You are on page 1of 93

Relativity, Symmetry and the

Structure of Quantum Theory I:


Galilean quantum theory
William H Klink
Department of Physics and Astronomy, University of Iowa, USA

Sujeev Wickramasekara
Department of Physics, Grinnell College, USA

Morgan & Claypool Publishers


Copyright ª 2015 Morgan & Claypool Publishers

All rights reserved. No part of this publication may be reproduced, stored in a retrieval system
or transmitted in any form or by any means, electronic, mechanical, photocopying, recording
or otherwise, without the prior permission of the publisher, or as expressly permitted by law or
under terms agreed with the appropriate rights organization. Multiple copying is permitted in
accordance with the terms of licences issued by the Copyright Licensing Agency, the Copyright
Clearance Centre and other reproduction rights organisations.

Rights & Permissions


To obtain permission to re-use copyrighted material from Morgan & Claypool Publishers, please
contact info@morganclaypool.com.

ISBN 978-1-6270-5624-3 (ebook)


ISBN 978-1-6270-5623-6 (print)
ISBN 978-1-6270-5626-7 (mobi)

DOI 10.1088/978-1-6270-5624-3

Version: 20150301

IOP Concise Physics


ISSN 2053-2571 (online)
ISSN 2054-7307 (print)

A Morgan & Claypool publication as part of IOP Concise Physics


Published by Morgan & Claypool Publishers, 40 Oak Drive, San Rafael, CA, 94903, USA

IOP Publishing, Temple Circus, Temple Way, Bristol BS1 6HG, UK


This book is dedicated to our wives, Judith and Tammy, who know what it is like to
live with a physicist.
Contents

Preface viii
Author biographies xi

1 Introduction 1-1
Bibliography 1-8

2 Newton relativity and one-particle Galilean quantum theory 2-1


2.1 Unitary representations of the Galilei group 2-4
2.2 Complete systems of commuting observables 2-15
Bibliography 2-19

3 Non-inertial transformations, fictitious forces and 3-1


the equivalence principle
Bibliography 3-13

4 Multiparticle systems and interactions 4-1


Bibliography 4-8

5 Internal symmetries 5-1

6 Conclusion 6-1
Bibliography 6-5

Appendices

A Transitive manifolds A-1

B Irreducible representations of the Galilei group and B-1


the origin of mass and spin
Bibliography B-15

C Decomposition of n-fold tensor products and Clebsch–Gordan C-1


coefficients
Bibliography C-8

vii
Preface

Quantum mechanics was developed in the early twentieth century; it arose from the
desire to understand the behavior of atoms and molecules, systems that surprisingly
did not follow the known rules of classical mechanics. So new rules had to be
created, rules that used classical mechanics as a starting point, but then deviated
significantly from most of the principal concepts underlying classical mechanics. The
history of how quantum mechanics was developed is a fascinating one and underlies
the focus of this book; namely, given the rules that the founders of quantum
mechanics developed, is it possible to find principles that lead to the structure of
quantum mechanics as it was historically formulated?
The answer given here is that principles incorporating relativity and symmetry are
sufficient to ground quantum mechanics. Relativity is a notion already present in
classical physics and, in particular, in Newtonian mechanics and Maxwell electro-
dynamics. The principle of relativity states that the laws of physics should be formulated
in such a way that their form remains unchanged (invariant) for large classes
of observers. Or, put differently, the laws of physics should not depend on specific
reference frames; rather, their form should be the same for whole classes of reference
frames. What sorts of reference frames and what classes lead to the notion of symmetry?
Symmetry as used in the sciences refers to transformations that leave a thing unchanged.
A glass rotated about its axis is unchanged relative to its original position. A ball
(sphere) with no markings looks unchanged after rotations about its center. The set of all
symmetry transformations of an object forms what is called a symmetry group, or just a
group. Groups are mathematical in origin and they, and their representations, will play
a key role in this work.
But the role that symmetry plays is not confined to material objects. Symmetries
can also refer to theories and, in particular, to quantum theory. For if the laws of
physics are to be invariant under changes of reference frames, the set of all such
transformations will form a group. Which transformations and which groups
depends on the systems under consideration.
The classes of transformations that leave the form of Newtonian mechanics
unchanged form what is called the Galilei group; we will call this form of relativity
(often called non-relativistic physics) Newton relativity. More precisely, Newton’s
second law, F = ma, is invariant under transformations that form the Galilei group.
But Newton relativity is not the only possible form of relativity. Maxwell’s
equations for electrodynamics generate another form of relativity, in that the
transformations that leave Maxwell’s equations invariant differ from those leaving
Newton’s equations invariant, and form a different group called the Poincaré group.
The structure of the quantum theory resulting from this differing relativity, which we
call Einstein relativity, will be discussed in a subsequent volume.
In chapter 1 we discuss the notions of relativity and symmetry in greater detail and
show how these ideas lead to different forms of quantum theory. In chapter 2 we focus
specifically on Newton relativity and the structure of the Galilei group. We show
that representations of the Galilei group lead to many properties of one-particle

viii
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

‘non-relativistic’ quantum theory and in particular, to the Heisenberg uncertainty


relations, the quantization of angular momentum, and the Schrödinger equation.
While the term relativity refers to those transformations that leave the form of a
law unchanged, there are also transformations between reference frames that change
the form of a law. A prime example involves transformations to accelerating reference
frames. Such transformations do not leave the form of Newton’s equations
unchanged; rather, they produce a modification of Newton’s laws in the form of
fictitious forces. But in the reworking of Newtonian classical mechanics to more
generalized formulations, known as the Lagrange and Hamilton formulations of
classical mechanics, the set of transformations leaving these formulations unchanged
is much larger than with Newton’s equations and, in particular, includes acceleration
transformations, so that fictitious forces arise naturally.
In chapter 3 we study the structure of quantum theory in the presence of acceleration
transformations. We show that the relativity principle can be broadened to include
a larger set of transformations that we call the Galilean line group. By studying
representations of this larger group, we show how quantum fictitious potentials arise,
and how they fit into a generalized Schrödinger equation.
In classical mechanics it is known that accelerated motion can be simulated by
gravitational fields; that is, there is a deep connection between acceleration and
gravity, called the equivalence principle, which Einstein used as a building block in
his general relativity theory. We conclude chapter 3 by showing that, unlike in
classical mechanics, there are many versions of quantum theory in accelerating
reference frames that violate the equivalence principle.
In chapter 4 we turn to multiparticle systems; here again, there is a natural way
of developing the quantum theory of multiparticle systems from the underlying
symmetry, namely through the use of tensor products of representations of the
underlying symmetry group. Chapter 4 begins with an analysis of two-particle
systems. These can be described in a variety of ways, depending on the application
at hand. The transformation coefficients between these different descriptions
are called Clebsch–Gordan coefficients, and these are worked out in detail for
two-particle systems.
Once these coefficients are known, it is possible to see how to introduce inter-
actions between particles in such a way that the Galilean symmetry is preserved. The
energy (Hamiltonian) operator is seen to be part of a larger algebraic structure, the
commutation relations of which serve to define a Galilean invariant structure with
interactions.
In going from two to many-particle systems, it is possible to generalize the
Clebsch–Gordan coefficients and so introduce interactions for many-body systems
in a Galilean invariant way. But since this can become quite complicated, there is an
alternative when dealing with identical particles such as many-electron systems;
this makes use of operators called creation and annihilation operators. Galilean
symmetry can be applied to these operators, from which a Galilean invariant many-
body theory can be developed. The chapter concludes by showing that in a Galilean
invariant theory it is not possible to create or destroy particles. Only by going to a

ix
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

different relativity, namely Einstein relativity, is it possible to describe particle


creation and destruction in an invariant way.
The next chapter, chapter 5, deals with symmetries not arising from transfor-
mations between reference frames; these are the so-called internal symmetries.
Internal symmetries often involve the group of unitary matrices, and in this chapter
representations of the unitary group are combined with the Galilei group repre-
sentations to formulate a quantum theory that combines both symmetries.
The book concludes with chapter 6 and three appendices, which work out in more
mathematical detail some of the concepts used in the various chapters. Our intention
has been to write the chapters in such a way that someone with knowledge of
undergraduate quantum mechanics can follow the presentation; more generally,
anyone with some background in quantum theory should be able to recognize how
the notions of relativity and symmetry lead to many of the familiar results given in
textbooks on quantum mechanics.
The mathematical level needed to read the appendices assumes a good background
in linear algebra and vector spaces. We try to develop the necessary group theory, or
give references for more detailed information. We use units where Planck’s constant is
one in order to keep mathematical expressions as simple as possible. The first
appendix deals with spacetime manifolds and the relationship between such manifolds
and groups of transformations that leave the structure of the manifold unchanged.
The second gives a detailed analysis of the irreducible representations of the Galilei
group and ends with a discussion as to how the concept of mass arises in Galilei group
representations. Finally, the third appendix deals with representations of the unitary
groups, in particular the group of unitary two by two matrices and the notion of
Clebsch–Gordan coefficients for the groups discussed in the previous chapters.

x
Author biographies

William H Klink
William Klink received his PhD from Johns Hopkins University. He was professor
of physics for many years at the University of Iowa, and is now professor emeritus.
His main research interests have centered about the study of symmetry, both
in its mathematical formulations, and in its physical consequences.

Sujeev Wickramasekara
Sujeev Wickramasekara received his PhD in theoretical physics from the University
of Texas at Austin. He held the Weiss Instructorship in Physics and Astronomy
at Rice University before moving to Grinnell College where he is now associate
professor of physics. His main research interests include representation theory
of groups and its applications to physics, resonance scattering and decay
phenomena, and rigged Hilbert spaces.

xi
IOP Concise Physics

Relativity, Symmetry and the Structure of Quantum Theory I:


Galilean quantum theory
William H Klink and Sujeev Wickramasekara

Chapter 1
Introduction

Quantum mechanics is now one of the most successful of all known physical
theories. Not only have there been innumerable tests of the theory, all confirming
its essential correctness; equally convincing is that our world is dominated by
devices that function because of the knowledge of the quantum world. Well-known
examples include lasers, transistors, magnetic imaging devices and nuclear energy.
This work is concerned with how quantum mechanics is grounded; that is to say,
why it is that quantum mechanics has the structure that it does. When quantum theory
was created in the 1920s by physicists such as Heisenberg, Born, Pauli, Dirac and
Schrödinger, there were no known underlying principles that guided its development;
rather, it was the genius and physical intuition of these physicists that led to its
development. But, given the theory they developed, it is possible in retrospect to ask
what basic ideas might be available to see why quantum theory has the structure that it
does. There are different ways of carrying out such an enterprise, some of which are
given in [1–5]. The point of view taken in this work is that the structure of quantum
mechanics can be grounded in the notion of symmetry, and in particular the symmetries
associated with spacetime transformations. These transformations are typically
envisioned as transformations between coordinate reference frames that a class of
different observers use to map out the spacetime. Such transformations between
different reference frames make up a principle of relativity, wherein physical theories
must be formulated in such a way as to be valid in any choice of reference frame.

Relativity. Einstein made the term relativity famous by including it in his two great
theories, the special theory of relativity (1905) and the general theory of relativity
(1916). In a sense, this is the logical completion of an idea that began with Galileo
and Newton, namely that it should be possible to describe some given physical
phenomena with respect to different observers, i.e., with respect to different refer-
ence frames. Physical theories should not be tied to one or another reference frame,

doi:10.1088/978-1-6270-5624-3ch1 1-1 ª Morgan & Claypool Publishers 2015


Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

but rather have a form that makes them valid in any of a large class of reference
frames. To take a simple example, if one observes a ball thrown upward from a
moving train, the description of the trajectory of the ball (a parabola) will differ
from the trajectory described by an observer sitting in the train, who sees the ball go
upward to a maximum height and fall downward in a straight vertical line. Both
descriptions of the trajectory are equally valid and can be related to one another
by a transformation that connects the one reference frame with the other. More
importantly still, both trajectories are determined by the same law of motion F = ma
with F = mg. In this work, by the term ‘relativity’ we mean the choice of the set of
preferred reference frames and the rules of coordinate transformations among them.
By ‘the principle of relativity’, we mean that the laws of physics should have the
same form in any of these reference frames.
As such, the principle of relativity can be considered a meta-principle of physics.
It is antecedent to any dynamical theory, be it classical or quantum, and it puts
constraints on a dynamical theory in that, once the set of preferred reference frames
and its relativity are agreed upon, the fundamental equations of the theory must
have the same form when written in the coordinates of any of the chosen set of
reference frames. It follows that while the principle of relativity is a meta-principle,
the fact that a given dynamical theory is constrained by it is meaningful only after a
particular type of relativity is chosen, i.e., a preferred set of reference frames and the
rules of coordinate transformations among them. Thus, there are many types of
relativity and in this series of books we are primarily concerned with three, which
may be called Newton, Einstein and de Sitter relativities. Newton relativity underlies
Newtonian mechanics. By Einstein relativity, we mean the relativity of Einsteinʼs
special theory of relativity. De Sitter relativity applies to certain curved spacetimes
associated with Einsteinʼs general relativity. We aim to show that each of these
relativities underlies a different type of quantum theory.
The principle of relativity did not apply to pre-Newtonian theories. The Ptolemaic
model of the motion of the planets was Earth centered; the reference frame where the
Earth remains at rest at the origin is the single, preferred reference frame. Although
one could imagine other reference frames, such as one in which the Sun or one of the
other planets is at rest, the Ptolemaic system of cycles and epicycles was formulated
in such a way that the Earth reference frame was the only allowed one. Even the
Copernican revolution, putting the Sun as the preferred reference frame, did not
allow for a principle of relativity. For while the motion of the planets was much
simpler in the Sun reference frame, it simply replaced the Earth reference frame as
the preferred one.
It is with Galileo and Newton that we see for the first time a principle of relativity
being formulated, a decisive point of departure for the new science they were
creating from the scholastic physics and worldview. Galileo recognized that there is
no substantive difference between the states of rest and uniform motion. Newton
formalized the notion of relativity further in his celebrated laws of motion. In
particular, Newtonʼs second law F = ma has the same form in any of his preferred set
of reference frames, called inertial reference frames. In Principia, Newton clearly
states his view on the matter [6]:

1-2
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

Corollary V: When bodies are enclosed in a given space, their motions in


relation to one another are the same whether the space is at rest or whether it is
moving uniformly straight forward without circular motion.

Newton’s ‘spaces moving uniformly straightforward without circular motion’ are


precisely the reference frames that can be transformed to his absolute space by
means of what we now call Galilean transformations, to be discussed below. His first
law is an assertion of the existence of these frames, in particular his absolute space, a
contentious notion that has divided philosophers for centuries [7].
Setting aside the metaphysical intricacies of absolute space, in practice, physicists
work with reference frames that are ‘sufficiently inertial’. For instance, for the trajec-
tories of the thrown ball considered above, the Earth is a sufficiently inertial reference
frame, as is the train moving with constant velocity with respect to the Earth. The
above-mentioned transformation that connects the trajectories in the two frames is
among the Galilean transformations and is called a boost. By taking the simpler motion
of the ball as seen from the train reference system and transforming (boosting) this
trajectory to the Earth reference frame, it is possible to show that the motion in the
Earth reference frame is indeed a parabola. The fact that the motion of an object in one
frame of reference can be used to describe the motion in another reference frame is tied
to the fact that classical Newtonian physics satisfies a relativity principle, even though,
somewhat misleadingly, it often goes under the heading of non-relativistic physics.
Although the principle of relativity may be considered prior to any dynamical
theory and this work is about the relationship between relativity and quantum theory,
the types of relativity that we consider were themselves historically discovered from
the transformation properties of dynamical theories, all classical. Newton derived his
corollary V, quoted above, from his theory of mechanics. Likewise, Einstein deduced
his special theory of relativity from Maxwellʼs equations of electrodynamics. Both of
these relativities deal with inertial reference frames in flat spacetimes and differ from
each other only in the way coordinate transformations are defined, in particular boost
transformations. Similarly, de Sitter relativity arose from a certain cosmological
solution to the field equations of Einsteinʼs general relativity. Unlike the first two,
de Sitter relativity applies to a certain kind of curved spacetime. Thus, the rules of
transformation between different reference frames are not determined by quantum
theory, but deduced from how classical observers relate to one another. One of the
pioneers of quantum mechanics, Niels Bohr, maintained that quantum mechanics is
always done with classical observers, and in this work this principle is manifested by
the choice of the specific types of relativity transformations. We will come back to
question whether the relevant relativity can be deduced from a given quantum theory
in our conclusion.
An important property common to all the above relativities is that the coordinate
transformations defining them form a group1. It is this group structure that turns out

1
A group is a set G = {g} with an associative composition rule G × G → G , which may be denoted by
(g2, g1) ↦ g2 g1, such that: (1) there exists an identity element e ∈ G so that for every g ∈ G , ge = eg = g and
(2) for every g ∈ G , there exists another g −1 ∈ G so that gg −1 = g −1g = e .

1-3
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

to be the key property of relativity that enables quantum theory to be grounded in it.
The groups of Newton, Einstein and de Sitter relativities are called, respectively,
the Galilei group, Poincaré group and de Sitter group. The structures of these
groups reveal a certain hierarchical order between the three types of relativity in that
the Galilei group can be obtained from the Poincaré group, which in turn can be
obtained from the de Sitter group, by means of a certain limiting process called group
contraction [8, 9]. Group contraction provides precise mathematical content to
statements such as ‘Newton relativity is the low velocity limit of Einstein relativity’
and ‘Einstein relativity is the zero-curvature limit of de Sitter relativity’. We will take
up the subject of group contraction and its role in quantum theory in a later work.
As is evident from the discussion, all of these relativities involve transformations on
a four-dimensional spacetime, but the precise geometric and topological structures
of the spacetime are different for the different relativities. In fact, another way of
formulating a relativity principle is to recognize that the transformations in spacetime
can be manipulated to form a manifold, and the allowed group of transformations
leave the manifold invariant. Appendix A discusses the precise mathematical meaning
of this statement. Thus, the manifold for Newton relativity is the Cartesian product of
three-space with the real timeline, while for Einstein relativity, it is the Minkowski
spacetime. Both of these manifolds are flat. By contrast, for de Sitter relativity, the
manifold is curved and called de Sitter space.
There are two different but equivalent ways of relating a given manifold to its
corresponding relativity group. First, we can consider the relativity group as con-
sisting of a set of transformations that map points on the manifold to different points,
also on the manifold. These transformations are called active transformations and this
way of understanding the action of a relativity group on its manifold is referred to as
the active point of view. A detailed discussion of active transformations of a group on
a manifold is given in appendix A. Second, we can introduce a coordinate basis for the
manifold and consider the relativity group as a set of transformations that map bases
into other bases. The relativity group does not move points on the manifold, but
simply provides the relationship between different systems of coordinates describing a
given, fixed point. Viewed this way, the relativity transformations are called passive
transformations and the action of the group is referred to as the passive point of view.
In either case there is an intimate relationship between a spacetime manifold and the
transformations that leave it invariant.

Symmetry. As mentioned above, each relativity is encoded in a particular group, the


key feature that makes relativity so fundamental to the approach we take to quantum
mechanics. Given the importance of the group structure of relativity more generally,
we can also consider other, more abstract transformations that form groups. Trans-
formations that form a group are called symmetry transformations because they leave
a particular object or a property invariant. For instance, rotations leave the length
between two points invariant. Therewith, a two-sphere, defined by the set of points
equidistant from a single point, is invariant under the group of three-dimensional
rotations. The form invariance of equations of a dynamical theory under its relativity
group is another, more sophisticated example of symmetry.

1-4
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

The symmetry associated with a given relativity is the most basic one in quantum
theory. As discussed below and expanded on in greater detail in chapter 2, the
kinematic properties of mass and spin of a particle may be directly attributed to the
relevant symmetry group. However, there are other properties of quantum systems,
such as isospin, a symmetry associated with the strong interaction, that are not
related to the spacetime transformation symmetry. Instead, they arise from internal
symmetries and the quantum states that possess such internal symmetries are
characterized by additional parameters with well-defined transformation properties
under an internal symmetry group. Among the examples are the SU(2) isospin and
SU(3) color groups. We will consider internal symmetries in chapter 5.

