − gµν p2 )
i
(4π)20 0 (α1 + α2 ) 4
!
m2i
α1 α2 2 i
+4gµν p + +
(α1 + α2 )4 (α1 + α2 )3 (α1 + α2 )2
α1 α2 2
× exp −i(α1 + α2 )m2i − i p
α1 + α2
=: (−gµν p2 + pµ pν )2Πgi 2 gd
Λ (p ) + 2ΠµνΛ (p).
193
Define
Πren (p2 ) Πgi 2 gi
:= lim Λ (p ) − ΠΛ (0) (6.45)
Λ→∞
1
e2 (1 − v 2 )p2
Z
2
= dv(1 − v ) log 1 + − i0 .
(4π)2 0 4m2
Denote the vacuum energy function for neutral bosons, introduced in (2.137),
by πbren . Let Πren
b denote the vacuum energy function for charged bosons (5.52)
and Πren
f for charged fermions (6.45). Let us note the following identity:
2Πren 2 ren 2 2 ren 2
b (p ) + Πf (p ) = 4e π (p ). (6.46)
This identity can be also derived from (6.44), (5.49) and (2.135).
194
Thus (6.47) is replaced by
1 ∞
2d/2 µ4−d Ωd
Z Z
ΠE,d
µν (p) = e2 dv |q|d−1 d|q|
(2π)d 0 0
2
(2/d − 1)gµν q 2 − 21 pµ pν − gµν (− 14 p2 + m2 ) + v 2 ( 12 pµ pν − gµν p4 )
× 2 2
q 2 + p4 (1 − v 2 ) + m2
Z 1
4e2 µ2 2π 2−d/2
= dv 2 Γ(2 − d/2)
(4π)2 0 p 2 2
4 (1 − v ) + m
!
p2 1 1 p2
2 1
2 2 2 2
× gµν (1 − v ) + m − pµ pν − gµν − p + m + v pµ pν − gµν
4 2 4 2 4
1
2e2 µ2 2π
Z 2−d/2
= 2
dv p2 Γ(2 − d/2)(v 2 − 1)(pµ pν − gµν p2 )
(4π) 0 (1 − v 2 ) + m2
4
Z 1
2e2 2
p2
2 2
' dv − γ + log(µ 2π) − log (1 − v ) + m (v 2 − 1)(pµ pν − gµν p2 )
(4π)2 0 4
4e2
+ (pµ pν − gµν p2 ). (6.49)
3(4π)2 (2 − d/2)
We can now renormalize (6.49):
ΠE,ren (p2 )(pµ pν − gµν p2 )
= lim ΠE,d 2 E,d
µν (p ) − Πµν (0)
d→4
Z 1
1 p2
= 2
dv(1 − v 2 ) log 1 + 2
(1 − v 2 ) (pµ pν − gµν p2 ).
(4π) 0 4m
Again, this coincides with the Wick rotated result obtained by the Pauli-Villars
method.
Remark 6.3 In the above computations we first try to eliminate gamma matri-
ces. The only remnant of gamma matrices is tr1l, where 1l is the identity on the
space of Dirac spinors, to which we apply the rule (6.48). However, we would
have obtained the same final result if we used eg. the rule Tr1l = 4, since at the
end we apply the normalization condition Πren µν (0) = 0. We use the condition
(6.48), since it is the usual choice in the literature.
Note, however, that in more complicated situations the dimensional renor-
malization can be problematic, especially for fermions in the presence of γ 5 .
195
The naive interacting Hamiltonian is
Z
~ (−i∂~ + eA(~
d~x:ψ̂ ∗ (~x) α ~ x)) + mβ + eA0 (~x) ψ̂(~x):.
Ĥ =
We apply (A.19) to compute the difference between the ground state energies
of Ĥ and Ĥfr , obtaining
Tr − α ~ (−i∂~ + eA(~
~ x)) + mβ + eA0 (~x) + α ~ + mβ)
~ (−i∂)
∞
X
= e2n En (A). (6.50)
n=1
Note that we could have assumed that Ĥfr and Ĥ are given by the antisym-
metric quantization, used the formula (A.17), and we would have obtained the
same result for the energy shift. Indeed, formally, the Wick and Weyl quantized
versions of Ĥfr and Ĥ differ by the same (infinite) constant (which was not true
in the bosonic case).