Quantum theory. It is an empirical fact that quantum mechanics obeys the super-
position principle. This means that different quantum states can be combined to form
other states. The simplest mathematical structure that enables the implementation of
this principle is a vector space. Further, the probability interpretation of quantum
theory requires that this vector space be an inner product space. Therefore, it is
common to take the space of states of a quantum system to be a Hilbert space / ,
a topologically complete inner product space [2]. We will denote a typical element of
the Hilbert space / by ϕ and the inner product between two elements by ψ ϕ .
If the principle of relativity is to hold for quantum as well as classical theory, then
it must be the case that laws of quantum theory have the same form in any of a set of
preferred reference frames related by a specific relativity. In this sense, we can talk
about a particular quantum mechanics that respects Newton relativity, Einstein
relativity or de Sitter relativity. Acknowledging the key role that the relativity
groups play in their respective quantum theories, we refer to these three quantum
theories as Galilean quantum theory, Poincaré quantum theory and de Sitter
quantum theory, respectively2. In each case, the key idea is that two observers need
not describe a given quantum state by the same element ϕ of the Hilbert space, but
by two different vectors, say ϕ and ϕ′ , such that the physical description of the
system in terms of ϕ has the same form as that in terms of ϕ′ . For instance, both
ϕ and ϕ′ would satisfy Schrödingerʼs equation of time evolution, much like the
trajectories of the ball thrown up in the train are described by the same law of
Newton in both the train and Earth reference frames.
Thus, the fundamental assumption is this: if two observers who describe a given
quantum state by two state vectors ϕ and ϕ′ can be related to each other by a
relativity transformation g, then there must exist an operator U(g) in the Hilbert space
/ that corresponds to g such that ϕ′ = U (g ) ϕ .
In other words, we are considering a mapping g ↦ U (g ) from the appropriate
relativity group G into the set of operators in the Hilbert space, that is to say a
representation of the spacetime transformations that define the relativity in the
Hilbert space of quantum states [10]. As discussed in chapter 2 below, a powerful
theorem of Wigner shows that the operators U(g) have to be linear and unitary
(or anti-linear and anti-unitary). Furthermore, for the operators U(g) to be

2
We use the terms ‘quantum theory’ and ‘quantum mechanics’ interchangeably.

1-5
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

representative of the relativity transformations g, the set of operators {U (g ): g ∈ G}


must have the same group structure as the relativity group G. Hence, we consider a
mapping g ↦ U (g ) such that
i. for every g ∈ G , U(g) is a linear, unitary operator in / ;
ii. if e is the identity element of G, then U (e ) = I , the identity operator in / ;
iii. for any two g2, g1 ∈ G , U (g2 ) U (g1 ) = U (g2 g1 ).

The last property is known as a homomorphism. The last two conditions state that the
mapping g ↦ U (g ) preserves the group structure. Finally, we call a representation
irreducible if there exists no proper, closed subspace Φ of the Hilbert space / that
remains invariant under the action of all the operators U(g). It is a mathematical
theorem that all other representations are built from irreducible representations.
Although our description above may give the impression that a representation of a
relativity group is defined on a preexisting Hilbert space, it is equally possible to think
of the Hilbert space as being generated by the representation, i.e., as the vector space
that carries the representation. In fact, in almost all cases the mathematical con-
struction develops the representation and the Hilbert space that it acts on together.
This may be seen explicitly in appendix B. The situation here runs parallel to the
relationship between relativity groups and spacetime manifolds. One can take the
point of view that spacetime is primary and the relativity group is defined in terms of
the transformations of that spacetime. However, it is equally valid as a starting point
for a physical theory to have a notion of the relativity principle, from which one can
determine the nature of the underlying spacetime manifold.
We are now ready to state the main claim, due to Wigner, that enables quantum
theory to be grounded in relativity: the mathematical image of an elementary
quantum system, i.e., a particle, is a unitary, irreducible representation of the relevant
relativity group. In other words, a quantum particle is described by the collection of
quantum states that can be mapped to one another by a unique, irreducible, unitary
representation of the relativity group. Note that even though Newton relativity
emerges from Newtonian mechanics, we do not develop Galilean quantum mechanics
by quantizing a classical theory, as is traditionally done. Rather, we develop Galilean
quantum theory from the unitary irreducible representations of the Galilei group, a
very different procedure.
An immediate corollary to grounding quantum mechanics in relativity is that all
possible unitary irreducible representations of a given relativity group determine
the types of quantum particles, which may be called forms of matter, allowed in the
corresponding spacetime. In chapter 2, we show that Newton relativity allows only
particles of positive mass and integer or half-odd-integer spin, the parameters that
define all unitary, irreducible representations of the Galilei group. That is, the
irreducible representations of the Galilei group are labeled by mass and spin, and
these labels characterize matter in Galilean quantum theory in an invariant way.
Thus, an electron is characterized by a representation of the Galilei group labeled by
1
mass 0.511 MeV and spin 2 , whereas a proton corresponds to a representation with
1
mass 0.938 GeV and spin 2 .

1-6
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

Einstein relativity allows for more types of matter. As will be seen in volume 2 of
this work, there are four classes of irreducible representations of the Poincaré group.
One class again deals with particles of positive mass and integer or half-odd-integer
spin. The second class corresponds to massless particles, still with integer or half-
odd-integer spin. The particles of this second class, which include the photon,
graviton, gluon and possibly some neutrinos, have properties that are very different
from those of massive particles. For example, in Einstein relativity particles with mass
always have velocities that are lower than than the speed of light, whereas massless
particles always travel at the speed of light. The third class of particles are called
tachyons and are characterized by an imaginary mass, which means they always
travel faster than light. It is not clear whether such particles exist in nature, and if they
did, they would have some very peculiar properties. The fourth class consists of
‘particles’ with no energy or momentum. Finally, the de Sitter group has its own
classes of irreducible representations and corresponding forms of matter, which will
be discussed in a later volume.
In this first volume, we will focus on Newton relativity and show how the
underlying spacetime symmetry group, the Galilei group, leads to Galilean quantum
theory. We devote chapter 2 to an analysis of the unitary irreducible representations
of the Galilei group. In particular, we show how unitary irreducible representations of
the Galilei group give rise to one-particle Hilbert spaces, ‘spacetime-based’ obser-
vables, as well as the Schrödinger equation. In chapter 3, we examine how Newton
relativity may be expanded to include transformations into accelerating, both linearly
and rotationally, reference frames and how quantum mechanics may be grounded in
this expanded notion of relativity. We show that the set of transformations amongst
all inertial and accelerating reference frames gives rise to an infinite dimensional
group that we call the Galilean line group. The unitary irreducible representations
of this group provide a definition of one-particle quantum states that holds for
accelerating observers. Furthermore, these representations also allow us to under-
stand the nature of many non-inertial effects in quantum mechanics, such as fictitious
forces, the equivalence principle and the emergence of synthetic magnetic fields in
rotating reference frames. The goal of chapter 4 is to extend the results of chapter 2 to
multiparticle systems. A multiparticle quantum system is modeled by the tensor
product of irreducible representations that correspond to the single particles that
make up the system. Therefore, multiparticle states also have well-defined properties
under Galilean transformations. Interactions in multiparticle systems are constrained
by properties of these tensor product representations, though, unlike the case with
Einstein relativity, these constraints are not very severe. Perhaps the most striking
difference between the two relativities is that in Newton relativity all quantum
interactions must preserve particle number, whereas this is decidedly not the case in
Einstein relativity. Chapter 5 deals with more abstract, internal symmetries. Again,
the group structure of these internal symmetries is the key and state vectors must have
well-defined transformations under unitary representations of internal symmetry
groups. We devote the last chapter to synthesizing the main ideas and mathematical
results that underlie this work.

1-7
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

Bibliography
[1] Dirac P A M 1999 The Principles of Quantum Mechanics 4th edn (Oxford: Oxford University
Press)
[2] von Neumann J 1983 Mathematical Foundations of Quantum Mechanics (Princeton Land-
marks in Mathematics) (Princeton, NJ: Princeton University Press)
[3] Haag R 1996 Local Quantum Physics 2nd edn (New York: Springer)
[4] Bohm D 1989 Quantum Theory (Dover Books on Physics) (Mineola, NY: Dover)
[5] Omnès R 1994 The Interpretation of Quantum Mechanics (Princeton Series in Physics)
(Princeton, NJ: Princeton University Press)
[6] Newton I 1999 The Principia: Mathematical Principles of Natural Philosophy Engl. transl.
ed I B Cohen and A Whitman (Berkeley, CA: University of California Press)
[7] Sklar L 1977 Space, Time and Space Time (Berkeley, CA: University of California Press)
[8] Inönü E and Wigner E P 1953 On the contraction of groups and their representations Proc.
Natl Acad. Sci. USA 39 510
[9] Saletan E J 1961 Contraction of Lie groups J. Math. Phys. 2 1
[10] Tung W K 2003 Group Theory in Physics (Singapore: World Scientific)

1-8
IOP Concise Physics

Relativity, Symmetry and the Structure of Quantum Theory I:


Galilean quantum theory
William H Klink and Sujeev Wickramasekara

Chapter 2
Newton relativity and one-particle
Galilean quantum theory

As stated in the introduction, the principle of relativity does not determine what
particular relativity is applicable to a given dynamical theory. Of the three possi-
bilities we consider in this series of works, namely Newton, Einstein, and de Sitter
relativities, Newton relativity is the simplest and was the first to be formulated.
Perhaps for that reason, it is also the most intuitive. Its origin is to be found in
classical Newtonian mechanics, in the first and second laws, the first dealing with
inertial reference frames, the second with F = ma . As noted in the introduction, the
set of transformations (excluding space and time inversions) between inertial
reference frames that leave Newton’s laws invariant form the Galilei group. The
main purpose of this chapter is to examine the unitary irreducible representations of
the Galilei group and see how they provide the framework for Galilean quantum
theory, which is often called, rather misleadingly, non-relativistic quantum theory.

Galilei group. We begin by considering the transformations between inertial refer-


ence frames that comprise the Galilei group. As alluded to in the introduction, it is
rather difficult to give a precise definition of what inertial reference frames are at this
level of our discussion. However, as a good operational definition, we consider a
reference frame in which distant stars are at rest and define inertial frames as those
that are not accelerating with respect to that frame. Such a starting point is not
sufficiently general to deal with cosmology, and we defer a better definition to a later
work on de Sitter relativity.
Let the origin of the reference frame tied to the distant fixed stars be used to define a
point x in space, where it is assumed that the space is a flat three-dimensional one, 3.
Let us also introduce a time parameter, which may be modeled by the real line . To
specify the trajectory of an object relative to a given reference frame is to state where

doi:10.1088/978-1-6270-5624-3ch2 2-1 ª Morgan & Claypool Publishers 2015


Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

an object is as a function of time, x = f (t ). In fact, the Newtonian equation F = ma is


precisely the equation, which, if the force F is given, allows one to calculate the
trajectory.
Given these preliminaries, using the passive point of view considered in the
introduction, we identify four classes of transformations of the coordinates of that
fixed point, relating the reference frame fixed to distant stars to all other inertial
reference frames.

1. The first class of transformations, called spatial translations, simply shifts the
origin of the original reference frame to a new point. Mathematically, this is
given by x′ = x + a , where a is the amount of the translation. Since there are
three directions in space, a translation can shift the origin in three different
directions, hence the vector character of the translations. As will be seen, a
transformation that shifts the origin leads to a conjugate quantity, the
observable called momentum.

2. The second class of transformations, called rotations, rotates a given refer-


ence frame by a fixed angle about a fixed axis in such a way that the origin of
the reference frame is unchanged. (It is important to emphasize that the two
reference frames are rotated with respect to each other by a fixed angle, i.e,
they are not rotating with respect to each other as the Earth or a merry-go-
round does. These are examples of accelerating reference frames.) Such
transformations can be written mathematically as x′ = Rx, where R is a 3×3
real matrix which, together with its transpose RT, satisfies RT R = I. Matrices
fulfilling this property have columns and rows that are of unit length and
orthogonal to all other columns and rows, respectively. A rotation leaves the
length of a vector unchanged, a fact from which the above enumerated
properties of rotation matrices may be deduced. All rotation matrices can be
written in terms of three parameters, such as the three Euler angles. It will be
shown that rotations are related to the observable quantity called angular
momentum.

3. The third class involves translations in time, namely that the time parameter
in one inertial frame t may differ from that in another by a translation
t′ = t + b. It simply accounts for the fact that different observers may choose
to measure time from different moments, i.e., their clocks may not be
synchronized. The fundamental characteristic of Newtonian time is, in fact,
that ideal clocks all record time at the same rate and can differ only in their
origin: clocks on the west coast of the United States differ only from clocks in
the midwest because of a shift of two hours between the two time zones. To
paraphrase Newton, ‘Absolute, true and mathematical time, from its own
nature, passes equably without relation to anything external, and thus
without reference to any change or way of measuring of time’. Relative times
in different frames may differ from one another as well as from the absolute
mathematical time only by a translation. As is well known, this is decidedly

2-2
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

not the case in Einstein relativity. In both relativities time translations are
defined the same way, but in Einstein relativity, a translation is not the only
way that two time parameters may differ. Yet for both Newton and Einstein
relativity, time translation as defined by the differing origins of clocks leads
to a conjugate quantity, the observable called energy.

4. The last class of transformations, called boosts, compare spatial coordinates


of two inertial reference frames that move with constant velocity with respect
to each other. While a part of both Newton and Einstein relativities, boosts
differ in form in the two relativities. In the Galilean case, boosts are given by
x′ = x + vt , where v is the velocity of the ‘primed’ frame with respect to the
other. To the extent that the Earth is an inertial frame, a train moving with
constant velocity with respect to the Earth constitutes a boost. Since this
velocity can have three independent components in the three spatial direc-
tions, boosts are characterized by three parameters. Thus, unlike time, we see
that spatial coordinates in two reference frames may differ from one another
in two ways: a boost or a translation. The conjugate to a boost leads to the
observable quantity intimately related to position.

There are now ten transformations that connect different inertial reference
frames: three spatial translations, three rotation angles, three boosts and one time
translation. As will be seen, this is true of all three types of relativity. As such, each
provides some notion of momentum, energy, angular momentum and position,
albeit some of the properties of these physical notions are quite different in different
relativities.
For Newton relativity, the set of ten transformations may be collectively written
as g = (R, v, a, b ) and its action as

(R , v , a , b ) : ( xt ) → ⎛⎝ xt′′⎞⎠ = ⎛⎝ Rx t++vtb+ a ⎞⎠.


⎜ ⎟ ⎜ ⎟ (2.1)

Implicit in these equations is the choice that we apply rotations first and then boosts
and translations. If we were to apply, say, translations first, then (2.1) would look
different, but the resulting group would be isomorphic to the group defined below.
Let us now consider the set of all Galilean transformations, . := {(R, v, a, b )}.
The application of two Galilean transformations, g1 = (R1, v1, a1, b1 ) followed by
g2 = (R 2, v2 , a 2, b2 ), on a spacetime point (x , t ) using (2.1) shows that the combined
transformation is again a Galilean transformation. This property allows us to define
a composition rule . × . → . , which we denote by (g2, g1 ) ↦ g2 g1, by

( R 2 , v2 , a 2, b2 )( R1, v1, a1, b1) = ( R 2 R1, v2 + R 2 v1, a 2 + R 2 a1


+ b1v2 , b 2 + b1) . (2.2)

It is straightforward to verify that (2.2) is associative: g3 ( g2 g1 ) = ( g3 g2 ) g1. Further,


e = (I , 0, 0, 0), where I is the 3×3-identity matrix, is the identity element for (2.2)

2-3
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

in that eg = ge = g for all g ∈ . . Finally, each element of g ∈ . has an


inverse g −1,
(
(R , v , a , b)−1 = R−1, −R−1v , −R−1(a − bv), − b , ) (2.3)
such that g −1g = g −1g = e under (2.2). These observations lead us to the conclusion
that the set of Galilean transformations . is a group under (2.2). We call this group
the Galilei group.
The Galilei group . can also be written as a group of 5×5 matrices by writing
(2.1) in the following form:
⎛ x′⎞ ⎛ R v a⎞ ⎛ x⎞
⎜ t′ ⎟ = ⎜ 0 1 b ⎟⎟ ⎜⎜ t ⎟⎟ . (2.4)
⎜ ⎟ ⎜
⎝1⎠ ⎝0 0 1⎠ ⎝ 1⎠
In fact, the rule of composition of two Galilei group elements given in (2.2) and the
inverse of a group element in (2.3) are perhaps more easily derived using the matrix
realization of (2.4).

2.1 Unitary representations of the Galilei group


As stated in the introduction, the key step in grounding quantum theory in Newton
relativity is constructing the unitary, irreducible representations of the Galilei group.
These representations furnish not only the relevant one-particle Hilbert spaces, but
also the observables and their commutation relations that characterize Galilean
quantum theory. This assertion includes the claim that the irreducible representa-
tions of the Galilei group give the candidates for possible forms of matter in Galilean
quantum theory. We will see that a form of matter characterized by positive mass
and integer or half-odd-integer spin is the only possibility. Furthermore, the Galilei
group is a non-compact group1, a property it shares with the symmetry groups of
Einstein and de Sitter relativities. A theorem in representation theory states that any
unitary representation of a non-compact group is infinite dimensional [1]. It follows
that all three types of quantum theory are inherently infinite dimensional.
Once a representation is constructed, the elements of the Hilbert space, called
vectors or state vectors, that carries the representation are interpreted as possible
states of the physical system. In this regard, the mathematical model of an ele-
mentary quantum system is simply the collection of states that can be mapped to one
another by means of the operators that furnish the irreducible representation of the
symmetry group. Nevertheless, the state vectors themselves do not correspond to
directly measurable quantities. Rather, measurable quantities are the probabilities
and it is an axiom of quantum mechanics that probabilities are given by the square
of the absolute value of an inner product between state vectors. Thus, if a system is
prepared in a state given by a vector ∣ ψ 〉, the probability that a later measurement
will find it in a state ∣ ϕ〉 is given by 7ψ → ϕ = 〈ϕ∣ψ 〉 2 . Now, clearly the entire families

1
As a set, the group is not compact. That is, given an open cover, it is not always possible to find a finite
subcover.

2-4
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

of vectors 9 ψ := {eiα ∣ ψ 〉, α ∈ } and 9 ϕ := {e iβ ∣ ϕ〉, β ∈ }, called rays, give the


same value for the probability 79 ψ → 9 ϕ . Therefore, we represent a physical state by a
ray, rather than a single vector, in the Hilbert space.
Just as is the case in classical physics, by a symmetry transformation in quantum
mechanics we mean a transformation that does not change the outcome of experi-
ments. In particular, if observers in two reference frames that are related to each
other by a Galilean transformation (2.1) view the same physical system, they need
not necessarily describe it by the same ray, which is after all fundamentally inac-
cessible by experiments, but they must find the same probabilities:
79 ψ →9 ϕ = 79′ψ →9′ϕ (2.5)

where the primed and unprimed quantities refer to the two different reference
frames. A powerful theorem proved by Wigner [2, 3] states that when this equality of
probability holds, then there exists an operator U : ∣ ψ 〉 ∈ 9 ψ → U ∣ ψ 〉 ∈ 9′ψ and
U : ∣ ϕ〉 ∈ 9 ϕ → U ∣ ϕ〉 ∈ 9′ϕ , where U is either unitary and linear,
Uϕ Uψ = ϕ ψ
U ( c1 ϕ + c2 ψ ) = c1U ϕ + c2 U ψ ,

or antiunitary and antilinear,

Uϕ Uψ = ϕ ψ * = ψ ϕ
U ( c1 ϕ + c2 ψ ) = c 1*U ϕ + c *2U ψ ,

for arbitrary complex numbers c1 and c2.


Applied to Newton relativity, the transformations such as 9 → 9′ are the
mathematical images of the Galilean transformations g ∈ . between different
observers. Therefore, in view of Wigner’s theorem, we must construct representa-
tions of the Galilean transformations g ∈ . by operators that are either unitary or
antiunitary. Now, as stated in the introduction, in a group representation we require
that the identity element e ∈ G be mapped to the identity operator, U (e ) = I , which
is unitary rather than antiunitary. Since all elements of the group (2.2) can be
continuously mapped to the identity, the continuity of the representation requires
that all operators representing Galilean transformations (or any connected sym-
metry transformation group) be given by unitary operators. This is the reason for
considering unitary representations. Symmetries in quantum physics that are
represented by antiunitary operators all seem to involve time reversal, a discrete
operation. Antiunitarity here is the result of the requirement that the sign of energy
should not change under time reversal.
Let us denote the operators furnishing a representation of the Galilei group by
U(g ), g ∈ . . They act on a Hilbert space / , a typical element of which we denote
by ∣ ϕ〉. This abstract Hilbert space may be realized in different ways, such as the
space of square integrable functions over the position x ∈ 3 or velocity (equiva-
lently, momentum) q ∈ 3. In that case, we refer to Hilbert space elements as

2-5
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

wavefunctions and denote them by, for instance, ψ (x ) or ϕ(q ). In these realizations,
the inner product is defined by integration in the usual way. For instance, for
position wavefunctions,

ψ ψ′ = ∫ dxψ *(x) ψ ′(x).


These wavefunction realizations of the Hilbert space are called L2-spaces. It is
common to indicate the domain of the wavefunctions, too, so that the position or
velocity realizations are written as L2(3). More generally, we denote the wave-
functions by ψ (x , σ ) or ϕ(q, σ ), where σ stands for the remaining degrees of freedom
needed to specify wavefunctions, and the Hilbert space by L2(3 × [σ ]), where [σ ] is
the range over which the σ vary2. It can be shown that such a Hilbert space is also
equivalent to the tensor product L2(3) ⊗ V , where V is the Hilbert space of
functions over σ. The inner product in this case becomes

ψ ψ′ = ∑ ∫ dxψ *(x, σ ) ψ ′(x, σ ),


σ

where the summation over σ stands for additional integrals or sums over discrete
parameters. Similar considerations also apply to the momentum or velocity wave-
functions. As seen from the discussion in appendix B, the representations of the
Galilei group are constructed in the Hilbert space realization by velocity wave-
functions. Since, as seen below, an irreducible representation of the Galilei group is
characterized by a definite, invariant value of mass m, velocity and momentum
variables are simply proportional to each other. For this reason, it is possible to
construct these representations equally well using either velocity wavefunctions
or momentum wavefunctions. While the momentum wavefunctions are more
commonly used, we have opted for velocity wavefunctions. In the transition to
quantum mechanics in non-inertial reference frames, carried out in chapter 3, this
choice is particularly useful.

Cocycle representations. Consider a unitary representation of a group G. As stated in


the introduction, along with the condition U(e ) = I , we must require that the
mapping g ↦ U(g ) be a homomorphism, U(g2 ) U(g1 ) = U(g2 g1 ). However, if we
consider operators U(g) to yield transformations of rays 9 ϕ into other rays 9 ϕ′,
rather than vectors ∣ ϕ〉 into other vectors ∣ ϕ′〉, then we must allow for a phase
freedom in the definition of homomorphism. That is,
U( g2 ) U( g1) ϕ = eiωU( g2 g1) ϕ , g2, g1 ∈ ., ϕ ∈ / ,

where ω is a real number. For such a phase choice to be a realistic possibility, we


must understand on what parameters ω may depend. We must certainly expect ω to
depend on the two group elements g2 and g1. The linearity of the operators U(g)

2
We use ψ for wavefunctions in the position representation and ϕ for wavefunctions in the velocity (or
momentum) representation.

2-6
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

prevents ω from being dependent on the state vector ∣ ϕ〉 on which the operators act,
but it can depend on the arguments of the wavefunctions, such as position or
velocity, in a particular realization of / . Thus, for instance, on the velocity
wavefunctions, we may have
U( g2 ) U( g1) ϕ(q ) = e iω( g 2, g1, q ) U( g2 g1) ϕ(q ). (2.6)

A representation that fulfills (2.6) is called a two-cocycle representation, or more


generally, a cocycle representation. Sometimes, it is also referred to as a ray
representation [4]. Likewise, a representation that is a true homomorphism is called
a vector representation.
When the phase ω does not depend on the argument of the wavefunctions, i.e.,
U( g2 ) U( g1) ϕ = e iω( g 2, g1) U ( g2 g1) ϕ (2.7)

it is called a projective representation. In that case, ω is a real-valued function on


G × G. Being linear operators, the products of U(g) must be associative,
U(g3 )(U(g2 ) U(g1 )) = (U(g3 ) U(g2 )) U(g1 ), and this condition imposes a constraint on
ω : G × G → :
ω( g3, g2 g1) + ω( g2, g1) = ω( g3, g2 ) + ω( g3 g2, g1). (2.8)

This condition is called the two-cocycle condition and any function G × G →  that
fulfills it is called a two-cocycle. As shown in appendix B, some two-cocycles are
trivial. Trivial two-cocycles always exist and whether and how many non-trivial two-
cocycles exist depends on the structure of the group under consideration. It is
common to identify all two-cocycles that differ from one another by a trivial cocycle
and this identification defines equivalence classes in the set (in fact, the group) of all
two-cocycles. Each equivalence class of two-cocycles gives rise to an inequivalent
projective representation.
For the Galilei group, the two-cocycles may be written as
m
ω( g2, g1) =
2
( a 2 · R 2 v1 − v2 · R 2 a1 + b1v2 · R 2 v1), (2.9)

where m is a real number. For each positive value of m, (2.9) defines a representative
element of a different equivalence class of two-cocycles. Thus, for each m there exists
a different projective representation of the Galilei group. Further, as seen from
appendix B, this parameter m has interpretation as mass. While vector representa-
tions of the Galilei group exist, only the projective representations are physically
meaningful in that it is only these representations that lead to a notion of mass as
well as a position operator canonically conjugated to momentum [5].

Irreducibility. Finally, recall that we defined in the introduction a representation as


irreducible if there exists no proper, closed subspace of the Hilbert space that
remains invariant under the action of all of the operators U(g ), g ∈ . . If such a
subspace Φ ⊂ / were to exist, then we would have U(g ) Φ ⊂ Φ for all g ∈ . , and no
element of Φ would be transformed out of that subspace by the action of {U(g )}.

2-7
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

This means the elements of Φ remain separate and distinct from the rest of the space
and the representation U(g) initially defined in / reduces to one in Φ. But, if the
quantum system under consideration is elementary, then there must not exist such an
invariant subspace of state vectors, which leads us to the association of the physical
notion of elementarity with the mathematical notion of irreducibility. As seen from
the discussion below and in appendix B, irreducible representations of the Galilei are
precisely characterized by a definite value of mass and spin. (This is also the case for
the Poincaré group.) Thus, insofar as symmetry transformations are concerned,
irreducibility can be taken as the definition of elementarity and even a composite
object such the proton is elementary when considered as an entity with a definite
mass and spin.

Unitary, irreducible, projective representations of the Galilei group. A detailed dis-


cussion of the unitary, irreducible, projective representations of the Galilei group is
given in appendix B and in [6, 7]. Here we summarize the results of that analysis. The
Hilbert space is a tensor product / = L2 (3) ⊗ V s , where L2 (3) is the Hilbert
space of wavefunctions over velocity q ∈ 3 and V s is the 2s + 1-dimensional
representation space of the rotation group. If ϕ (q, σ ) is an element of / , where σ is
now the spin projection, then the action of the four types of elements of the Gaililei
group is given by
(U (a ) ϕ)(q , σ ) = e ima · qϕ(q , σ ) (2.10a )

(U (R ) ϕ)(q , σ ) = s
∑D σσ′(R ) ϕ( R−1q , σ ′) (2.10b)
σ′

(U (v) ϕ)(q , σ ) = ϕ(q − v , σ ) (2.10c )

mq · q b
(U (b) ϕ)(q , σ ) = e−i 2 ϕ(q , σ ). (2.10d )

Each of the four types of group elements of the Galilei group gives rise to a
unitary group action on the Hilbert space. The meaning of these group actions is as
follows: if ϕ is the wavefunction for a quantum system in a given reference frame,
then U(a ) ϕ is the wavefunction in the reference frame translated by an amount a .
Similarly, U(R ) ϕ and U(v ) ϕ are the wavefunctions in the reference frames rotated by
R and boosted by v, respectively. Finally, U(b ) ϕ is the wavefunction in the reference
frame in which the origin of time is shifted by the amount b.
These results for the four types of group elements can be combined to obtain the
unitary operator corresponding to an arbitrary element (R, v, a, b ) of the Galilei
group:
m s
(U (R , v , a , b) ϕ)(q , σ ) = e−i 2 v · ae i(ma · q − Eb) ∑D σσ −1
( )
′(R ) ϕ R q − v , σ ′ (2.11)
σ′
1
where E = 2 mq 2 . The use of (2.11) twice on a wavefunction shows that this repre-
sentation is indeed projective, with the phase factor equal to (2.9). The Ds in (2.11)

2-8
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

and (2.10b) are the 2s + 1-dimensional unitary matrices that furnish a unitary,
irreducible representation of SU (2), the covering group of the group of rotation
matrices R. The notion of covering group is introduced in appendix B. Unitary
irreducible representations of SU (2) are discussed in appendix C and further details
may be found in [8].

Generators. The Galilei group (2.2) is a ten-parameter group, each parameter run-
ning over the real line (modulo 2π for rotations). The identity element corresponds
to the case when all parameters assume the value zero. It is a property of this type of
group that when all other parameters are set to zero, each parameter defines a one
parameter subgroup of the group and this subgroup is always commutative. Groups
in which all elements commute with one another are called Abelian. This property
holds true for representations of the group as well. Therefore, if we restrict (2.11) to a
single parameter which we denote by α, then we have the operator identity
U (α2 ) U (α1) = U (α2 + α1), U (0) = I . (2.12)

The commutativity of U( α2 ) and U( α1 ) is obvious. The composition rule (2.12) can


be used to differentiate U (α ):
d U (α + ϵ ) − U (α )
U (α ) = lim
dα ϵ→0 ϵ
U (α )(U (ϵ ) − I ) (U (ϵ ) − I ) U (α )
= lim = lim
ϵ→0 ϵ ϵ→0 ϵ
= U (α ) A = AU (α ). (2.13)
The operator
dU ( α )
A := (2.14)
dα α=0

is called the generator of the one parameter group of operators U (α ). The content of
the differential equation (2.13) is that the entire family of operators (2.12) is
dU (α )
determined by the derivative dα at the identity. Since the familiar exponential
function has this property, it is customary to denote the solution of (2.13) as3
U (α ) = e Aα . (2.15)
In view of (2.15), the operator A is referred to as the generator of U (α ). From (2.13),
we see that the generator of each one parameter group commutes with the group
operators.
In addition, the calculation leading to (2.13) can be repeated to show that U †(α ),
the adjoint of U (α ), is also a one parameter group whose generator is A†, the adjoint

3
However, unlike the exponential function of real or complex numbers, U (α ) ϕ does not necessarily have a
Taylor expansion for every ϕ ∈ / . There does exist a dense subspace vector, called the analytic vector,
where the Taylor expansion holds. In fact, the generator A is itself defined, in general, only on a dense subspace
of the Hilbert space, although the group operators U(α ) are defined on the whole of the Hilbert space [9, 10].

2-9
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

of the generator of U(α ). What is most important is the case that the U(α ) are unitary
operators. Then,
U †(α ) ≡ (U (α ))† = (U (α ))−1 = U ( −α ), (2.16)
and it follows that A† = −A, i.e., A is a skew adjoint operator. In this case, it is
common in the literature to define the generator instead as
dU ( α )
A := −i , (2.17)
dα α=0

so that it is a self-adjoint operator A = A† and U(α ) = eiAα . The generator of the


dU † (α )
dual operator group U † (α ) = e−iAα is then given by i dα
, which is again equal
α=0
to A.
These general results for one parameter subgroups and their generators can be
applied to the Galilei group representation (2.11). Thus, we let H, Pi, Ki and Ji be the
generators of one parameter subgroups, respectively, of time translations U(b), space
translations U(a i ), Galilean boosts U(v i ) and rotations U (R(θ i )) along and about the
i
ith -axis for i = 1, 2, 3. For instance, U (b ) = e−iHb and U (a i ) = e iPi a (no summation
over i). Similar expressions hold for other one parameter subgroups.
The action of these generators on the representation space can be obtained from
their definition as the derivative of the corresponding one parameter group at the
group identity. Thus, for Ki, we obtain

dU ( v i )
( K i ϕ )( q , σ ) = − i dv i
ϕ(q , σ )
v i =0

d
= −i ϕ(q − v , σ )
dv i v i =0

= i i ϕ(q , σ ). (2.18a )
∂q

In the same way, we obtain


1
(Hϕ)(q , σ ) = mq 2ϕ(q , σ ) (2.18b)
2

( Pi ϕ)(q, σ ) = mqi ϕ(q, σ ) (2.18c )


( Ji ϕ)(q, σ ) = iϵijk ∂q q kϕ(q , σ ) − ( Si )σσ′ϕ(q , σ ′). (2.18d )
j

In the last expression, the first term comes from the way rotations act on the
argument q of the wavefunction, given by (2.11). The second term comes from

2-10
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

the Ds matrices of (2.11), which provide a representation of the rotation subgroup


of the Galilei group (more, precisely, the covering group SU (2) of the rotation
group, see appendix B.) In order to find the explicit form of operators Si, we need to
know the Ds-matrices explicitly, which may be found in appendix C.

Lie algebra and enveloping algebra. The Galilei group is a Lie group. In addition to
the algebraic group structure, a Lie group also has a certain topological structure
that is consistent with the group structure. In particular, a Lie group is an analytic
manifold and the composition rule that makes it a group is an analytic, invertible
map on this manifold. There are two such maps associated with every element g
of the group G: a left-translation, L g : h ↦ gh, h ∈ G and a right translation,
R g : h ↦ hg, h ∈ G . A more precise definition and detailed studies of Lie groups
can be found in many textbooks, such as [11].
An important property of Lie groups is that they have an associated Lie algebra,
which is of the same dimension as the group. The abstract definition of a Lie algebra
is that it is a vector space that is closed under an antisymmetric bilinear operation,
which may be denoted by (V , W ) ↦ [V , W ], called a Lie bracket. In addition to
antisymmetry, [V , W ] = − [W , V ], a Lie bracket fulfills the Jacobi identity:
[U , [V , W ]] + [V , [W , U ]] + [W , [U , V ]] = 0. In the case of a Lie group, the
associated Lie algebra may be identified with Te, the vector space of all tangent
vectors to the group manifold at the identity e of the group. Each member of Te is a
tangent to a vector field that is invariant under, say, left translations. Each such
vector field is completely determined by the corresponding tangent vector element of
Te, with which it has the exact same relationship that a one parameter operator
group has with its generator, given in (2.15) above.
Thus, in a Lie algebra of a Lie group, the Lie bracket is realized by the com-
mutator of vector fields. Now, since a Lie algebra is a vector space, which we take
here to be finite dimensional, we can choose a basis and write each element of the Lie
algebra as a linear combination of these basis elements. If e¯i , i = 1, 2, 3, …, is such
a basis for Te, the closure under the Lie bracket means that there must exist a set of
numbers c ijk , called structure constants, such that
⎡⎣e¯i , e¯ j⎤⎦ = c k e¯k , (2.19)
ij

where the summation over repeated indices is implied4.


A noteworthy property of a unitary representation of a Lie group G is that the
generators of one parameter operator subgroups, defined by (2.14), furnish a
representation of the Lie algebra of G. What we mean by this is that the when the
one parameter operator subgroups correspond to the vector fields generated by the
basis vectors ēi , the generators Ai fulfill the same Lie bracket identity as (2.19):
⎡Ai , A j⎤ = c k Ak .
⎣ ⎦ ij (2.20)

4
Though they are called structure constants, these numbers clearly depend on the choice of the basis for the Lie
()
algebra. What is important is that c ijk transform as a 1 tensor under changes of the basis. Therefore, it is
2
more precise to consider the structure tensor C that completely determines a given Lie algebra.

2-11
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

Hence, a group representation automatically furnishes a Lie algebra representation,


i.e., a linear map e¯i ↦ Ai that preserves the Lie algebra structure of Te. The
Lie bracket is now simply given by the commutator of operators, [Ai , A j ] =
Ai A j − A j Ai . Recall that in an abstract Lie algebra only a Lie bracket [V , W ] is
defined and is not necessarily a ‘product’ VW. For an operator realization, such a
product always exists, defined by composition the usual way, (AB ) ∣ ϕ〉 := A (B ∣ ϕ〉 ).
Therefore, we may also consider arbitrary products of the generators Ai and their
linear combinations over complex numbers, not just the commutators [Ai , A j ].
This gives us an operator algebra, not simply a Lie algebra, spanned by the Ai. This
associative operator algebra is called the enveloping algebra of the Lie algebra.
Given these preliminaries, we can now consider the Lie algebra representation
that follows from the unitary group representation (2.11). We have already deter-
mined the generators that form a basis for the operator algebra. With the explicit
expressions for the generators, we can calculate the commutation relations that
define the Galilean Lie algebra representation:
⎡Ji , J j ⎤ = iϵijk Jk ⎡ Ji , K j ⎤ = iϵijk K k ⎡ Ji , Pj ⎤ = iϵijk Pk
⎣ ⎦ ⎣ ⎦ ⎣ ⎦
⎡⎣K i , K j⎤⎦ = 0 ⎡⎣ Pi , Pj ⎤⎦ = 0 ⎡⎣K i , Pj⎤⎦ = imδ ij I . (2.21)
⎡⎣ Ji , H ⎤⎦ = 0 ⎡⎣ Pi , H ⎤⎦ = 0 ⎡⎣K i , H ⎤⎦ = iPi .

In fact, these commutation relations contain the entire content of Galilean invariance
in quantum mechanics. We will return to this point in the next section of this chapter.

Consequences. A number of familiar features of Galilean quantum theory, often


considered emblematic of it, follow from the above results:
1. The second and third equations of (2.18) show that there is a constraint
between H and P:
P2
H= , (2.22)
2m
where P denotes the three operators Pi , i = 1, 2, 3 collectively. This is the
expression between kinetic energy and momentum for a single particle,
familiar from elementary Newtonian physics, which we have obtained here as
a consequence of Newton relativity. Therefore, we identify H and P as the
Hamiltonian (energy) and momentum operators, respectively. (If we were
to insert an ℏ explicitly so that all exponents are unitless, we would see that H
and P have the units that further justify these labels.)
2. Consider the commutation relations between the generators of boost trans-
formations and space translations:
⎡⎣ K i , Pj ⎤⎦ = imδ ij I .

If we divide this equation by m, we obtain


⎡ Ki ⎤
⎢⎣ , Pj ⎥⎦ = iδ ij I , (2.23)
m

2-12
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

the well-known Heisenberg commutation relations. Therewith, we identify


the generator of boost transformations, scaled by mass, as the position
operator canonically conjugated to momentum: X := m1 K .
3. Consider (the dual of) (2.13) for the time translation subgroup:
dU ( b ) dU ( b )
i = HU (b), H := i .
db db b=0

Applying this operator identity on an arbitrary wavefunction (or state vector)


and denoting the action of U(b) by
ϕ(b) ≡ U (b) ϕ
we obtain
d
i ϕ(b) = Hϕ(b). (2.24)
db
This is of course the equation that goes by the name of Schrödinger’s
equation. Its content is simply that the Hamiltonian is the generator of
time translations and, just as for the Heisenberg commutation relations, it
follows as a consequence of Newton relativity. Since we can always identify
time t with a translation b = t from the origin, we may write (2.24) as
d
i dt ϕ(t ) = Hϕ(t ), the more commonly known form of the Schrödinger
equation.
4. One of the most important consequences of the symmetry approach to
quantum theory is the prediction of an intrinsic spin characterizing all
particles. From (2.18) it is seen that the angular momentum generator has
two terms, the orbital angular momentum operator and the intrinsic spin
operator. Whereas the orbital angular momentum operator acts on functions
in the L2 space, the intrinsic spin operator acts only on elements in the spin
space, V s; for a spin s particle, the dimension of this space is 2s + 1.
Moreover, the intrinsic spin of a particle is a Galilean invariant; under
arbitrary Galilei transformations its value does not change, because it is the
eigenvalue of the spin Casimir operator, discussed below.
5. Note that the parameter m that appears in (2.22) and the commutation
relations (2.23) enters the theory by way of Galilean two-cocycles (2.9),
where it is introduced as an arbitrary real number. Along with the identifi-
cation of H and P as Hamiltonian and momentum, (2.22) leads us to the
interpretation of m as mass. Therefore, the important inference that we draw
is that it is the projective representations of the Galilei group that allow us to
define a mass for the physical system under consideration, as well as a
position operator fulfilling (2.23), the Heisenberg commutation relations [4].
For a further discussion of this point, see appendix B. Moreover, just like
spin, mass defined this way is also a Galilei invariant notion.
6. Now consider the enveloping algebra spanned by the operators (2.18),
fulfilling (2.21). Recall that this is the collection of linear combinations of

2-13
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

products of the Galilean generators (2.18). It readily follows from (2.21) that
there are three operators, called Casimir operators, that commute with the
entire enveloping algebra, all of which in the representation space of (2.11)
are proportional to the identity operator:

C1 = M = mI , m ∈ + (2.25a )

⎛ 1 ⎞2
C2 = ⎜ J − K × P⎟ = s(s + 1) I , s = 0, 1/2, 1, 3/2, … (2.25b)
⎝ M ⎠

1 2
C3 = H − P = wI , w ∈ + . (2.25c )
2M

The mutually commuting Casimir operators have common eigenvectors.


Since Casimir operators commute with all operators in the algebra, under
the action of any other operator an eigenvector of these operators trans-
forms into another eigenvector with the same set of eigenvalues. Further, in
an irreducible representation, the m, s and w are fixed numbers chosen from
their allowed ranges, making all Casimior operators proportional to the
identity, (2.25). Therefore, all vectors of the representation space are
eigenvectors of these operators. Their eigenvalues (m, s, w ) are invariant.
That is to say, observers in all inertial frames record the same values in all
experiments to measure the Casimir observables, m for C1, s(s + 1) for C2
and w for C3. This is in sharp contrast to other observables, such as
momentum P , position X or spin projection S3. In this light, it is the
eigenvalues of the Casimir operators that provide the intrinsic properties of
the system, properties which characterize the system, whereas properties
such as momentum or position are meaningful only as a relationship
between the observer and the system. Thus, we speak of a mass m, spin s and
an internal energy w of an elementary quantum system in Galilean quantum
mechanics. In fact, for the one-particle Hilbert spaces, the internal energy
eigenvalue is trivial and can be chosen to be zero5, leaving us with only mass
and spin as the invariant physical parameters defining a particle. However,
in the case of multiparticle systems this is no longer the case. As is shown
in chapter 4, the internal energy is crucial for multiparticle systems,
particularly when they are interacting.

In summary, we emphasize that all of the operators corresponding to kinematic


or ‘spacetime’ observables of a one-particle system, such as energy, momentum,

5
This means that a unitary, irreducible, projective representation of the Galilei group characterized by the
three numbers (m, s, w ) is equivalent to any other characterized by (m, s, w′). Therefore, without loss of
generality, we may set w = 0 [6]. This property can also be anticipated on the basis of elementary Newtonian
physics: the energy of a system is defined only up an arbitrary constant.