All the terms in (6.50) with n ≥ 2 are well defined. The term with n = 1
needs renormalization. The renormalized energy shift is
Z ∞
ren 2 d~p X
E = e Πren (~
p2 )Fµν (~
p)F µν (~
p) 3
+ e2n En (A),
(2π) n=2
7 Majorana fermions
In this section we consider again the Dirac equation
(−iγ µ ∂µ + m)φ(x) = 0.
We will quantize the space of its solutions satisfying the Majorana condition.
We obtain a formalism that describes neutral fermions.
In the bosonic case we first treated the neutral case and only then the charged
case. In the fermionic case it is convenient to reverse the order.
196
7.1.2 Space of solutions
If a function ζ satisfies the Dirac equation
(−iγ µ ∂µ + m)ζ(x) = 0,
then κζ also satisfies the Dirac equation. Therefore, we can restrict the Dirac
equation to functions ζ satisfying the Majorana condition
κζ = ζ. (7.1)
The space of smooth space compact solutions of the Dirac equation satisfying
(7.1) will be denoted YD . Note that it is a real vector space equipped with a
nondegenerate scalar product
Z
ζ 1 · ζ2 = ζ1 (t, ~x)ζ2 (t, ~x)d~x.
hφ(x)|ζi = ζ(x).
hφ((ζ))|ρi := ζ · ρ, ρ ∈ YD .
197
7.1.4 Plane waves
Since we consider neutral fields, the generic name for the momentum variable
is again k, instead of p.
Recall that in the Dirac representation we defined the plane waves u(k, s)
given by (6.5). These plane waves are compatible with the Majorana condition
in the following sense:
κu(k, s) = u(−k, −s). (7.3)
We can introduce the plane wave functionals, where k 0 > 0,
Note that
We have
XZ d~k
u(k, s)eikx a(k, s) + u(−k, −s)e−ikx a∗ (k, s)
φ(x) = p
s (2π)3
XZ
d~k |k, s)a(k, s) + | − k, −s)a∗ (k, s) .
=
s
7.1.5 Quantization
To quantize the Dirac equation with the Majorana condition we use the formal-
ism of quantization of neutral fermionic systems [15].
We want to construct (H, Ĥ, Ω) satisfying the standard requirements of QM
(1)-(3) and a distribution
R1,3 3 x 7→ φ̂(x), (7.4)
with values in C4 ⊗ B(H), satisfying the Majorana condition
198
The above problem has an essentially unique solution, which we describe
below.
Let ZD ' L2 (R3 , C2 ) denote the fermionic positive frequency Hilbert space
defined in Subsubsect. 6.1.8. We set H := Γa (ZD ). Creation/annihilation oper-
ators on ZD will be denoted â∗ /â. In particular, for k on shell and s = ± 21 , we
have creation operators, written below in both physicist’s and mathematician’s
notation:
â∗ (k, s) = â∗ |k, s) .
(7.6)
The quantum field is
XZ d~k
u(k, s)eikx â(k, s) + u(−k, −s)e−ikx â∗ (k, s) .
φ̂(x) := p
s (2π)3
˜
The whole R1,3 o Spin↑ (1, 3) acts unitarily on H. Moreover, if we set φ̂(x) :=
β φ̂∗ (x), then
˜
[φ̂a (x), φ̂b (y)]+ = 2Sab (x − y). (7.7)
We have
˜ (+)
(Ω|φ̂a (x)φ̂b (y)Ω) = 2Sab (x − y),
˜ c
(Ω|T(φ̂a (x)φ̂b (y))Ω) = 2Sab (x − y).
The observable algebra A(O) is the even subalgebra of F(O). The nets of alge-
bras F(O) and A(O), O ⊂ R1,3 , satisfy the Haag-Kastler axioms.
199
7.1.6 Quantization in terms of smeared fields
There exists an alternative equivalent formulation of the quantization program,
which uses the smeared fields instead of point fields. We look for a linear function
YD 3 ζ 7→ φ̂((ζ))
satisfying
suppS ± ⊂ {x, y : x ∈ J ± (y)}.