2-14
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

spin, angular momentum and position, are contained in the enveloping algebra
spanned by the generators of an irreducible representation of the Galilei group.
Charges, such as electric charge and lepton number, do not emerge from the
representations of the spacetime symmetry group. They are, however, also defined
by symmetry groups, only these symmetry groups act on a set of internal indices that
characterize states of quantum systems with these charges. We will discuss internal
symmetries in chapter 5.

2.2 Complete systems of commuting observables


In the analysis we have presented, the operator algebra defined by the commutation
relations (2.21) was obtained from a unitary representation of the Galilei group.
Alternatively, we may start from the operator algebra (2.21) as the mathematical
expression of Galilean symmetry and construct a Hilbert space in which the Galilean
generators are defined as self-adjoint operators. This would give us a representation
of the Galilean Lie algebra, rather than the Galilei group. We can then integrate this
operator Lie algebra and obtain a unitary representation of the Galilei group. Such a
construction is the converse of what we have done in the preceding section and it
begins with the choice of what is called a complete system of commuting operators
(CSCO), a sufficiently large, mutually commuting set of operators chosen from the
enveloping algebra of (2.21). The Hilbert space can then be defined as the space of
complex square intergrable functions defined on the Cartesian product of the spectra
of the CSCO. There are some technical subtleties one must be mindful of in carrying
out this process, such as the choice of the integration measure to define the inner
product; if the operators of CSCO do not have singularly continuous spectra, then
the usual Lebesgue measure may be used. Since the Casimir operators commute with
all the Galilei generators, it is natural to include these in any choice of CSCO. As
stated above, if the unitary representation of the Galilei group carried by the Hilbert
space is irreducible, then the Casimir operators are all proportional to the identity.
The converse of this statement is also true and, with that fact, we must start with a
set of numbers m, s and w that define the action of the three Gallean Casimir
operators.
We will now discuss several examples of CSCO and how they lead to a Hilbert
space realized by L2-functions on their spectra.
1. The choice of CSCO that is completely equivalent to the induced unitary
representation that we have given above consists of, along with the
Casimir operators (2.25), the energy H, momentum P and the spin pro-
jection S3. As evident from the analysis above, operators H and P will
become the generators of a unitary representation of the largest Abelian
invariant subgroup {(I , 0, a, b )} of the Galilei group, the inducing sub-
group for the representation given above. (See appendix B for details.)
Operators H and P have continuous spectra, while S3 has a point spectrum
consisting of s, s − 1, …, −s + 1, −s , where s is an integer or a half-odd
integer determined by the Casimir operator C2. The Hilbert space can be
defined as linear combinations of generalized eigenvectors (‘generalized’

2-15
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

because they are not square integrable) of the CSCO, which we denote by
q, E , σ [m, s, w ] :
C1(= M ) q , E , [m , s , w ] = m q , E , σ , [m , s , w ]
C 2 q , E , [m , s , w ] = s(s + 1) q , E , σ , [m , s , w ]
C3 q , E , [m , s , w ] = w q , E , σ , [m , s , w ]
H q , E , [m , s , w ] = E q , E , σ , [m , s , w ]
P q , E , [m , s , w ] = mq q , E , σ , [m , s , w ]
S3 q , E , [m , s , w ] = σ q , E , σ , [m , s , w ] . (2.26)

These generalized eigenvectors furnish a basis for sufficiently smooth


L2-functions which are regular at infinity, in the sense that if
ϕ(E , q, σ , [m, s, w ]) = E , q, σ , [m, s, w ] ϕ is such a function, then

ϕ= ∑
σ
∫ d3p q , E , σ , [m , s , w ] q , E , σ , [m , s , w ] ϕ . (2.27)

The coefficients
ϕ(q , E , σ , [m , s , w ]) := q , E , σ , [m , s , w ] ϕ (2.28)
are precisely the wavefunctions that we used to obtain the unitary repre-
sentation of the Galilei group discussed above. The transformation formula
(2.11) for the wavefunctions can be obtained from the transformation
properties of the generalized eigenvectors of CSCO q, E , σ , [m, s, w ] under
a general element of the Galilei group:
U (g ) q , E , σ , [m , s , w ] = e im(q′·a− 2 v·a )−ibE ′
1

× ∑D σs′σ(R ) q′ , E ′ , σ ′ [m , s , w ] , (2.29)
σ′

where
q′ = Rq + v
1 2
E ′ = E + mRq · v + mv . (2.30)
2
It follows that
1 1
E′ − mq′2 = E − mq 2. (2.31)
2 2
That is to say, the difference between the total energy and kinetic energy,
generally understood as the internal energy, is a Galilean invariant; this result
is as anticipated from (2.25c) as well as classical Newtonian physics.
2. Next, we consider what is known as the angular momentum basis, which is
often used in partial wave analyses in scattering experiments [12]. On this

2-16
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

basis, a CSCO includes, along with the Casimir operators, the energy,
total angular momentum, projection of the total angular momentum,
and the total orbital angular momentum: (H , J 2, J3, L2 ), with eigenvalues6
q2
( 2m , j (j + 1), m j , l (l + 1)) and wavefunctions ϕ(q, j , m j , l ). Recall that
1
orbital angular momentum is defined by L = M K × P and is an element of
the enveloping algebra of the Galilean algebra.
The generalized eigenvectors for this CSCO can be obtained from those of
the energy-momentum operators given above by transforming the momen-
tum direction via spherical harmonics to orbital angular momentum and
projection of orbital angular momentum, and then coupling the orbital
angular momentum with the intrinsic spin angular momentum to form the
total angular momentum. See chapter 4 and appendix C for a discussion of
the Clebsch–Gordan coefficients that do this.
3. The most familiar basis is the position basis. As seen above, the Galilean
boost operators K , when scaled by the mass, fulfill the canonical commu-
tation relations with the momentum operators. This property, as well as the
related transformation properties of their generelized eigenvectors under
1
spatial translations, justify the interpretation M K as the position operator.
1
Now the CSCO consists of the three operators X := M K , the internal spin
projection S3 and the three Casimir operators. Suppressing the Casimir
operators and their eigenvalues for simplicity, let us denote generalized
eigenvectors of this CSCO by x, σ . The eigenvalue equations read
X x, σ = x x, σ
S3 x , σ = σ x , σ . (2.32)

As before, these eigenvectors furnish a basis for sufficiently smooth, regular


elements of the L2-space:

ψ = ∑∫ d3x x , σ x, σ ψ . (2.33)
σ

The coefficients
ψ (x , σ ) := x , σ ψ (2.34)
are the well-known ‘position basis wavefunctions’, the usual starting point
for so-called non-relativistic quantum theory found in many textbooks,
such as [12] or [13].
Just as was the case for the angular momentum basis vectors, the position
basis generalized eigenvectors x, σ can be related to the momentum basis
vectors q, σ , or equivalently, the position basis wavefunctions ψ (x , σ ) can

6
We normally use σ for the spin projection. However, when we want to distinguish between total and orbital
angular momentum projections, we will revert to the notation common in the literature, mj, ml, etc.

2-17
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

be related to the momentum basis wavefunctions ϕ(q, σ ). The key idea here
is to recognize that the momentum and position bases diagonalize two sets of
operators that are related to one another by the canonical commutation
relations. Therefore, the mapping we seek between the two sets of eigen-
vectors must have the property that it maps two canonically conjugated
operators into each other. The Fourier transform has precisely this prop-
erty7. Therefore,
m3
ψ (x , σ ) = 3
(2π ) 2
∫ d3q eimq·xϕ(q, σ ). (2.35)

That these functions are precisely the same as the functions obtained by
solving the eigenvalue problem for the CSCO involving position can be seen
by their transformation properties under Galilean boosts and by the action of
the boost generator on them:
m3
U (v ) ψ (x , σ ) =
(2π ) 2
3 ∫ d3q eimq·xϕ(q − v, σ )
= eimv·xψ (x , σ ) (2.36)


Kψ (x , σ ) = −i U (v ) ψ (x , σ )
∂v v=0

= mxψ (x , σ ). (2.37)
K
This is indeed the basis in which the position operator X = m
is diagonal.
For space translations, we get
1
U (a ) ψ (x , σ ) =
(2π )
3
2
∫ d3q eimq·(x+a) ϕ(q, σ )
= ψ (x + a , σ ) (2.38)


Pi ψ (x , σ ) = −i U ( ai ) ψ (x , σ )
∂a i a i =0

1 ∂
= ψ (x , σ ), (2.39)
i ∂x i
which is the expression found in most quantum mechanics textbooks for the
momentum operator.

7
As stated in footnote 2, we use ψ for the position representation and ϕ for the velocity representation. Both
functions are realizations of the same abstract Hilbert space vector. Thus, for some φ ∈ / , we have
ψ (x, σ ) = x, σ φ and ϕ(q, σ ) = q, σ φ .

2-18
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

As in the momentum representation, the angular momentum operator will


have two terms, corresponding to orbital and intrinsic spin angular momentum:

m3
U (R ) ψ (x , σ ) =
(2π )
3
2
∫ d3q eimq·xU(R) ϕ(q, σ )
(
= D σσs ′(R ) ψ R−1x , σ ′ ) (2.40)


Ji ψ (x , σ ) = iϵijk x j ψ (x , σ ) + ( Si )σσ′ψ (x , σ ′). (2.41)
∂x k
In the position basis, the free particle energy operator is not diagonal. To
obtain its action, let us start with the time translation operator:
m3
U (b ) ψ (x , σ ) =
(2π ) 2
3 ∫ d3q eimq·xU(b) ϕ(q, σ )
m3 q 2b
=
(2π )
3
2
∫ d3q eimq·xei 2m ϕ(q , σ )

= ∫ d3x′K b(x, x′) ψ (x′, σ ), (2.42)

where
3
⎛ m ⎞ 2 i m (x − x ′ )2
K b(x , x′) = ⎜ ⎟ e 2b .
⎝ 2πib ⎠
Differentiating (2.42) and evaluating the derivative at b = 0, we obtain

Hψ (x , σ ) = i U (b ) ψ (x , σ )
∂b b=0

3
1 ∂ ∂
=− ∑ i i
ψ (x , σ ), (2.43)
2m i=1
∂x ∂x

which gives the usual expression for the Hamiltonian of a free particle in the
position basis.

These examples show that Newton relativity and its symmetry group, the Galilei
group, are at the heart of all the well-known expositions of what is commonly called
non-relativistic quantum mechanics.

Bibliography
[1] Taylor M E 1986 Noncommutative Harmonic Analysis (Mathematical Surveys and Mono-
graphs, no. 22) (Providence, RI: American Mathematical Society)
[2] Wigner E P 1959 Group Theory and Quantum Mechanics (New York: Academic)

2-19
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

[3] Weinberg S 1995 The Quantum Theory of Fields vol 1 (Cambridge: Cambridge University
Press)
[4] Bargmann V 1954 On unitary ray representations of continuous groups Ann. Math. 59 1
[5] Hamermesh M 1962 Group Theory and its Applications to Physical Problems (Boston, MA:
Addison-Wesley)
[6] Levy-LeBlond J-M 1963 Galilei group and nonrelativistic quantum mechanics J. Math.
Phys. 4 776
[7] Warren R E and Klink W H 1970 Model independent analysis of nonrelativistic multi-
particle reactions J. Math. Phys. 11 1155
[8] Edmunds A R 1974 Angular Momentum in Quantum Mechanics (Princeton Landmarks in
Physics) (Princeton, NJ: Princeton University Press)
[9] Rudin W 1991 Functional analysis 2nd edn (New York: McGraw-Hill)
[10] Nelson E 1959 Analytic vectors Ann. Math. 70 572
Flato M, Simon J, Snellman H and Sternheimer D 1972 Simple facts about analytic vectors
and integrability Ann. Sci. Ec. Norm. Sup. 5 423
[11] Warner F W 1983 Foundations of Differentiable Manifolds and Lie Groups (New York:
Springer)
[12] Sakurai J J and Napolitano J J 2014 Modern Quantum Mechanics 2nd edn (Harlow: Pearson)
[13] Griffiths D J 2014 Introduction to Quantum Mechanics 2nd edn (Harlow: Pearson)

2-20
IOP Concise Physics

Relativity, Symmetry and the Structure of Quantum Theory I:


Galilean quantum theory
William H Klink and Sujeev Wickramasekara

Chapter 3
Non-inertial transformations, fictitious forces
and the equivalence principle

In the preceding chapter, we showed how one-particle quantum states can be obtained
from unitary irreducible representations of the Galilei group, the transformation group
of Newton relativity. Since the Galilei group defines transformations amongst inertial
reference frames, this description of one-particle quantum states is limited to inertial
observers. In particular, observers who are accelerating are excluded in this description.
This motivates the question of whether it is possible to provide a description of
one-particle quantum states that holds for non-inertial observers by expanding the
notion of Newton relativity to include non-inertial reference frames.
As follows from Newton’s corollary quoted in the introduction, his formulation
of mechanics is not covariant under transformations into accelerating reference
frames. In the language of differential geometry, Newtonian acceleration is not a
vector under coordinate transformations more general than Galilean transform-
ations and, consequently, what are known as fictitious forces appear when Newton’s
second law is written in terms of coordinates of an accelerating reference frame.
Fictitious forces, unlike physical forces, do not originate in physical bodies. More-
over, they lead to trajectories that are independent of the mass of a test particle, also
a signature feature of motion determined by the gravitational force. This means that
the motion of an object as seen by an accelerating reference frame can be simulated
by a suitable gravitational field, a property encapsulated in the equivalence principle.
Along with the notion of one-particle states, we expect a consistent formulation of
quantum mechanics in non-inertial reference frames to provide us with the means to
understand the nature and role of non-inertial effects in quantum mechanics,
including the equivalence principle and fictitious forces.
While Newton’s formulation of mechanics is covariant only under Galilean
transformations, the more general formulations of mechanics by Lagrange and

doi:10.1088/978-1-6270-5624-3ch3 3-1 ª Morgan & Claypool Publishers 2015


Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

Hamilton are in fact covariant under a much larger set of transformations, including
transformations into accelerating reference frames [1]. As is well known, the equations
of motion arising from Lagrange’s equations in accelerating reference frames auto-
matically generate the appropriate fictitious forces. We will draw from this property
of Lagrange and Hamilton formulations in developing quantum mechanics in
non-inertial reference frames, namely grounding it in an expanded notion of relativity.
The generalization we seek is straightforward: construct the (smallest) group of
transformations that includes transformations into accelerating reference frames and
construct its unitary, irreducible representations. Naturally, we want to construct
these representations under the requirement that they contain as a subrepresentation
a unitary, irreducible, projective representation of the Galilei group so that when
transformations are restricted to be between inertial reference frames, we recover the
usual Galilean notion of a particle, developed in chapter 2. These representations are
to act on the one-particle Hilbert spaces, leading to a quantum theory that holds in
non-inertial reference frames.
To that end, consider again the Galilean transformations (2.1). Unlike Lorentz
transformations, the Galilean boosts do not act on the time coordinate and their
action on space coordinates, vt , has the same form as a space translation. Conse-
quently, boosts and translations may be combined into time-dependent space
translations, a(t ) = a + vt . This structure suggests that transformations to linearly
accelerating reference frames may be implemented by space translations a(t ) with
arbitrary time dependence. For instance, a transformation to a reference frame with
1
constant acceleration would be given by a(t ) = 2 at 2 . Likewise, transformations to
rotating reference frames can be achieved by letting the arguments of rotation
matrices R be functions of time. Hence, let us consider spacetime transformations

⎛ ⎞
(R , a , b): ( tx) → ⎛⎝ tx′′⎞⎠ = ⎜⎝ R (t)tx++ba (t) ⎟⎠,
⎜ ⎟ (3.1)

where R and a are time-dependent rotations and space translations which get eval-
uated at the time coordinate t when acting on a spacetime point (x , t ). We take
R and a to be analytic functions. Unlike in (2.1), now the velocity boosts are not
independent parameters, but given by the time derivatives of a .
It follows from (3.1) that the transformations (R, a, b ) compose as

( R 2 , a 2, b2 )( R1, a1, b1) = ( ( Λ b1R 2 ) R1, Λ b1a 2 + ( Λ b1R 2 ) a1, b2 + b1), (3.2)

where Λ b is the shift operator (Λ b f )(t ) = f (t + b ). It accounts for the fact that in
a successive application of (3.1), the group elements R2 and a 2 get evaluated at
t + b1, whereas R1 and a1 get evaluated at t. The mapping b → Λ b is in fact a
homomorphism, a property which in turn ensures that (3.2) is associative. Noting
that each (R, a, b ) has a unique inverse (Λ−bR−1, −Λ−b (R−1a ), − b ), we conclude that
the set of elements  := {(R, a, b )} is a group under (3.2). It contains the Galilei
group as a subgroup because (3.2) reduces to (2.2) under constant rotations and

3-2
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

translations of the form a(t ) = a (0) + vt . In [2, 3], this group was introduced as the
Galilean line group. We denote it by .
In much the same way that the Galilei group ties together all inertial reference
frames in a Galilean spacetime, the Galilean line group  ties together all inertial
and accelerating, both linearly and rotationally, reference frames. Therefore, in the
spirit of the philosophy we articulated in the introduction, Galilean quantum
mechanics in non-inertial reference frames must follow from the unitary, irreducible,
possibly cocycle representations of the Galilean line group.

Properties of the Galilean line group. Before discussing representations of , let us con-
sider some of its important structural properties. The subsets ) := {(I , 0, b )} = ,
( := {(I , a, 0)} and 9 := {(R, 0, 0)} of time translations, time-dependent space
translations and rotations, respectively, are all subgroups of . The semidirect product [4]
of the latter two, {(R, a, 0)}, is in fact the group of mappings on the real line t ∈  taking
values in the Euclidean group in three dimensions. Therefore, we call this subgroup the
Euclidean line group and denote it by ,(3) [2]. The mapping b ↦ Λ b is a homomorphism
from ) into the automorphism group of ,(3) and  is the semidirect product of ,(3)
and ) under the mapping to Λ b:
 = ,(3) ⋊Λ ) . (3.3)
In light of the map group structure  → E(3) of , note that our approach bears a
resemblance to gauge theories and, in a sense, the Galilean line group may be
considered a gauged Galilei group.
Since analytic functions are uniquely defined by their Taylor coefficients, we may
consider  an infinite dimensional topological group parametrized by the Taylor
dna dθ dR (θ (t ))
coefficients of a and R: ai(n) = ni ∣t=0 and R i(n) = dt idθ ∣t=0 , where i = 1, 2, 3
dt
indicates the three spatial directions and R i (θ ) denotes a rotation about the xi-axis
by angle θ (t ). Hence, denoting the three nth Taylor coefficients of translations and
rotations collectively by a (n) and R (n) , a typical element of  may be written as
({R (0), R (1), …}, {a (0), a (1), …}, b ). This shows that  has a countable basis.
A closed-form expression can also be obtained for the action of the automorphism
Λ b on ({R (0), R (1), …}, {a (0), a (1), …}, 0) [3].
As usual, we can define the Lie algebra of  as, say, the left invariant vector fields
on . With the Taylor coefficient parametrization of , a basis for these vector fields
may be chosen to consist of K i(n) , Ji(n) and H, the generators of the one parameter
subgroups defined by ai(n) , R i(n) and b, respectively. They fulfill the commutation
relations [2, 3]:
⎡K (n) , K (m )⎤ = 0
⎣ i j ⎦ (3.4a )

⎡H , K (n)⎤ = −iK (n−1) ⎡ H , K (0) ⎤ = 0


⎣ i ⎦ i for n ⩾ 1; ⎣ i ⎦ (3.4b)

⎡J (n) , K (m )⎤ = i (m + n ) ! ϵ K (m+n)
⎣i j ⎦ ijk k (3.4c )
n ! m!

3-3
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

⎡J (n) , J (m )⎤ = i (m + n ) ! ϵ J (m+n)
⎣i j ⎦ ijk k (3.4d )
n ! m!

⎡H , J (n)⎤ = iJ (n−1) ⎡ H , J (0) ⎤ = 0.


⎣ i ⎦ i for n ⩾ 1; ⎣ i ⎦ (3.4e )

Evidently, these commutation relations have the typical structure of a Kac–Moody


algebra associated with a map group [5]. It also bears a strong resemblance to
Galilean conformal algebras [6], albeit here it is infinite dimensional.

Unitary, irreducible representations of . Remarkably, even though the Galilean line


group  is an infinite dimensional group, its unitary irreducible representations can be
constructed by the method of induced representations, discussed in appendix B. Since
these representations are required to contain a physically meaningful representation
of the Galilei group as a subrepresentation and since all physically meaningful
representations of the Galilei group are projective representations, all physically
relevant representations of  must be cocycle representations. As seen in chapter 2,
the construction of such representations begins with the choice of a two-cocycle. In
the case of the Galilei group, these are real-valued functions ω : . × . →  fulfilling
the constraint (2.8). In the present case, since the group elements of  are functions of
time, the relevant two-cocycles must not be real-valued functions on  × . Rather,
they must take values in the set of functions—say, analytic functions—on the real-
line. This set of functions is naturally an Abelian group under point-wise addition and
we denote it by -(), or more simply by - . These functions must also fulfill the
requirement that they reduce to (2.9), or a two-cocycle equivalent to it, when the
transformations of  are restricted to the Galilean transformations.
An increasing number of general classes of unitary, irreducible, cocycle repre-
sentations of  have been constructed [2, 3, 7, 8]. They are completely defined by
the transformation properties of the velocity eigenkets ∣ q〉. To begin we present the
transformation formula for ∣ q〉 that defines the simplest class of representations:
U (g ) q = eimξ(g;q ) Λ −bq′ (3.5)
where m is the mass and
⎛ 1 1 ⎞
ξ (g ; q ) := ⎜ q′ · a − a · a˙ + ( Λ −b − 1) q′ · a q′ ⎟
⎝ 2 2 ⎠

d
a q := ∫ dt q , q=
dt
aq (3.6)

d d
q′ = Rq + R˙ a q + a˙ =
dt
(
Ra q + a = a q′
dt
) (3.7)

Here, aq is that time-dependent spatial translation which transforms the particle


from rest q = 0 to velocity q . Therefore, it is a boost, the standard one used to define
the velocity eigenstates ∣ q〉 by applying the operator U (aq ) on the rest states ∣ 0〉.

3-4
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

For simplicity, we consider only spin zero representations. The extension to spin is
straightforward.
This is a cocycle representation in that U (g2 ) U (g1 ) ≠ U (g2 g1 ). Rather,

U( g2 )U( g1) ∣ q〉 = eimξ2( g2,g1;q ) U ( g2 g1) ∣ q〉 (3.8)


where
1 d
ξ2( g2, g1; q ) =
2
( )
Λ −b1 − 1 Λ b1( a 2 R 2 ) ·
dt
(
a1 + R1a q )
1

2
( ) (
Λ −b1 − 1 Λ b1( a˙ 2R 2 ) · a1 + R1a q )
( ) ( (
+ Λ b1ω2 · R1a q × a1 − Λ b1R 2 a1 × Λ b1a 2 ) ) (3.9)

and ω2 is the angular velocity associated with the time-dependent rotation R 2(t ) by
the well-known equality ω × r = RR ˙ T r . Recall that the phase (2.7) which appears in
the operator composition for the Galilei group depends only on the group elements
g2 and g1, making the relevant representations projective. By contrast, here we see
that the phase factor of (3.8) depends not only on the two group elements g2 and g1,
but also on q , the argument of the state vectors, making the representations belong
to a more general class of cocycle representations.
A second important difference between the representations of the Galilei group
and those of the Galilean line group is that the state vectors for the former are
defined over velocities q which are constant in time, whereas the state vectors for the
latter are defined over time-dependent velocity vector fields. This is evident from
the transformed velocity q′ defined by (3.7). That is, unlike the Galilei and Poincaré
groups, the Galilean line group transforms the momentum (or, equivalently, velo-
city) of a particle into a time-dependent vector field. We can certainly anticipate such
a time dependence from the transformation properties of a vector field under general
coordinate transformations, now applied to (3.1). Recall that if x′μ = f μ(x ),
μ = 0, 1, 2, 3 is a general coordinate transformation on a four-dimensional space-
time manifold, then a vector field V on that manifold transforms as
∂x′ μ ν ∂f μ ν
V ′ μ (x′) = V ( x ) = V (x ) (3.10)
∂x ν ∂x ν
where we have adopted the four vector notation for simplicity. Letting μ = 0 stand
for time t and using (3.1) in (3.10), we obtain
∂t′ ν
V ′0 (x′) = V (x ) = V 0(x )
∂x ν
∂x′i ν
V ′i (x′) = V (x )
∂x ν
i
( )
= R ji V j (x ) + R˙ j x j + a˙ i V 0(x ). (3.11)

3-5
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

Not surprisingly, we see from the first equality of (3.11) that the time-component of
any vector field must transform trivially under the Galilean line group, an immediate
consequence of the universal time of a Galilean spacetime. Therefore, we can simply
set V 0 = 1. If we now take Vi to be velocity qi and evaluate (3.11) along an integral
curve of qi, we get the velocity transformation formula of (3.7), which we obtained
from the induced representations of .

Generators and observables. The operators (3.5) reduce to a one parameter group of
uinitary operators when the line group transformations are restricted to time
translations g = (I , 0, b ). The same is true for constant spacial translations,
g = (I , a (0), 0). Therefore, the Hamiltonian and momentum operators may be
obtained as generators of these one parameter subgroups in the usual way:
dU (b ) (0)
H=i db b=0
∣ and P i = −i dU (a(0) ) ∣a (0) =0 . Using the explicit form (3.5) of U(g ), we
da i i

can obtain the action of P i and H on the velocity wavefunctions ϕ(q ) = 〈q ∣ ϕ〉:

( P iϕ) (q ) = mqiϕ (q )
⎛ P2 ⎛ 1 ⎞⎞
(Hϕ)(q ) = ⎜ + mq˙ · ⎜ X + a q ⎟ ⎟ϕ (q ), (3.12)
⎝ 2m ⎝ 2 ⎠⎠

where X is a position operator that is canonically conjugated to momentum,


[X i , P j ] = iδ ijI .
The first expression of (3.12) justifies the interpretation of q as velocity. Unlike
the inertial case, it is now a time-dependent vector field, q(t ). The first term of the
Hamiltonian shows the usual contributions of the kinetic energy. The second term
encodes the non-inertial effects and may be interpreted as a fictitious potential
energy. As one might expect on the basis of the equivalence principle, it is pro-
portional to the inertial mass. Hence, it may be understood as a gravitational
potential energy, arising here as a purely quantum mechanical result rather
than, say, from ‘quantization’ of a preexisting classical field. The main limitation of
this result is that only gravitational potentials that are linear in position can be
accommodated, though these fields may have arbitrary time dependence. On the
other hand, as seen below, the fictitious potential term is not proportional to inertial
mass for more general representations of , leading to the inference that the
equivalence principle can be violated at the quantum level if there exist physical
systems that are the realizations of these general representations.
It was mentioned above that the Galilean line group  can be considered a
‘gauged Galilei group’. We can appeal to this gauge structure to obtain an alter-
native derivation of the Hamiltonian in accelerating frames. The key observation is
that the unitary operators U(g) do not commute with the derivative operator d
dt

3-6
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

because  is parametrized by analytic functions of time. Thus, if we apply U(g) to


both sides of the Schrödinger equation, we get

iU (g ) = U (g ) Hϕ = U (g ) HU −1(g ) U (g ) ϕ
dt
⎛ d(U (g )ϕ) dU ( g ) ⎞
i⎜ − ϕ⎟ = H ′U (g ) ϕ
⎝ dt dt ⎠

dϕ′
i
dt
(
= H ′ + H f ϕ′ ,) (3.13)

where ϕ′ := U(g ) ϕ is the transformed wavefunction, H ′ := U (g ) HU −1(g ) is unitarily


dU (g )
equivalent to H, the Hamiltonian in the inertial frame, and H f := i dt U −1(g ) is the
Hamiltonian corresponding to the fictitious forces. Thus, we see that the Hamilto-
nian in the accelerating frame is not unitarily equivalent to the Hamiltonian in the
inertial frame because it is the sum of a term unitarily equivalent to the inertial frame
Hamiltonian and the ‘fictitious’ Hamiltonian. This means that while probabilities
are conserved in going from inertial to accelerating reference frames, the meaning
of these probabilities has changed because the Hamiltonian in the accelerating frame
is not unitarily equivalent to the Hamiltonian in the inertial frame. Hence, we see
that the Hamiltonian that follows from the unitary representations of  has exactly
the form (3.13) that we would expect from the gauge structure of acceleration
transformations.

Synthetic magnetic fields. In addition to leading to a mathematically sound defi-


nition of one-particle quantum states in non-inertial reference frames, the unitary
cocycle representations of the Galilean line group provide us with the means to
systematically study the role of non-inertial effects in quantum systems. Among
these are the synthetic or simulated magnetic fields that appear in rotating
reference frames. To see how this comes about, consider a transformation from
an inertial reference frame to a rotating reference frame. In an inertial frame,
the velocity is time independent, which we denote by q (0) . Therefore, from (3.5)
and (3.7),

∣ q〉 = U (0, R , 0 , 0) q 0 , (3.14)

where

q = Rq 0 + R˙ a q 0
d
=
dt
(Ra q 0 )
= Rq 0 + ω × a q. (3.15)

3-7
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

Differentiating this expression gives

q˙ = R˙ q 0 + ω
˙ × aq + ω × q

(
˙ T q − ω × aq + ω
= RR )
˙ × aq + ω × q

˙ × a q + 2ω × q − ω × ω × a q .
= ω ( ) (3.16)

It is clear from (3.16) that we recover the centrifugal term −ω × (ω × aq ), the


Coriolis term 2ω × q and, for time-dependent angular velocities, the Euler term
ω̇ × aq . We can obtain the contributions of these non-inertial effects to the
Hamiltonian by substituting (3.16) in (3.12), where all three terms of (3.16) are
multiplied by the mass m and position operator X , leading to fictitious potential
energy terms of the expected form. After a bit of algebra, we obtain
⎛ 1 ⎞
H=⎜ (P − A)2 + A0 ⎟ , (3.17)
⎝ 2m ⎠

where

⎛ 1 ⎞
A = 2mω × ⎜ X + a q ⎟
⎝ 2 ⎠
⎛ ⎛ 1 ⎞⎞
A0 = −2m ⎜ ω × ⎜ X + a q ⎟ ⎟ · (ω × X ) − ma q · ( ω
˙ × X ). (3.18)
⎝ ⎝ 2 ⎠⎠

Thus, we see that the Coriolis term gives rise to a vector potential A, while the
centrifugal and Euler terms lead to a scalar potential A0. As done in electrodynamics,
we can define fields from these potentials. In particular, the vector potential A leads
to a synthetic magnetic field that is proportional to angular velocity ω. It is synthetic
in that a particle would couple to it by way of its mass, not electric charge. Much like
the wavefunction of a charged particle in a real magnetic field, the wavefunction of a
massive quantum particle, charged or neutral, in such a synthetic magnetic field
would undergo a phase shift which, for instance, in turn could lead to an interference
effect in an interferometry experiment. This effect was observed in the beautiful
experiment of Werner, Staudenmann and Colella [9] that measured a phase shift
in the neutron wavefunction due to the rotation of the Earth. The subsequent
theoretical analyses [10–12] of this experiment relied on various analogies, such as
that between the Coriolis force 2mv × ω and the Lorentz force qv × B , to associate a
simulated magnetic field to a rotating reference frame. By contrast, here we see that
this synthetic U (1)-gauge field can be derived as a rigorous mathematical result from
the representations of the Galilean line group. Furthermore, (3.18) shows that there
is an additional contribution to the phase shift coming from the Euler term when the
rotating reference frame has an angular velocity that is time-dependent. As shown in
[13], this term would be missed if one were to start with the classical Hamiltonian

3-8
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

and transition to the quantum theory, highlighting a sharp difference between a


quantum theory based on group representations and one obtained by ‘quantizing’ a
classical theory.

Equivalence principle. The representation given by (3.5) is the simplest of a large class
of inequivalent representations of . As mentioned above, the construction of these
representations must begin with the choice of a suitable function  ×  → -() that
reduces to the Galilean two-cocycle (2.9) when the transformations are restricted to
inertial frames. This is the mathematical condition that ensures that the resulting
representation of the line group  contains a projective representation of the Galilei
group as a subrepresentation.
The simplest function  ×  → - with this reduction property is
m
ω( g2, g1) =
2
( (Λ a )Λ b1 2 b1 ( )
· R 2 a˙1 − Λ b1a˙ 2 · Λ b1R 2 a1 , ) (3.19)

and the representation (3.5) follows from this choice. More generally, the required
function  ×  → - can be of the form

m

⎛ dk a 2 ⎞ dl a1
ω( g2, g1) =
2
∑ ( βk γl − γk βl ) ⎜⎝ Λ b1
dt k ⎠
⎟ · Λ (
b1R 2
dt l
) , (3.20)
k, l = 0

where the constants βk and γk are arbitrary, except for the constraint β0 γ1 − γ0 β1 = 1,
which ensures that (3.20) reduces to (2.9). We mention here that neither (3.19) nor
(3.20) is a two-cocycle. Rather, they are two cochains, which are necessarily three-
cocycles. The failure of (3.19) or (3.20) to be two-cocycles means that these
representations of the Galilean line group to do not correspond to central (or, for
that matter, non-central) group extensions of  in the way that the projective
representations of the Galilei group correspond to central extensions. We discuss
group extensions in appendix B. We will return to the absence of two-cocycles on 
and its implications for the representations in the last section of this chapter.
A more detailed analysis is given in [7].
The representations of  that follow from (3.20) lead to a Hamiltonian of
the form
⎛ 2 ⎛ ∞
β k dk ( a q ) ⎞
P
(Hϕ)(q ) = ⎜⎜ + mq˙ · ⎜⎜ X + a q I + ∑ k ⎟

⎝ 2m ⎝ k=1
m d t ⎠

1

1

dk a q ∞
dk a q ⎞ ⎞
+ P · ⎜⎜ ∑ βk − ∑ γk ⎟ ⎟ϕ (q ). (3.21)
2 ⎝ m k=1 dt k k=2 dt k ⎟⎠ ⎟⎠

By comparison with the Hamiltonian for the simplest case, (3.12), we now see that
the fictitious potential energy term is not simply proportional to the inertial mass.
Further, the last term is proportional to the momentum, a property that we never see
in fictitious forces in classical theory.

3-9
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

Let us now consider the implications of the form of this general Hamiltonian for
the equivalence principle. Classically, it is an empirical fact that the trajectory of a
particle moving under the influence of gravity is independent of its mass, which leads
to the inference that the ratio of gravitational mass mg to inertial mass m is a universal
constant. Further, the trajectory of a particle is necessarily independent of its mass m
when its motion is governed by fictitious forces, suggesting that gravitational forces
may be simulated by transforming to suitable (possibly local) non-inertial reference
frames. Quantum mechanically, the situation is rather different in that the inertial
mass in the fictitious potential term of (3.12) (or more generally, (3.21)) does not drop
out in the Schrödinger equation and the evolution of a quantum state governed
by (3.12) depends on = .1 Hence, non-inertial effects are measurably different for
m
quantum systems with different masses. Likewise, the evolution of a particle in a
gravitational potential ϕg , with Hamiltonian
P2
H= + m g ϕg , (3.22)
2m
= = mg
depends on both and , and not solely on the ratio . In fact, it is known that
m mg m

the wave vector of a state evolving under (3.22) depends on (mm g )1/3, while the
mg
center of mass motion (semi-classically) depends only on the ratio [14]. Therefore,
m
unlike the classical case, what we mean by the equivalence principle in quantum
mechanics is not whether mass disappears in equations of motion, but whether the
evolution under (3.12) or (3.21) may be simulated by (3.22) with a suitable grav-
mg
itational potential ϕg . Provided the universality of holds, it is clear that this is
m
always possible for (3.12), but not for (3.21) due to the term linear in momentum.
Therefore, (3.21) shows that the equivalence principle would not be upheld for
physical systems governed by the general representations of . Such violations are
completely a quantum effect because, as shown in [8], they disappear in the classical
limit. It remains an interesting open problem, which must be settled by experiment,
whether these equivalence principle violating representations are realized in nature.

Cohomology and loop prolongations. As mentioned in chapter 2, the physically


relevant unitary representations of the Galilei group . are the projective repre-
sentations. A projective representation of a group G, defined by a function
ω : G × G → , is equivalent to a true, vector representation of a central extension
of G by the real line , a theorem proved by Bargmann [15]. In appendix B,
we denote this central extension by G˜ := {g˜ = (φ , g ) : φ ∈ , g ∈ G}. The com-
position rule under which G̃ becomes a group is
(
( φ2 , g2 )( φ1, g1) = φ2 + φ1 + ω(g2, g1) , g2 g1 . ) (3.23)
The group G̃ is called a central extension because the subgroup (φ , e ) ≈ 
that extends G into G̃ is a central subgroup, meaning that it commutes with all

1
In this section, we include the ℏ explicitly so that its role in the theory is fully emphasized.

3-10
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

elements of G̃ . The two-cocycle condition (2.8), which follows from the associa-
tivity of the operators U(g) that furnish a projective representation of G, is precisely
the condition that ensures that the composition rule (3.23) is associative.
As mentioned in chapter 2, every inequivalent two-cocycle (2.8) leads to an
inequivalent projective representation and, in light of the Bargmann result,
an inequivalent central extension of G by .
We emphasized above that, if we are to formulate a notion of quantum particle
that holds for non-inertial observers by means of unitary irreducible representations
of the Galilean line group , the relevant representations of  must contain a
projective representation of the Galilei group as a subrepresentation so that this
more general notion of one-particle states reduces to the usual Galilean notion.
Now, given the relationship between projective representations and central exten-
sions, we can state this embedding condition as follows:

Construct all extensions of the Galilean line group that contains


a given central extension of the Galilei group. (3.24)

An interesting property of the Galilean line group  proved in [3] is that, while it
contains the Galilei group . as a subgroup, it has no group extensions that fulfill this
embedding requirement. It was pointed out above that, because the group elements
of  are parametrized by functions of time, rather than real numbers as is the case
for . , the possible extensions of  must be by the Abelian group of analytic (scalar)
functions -(), rather than by the real line . These general extensions, if they exist,
are again determined by functions  ×  → -(), which fulfill a two-cocycle
condition that is a generalization of the two-cocycle condition (2.8) for the central
extensions of the Galilei group [3]:

δω( g3, g2, g1) := Λ b1ω( g3, g2 ) + ω( g3 g2, g1) − ω( g2, g1) − ω( g3, g2 g1) = 0, (3.25)

where Λ b is the shift operator used to define the composition rule (3.2) of . It is a
particular example of the homomorphism σ : G → Aut (A) from a group G to the
automorphism group of an Abelian group A that defines the group extensions of
G by A, given in (B.9).
Therefore, the task at hand is to construct a function  ×  → -() subject to
the dual requirements that it fulfill the two-cocycle property (3.25) and reduce to the
Galilean two-cocycle (2.9). Such a function can be used to construct an extension
of  of the form (B.9) that solves the embedding condition (3.24). Now, since
the Galilean two-cocyle (2.9) (or any other Galilean two-cocycle) involves velocities,
a two-cocycle of  must involve the derivatives ȧ of translations a if it is to reduce
to (2.9). This leads to an additional complication: under time-dependent rotations,
a and ȧ do not transform the same way owing to the inhomogeneous Ṙ -term in
d
dt
(Ra ) = Ra˙ + R˙ a . This difficulty is a familiar one from gauge theories. In the
present case, the trouble is algebraic: any function ω on  ×  taking values in
-() and containing derivatives of translations fails to fulfill the two-cocycle

3-11
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

condition (3.24)! For example, the substitution of (3.19) into the right-hand-side of
(3.25) shows
Λ b1ω( g3, g2 ) + ω( g3 g2, g1) − ω( g2, g1) − ω( g3, g2 g1)
1
=
2
( (
Λ b1ω2 · Λ b1R 2 a1 × Λ b2 + b1 R 3T a 3 ))
1
(
− Λ b2 + b1ω3 · Λ b1a 2 × Λ b1R 2 a1 = 0.
2
) (3.26)

It is evident from this that it is precisely the time dependence of rotation matrices, by
way of the inhomogeneous R˙ a in the derivatives ddt (Ra ), that violates the two-cocycle
condition (3.25) and precludes group extensions. As shown in [3], if only linear
accelerations are considered, then there do exist group extensions that fulfill (3.24),
albeit these extensions are not central. Therefore, we see that there is a rather
striking difference between linear and rotational accelerations that we must contend
with when trying to ground quantum mechanics in the unitary representations of the
Galilean line group.
In light of the difficulty with time-dependent rotations, the embedding require-
ment (3.24) is fulfilled not by group extensions of , but by certain non-associative
extensions that fit very nicely into the theory of loop prolongations developed by
Eilenberg and MacLane as a part of their of algebraic cohomology of groups [16, 17].
A loop is a set with a binary operation that fulfills all the axioms of a group except
associativity, and therewith also the existence of a two-sided inverse for every element
(the left and right inverses may be distinct). Further, given three elements a, b, c
of a loop L, there exists a unique element A(a, b, c ) ∈ L , called an associator,
such that
a (bc ) = A (a , b , c ) [ (ab) c ] . (3.27)
Associators measure deviations from associativity, much like commutators measure
the lack of commutativity. We will not review the general theory of reference [16]
here, but only mention that the construction of loop prolongations runs parallel
to that of group extensions, with a little additional care required to handle the
complications resulting from the failure of (3.25).
Since (3.24) is the key requirement, the construction of a loop prolongation of 
must start with a function, called a two-cochain, ω(g2, g1 ) :  ×  → -() that
reduces to (2.9). These functions are of the general form (3.20), of which (3.19) is the
¯ := {(φ , g ): φ ∈ -(), g ∈ }, together with
simplest. It can be shown that the set 
the composition rule

(
g¯2 g¯1 = Λ b1φ2 + φ1 + ω( g2, g1), g2 g1 ) (3.28)

fulfills all axioms of a loop prolongation of a group by an Abelian group [7].