We also set
S(x, y) := S + (x, y) − S − (x, y).
Clearly
suppS ⊂ {x, y : x ∈ J(y)}.
The “classical” Majorana field coinciding with the free field at time t = 0 is
defined as
Z
φ(t, ~x) = S(t, ~x, 0, ~y )βφfr (0, ~y )d~y .
200
7.2.2 Lagrangian and Hamiltonian formalism
The Lagrangian density that yields (7.9) is
1
φ̃(x)γ µ (−i∂µ )φ(x) + i∂µ φ̃(x)γ µ φ(x) + φ̃(x) m + σ(x) φ(x),
L(x) = −
2
where φ(x) are off-shell fields satisfying the Majorana condition (7.2).
We can introduce the Hamiltonian density
L(x)
H(x) = φ̇(x) − L(x)
∂φ̇(x)
1 ∗ ~
~ ∗ (x)φ(x) + φ∗ (x) m + σ(x) β φ(x),
= φ (x)~
α(−i∂)φ(x) + i∂φ
2
and the Hamiltonian Z
H(t) = H(t, ~x)d~x.
201
7.2.4 Path integral formulation
The generating function (and hence all the other quantities introduced above)
can be computed exactly. It equals
!1
− iγ∂ + m + σ
1 2
Z(f ) = det exp −σ
− iγ∂ + m + σ − i0 −iγ∂ + m − i0
i
× exp f (−i∂γ + m + σ − i0)−1 f
2
c c
= det 1l + σSfr exp −σSfr
i c c −1
× exp f S (1l + σSfr ) f . (7.12)
2 fr
Note that En = −i D n c n
2n , where Dn = (−1) Tr(σS ) is the value of the loop with
n
n vertices, similarly to the bosonic case (2.135) except for a different sign.
202
For n = 2, 3, 4, En are divergent and need renormalization.
Using the Pauli-Villars method we define for n = 1, 2, 3 the renormalized
vacuum energy functions
Thus
Enren
Z
dk1 dkn−1
= π ren (k1 , . . . , kn−1 )σ(k1 ) · · · σ(kn−1 )σ(−k1 · · · − kn−1 ) ···
(2π)4 (2π)4
Z
dk1 dkn−1
= lim πΛ (k1 , . . . , kn−1 )σ(k1 ) · · · σ(kn−1 )σ(−k1 · · · − kn−1 ) 4
···
Λ→∞ (2π) (2π)4
Z !
−πΛ (0, . . . , 0) σ(x)n dx .
203
4πΛ (k 2 )
− (q + 12 k)γ + mi − (q − 12 k)γ + mi
d4 q X
Z
= −i Ci tr
(2π)4 i ((q + 21 k)2 + m2i − i0)((q − 12 k)2 + m2i − i0)
d4 q X − 4q 2 + k 2 + 4m2i
Z
= −i C i
(2π)4 i ((q + 12 k)2 + m2i − i0)((q − 21 k)2 + m2i − i0)
Z ∞ Z ∞ !
1 X 4α1 α2 k 2 4m2 2i
= − dα1 dα2 Ci + +
(4π)2 0 0 i
(α1 + α2 )2 (α1 + α2 )2 (α1 + α2 )3
α1 α2 2
× exp −i(α1 + α2 )m2i − i k
α1 + α2
Z 1 Z ∞ !
1 dρ X 2 2 2 2i
= − dv Ci (1 − v )k + 4m +
(4π)2 0 0 ρ i ρ
1 − v2 2
× exp −iρ m2i + k
4
Z 1 X
(1 − v 2 ) 2
1 2 2 2
2
= dv Ci (1 − v )k + 4mi log mi + k − i0 .
(4π)2 2 0 i
4
Note that at the end we use (A.24) besides (A.25), because of the quadratic
divergence.
Finally, the renormalized vacuum energy function is defined as
π ren (k 2 ) = lim πΛ (k 2 ) − πΛ (0) (7.14)
Λ→∞
Z 1
m2 1 (1 − v 2 )k 2 (1 − v 2 )k 2
= − + log 1 + − i0 dv.