Further, the associators of ̄ are of the form

(
A( g¯3, g¯2, g¯1) = (δω)(g3, g2, g1) , e , ) (3.29)

3-12
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

where δω is the function defined by the right-hand-side of (3.25). If (3.25) holds, i.e.,
δω = 0, we see that all associators vanish and we have a group extension. We also
see from (3.29) that associators all belong to the Abelian subgroup -() of ̄, a
property that makes the breaking of associativity of the loop ̄ systematic and
controlled.
The theoretical framework on which the above analysis is based is called algebraic
cohomology of groups [16, 18]. A summary of the results important for the appli-
cations we are considering can be found in [7]. As shown there, the function
δω(g3, g2, g1 ) that defines the associators of the loop  is in fact a three-cocycle, an
element of the third cohomology group HΛ3b (, -()). (Similarly, the two-cocycles
are elements of the second cohomology group, to which the group of group
extensions is isomorphic.) Three-cocycles are not very common in physics, but they
have been considered, particularly in connection with magnetic monopoles [19],
although many of these studies are largely heuristic or formal. The physical content
of the present case is also quite different from these previous studies.
A class of unitary representations of loop prolongations that correspond to
cochains (3.19) has been constructed in [7] and a similar construction also holds for
the cochains (3.20). It is important to note that the unitary operators U(g¯ ) that
furnish these representations do uphold associativity, as linear operators necessarily
do, even though they furnish a representation of a non-associative loop. Thus,
the physically relevant representations of the Galilean line group correspond to
representations of its loop prolongations, not central group extensions. This is a
generalization of the Bargmann result [3] on projective representations and central
extensions. To the best of our knowledge, the study of [7] is the first time that
representations of loop prolongations of an infinite dimensional Lie group have been
constructed and utilized in a physical theory.
Therefore, summarizing the above discussion, we state our broad theoretical
claim as follows: to ground quantum theory in the expanded relativity principle
that includes non-inertial reference frames in a Galilean spacetime, we must go
beyond groups and two-cocycles, the familiar framework within which symmetry
transformations are understood, and allow for loop prolongations based on
three-cocycles.

Bibliography
[1] José J V and Saletan E J 2002 Classical Dynamics: A Contemporary Approach (Cambridge:
Cambridge University Press)
[2] Klink W H 1997 Quantum mechanics in noninertial reference frames: I. Nonrelativistic
quantum mechanics Ann. Phys. 260 27
[3] MacGregor B A, McCoy A E and Wickramasekara S 2012 Unitary cocycle representations
of the Galilean line group: quantum mechanical principle of equivalence Ann. Phys. 327 2310
[4] Taylor M E 1986 Noncommutative Harmonic Analysis (Mathematical Surveys and Mono-
graphs no. 22) (Providence, RI: American Mathematical Society)
[5] Moody R V 1967 Bull. Amer. Math. Soc. 73 217
Kac V G 1985 Infinite Dimensional Lie Algebras (Cambridge: Cambridge University Press)

3-13
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

[6] Negro J, del Olmo M A and Rodríguez-Marco A 1997 Nonrelativistic conformal groups
J. Math. Phys. 38 3786
Aizawa N Galilean conformal algebras in two spatial dimensions (arXiv:1112.0634)
[7] Klink W H and Wickramasekara S 2013 Quantum mechanics in noninertial reference
frames: time-dependent rotations and loop prolongations Ann. Phys. 336 261
[8] Klink W H and Wickramasekara S 2013 Quantum mechanics in noninertial reference
frames: violations of the nonrelativistic equivalence principle Ann. Phys. 340 94
[9] Werner S A, Staudenmann J-L and Colella R 1979 Effect of Earth’s rotation on the quantum
mechanical phase of the neutron Phys. Rev. Lett. 42 1103
[10] Sakurai J J 1980 Comments on quantum-mechanical interference due to the Earth’s rotation
Phys. Rev. D 21 2993
[11] Mashhoon B 1988 Neutron interferometry in a rotating frame of reference Phys. Rev. Lett.
61 2639
[12] Anandan J 1992 Comment on spin-rotation-gravity coupling Phys. Rev. Lett. 68 3809
[13] Klink W H and Wickramasekara S 2013 Fictitious forces and simulated magnetic fields in
rotating reference frames Phys. Rev. Lett. 111 160404
[14] Kajari E, Harshman N L, Rasel E M, Stenholm S, Süssmann G and Schleich W P 2010
Inertial and gravitational mass in quantum mechanics Appl. Phys. B 100 43
[15] Bargmann V 1954 On unitary ray representations of continuous groups Ann. Math. 59 1
[16] Eilenberg S and MacLane S 1947 Algebraic cohomology of groups and loops Duke Math. J.
14 435
See also Smith J D H 2006 An Introduction to Quasigroups and Their Representations
(London: Taylor & Francis)
[17] De Azcárraga J A and Izquierdo J M 1998 Lie Groups, Lie Algebras, Cohomology and Some
Applications in Physics (Cambridge: Cambridge University Press)
[18] Eilenberg S and MacLane S 1947 Cohomology theory of abstract groups I Ann. Math. 48 51
Eilenberg S and MacLane S 1947 Abstract theory of groups and loops II Ann. Math. 48 326
[19] Jackiw R 1985 Three-cocycle in mathematics and physics Phys. Rev. Lett. 54 159
Hou B Y, Hou B Y and Wang P 1986 How to eliminate the dilemma in 3-cocycle Ann. Phys.
171 172
Grossman B 1985 The meaning of the third cocycle in the group cohomology of nonabelian
gauge theories Phys. Lett. 152 93
Nesterov A I 2004 Three-cocycles, non associative gauge transformations and Dirac’s
monopole Phys. Lett. A 328 110
Wess J and Zumino B 1971 Consequences of anomalous ward identities Phys. Lett. 37B 95

3-14
IOP Concise Physics

Relativity, Symmetry and the Structure of Quantum Theory I:


Galilean quantum theory
William H Klink and Sujeev Wickramasekara

Chapter 4
Multiparticle systems and interactions

Thus far we have discussed only single-particle systems. There is, however, a
natural group theoretical procedure for obtaining states of multiparticle systems:
multiparticle states are tensor products of single particle states. But the tensor
product of irreducible single particle states is no longer irreducible. That means
there are different ways of writing multiparticle states, with transformation
coefficients connecting the different possibilities. The coefficients that transform
between tensor product states and irreducible multiparticle states are called
Clebsch–Gordan coefficients. In this chapter we show how the structure of the
Galilei group representations enables one to deal with multiparticle states and, in
particular, how to introduce interactions in many-particle systems in a Galilean
invariant way.
To begin, we analyze two particle states. A natural basis for a two-particle
system is the two-fold tensor product of single particle basis states, which, in
a momentum basis is written as ∣q1, σ1 〉 ⊗ ∣ q2, σ2 〉; here the mass and spin
labels have been suppressed. Then a two particle state can be written as
∣ ψ2 〉 : = Σ ∫ ψ2 (q1, σ1; q2, σ2 ) ∣ q1, σ1 〉 ⊗ ∣ q2, σ2 〉, with norm ∣∣ψ2∣∣2 = Σ ∫ d3q1 d3q2 ∣ψ2∣2 .
The tensor product basis state transforms reducibly under the Galilei group. The
transformation to an irreducible basis begins by defining a new two particle state
∣ Q , q, s, σ 〉 : = Σσ1σ 2 ∣ q1, σ1 〉 ⊗ ∣ q2, σ2 〉〈s1 σ1, s2 σ2 ∣s, σ 〉, where MQ = m1 q1 + m 2 q2,
q = q1 − q2 , with M = m1 + m 2 . Note that Q and q are the derivatives of the usual
definitions given for the center of mass and relative coordinates in classical
m1 m 2
mechanics; also, as in classical mechanics, the reduced mass is μ = M . Further, the
two spins have been coupled together to form an overall spin state, using the SU (2)
Clebsch–Gordan coefficients discussed in appendix C. The action of the Galilei
group elements on the two particle state follows from the product action on one-
particle states:

doi:10.1088/978-1-6270-5624-3ch4 4-1 ª Morgan & Claypool Publishers 2015


Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

U (a ) Q , q , s , σ = e imQ·a Q , q , s , σ
U (v ) Q , q , s , σ = Q + v , q , s , σ

U (R ) Q , q , s , σ = ∑ RQ, Rq, s , σ′ D σs′σ(R )

⎛ MQ 2 μq 2 ⎞
−i ⎜ + ⎟b
U (b ) Q , q , s , σ = e ⎝ 2 2 ⎠ Q, q, s , σ . (4.1)

The action of time translation and rotations indicates this state is not irreducible; two
steps are needed to make the action of rotations on the two particle state irreducible.
First, the internal velocity q is written as a magnitude times a unit vector, q = qqˆ , and
the unit vector transformed to orbital angular momentum l, and projection ml using
spherical harmonics. Then, the total spin angular momentum is coupled to the orbital
angular momentum to give the total angular momentum j, with projection mj.
The resulting two particle state is now written as ∣ Q , q, j , m j , l , s〉 and transforms as

U (a ) Q , q , j , m j , l , s = e iMQ·a Q , q , j , m j , l , s

U (v ) Q , q , j , m j , l , s = Q + v , q , j , m j , l , s
j
U (R ) Q , q , j , m j , l , s = ∑ RQ, q, j, m j′, l , s D
m ′j m j
(R )

⎛ MQ 2 μq 2 ⎞
U (b ) Q , q , j , m j , l , s = e − i ⎝
⎜ + ⎟b
2 2 ⎠ Q, q, j, m j , l , s
⎛ MQ 2 μq 2 ⎞
H0 Q , q , j , m j , l , s = ⎜ + ⎟ Q, q, j, m j , l , s . (4.2)
⎝ 2 2 ⎠

The labels Q , j and mj correspond to one-particle labels, for a particle of mass M and
spin j. But the additional labels indicate a particle with internal structure, with the
μq 2
label q giving an internal kinetic energy 2 , and the spin labels l and s related to
internal orbital and spin angular momentum. Note also that the Hamiltonian H0 in
M
(4.2) is the sum of kinetic energy of the ‘particle’ as a whole, 2 Q 2 , plus the internal
μ
kinetic energy, 2 q 2 . Here and below, we use the subscript 0, as in the two-particle
Hamiltonian H0 in (4.2), to indicate that there are no interactions between the particles.
The coefficients that transform between two-particle basis states and irreducible
two-particle states are called the Galilei group Clebsch-Gordan coefficients. They
are given by

q1, σ1 ⊗ q2, σ2 Q , q , j , m j , l , s = ∑ ∫ d2qY l( )( 2 2)


* qˆ MQ − m q − m q δ 3
ˆ lm 1 1

× δ 3(q − q1 + q2 ) l , ml ; s , m s j , m j

× s1, σ1; s2 , σ2 s , m s , (4.3)

4-2
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

where 〈· ·〉 are SU (2) Clebsch–Gordan coefficients, which couple the various angular
momenta together. They are discussed in detail in appendix C.
The Hamiltonian given in (4.2) is the Hamiltonian for two non-interacting
particles. The key idea for adding interactions between the two particles is to require
that the generators of the Galilei group still respect the Galilean commutation
relations (2.21) when the generators are modified to include interactions. For
Newton relativity this is generally done by modifying the Hamiltonian while
leaving the other generators unchanged. As will be seen in the succeeding volume,
for Einstein relativity this is no longer possible. It is necessary to modify more than
one generator in order to have a Poincaré invariant theory. In any event, the
commutation relations for the non-interacting Hamiltonian can be written in the
following form:
⎡⎣ H0, P ⎤⎦ = 0 (4.4a )

UR H0 UR−1 = 0 (4.4b)

m 2
Uv H0 Uv−1 = H0 + P · v + v I. (4.4c )
2
This last equation is most readily obtained by computing Uv Ub Uv−1 and letting b be
infinitesimally small.
It is now straightforward to introduce interactions into the Hamiltonian in a
Galilean invariant way. Simply modify the free Hamiltonian (4.2) by adding to it an
operator whose matrix elements depend only on q, l , s and which is diagonal in j.
That is, the operator depends only on ‘internal’ variables, and not on the ‘one-
particle’ variables. While this indeed generates Galilean invariant interactions, a
more familiar procedure for generating Galilean invariant interactions is to make
use of the intermediate basis states, whose transformation properties are given
in (4.1). Consider a potential operator, V, which is added to the Hamiltonian,
H = H0 + V and whose kernel is written as Q′, q′, s′, σ ′ V Q , q, s, σ . In order
that the commutation relations (2.21) remain unchanged when the potential is added
to the free Hamiltonian, V must commute with the momentum, angular momentum
and boost generators. This then constrains the kernel of V to have the following
general form:
Q′ , q′ , s′ , σ ′ V Q , q , s , σ = δ 3(Q′ − Q ) f (q′ , q , q′ · q ) δ s′s δ σ′σ . (4.5)
The delta function is required so that V commutes with momentum, and the arbi-
trary function f can be a function of internal velocities only as a consequence of
commuting with the boost generator. Then rotational invariance requires that the
initial and final spin and spin projections be identical.
As an example of such interactions, any local interaction such as the Coulomb or
Yukawa forces the function f to have the form f ( q − q′ ). For example, the
k
Coulomb interaction has the form f = 2
, with k a constant.
(q − q ′)

4-3
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

It is also possible to introduce external forces, such as an external electromagnetic


field acting on a charged particle, by letting the mass of, say, particle 2 go to infinity,
while the mass of particle 1 stays finite; for then the total mass M goes to infinity,
while the reduced mass becomes the mass of particle 1. Thus Galilean invariance can
be used to obtain the form for external interactions, even though the Galilean
invariance is now broken.
The procedures outlined above readily generalize to multiparticle systems. An
n-particle state is the n-fold tensor product of one-particle states, written
∣ q1, σ1 〉 ⊗ ∣ q2, σ2 〉 ⊗ … ⊗ ∣ qn, σn 〉. However, the reduction to an irreducible state
with internal degrees of freedom is now much more ambiguous. One possibility
involves stepwise coupling, where particle 1 is coupled to 2, then 1–2 coupled to 3,
and so forth; the internal variables are the q, l, s variables for each of the subsystems.
The problem with such a procedure is particularly evident for identical particles,
where particle labels are interchanged. Coefficients transforming between different
coupling schemes are called Racah coefficients; they can be calculated by using
another procedure for obtaining irreducible states.
This procedure is called simultaneous coupling, wherein all the one-particle states
are coupled simultaneously, rather than in a stepwise fashion. In this case the
internal variables are sets of dot products of one-particle velocities, along with the
total orbital angular momentum of the n-particle system and the spins of the indi-
vidual systems. As in the two-particle case, these spins can be coupled to give the
total intrinsic spin, which in turn is coupled to the overall orbital angular momentum
to give the total angular momentum. For more details, see [1]. Given the simulta-
neous procedure for reducing n-particle states to irreducible states, it is possible to
find coefficients that transform between any stepwise scheme and the simultaneous
scheme. Racah coefficients then are products of such coefficients.
Similarly the Hamiltonian for non-interacting n-particle systems can be modified
to include interactions between the particles in such a way that the commutation
relations of the n-particle potentials commute with the other Galilei generators. The
procedure is analogous to that given for two particles, but is notationally much more
cumbersome.

Creation and annihilation operators. For systems of identical particles, such as an


ensemble of electrons or photons, there is a more elegant method for dealing with
many-particle systems that makes use of operators called creation and annihilation
operators. These are defined on Hilbert spaces called Fock spaces, which are built
out of direct sums of n-fold tensor product spaces, with n running from zero to
infinity. Two types of such spaces are of particular importance, corresponding to
fermionic or bosonic types of particles. Fermionic particles, such as electrons or
protons, have half integral spin and their Fock spaces are given by direct sums of
antisymmetric n-fold tensor product spaces. Bosons—for example, photons or pi
mesons—are particles with integral spin values, and their Fock spaces are direct
sums of symmetrized n-fold tensor product spaces. In the following we deal only
with bosonic particles, but the results are easily generalized to fermions [2]. The
theorem connecting the spin of a particle with its statistics is also discussed in [2].

4-4
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

Consider, then, the n-fold symmetrized tensor product of one-particle irreducible


Hilbert spaces, / n = (/ ⊗ … ⊗ / )sym n and add these all up to define the symmetrized
Fock space - = Σ n=0 / n ; n = 0 corresponds to the vacuum, the space of no particles,
and is denoted by ∣ 0〉 (that is, ∣ 0〉 = /0). ‘Symmetrized’ means that for any wave
function in the n-fold tensor product space, it is unchanged under any particle label
exchange. For example, a symmetrized two-particle wave function has the property
that ϕ (q1, σ1; q2, σ2 ) = ϕ (q2, σ2; q1, σ1 ); the symmetrized Fock space is then the sum of
all such symmetrized wave functions, with the sum ranging from zero to infinity.
Creation operators are operators that act in - by increasing the particle number
by one, and annihilation operators decrease the particle number by one. In particular,
the annihilation operator acting on the vacuum gives 0 (annihilates the vacuum),
while the creation operator acting on the vacuum produces a one-particle state.
More precisely, creation and annihilation operators satisfy certain algebraic
properties that make them ideally suited for dealing with identical many-particle
systems. If c(q, σ ) is an annihilation operator indexed by particle velocity q and spin
projection σ, and its adjoint is the creation operator c†(q, σ ), then these operators
satisfy the following properties
⎡⎣ c(q , σ ), c †( q′ , σ ′) ⎤⎦ = δ 3( q − q′) δ σσ′ (4.6a )

⎡⎣ c( q , σ ), c( q′ , σ ′) ⎤⎦ = 0 (4.6b)

c( q , σ ) 0 = 0 (4.6c )

c †( q , σ ) 0 = q , σ . (4.6d )

Since mass and spin are all the same for identical particles, these labels have been
suppressed. Many-particle states are produced by products of creation operators
acting on the vacuum:

∣ q1, σ1〉 ⊗ … ⊗ ∣ qn, σn 〉 = c †( q1, σ1) … c † ( qn, σn ) 0 ; (4.7)

ϕn = ∑∫ d3q1 … d3qn ϕ ( q1, σ1, … , qn, σn )


σi

c † ( q1, σ1) … c † ( qn, σn ) 0 (4.8)

is a square-integrable many-body wavefunction that is automatically correctly


symmetrized because of the properties of the creation operators. That is, the adjoint
of (4.6b) states that products of creation operators commute among themselves, so
that the order of creation operators in a many-particle state is irrelevant. The
symmetrized Fock space can now be defined via products of creation operators
acting on the vacuum and is equivalent to the original definition as an infinite direct
sum of symmetrized n-fold tensor products.

4-5
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

The vacuum state corresponds to the identity representation of the Galilei group,
so that the action of Galilei group elements is trivial (that is, every element of the
Galilei group multiplies the vacuum state by one). Since c†(q, σ ) generates a one-
particle state from the vacuum, it must transform as a one-particle operator; that is,
under a spatial translation,
U ( a ) c †( q , σ ) 0 = U ( a ) c †( q , σ ) U −1( a ) U ( a ) 0
= e imq·ac †( q , σ ) 0 (4.9)

U ( a ) c †( q , σ ) U (a )−1 = e imq·ac †( q , σ ), (4.10)

with analogous results for the other Galilei group elements. Similarly, an n-particle
state will have well-defined transformation properties under the Galilei group; for
example,

U ( a ) c † ( q1, σ1) … c † (qn, σn ∣ 0〉 = e im∑qi·ac † ( q1, σ1) … c † (qn, σn ∣ 0〉. (4.11)

Creation and annihilation operators also function as bases for operators. Consider
first a one-body operator defined by

O1 := ∑ ∫ d3q d3q′ c †(q , σ ) q , σ O q′ , σ ′ c(q′ , σ ′), (4.12)

where O is the one-body matrix element of the operator O. Prominent examples


are the free Hamiltonian and momentum operators:

H1 = ∑∫ d3q d3q′ c †( q , σ ) q , σ H0 q′ , σ ′ c(q′ , σ ′)

mq 2 †
= ∑∫ d3q
2
c ( q , σ ) c( q , σ ) (4.13a )

P1 = ∑∫ d3q mqc †( q , σ ) c( q , σ ). (4.13b)

From these expressions, with the use of (4.6) and the transformation properties of
creation and annihilation operators such as (4.10), we can easily obtain the com-
mutation relations and transformation formulas of these operators:
⎡⎣ H1, P1 ⎤⎦ = 0, ⎡⎣ H1, U (R ) ⎤⎦ = 0 (4.14a )

mq 2 †
U (v) H1U −1(v) = ∑∫ d3q c (q − v , σ ) c (q − v , σ )
2
m(q + v)2 †
= ∑∫ d3q c (q , σ ) c (q , σ )
2
mv 2
= H1 + P1 · v + N, (4.14b)
2

4-6
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

which agrees with (4.4) for N = 1. Here N is the number operator,


N := Σ ∫ d3q c†(q, σ ) c(q, σ ). But now these one-body operators are well-defined on
any n-particle subspace of the full Fock space and not just on a one- or two-particle
space, as before.
Generalizing, n-body operators are defined in terms of n creation and n annihi-
lation operators. The most important is the two-body operator that produces
interactions when added to the free Hamiltonian:
H = H1 + H2

H2 : = ∑∫ d3q1 d3q2 d3q1′ d3q2′ q1, σ1; q2, σ2 V q1′, σ1′; q2′, σ2′

c † ( q1, σ1) c † ( q2, σ2 ) c ( q1′, σ1′ ) c ( q2′, σ2′ ) . (4.15)

Here, V is a two-body kernel, which must satisfy certain properties in order that
the Galilean commutation relations (2.21) be preserved. These properties arise
from requiring that H2 commute with P , U (v ), and U(R). Commuting with P means
the two-body kernel is a product of a delta function, δ (q1 + q2 − q1′ − q2′ ), and other
variables. Commuting with Galilei boosts implies that the kernel is only a function of
differences of velocities. And commuting with rotations means that only magnitudes
of velocity differences are allowed. Further, the spins must be coupled together to form
a scalar. Putting these requirements together gives 〈∣V ∣〉 = δ (q1 + q2 − q1′ − q2′ )
f ( ∣ q1 − q2 ∣, ∣ q1′ − q2′ ∣ ) g (σ1, σ2, σ1′, σ2′ ), where f is an arbitrary function of the indicated
variables and g is a scalar in the spin variables. Any kernel satisfying these properties
will preserve the Galilean commutation relations. Such mild restrictions on the form of
the two-body kernel mean that many different types of interactions are Galilean
invariant. This is to be contrasted with the case of Einstein relativity, where it is much
more difficult to introduce interactions in a manner that preserves Einstein relativity.
It is possible to continue adding terms to the Hamiltonian that correspond to
three-body, four-body and higher order terms simply by taking products of creation
operators followed by the same number of annihilation operators. Again, there will
be mild restrictions on the form of the many-body kernels required for Galilean
invariance. But it is not possible to add many-body forces that have products of
creation operators, the number of which differs from the number of annihilation
operators. That is to say, it is impossible to have Galilean invariant interactions that
create or destroy particles. To see this, consider the example of a many-body force
with two creation and three annihilation operators. This can be written as

H2−3 = ∑∫ d3q1 d3q2 d3q1′ d3q2′ d3q3′ 1, 2 V 1′ , 2′ , 3′

c † ( q1, σ1) c † ( q2, σ2 ) c ( q1′, σ1′ ) c ( q2′, σ2′ ) c ( q3′, σ3′ ) (4.16)

and must commute with P and U(v). But, commuting with P means
q1 + q2 = q1′ + q2′ + q3′, while commuting with U(v) means q1 − v + q2 − v = q1′ −
v + q2′ − v + q3′ − v , which is impossible. This generalizes to arbitrary production

4-7
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

processes and thus there can be no Galilean invariant potentials that change particle
number.
The consequence of this result is that the fundamental forces of nature cannot be
understood in Galilean quantum theory as arising from particle exchange. While
there are many interaction kernels that give rise to forces between quantum systems,
these forces cannot be due to particle exchange, which is the modern way for
understanding the nature of the fundamental forces. Einstein relativity is required to
see how the fundamental forces of nature arise from particle exchange. And, as will
be shown, the kernels that generate the interactions in Einstein relativity are severely
restricted in their general form, unlike in Newton relativity.