(4π)2 0 2 8m2 4m2
A Appendix
A.1 Second quantization
A.1.1 Fock spaces
Let Z be a Hilbert space. Let Sn denote the permutation group of n elements
and σ ∈ Sn . Θ(σ) is defined as the unique operator in B(⊗n Z) such that
204
In what follows we will consider in parallel the symmetric/antisymmetric, or
bosonic/fermionic case. To facilitate notation we will write s/a for either s or
a.
Θns/a are orthogonal projections. The n-particle bosonic/fermionic space is
defined as
⊗ns/a Z := Θns/a ⊗n Z. (A.1)
The bosonic/fermionic Fock space is
∞
Γs/a (Z) := ⊕ ⊗ns/a Z. (A.2)
n=0
We will call the notation on the left of (A.3) and (A.4) “mathematician’s nota-
tion” and on the right “physicist’s notation”.
Sometimes one introduces formal symbols |ξ) treated as vectors, possibly
nonnormalizable, such that for g ∈ L2 (Ξ) we can write
Z
g = |ξ)g(ξ)dξ, g(ξ) = (ξ|g).
205
We have the following dictionary between creation operators written in the
“physicist’s notation” (on the left) and the “mathematician’s notation” (on the
right):
Let [·, ·]− , resp. [·, ·]+ denote the commutator, resp. anticommutator. Bosonic/fermionic
creation and annihilation operators satisfy the canonical commutation/anticommutation
relations, which in the “mathematician’s notation” read
then (A.8) has the meaning of a polynomial function. It is common to use the
name a polynomial for (A.8) also in the antisymmetric case.
The Wick quantization of (A.8) is the operator on the Fock space given by
the same expression, except that we put the “hats” on a and a∗ . Note that the
creation operators are on the left and annihilation operators are on the right:
Z
b(ξ1 , · · · ξm , ξn0 , · · · , ξ10 )
206
In practice we often have some fields, say ϕ1 (ξ), ϕ2 (ξ), that can be written
as linear combinations of a(ξ) and a∗ (ξ), eg.
Z Z
ϕi (ξ) = Ai (ξ)a(ξ) + Bi (ξ)a∗ (ξ).
For 1st order polynomials their Wick quantization obviously coincides with
their Weyl/antisymmetric quantization:
Z Z
: f (ξ)ϕ̂(ξ)dξ: = f (ξ)ϕ̂(ξ)dξ.
207
A.1.4 Second quantization of operators
For a contraction q on Z we define the operator Γ(q) on Γs/a (Z) by
Γ(q) = q ⊗ ··· ⊗ q .
⊗n
s/a
Z ⊗n
s/a
Z
iff Z
|f (ξ)|2 dξ < ∞.
Up to a phase factor
Z
U = exp â(ξ)f (ξ) − â∗ (ξ)f (ξ) dξ .
208
A.1.6 Implementability of Bogoliubov rotations
We will treat simultaneously the bosonic and fermionic case. The upper signs
will always correspond to the bosonic case and lower to the fermionic case.
Let p, q be operators with the integral kernels p(ξ, ξ 0 ), q(ξ, ξ 0 ). We assume
that q(ξ, ξ 0 ) = ±q(ξ 0 , ξ). Set
Z
∗
p(ξ, ξ 0 )â∗ (ξ) + q(ξ, ξ 0 )â(ξ 0 ) dξ 0 ,
â1 (ξ) = (A.11)
Z
q(ξ, ξ 0 )â∗ (ξ 0 ) + p(ξ, ξ 0 )â(ξ) dξ 0 .
â1 (ξ) = (A.12)
Assume that
p∗ p ∓ q # q = 1l, p∗ q ∓ q # p = 0,
pp∗ ∓ qq ∗ = 1l, pq # ∓ qp# = 0,
Theorem A.2 There exists a unitary U on the Fock space such that
The above theorem is called the Shale criterion [48] in the bosonic and the
Shale-Stinespring criterion [49] in the fermionic case. See also eg. [15].
Such Hamiltonians are sometimes called van Hove Hamiltonians [13, 15]. As-
sume that ε is positive. We would like to compute the infimum of the spectrum
of H, denoted inf H.