Bibliography
[1] Warren R E and Klink W H 1970 Model independent analysis of nonrelativistic multiparticle
reactions J. Math. Phys. 11 1155
[2] Weinberg S 1995 The Quantum Theory of Fields vol I (Cambridge: Cambridge University
Press)

4-8
IOP Concise Physics

Relativity, Symmetry and the Structure of Quantum Theory I:


Galilean quantum theory
William H Klink and Sujeev Wickramasekara

Chapter 5
Internal symmetries

The only symmetry discussed thus far has been the spacetime symmetry associated
with the Galilei group and its generalization to accelerating reference frames. But
there are also symmetries used in quantum mechanics that are not associated with
spacetime, the so-called internal symmetries. They appear when particles are seen to
have similar properties with respect to a given interaction. A prime example is
isospin symmetry, in which the proton and neutron are seen to have similar prop-
erties under strong interactions. The fact that such particles can be grouped into
multiplets is an indication that internal symmetries are generated by compact groups
(and their Lie algebras), because only such groups have finite-dimensional unitary
irreducible representations. And of the possible compact groups, the unitary ones
seem to play a particularly important role; it is these groups and their representa-
tions that we discuss in this chapter. Notable examples of internal symmetries
generated by the unitary groups include the previously mentioned isospin symmetry
SU(2), color and flavor symmetries SU(3) and flavor-spin symmetries SU(6) (where
the S in SU(N) stands for the subgroup of unitary matrices with determinant one).
The unitary groups are the groups of unitary matrices in N dimensions,
U (N ) := {u ∈ GL(N , C ), u †u = I }, where GL(N , C ) is the general linear group of
N × N non-singular matrices over the complex numbers and I is the N-dimensional
identity matrix. The Lie algebra of U(N) is the vector space of Hermitian matrices
and has a basis that can be chosen to have the form h m,n(i , j ) = δim δ jn ; here m, n are
the matrix indices and i , j label the Lie algebra basis elements. The commutation
relations of such Lie algebra basis elements are given by

⎡⎣ h(i , j ), h(k , l )⎤⎦ = ∑ h m,p(i, j )h p,n(k, l ) − h m,p(k, l )h p,n(i, j )


m,n
p
= h m,n(i , l )δ jk − h m,n(k , j )δ il . (5.1)

doi:10.1088/978-1-6270-5624-3ch5 5-1 ª Morgan & Claypool Publishers 2015


Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

There are many different realizations of these unitary algebra commutation


relations. Since we want to include internal symmetries in a many-particle quantum
theory, the most natural choice is to make use of the creation and annihilation
operators introduced in the previous chapter. We start here with creation and
annihilation operators indexed by i , j , … running from 1 to N. From the fact that
creation and annihilation operators satisfy [a(i ), a†( j )]± = δij , it follows that
⎡⎣a †(i )a( j ), a †(k )a(l )⎤⎦ = a †(i )a(l )δ − a †(k )a( j )δ il (5.2)
jk

holds for both fermionic and bosonic creation and annihilation operators, as indicated
by the plus-minus sign in the commutator. But these commutation relations are the
same as those given by the unitary algebra commutation relations (5.1), so that
bilinears in creation and annihilation operators provide a representation of the
unitary algebra.
The representation spaces on which these bilinears act is generated through a
vacuum state, defined by a(i )∣ 0〉 = 0 . The vacuum state is essentially the one-
dimensional (trivial) representation of the unitary algebra. All the other symmetric
(bosonic) or antisymmetric (fermionic) irreducible representations are given by
products of creation operators acting on the vacuum: a†(i1 )... a†(i n )∣ 0〉; and a general
state in the n-body internal symmetry space is given by
ϕ = ∑ f ( i1, ⋯, in )a†( i1)... a†( in ) 0 , (5.3)

where f is an arbitrary function of the discrete variables.


As a first example, consider isospin symmetry generated by the group SU(2)
and write the three isospin generators in terms of the U(2) algebra, namely
1
I+ = a†(2)a(1), I − = a†(1)a(2) and I3 = 2 (a†(2)a(2) − a†(1)a(1)), with commutation
relations
⎡⎣I3, I±⎤⎦ = ±I±,
⎡⎣I+, I −⎤⎦ = 2I3, (5.4)

the usual commutation relations for isospin. The two-dimensional representation


space generated by a†(i )∣ 0〉 gives the representation for the proton state (i = 2) and
neutron state (i = 1). The generator I := a†(2)a(2) + a†(1)a(1) is a Casimir operator
(since it commutes with all the generators) and can be used to label the irreducible
representations.
As a second example consider the pion triplet, an isospin 1 representation of
SU(2). Since there are now three pion states, we start with the group U(3), indexed
for convenience in this case by i = 1, 0, −1. Since pions are bosons, we write
the creation and annihilation operators in terms of c†(i ), c(i ) with commutation
relations, and define the isospin generators by

I+ = (
2 c †(1)c(0) + c †(0)c( −1) )
† †
I3 = c (1)c(1) − c ( −1)c( −1), (5.5)

5-2
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

which again satisfy the isospin commutation relations (5.4). The procedures used
here can be generalized to any of the representations of any unitary groups.
Given these results for isospin (or any other internal symmetry) in terms of
creation and annihilation operators, it is possible to adjoin isospin labels to space-
time labels and so combine spacetime symmetries with internal symmetries by
simply including the labels for both symmetry types in the creation and annihilation
operators. Consider the pion isospin triplet as an example, so that the creation
operator for creating a pion with velocity q and isospin i = 1, 0, −1 is c†(q, i ), and a
one-particle state is given by ∣ q, i 〉 = c†(q, i )∣ 0〉. Then the isospin generators can be
written as

I+ = ∫ d3q( c†(q, 1)c(q, 0) + c†(q, 0)c(q, −1))


2

I3 = ∫ d3q( c †(q , 1)c(q , 1) − c †(q , −1)c(q , −1)) (5.6)

and it is easy to check that the isospin commutation relations are again satisfied.
The Hamiltonian (and other generators) of the Galilei group is now written as
mq 2 †
H1 = ∑ ∫ d3q c (q , i )c (q , i ) (5.7)
i
2

and it commutes with the isospin generators: [H1, I+ ] = [H1, I3 ] = 0.


The many-body operators are also given in terms of creation and annihilation
operators indexed by velocity and isospin. For example, a two-body operator such
as the two-body Hamiltonian is written as

H2 = ∑∫ d3q1 d3q2 d3q1′ d3q2′ q1, i1, q2, i 2 V q1′, i1′, q2′, i 2′

c †( q1, i1)c †( q2, i 2 )c( q1′, i1′ )c( q2′, i 2′ ) , (5.8)

where the kernel must again satisfy conditions arising from Galilean invariance.
But now there are also the isospin indices and, depending on how isospin is being
used, there will be constraints on the kernel due to isospin. For example, if V is the
potential for the strong interaction, H2 should commute with the isospin generators,
whereas if V arises from electromagnetic interactions, the kernel need only commute
with I3.
Although internal symmetries play an important role in Newton relativity, as will
be shown in the next volume, they play an even more important role in Einstein
relativity. That is, map groups, defined in chapter 3, can be combined with internal
symmetry groups to generate the fundamental interactions through gauge trans-
formations and gauge invariance.

5-3
IOP Concise Physics

Relativity, Symmetry and the Structure of Quantum Theory I:


Galilean quantum theory
William H Klink and Sujeev Wickramasekara

Chapter 6
Conclusion

Feynman famously said that no one understands quantum mechanics [1]. Quite
likely, his statement is about the new perspective brought forth by quantum theory
that a mind engrained in the classical Newtonian perspective finds very strange.
In other words, Feynmanʼs statement is perhaps more about incompatibilities
between two perspectives than about the comprehensibility of quantum theory itself.
If so, can one understand quantum theory on its own terms?
It seems to us that this question is really about the structure of quantum theory—
in particular, why the theory has the structure that is does. Put another way, what
physical principles, if any, necessitate quantum theory to have the form it does. It is
our claim that this question is at the center of what it means to ‘understand quantum
mechanics’ and, if it is answerable, that we will have a sense of why the quantum
perspective is what it is. Note that it is not our objective to focus on the details of any
particular quantum models, such as the hydrogen atom or the standard model.
Rather, our goal is to attempt to understand the principles that dictate the nature of
the characteristic features of the theory common to all models, such as quantum
states, observables, commutation relations and the time evolution equation. It is in
this sense that we have repeatedly referred to our attempt as ‘grounding quantum
theory’ and it is these universal features that we have in mind when we use the phrase
‘structure of quantum theory’.
Clearly, the very phrase ‘principles that determine the form of quantum theory’
suggests that those principles that we seek must be antecedent to quantum theory.
As lofty a goal as the search for such principles appears, the history of physics seems
to tell us that physical theories are almost never derived from meta-principles as
such. Rather, it is the form of successful theories themselves that hints at its
grounding principles. In this book, we have closely followed this historical exemplar
and attempted to extract its grounding principles by examining the form of quantum
theory.

doi:10.1088/978-1-6270-5624-3ch6 6-1 ª Morgan & Claypool Publishers 2015


Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

The answer laid out in the preceding five chapters is that symmetry determines the
structure of quantum theory. Depending on what we take to be the starting point,
symmetry determines the structure of quantum theory either entirely or to a large
extent. To clarify this statement, recall that, as explained in the introduction, by
symmetry we mean a set of transformations on a suitable set, for example a spacetime
manifold, such that they form a group. Hence, we speak of various symmetry groups.
Chief among these are the relativity groups, which include the Galilei, Poincaré and de
Sitter groups, but other groups also exist, which involve internal symmetries, such as
isospin and quark flavors. The theme that synthesizes the preceding five chapters of
this book is that unitary representations of these symmetry groups entirely dictate the
form of quantum theory. Note the critical role of unitary representations. In other
words, even after we identify the symmetry groups by whatever means, such as by the
transformation properties of classical trajectories of physical objects, we still need to
consider their unitary representations in order to extract a quantum theory from that
symmetry. If the unitary representations of symmetry groups form the starting point,
the fundamental principle, then the structure of quantum theory entirely follows from
it. However, it is possible to consider groups as primary and their representations as a
secondary structure that follows from or is imposed on symmetry. That consideration
leads to the question of why, in the first place, must we consider representations of
symmetry groups, rather than the groups themselves, to get to quantum mechanics.
We addressed this in the introduction: representations are necessary because the
superposition principle requires that quantum states inhabit a vector space. Hence,
we must consider not simply the symmetry group itself, but its representations in
the vector space of quantum states. In this light, quantum theory is the synthesis of
two grand principles, symmetry and superposition, and we might say that symmetry
determines the structure of quantum theory ‘to a large extent’. Still, it is important to
emphasize, as we have in the introduction and appendix B, that in almost all cases a
representation of a group and the vector space on which it is defined are constructed
together, and we can consider the representation vector space as being generated by the
representation. In fact, all of the representations of groups that we have discussed
(Galilei group, Galilean line group and unitary groups) are obtained as induced
representations which, as shown in appendix B for the Galilei group, generate the
vector spaces on which they act. Viewed this way, symmetry is the primary principle
and the superposition principle is a secondary principle that follows from the linear
structure of the vector spaces that carry the representations of symmetry groups.
Let us return to the fact that there are two main types of symmetry: the symmetry
of spacetime transformations that define a relativity principle and the symmetry of
internal transformations that are related to various interactions, such as isospin.
While symmetry determines the structure of quantum theory, what particular types
of symmetry are relevant for a given theory are to be determined by examining
the fundamental equations of that theory, or empirical data. Now, since spacetime
symmetries are common to both classical and quantum theories, the group structure
of each relativity can be deduced from the corresponding classical theory.
For example, the transformation properties of Newtonʼs laws and the trajectories
of particles they predict can be used to deduce Newton relativity. But this does

6-2
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

not mean that we are starting with a classical dynamical theory and quantizing it,
as is traditionally done. In fact, as was pointed out in chapter 3, the traditional
method of quantizing a classical theory makes predictions for accelerating systems
that differ when compared with our procedure that uses the unitary representations
of the Galilean line group. As such, these differing predictions can only be settled
by experiments.
The procedure for getting spacetime symmetries from classical considerations
does not work for internal symmetries because a particular quantum theory dis-
playing an internal symmetry may not have a classical counterpart. An example is
the flavor symmetry of the eightfold way, which was deduced from the elementary
particle physics data. Our approach has been to use whatever information is
available to infer a particular symmetry group and then consider its representations
as the grounding principle for the quantum theory. In particular, we took coordinate
transformations between different classical observers as defining Newton relativity
and its symmetry group, the Galilei group. As may be evident, this is more a matter
of convenience than principle. One can envision extracting the group structure of the
Galilei or Poincaré groups from a sufficiently large set of quantum mechanical data,
in much the same way that internal symmetry groups are deduced from such data.
For instance, from the measurable uncertainty relations it should be possible to
deduce the commutation relations between position and momentum operators, a
part of the Galilean Lie algebra.
The role of symmetry in quantum theory articulated here can be supported by a
number of results developed in this book, which we summarize in the rest of this
chapter. In chapter 2, we showed that unitary, irreducible, projective representations
of the Galiei group give rise to the notion of a quantum particle. In particular, the
Hilbert space on which one such representation is defined can be identified with the
state space of an elementary quantum system, which we call a particle. The generators
of the representation produce all of the kinematic observables: momentum, energy,
angular momentum and position. Further, we saw that the Hamiltonian arises as the
2
generator of time translations. Its form for a free particle, H = 2Pm , also follows from
the Galilean algebra as a derived mathematical result, without having to appeal to the
classical expression first and then promote it to a quantum operator-valued quantity.
Likewise, we obtain the Heisenberg commutation relations and the Schrödinger time
evolution equation as consequences of the Galilei group representations.
The irreducible representation spaces are characterized by three numbers, which
are the eigenvalues of the three Casimir operators of the Galilean algebra. There-
fore, they are invariant, i.e., all observers find the same value when these observables
are measured. As such, these are precisely what we define as the intrinsic properties
of a particle and call mass, spin and internal energy. As a consequence, we con-
cluded that relativity determines what forms of matter are allowed in the resulting
quantum theory: in Newton relativity, only a form of matter characterized by a
positive mass and an integer or half-odd-integer spin is possible, perhaps the most
striking result that demonstrates the reach of symmetry as a grounding principle of
quantum theory.

6-3
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

Our undertaking in chapter 3 was to show that the notion of a quantum particle
could be extended to make it meaningful for observers who are accelerating. The
one-particle quantum theory based on the representations of the Galilei group,
like the group itself, is meaningful only for inertial observers. Therefore, the results
of chapter 3 extend those of chapter 2, but the main idea is the still the same: an
elementary quantum system is the physical realization of a unitary, irreducible
(cocycle) representation of the symmetry group that ties together all accelerating
(as well as inertial) reference frames. Owing to the fact that this group is infinite
dimensional and has a very rich cohomology, the mathematical construction of
the quantum theory in non-inertial reference frames is considerably more intricate
than for the Galilean case. Still, we saw in chapter 3 that there is a well-defined
one-particle quantum theory that holds in accelerating reference frames. This
construction also allows us to understand the role of what are classically known as
fictitious forces and the equivalence principle in quantum theory.
Given that the unitary, irreducible, projective representations of the Galilei
group generate the one-particle Hilbert spaces, from a symmetry point of view,
multiparticle systems are described by tensor products of one-particle representations.
In chapter 4, we considered such multiparticle tensor product states. The tensor
product of irreducible representations is reducible and resolving this reducible
representation into a direct sum (in fact, integral) of irreducible representations is an
important first step in the analysis of multiparticle systems, particularly for the
introduction of interactions into such a system. Interactions, which are introduced as
modifications to the multiparticle Hamiltonian, are constrained by the commutation
relations of the Galilean Lie algebra. How this is done under the constraint of
Galilean commutation relations is more clearly seen by using Clebsch–Gordan
coefficients to find quantities that are invariant under Galilei transformations.
Further, the discussion of chapter 4 illustrates an interesting point: while Galilean
symmetry rules out some conceivable interactions, it does not dictate what inter-
actions must occur in nature. More generally, relativity symmetry groups determine
the basic structure of quantum theory, but they do not tell us what particular
interactions, such as the harmonic oscillator or Coulomb potential, are realized
in nature. Interactions are determined by examining the empirical data within
the framework imposed by the general theory, which can then lead to further
constraints, such as those occuring in Einstein relativity with gauge theories.
For identical particles we have also shown how interactions can be introduced
in a Galilean invariant way with the help of creation and annihilation operators,
which automatically preserve the identical particle structure. When the energy and
momentum operators are given in terms of creation and annihilation operators, the
kernels that generate the two (or more) body interactions have only mild restrictions
imposed on them from Galilean invariance. Therefore, there is no hint as to what
might constitute fundamental interactions in Galilean quantum theory. As will be
seen in the sequel to this book, the quantum theory with Einstein relativity imposes
much more severe constraints on the form of interaction kernels. In fact, from the
theorem that particle production is not allowed in Newton relativity, we see that it is
impossible to understand the nature of the fundamental forces as being due to particle

6-4
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

exchange, which is the modern understanding of the nature of the fundamental forces.
Only in a quantum theory with Einstein relativity is it possible to make particle
exchange the mechanism that generates the fundamental forces.
A very interesting inference that can be drawn from the above analysis is that a
quantum theory that follows from relativity and symmetry, by way of the unitary
representations of the symmetry groups, is fundamentally a theory of particles,
not fields. It is perhaps worthwhile emphasizing again that unitary irreducible
representations of symmetry groups are the key: once the one-particle states are
obtained from these representations, as shown in chapter 4, there is a well-defined
procedure for generating the multiparticle states, as well as creation and annihilation
operators that map between these states, including the vacuum. Free quantum fields
are simply certain linear combinations of the creation and annihilation operators.
In chapter 5, we discussed internal symmetries, including how they lead to
additional degrees of freedom in the state vectors, as well as the creation and
annihilation operators. Included in the discussion were several examples involving
isospin symmetry.
Although we have treated relativity and internal symmetries as two independent
and distinct notions, we note as a final remark that there is at present great interest in
seeing whether certain internal symmetries can themselves be understood as space-
time symmetries, by enlarging the four-dimensional spacetime manifold to include
extra compactified dimensions [2].