209
By completing the square we can rewrite (A.13) as
|v(ξ)|2
Z Z
∗ v(ξ) v(ξ)
ε(ξ) â (ξ) + â(ξ) + dξ − dξ. (A.14)
ε(ξ) ε(ξ) ε(ξ)
It is easy to see that the infimum of the first term in (A.14) is zero. Hence
|v(ξ)|2
Z
inf H = − dξ. (A.15)
ε(ξ)
We assume that h(ξ, ξ 0 ) = h(ξ 0 , ξ), g(ξ, ξ 0 ) = ±g(ξ 0 , ξ). We will call (A.16)
Bogoliubov Hamiltonians. Note that (A.16) is the Weyl/antisymmetric quanti-
zation of the corresponding classical quadratic Hamiltonian. In the case of an
infinite number of degrees of freedom it is often ill defined, but even then it is
useful to consider such formal expressions.
We have the following formula for the infimum of H [15]:
12
h2 ∓ gg ∗ ∓hg ± gh#
1
inf H = ± Tr . (A.17)
2 g h − h# g ∗
∗
h#2 ∓ g ∗ g
Here, we write h for the operator with the integral kernel h(ξ, ξ 0 ) and g for the
operator with the integral kernel g(ξ, ξ 0 ).
Consider the Wick ordered version of (A.16):
Z
:H: := 2 h(ξ, ξ 0 )â∗ (ξ)â(ξ 0 )dξdξ 0
Z
g(ξ, ξ 0 )â∗ (ξ)â∗ (ξ 0 ) ± g(ξ, ξ 0 )â(ξ)â(ξ) dξdξ 0 .
+ (A.18)
(In the case an infinite number of degrees of freedom :H: has a better chance to
be well defined compared with H). The formula for the infimum of :H: is more
complicated, but is more likely to lead to a finite expression [15]:
1 !
h2 ∓ gg ∗ ∓hg ± gh# 2
1 h 0
inf :H: = Tr ± ∓ . (A.19)
2 g ∗ h − h# g ∗ h#2 ∓ g ∗ g 0 h#
210
A.2 Miscellanea
A.2.1 Identities for Feynman integrals
Z ∞
1
= i dα exp(−iαA), (A.20)
A − i0 0
Z
log(A2 − w2 )dw = w log(A2 − w2 ) − 2w
(A + w)
+A log , 0 < w < A; (A.26)
(A − w)
w3 2w3 2A2 w
Z
w2 log(A2 − w2 )dw = log(A2 − w2 ) − −
3 9 3
3
A (A + w)
+ log , 0 < w < A. (A.27)
3 (A − w)
211
A.2.2 Identities for the dimensional regularization
The Feynman identity:
Z 1
1 1 dv
= 2 . (A.28)
AB 2 1
+ B) + 12 (A − B)v
2 (A
−1
µ4−d Ωd ∞ |q|d+1
Z
d
(−1 + 2/d) 2 d|q|
(2π) 0 q 2 + A2
A2 µ2 4π 2−d/2
= Γ(2 − d/2)
(4π)2 A2
!
A2 µ2 4π 1
≈ 1 + (2 − d/2) log − γ
(4π)2 A2 2 − d/2
A2 µ2 4π 1
≈ − γ + log + . (A.34)
(4π)2 A2 (2 − d/2)
212
A.2.3 Operator identities
If A is a positive self-adjoint operator, then
Z
A dτ
A1/2 = 2
, (A.35)
(A + τ ) 2π
Z
1 dτ
A−1/2 =
(A + τ 2 ) 2π
Z
1 dτ
= −2 τ2 . (A.36)
(A + τ 2 )2 2π
In the following identity κ is a certain operator. It is useful when studying nth
order loop diagrams:
Z n−1 dτ
1 1
Tr κ κ τ2
(A + τ 2 )2 (A + τ 2 ) 2π
Z n dτ
1 1
= − Tr κ . (A.37)
2n (A + τ 2 ) 2π
e−m|~x−~y|
Z
2 −1
(m − ∆) ρ(~x) = ρ(~y )d~y . (A.40)
4π|~x − ~y |
If it is not necessarily transversal but sufficiently nice, its transversal and lon-
gitudinal part are defined as
~ tr (~x) := A(~
A ~ A(~
~ x) + (−∆)−1 ∂div ~ x), (A.41)
− 21 ~ x),
Alg (~x) := −(−∆) divA(~ (A.42)
213
We have the decomposition
~ x) = A
A(~ ~
~ tr (~x) − ∂(−∆)−1/2
Alg (~x).
We have the identities
Z Z Z
~ 2 ~
A(~x) d~x = Atr (~x) d~x + Alg (~x)2 d~x
2
Z Z Z
~ x) 2 d~x =
∂~ A(~ ~ tr (~x) 2 d~x +
∂~ A ~ lg (~x) 2 d~x
∂~ A
(A.43)
Z Z
2
~ ~ ~ x) 2 d~x,
∂ Alg (~x) d~x = divA(~ (A.44)
Z Z
∂~ A~ tr (~x) 2 d~x = 1 ~ x) 2 d~x.
rotA(~ (A.45)
2
214
By assumption, the second term on the rhs goes to zero. 2
Sometimes a function f does not have enough decay. We can then use the
so-called once substracted dispersion relations.
f (E)
and on the upper half-plane lim E = 0. Then for E ∈ R
|E|→∞
Z
1 1 1
fR (E + i0) = fR (0 + i0) + P fI (ξ + i0) − dξ,
π ξ−E ξ
Z
1 1 1
fI (E + i0) = fI (0 + i0) − P fR (ξ + i0) − dξ.
π ξ−E ξ
References
[1] Asmuß, B. 1990: Zur Quantentheorie geladener Klein-Gordon Teilchen in
äußeren Feldern, PhD Thesis at Fernuniversität Hagen
[2] Araki, H., 2000: Mathematical Theory of Quantum Fields Oxford University
Press
[3] Bär, C., Ginoux, N., Pfäffle, F., 2007: Wave equation on Lorentzian mani-
folds and quantization, ESI Lectures in Mathematics and Physics, EMS.
[4] Baez, J.C., Segal, I.E., Zhou, Z., 1991: Introduction to algebraic and con-
structive quantum field theory, Princeton University Press.
[5] Bialynicki-Birula, I., Bialynicka-Birula, Z. 1975: Quantum Electrodynamics,
Pergamon, Oxford 1975,
[6] Bleuler, K. 1950: Helv.Phys.Acta 23, 567586
[7] Bloch, F., Nordsieck, A., 1937: Note on the radiation field of the electron,
Phys. Rev. 52, 54-59.
[8] Bogoliubov, N.N., Logunov, A.A., Oksak, A.I., Todorov, I.T., 1990: General
Principles of Quantum Field Theory, Kluwer Academic Publishers
215
[9] Brunetti, R., Fredenhagen, K., 2009: Quantum field theory on curved back-
grounds, in Bär, C. and Fredenhagen, K., editors, Quantum Field Theory
on Curved Spacetimes, Lecture Notes in Physics 786, Springer
[10] Cardona, A., Ducourtioux, C., Paycha, S. 2003: From Tracial Anomalies
to Anomalies in Quantum Field Theory, Commun. Math. Phys. 242, 31-65
[11] Carey, A.L., Ruijsenaars, S.N.M. 1987: On fermion gauge groups, current
algebras and Kac-Moody algebras, Acta Applicandae Mathematica 10, 1-86
[12] Deckert, D.-A., Dürr, D., Merkl, F., Schottenloher, M., 2010: Time Evolu-
tion of the External Field Problem in QED J. Math. Phys. 51
[22] Haag, R., 1992: Local quantum physics, Texts and Monographs in Physics,
Springer.
[23] Haag, R., Kastler, D., 1964: An algebraic approach to quantum field theory,
Journ. Math. Phys. 5, 848-862.
[24] Heisenberg, W., Euler, H., 1936: Folgerungen aus der Diracschen Theorie
des Positrons Z. Phys. 98, 714
[25] Hochstenbach, W. J. M. A. 1976 Field theory with an external potential
Comm. Math. Phys. 51, 211–217
216
[26] Itzykson, C., Zuber, J.-B. 1980: Quantum Field Theory, McGraw Hill
[27] Jauch, J. M., Röhrlich, F., 1976: The theory of photons and electrons, 2nd
edition, Springer.
[28] Kibble, T. W. B., 1968: Coherent soft-photon states and infrared diver-
gences I. Classical currents, J. Math. Phys. 9, 315-324.
[29] Klaus, M. 1980: Nonregularity of the Coulomb potential in quantum elec-
trodynamics Helv. Phys. Acta 53 36–39
[30] Klaus, M. and Scharf, G. 1977: The regular external field problem in quan-
tum electrodynamics, Helv Phys. Acta, 50, 779–802
[31] Labzowski L.N. Klimchitskaya G. Dmitriev Y.: Relativistic Effects in
Spactra of Atomic Systems. IOP Publ. Bristol 1993
[32] Langmann, E. 2001: Generalized YangMills actions from Dirac operator
determinants, J. Math. Phys. 42, 5238
[33] Langmann, E, 1994: Fermion current algebras and Schwinger terms in
(3+1)-dimensions, Commun. Math. Phys. 162, 1-32
[34] Langmann, E., Semenoff, G.W. 1994: QCD1+1 with massless quarks and
gauge covariant Sugawara construction, Phys. Lett. B, 341, 195204
[35] Langmann, E., Mickelsson, J. 1996: Scattering matrix in external field
problems, J. Math. Phys. 37, 3933
[36] Lundberg, L.-E. 1973: Spectral and scattering theory for the Klein-Gordon
equation, Comm. Math. Phys. 31 243–257
[37] Lundberg, L.-E. 1973: Relativistic quantum theory for charged spinless
particles in external vector fields, Comm. Math Phys. 31 295–316
[38] Meissner, K.A. 2002: Classical field theory (in Polish), Wydawnictwo
Naukowe PWN
[39] Molinari, G.L. 2007: Another proof of Gell-Mann and Lows theorem, J.
Math. Phys. 48, 052113
[40] Nenciu, G. and Scharf, G., 1978: On regular external fields in quantum
electrodynamics Helv. Phys. Acta 51, 412–424
[41] Reed, M., Simon, B., 1980: Methods of Modern Mathematical Physics, I.
Functional Analysis, London, Academic Press.
[42] Reed, M., Simon, B., 1975: Methods of Modern Mathematical Physics, II.
Fourier Analysis, Self-Adjointness, London, Academic Press.
[43] Ruijsenaars, S. N. M. 1977: On Bogoliubov transformations for systems of
relativistic charged particles, J. Mathematical Phys. 18, 517–526
217
[44] Scharf, G. 1995: Finite Quantum Electrodynamics. The Causal Approach,
Springer
[45] Schroer, B., Seiler, R., Swieca, J. A. 1970: Problems of stability for quan-
tum fields in external time-dependent potentials Phys. Rev. D 2, 927–2937
[48] Shale, D., 1962: Linear symmetries of free boson fields, Trans. Amer. Math.
Soc. 103, 149-167.
[49] Shale, D. and Stinespring, W.F., 1964: States on the Clifford algebra, Ann.
Math. 80, 365-381.
[50] Srednicki, M., 2007: Quantum Field Theory, Cambridge University Press.
[51] Sternberg, S., 1994: Group Theory and Physics, Cambridge University
Press
[52] Streater, R.F, Wightman, A.S., 1964: PCT, spin and statistics and all that,
W.A.Benjamin, New York-Amsterdam.
[53] Thirring, W., Narnhofer, H., 1992: Covariant QED without indefinite met-
ric, Rev. Math. Phys. SI1 197-211
[54] Wald, R.M., 1994: Quantum field theory on curved space-time and black
hole dynamics, University of Chicago Press, Chicago.
[55] Weinberg, S., 1995: The quantum theory of fields, vol. I, Foundations,
Cambridge.
[56] Wightman, A.: oral communication (2000)
[57] Yennie, D., Frautschi, S. and Suura, H. 1961: The infrared divergence
phenomena and high-energy processes Ann. Phys., 13 379-452.
218