Bibliography
[1] Feynman R P 1965 The Character of Physical Law (New York: Penguin)
[2] Douglas M R and Kachru S 2007 Flux compactification Rev. Mod. Phys. 79 733

6-5
IOP Concise Physics

Relativity, Symmetry and the Structure of Quantum Theory I:


Galilean quantum theory
William H Klink and Sujeev Wickramasekara

Appendix A
Transitive manifolds

Quite often, symmetry groups appear in physics as groups of transformations on a


manifold. This is certainly the case for the relativity group considered in this book.
Thus, in general, consider an n-dimensional differentiable manifold M and let
x i , i = 1, 2, 3, …, n be a set of local coordinates for it. When G is a Lie group, as is
the case for the Galilei group, the elements g of G are described by a set of real
parameters g i , i = 1, 2, 3, …, r , where r is the dimension of the group. Then, the
action of G on M can be described as

(
x′i = f i g1, … , g r ; x1, … , x n ) (A.1)

We often also write (A.1) more succinctly as


x′ = f (g , x ), g ∈ G . (A.2)

We take the functions f i to be smooth, often analytic, functions of group parameters


g i and coordinates x j . The set of functions f i is called essential when they are
necessary and sufficient to determine the transformation x ↦ x′. We assume here
that this is the case.
That G is a group of transformations means that the transformations (A.2) fulfill
the following properties:

(
x = f g −1, x′ )
x = f (e , x )
x″ = f ( g2, x′)

(
= f g2, f ( g1, x ) )
= f ( g2 g1, x ) . (A.3)

doi:10.1088/978-1-6270-5624-3ch7 A-1 ª Morgan & Claypool Publishers 2015


Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

The last equality of (A.3) can be considered a definition of the composition rule
under which G is a group.
The above considerations can be made more formal and precise as follows.

Definition. A smooth mapping f : G × M → M is called a left action of a Lie


group G on a manifold M if f : (g, x ) ↦ f (g, x ) ∈ M fulfills the conditions
f (e, x ) = x and f (g2, f (g1, x )) = f (g2 g1, x ). In this case, we often write
f (g, x ) as gx.
A smooth mapping f : M × G → M is called a right action of a Lie group
G on a manifold M if f : (x , g ) ↦ f (x , g ) ∈ M fulfills the conditions
f (e, x ) = x and f (f (x , g1), g2 ) = f (x , g1 g2 ). In this case, we often write
f (x , g ) as xg.

The mappings f : G × M → M and f : M × G → M defining left and right


actions can also be thought of as a homomorphism and an anti-homomorphism,
respectively, from G into the set of smooth transformations on M. For left
actions, consider f as a mapping f : g → fg , where fg is a smooth mapping on M
defined by fg : x → fg (x ) ≡ f (g, x ). From the above definition of left action, it then
follows
fg ◦ fg = fg g (A.4)
2 1 2 1

Similarly, the right action defines an anti-homomorphism:


fg ◦ fg = fg g . (A.5)
2 1 1 2

In either case, ( fg )−1 = fg −1. Further, if fg is a left action, then fg −1 is a right action and
vice versa.
A key notion that appears in the analysis of a group action on a manifold is that
of an orbit. Suppose x ∈ M . Then, the f-orbit of x is the set { fg (x ) : g ∈ G} of all
points of M reached from x by, say, the left action of G. Orbits clearly introduce an
equivalence relation 9 into the manifold: x′ 9x if x′ = gx . The set of equivalence
classes M /9 is the set of f-orbits. The subgroup Sx := {g ∈ G : fg (x ) = x} of G that
leaves a given point x invariant is called the stability group of x.
Based on this notion of f-orbit, we identify several specific actions that have
proved to be useful.
• An action is called transitive if the orbit of any x ∈ M is the entire M.
• An action is called free if the stability group Sx of every point contains only
the identity element.
• An action is called effective if there exists no non-trivial subgroup of G that
leaves the entire manifold invariant. That is to say that the intersection of
stability groups for all the points of M, S := {g ∈ G : fg (x ) = x , ∀ x ∈ M }
contains only the identity.
• An action is called trivial if fg (x ) = x for all g ∈ G and for all x ∈ M , i.e, if
S = G.

A-2
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

It is straightforward to show that the subgroup S related to the definition of an


effective action is an invariant subgroup of G. From this property, it is easy to
conclude that if a given group action is not effective, the action restricted to G /S is
effective. Recall that G /S is the set of cosets with respect to the subgroup S. Since S
is an invariant (also called normal) subgroup, the set of cosets G /S is a group, called
the factor group.
If a given action is free, then every point of x ∈ M is transformed by every g ∈ G
that is not equal to the group identity. Hence, two points x′ and x are either
unrelated or related by a unique element g ∈ G, x′ = gx . In other words, there is a
one-to-one correspondence between points on each orbit and elements of the group,
and the mapping g → fg (x ) is one-to-one for each x ∈ M . On the other hand, if the
action is not free, then there exists a non-trivial stability group Sx for at least one
point x ∈ M . In that case, let h ∈ Sx and suppose y = g0 x for some g0 ∈ G . Then
g0 hg0−1 y = g0 hg0−1 g0 x = y . Therefore, Sg0 x = g0 Sx g0−1 and we see that stability
groups of points on the same orbit are isomorphic. In this case, there is a one-to-one
mapping between the cosets G /Sx and the points on the f-orbit of x.
If the action is transitive, then there is a one-to-one mapping between G /Sx for
any x ∈ M and the manifold M. In this case, M is called a homogeneous space of G
and it is possible to think of the manifold M as being generated by the group action.
In particular, given x0 ∈ M , every element g ∈ G has a coset decomposition
g = h (g ) k (g ) for h (g ) ∈ Sx0 and k (g ) identifiable with a unique point x (g ) = k (g )
of M. Therefore, the group action that transforms a point x ∈ M can also be
thought of as generating a transformation of the corresponding coset representative
to another: k → kg0 = h (g0 k ) k′ (g0 k ).
We now apply the above general considerations to the Galilei group and the
Newton/Galilei spacetime manifold, which is the Cartesian product of the real three-
dimensional space with one-dimensional time, 3 × . From the transformation
formula (2.1), we see that the Galilei group acts transitively on the spacetime
manifold. Furthermore, this action is not free, as can be seen by setting x = 0 and
t = 0 in (2.1) and noting that the stability group of the origin (0, 0) is the subgroup
S(0,0) := {(R, v )} of homogeneous Galilei transformations. This group is isomorphic
to E(3), the Euclidean group in three dimensions consisting of rotations and
translations. From the above discussion, we then see that the stability group of
any other point of the spacetime manifold is isomorphic to S(0,0) , therefore to E(3),
and that there is a one-to-one correspondence between the set of cosets ./E (3)
and the manifold: M = ./E (3). The coset decomposition g = h (g ) k (g ) of
the Galilei group with respect to the subgroup E(3) is given by the
correspondence g = (R, v, a, b ), h (g )(R, v ) and k (g ) = (x , t )(g ) = (a, b ), so that
g = h (g ) k (g ) ↦ (R, v, a, b ) = (R, v )(a, b ). Thus, as stated in the introduction, we
can either think of the Galilei group as generating the Newtonian spacetime
manifold, or start with the Newtonian spacetime manifold and show that the action
of the Galilei group leaves the manifold invariant.

A-3
IOP Concise Physics

Relativity, Symmetry and the Structure of Quantum Theory I:


Galilean quantum theory
William H Klink and Sujeev Wickramasekara

Appendix B
Irreducible representations of the Galilei
group and the origin of mass and spin

Here, we present a discussion of the construction of unitary, irreducible, projective


representations of the Galilei group. As we have argued throughout this book, it is
these representations that form the foundation of Galilean quantum mechanics.
In a monumental paper published in 1939 [1], Eugene Wigner constructed all
unitary irreducible representations of the Poincaré group. The method he intro-
duced, the method of induced representations, was further developed by George
Mackey [2] and can be also used to construct the unitary irreducible representations
of the Galilei group. Before we discuss this construction, let us briefly review some
fundamental group theoretical results.

Preliminaries. Projective representations and central extensions. By a unitary


representation of a group G on a Hilbert space / , we mean a mapping
U : G × / → / , U : (g, ϕ ) ↦ U (g ) ϕ ∈ / , such that
• for each g ∈ G , U (g ) is a unitary, linear operator in / ;
• for e ∈ G , the identity element of G, U (e ) = I ;
• for g2, g1 ∈ G , the mapping U : g → U (g ) is a homomorphism: U (g2 g1 ) =
U (g2 ) U (g1 ) .
A unitary projective representation is one for which the homomorphism property
holds only up to a phase factor:
U ( g2 ) U ( g1) = eiω( g2, g1) U ( g2 g1), (B.1)

where ω (g2, g1 ) is a real-valued function on G × G . In quantum mechanics, we must


allow for projective representations because, as mentioned in chapter 2 in

doi:10.1088/978-1-6270-5624-3ch8 B-1 ª Morgan & Claypool Publishers 2015


Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

connection with the discussion on Wignerʼs symmetry representation theorem,


quantum states are represented by rays on a vector space, rather than vectors. The
collection of rays on a vector space is called the projective space and a representation
on this projective space is then called a projective representation [3].
While the homomorphism property holds only up to a phase factor, the asso-
ciativity property U (g3 )(U (g2 ) U (g1 )) = (U (g3 ) U (g2 )) U (g1 ) must exactly hold
because the composition of linear operators is necessarily associative. This fact, in
turn, imposes the following restriction on the phase factor of (B.1):
ω (g3, g2 g1) + ω ( g2, g1) = ω ( g3 g2, g1) + ω ( g3, g2 ). (B.2)

The functions
ω:G×G→ (B.3)
that satisfy the constraint (B.2) are called two-cocycles. They give rise to projective
representations of the group. However, note that any function of the form
ω ( g2, g1) = ϕ ( g2 ) + ϕ ( g1) − ϕ ( g2 g1), (B.4)

where ϕ is any real valued function on G, necessarily fulfills (B.2). Such functions,
called two-coboundaries or trivial two-cocycles, do not lead to projective
representations because in this case the phase factor of (B.1) can be removed
by redefining the group operators as V (g ) := e iϕ(g ) U (g ), so that V (g2 ) V (g1 ) =
e iϕ(g2 ) + iϕ(g1) U (g2 ) U (g1 ) = e iϕ(g2 ) + iϕ(g1) e iω( g2, g1)U (g2 g1 ) = e iϕ(g2 g1) U (g2 g1 ) = V (g2 g1 ).
Therefore, (B.4) introduces an equivalence relation on the set of two-cocycles. Mem-
bers of each equivalence class, consisting of two-cocycles that differ from one another
by a two-coboundary (B.4), give rise to the same projective representation of G. The
number of such equivalence classes of two-cocycles depends on the structure of the
group G.
While the phase factor of (B.2) cannot be removed by a redefinition of the group
operators when the two-cocycle is non-trivial, it is possible to extend the group G by
the Abelian group  so that the projective representations of G are equivalent to
vector representations of the centrally extended group G̃ , a theorem first proved by
Bargmann [3]. By the extension of a group G by another group A, we mean the
existence of a group G̃ such that A is an invariant subgroup of G̃ and G is isomorphic
to the factor group G̃ /A. The extensions can be summarized in terms an exact
sequence of homomorphisms:

e → A → G˜ → G → e, (B.5)

where e denotes the trivial group consisting of just the identity. The sequence is
called exact because the kernel of each homomorphism is the image of the one
before. Our focus in this book is limited to the case where A is an Abelian group.
We call an extension central if the group A that extends G is a central subgroup
of G̃ . Recall that a subgroup of a group is central if all of the elements of the
subgroup commute with all elements of the group.

B-2
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

We can denote the elements of G̃ by g˜ = (φ , g ) with φ ∈ A and g ∈ G . If we


denote the composition rule for the Abelian group A additively, the composition rule
for G̃ is
g˜2 g˜1 = ( φ2 , g2 )( φ1, g1) = ( φ2 + φ1 + ξ ( g2, g1), g2 g1) (B.6)

for some function ξ: G × G → A. The constraint on this function that ensures the
associativity of the composition rule (B.6) is precisely (B.2). It is this fact that allows us
to construct a vector representation of G̃ that is equivalent to the projective repre-
sentation (B.1) of G. However, since ω ( g2, g1 ) of (B.1) must take values in  (so that
U (g ) would be unitary), we let A of (B.5) be  or isomorphic to a subgroup thereof,
such as Z2. Now, simply define operators U ( g˜ ) = U ((φ , g )) := e icφU (g ). Then,
1
U ( g˜2 ) U ( g˜1) = e ic( φ1 + φ 2 ) U ( g2 ) U ( g1) = e ic(φ1 + φ 2 + c ω( g2, g1) ) U ( g2 g1) (B.7)

and we see that U ( g˜ ) is a vector representation of G̃ , i.e.,


U ( g˜2 ) U ( g˜1) = U ( g˜2 g˜1) (B.8)

if we set ξ ( g2, g1 ) = 1c ω ( g2, g1 ).


In passing, we mention that the composition rule (B.6) is a special case of the
more general rule
( φ2 , g2 )( φ1, g1) = ( φ2 + σ ( g2 ) φ1 + ξ ( g2, g1), g2 g1) , (B.9)

where σ (g ) is an automorphism on A and σ : g → Aut (A) is a homomorphism. This


composition rule defines an extension of G by A, but now A is not a central subgroup
of the extended group G̃ . The central extension (B.6) corresponds to the trivial
automorphism σ (g ) = I . When the two-cocycle is trivial but the automorphism σ (g )
is not, we get the semidirect product extension G˜ = A ⋊ G . When both are trivial,
G̃ is simply the direct product of A and G.
The phase factor (B.1) leading to a projective representation can come from two
distinct properties of the group, one topological and the other algebraic. The two
cases are not mutually exclusive. In fact, both types of projective representations are
possible and physically meaningful for the Galilei group.

Topology and covering groups. In the topological case, the phase factor provides a
representation of the first (fundamental) homotopy group π (G ), which describes the
non-simply connectedness of the group manifold G. Recall that a manifold is called
simply connected if a curve joining any two points can be smoothly deformed into
any other curve joining the two points. Thus, 2 is simply connected, but an annulus
is not. Now, if we extend the group G by π (G ), with π (G ) serving as A in the exact
sequence (B.5), the resulting group G̃ is simply connected and called the universal
covering group of G. If π (G ) is discrete, as is the case for most if not all examples
encountered in physics, then the extension is central because a discrete invariant
subgroup of a connected group is central. The function ξ (g2, g1 ) of (B.6) takes

B-3
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

values in π (G ) in this case and the phase factor of (B.1) provides an irreducible
representation of π (G ). Thus, if the group G is not simply connected, there exist
projective representations of G that are equivalent to vector representations of the
central extensions of G by its first homotopy group π (G ). Perhaps the best known
example is SO (3), the group of rotation matrices in real three dimensional space.
It is not a simply connected manifold, with a the first homotopy group isomorphic to
Z2. The central extension of SO (3) by Z2, with ξ: SO (3) × SO (3) → Z2 , is SU (2),
the group of 2 × 2 unitary matrices with unit determinant. Thus, there exist pro-
jective representations of SO (3),
U ( g2 ) U ( g1) = e iω( g2, g1) U ( g2 g1) = ±U ( g2, g1). (B.10)

The two values ± clearly correspond to ξ ( g2, g1 ) taking values in Z2. This projective
representation is equivalent to a vector representation of SU (2).

Algebra and central generators. In the algebraic case, the phase factor of (B.1)
corresponds to a central extension by , a one-dimensional Lie group. Hence, the
parameter φ of (B.4) varies over  and the function ξ (g2, g1 ) takes values in . Unlike
the topological case, this extension expands the Lie algebra by a basis element. Again,
the projective representation (B.1) is equivalent to a vector representation of this
central extension, as in (B.8).
From the definition of group operators, we see that this one parameter central
subgroup (φ , e ) is trivially represented: U (φ , e ) = e icφU (e ) = e icφI . Its generator,
∂U (φ , e )
C := −i = cI , (B.11)
∂φ φ=0

which increases the dimension of the Lie algebra of G by one, clearly commutes with
the generators associated with all other one parameter subgroups of G. Therefore,
we often include cI among the Casimir operators of the enveloping algebra of the
representation, as discussed in chapter 2.
An important feature of the central extension G̃ is that, while the set of elements
{(φ , e )} which furnishes the extension is a subgroup of G̃ , the group G that is
extended is not a subgroup of the extension G̃ . It is easy to verify this statement:
(0, g2 )(0, g1 ) = (ξ (g2, g1 ), g2 g1 ) ≠ (0, g2 g1 ). At the level of the generators, this means
that the generators of one parameter subgroups of G̃ are in general different from
those of the corresponding one parameter subgroups of G. To illustrate this point,
suppose g i , i = 1, 2, 3, …, r , is a coordinate basis for G and let g˜ (0, …, g i , …, 0)
and g (0, …, g i , …, 0) be the one-dimensional subgroups of G̃ and of G, respectively,
parametrized by g i . If we denote the Lie algebra elements (say, the left invariant
vector fields) of these one parameter subgroups by X̃i and Xi, respectively, then
∂ξ (g′ , g )
X˜i = Xi + Xφ. (B.12)
∂g i g′=g =e

B-4
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

Further, if the Lie algebra of G is defined by the commutation relations


⎡⎣ Xi , X j ⎤⎦ = c k Xk , i , j , k = 1, 2, 3, … , r , (B.13)
ij

where ci kj are the structure constants in the basis defined by the choice of g i , the
commutation relations for the Lie algebra of G̃ are given by
⎡ X˜i , X˜ j ⎤ = c k X˜k + c φ Xφ
⎣ ⎦ ij ij

⎡⎣ X˜i , Xφ ⎤⎦ = 0, i = 1, 2, 3, … , r , (B.14)

where the additional structure constants ciφj are defined by


⎛ ∂ 2ξ ( g , g ) ∂ 2ξ ( g2, g1) ⎞
ciφj := ⎜⎜ ⎟
2 1
− . (B.15)
i
⎝ ∂g2 ∂g1
j
∂g2j ∂g1i ⎟⎠
g2,g1=e

The commutativity of Xφ with all the other generators or, equivalently, the vanishing
of the structure constants cφi j and cφφi , readily follows from the centrality of the
subgroup {(φ , e )}. These considerations also hold for the unitary representation
(B.8) of G̃ . For the operator Lie algebra, for instance, we have commutation
relations
⎡⎣ K˜ i , K˜ j ⎤⎦ = ic k K˜ k + ic φ C
ij ij

⎡⎣ K˜ i , C ⎤⎦ = 0, i = 1, 2, 3, … , r , (B.16)

where C , K i and K̃ i are, respectively, the self-adjoint operators that represent the Lie
algebra basis elements Xφ , Xi and X̃i .
As we showed in chapter 2, many of the physical observables of quantum
mechanics arise as the generators of unitary representations of the relevant space-
time symmetry group. What the above discussion highlights is that the meaning and
commutation relations of these observables depend on whether they follow from a
projective representation or a vector representation. As we see below, the mass and
position observables, as well as the Heisenberg commutation relations, are possible
in non-relativsitic quantum mechanics only if we consider unitary projective
representations of the Galilei group. Likewise, particles of half-odd-integer spin
are possible in both Poincaré and Galilean quantum mechanics only if we consider,
respectively, the projective representations of the Poincaré and Galilei groups that
correspond to their first homotopy group Z2.

Preliminaries. Induced representations. As mentioned above, unitary, irreducible,


projective representations of the Galilei group can be constructed by the method of
induced representations. In fact, it is known that all such representations can be
obtained by this method. Any other (i.e., reducible) unitary projective representation
for the Galilei group can be obtained as a direct integral of irreducible representations.

B-5
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

Moreover, all the representations discussed in this book are induced representations,
including the representations of the Galilei line group discussed in chapter 3 and the
representations of the unitary groups, discussed in appendix C.
Before describing this construction, we briefly review the method. In its full
generality, it is quite simple: it is always possible to induce a representation of a
group from a given representation of a subgroup of the group [4]. However, these
representations are in general not irreducible. Therefore, here we want to focus on
how the method can be used to obtain unitary irreducible representations. The key
idea in that case is to induce the representation of a group from an Abelian,
invariant subgroup.
To that end, suppose a Lie group G has an Abelian invariant subgroup A.
Suppose the factor group G /A is isomorphic to a group K so we may label elements
of G by g = (a, k ):
G = {(a , k ) : a ∈ A , k ∈ K }. (B.17)
It is known that unitary irreducible representations of an Abelian group are all one-
dimensional and are furnished by the elements of its Pontrjagin dual Ǎ [4]. Recall
that this dual group consists of homomorphisms of A into . Hence, if ϕ ∈ Ǎ, then
ϕ : a ∈ A → ϕ (a ) ∈  and a unitary representation of A can be obtained by the
mapping
U : a → U (a ) = eiϕ(a ). (B.18)
Since A is an invariant subgroup of G, the elements of G act on A by conjugation:
g : a → gag −1. Since A is Abelian, only the elements of the factor group K = G /A
have a non-trivial action on A under conjugation. Hence, conjugation establishes a
natural homomorphism from K to the automorphism group of A, k → [k ] ∈ Aut (A),
where

[k ] : a → [k ] a := kak −1. (B.19)


By duality, there exists a homomorphism from K to the automorphism group of the
dual group Ǎ:

( ⎡⎣kˇ ⎤⎦ϕ) (a) := ϕ ( ⎡⎣ k ⎤⎦ a),


−1
for all a ∈ A and ϕ ∈ Aˇ . (B.20)

For notational simplicity, let us also denote the automorphisms on Ǎ defined by


(B.20) by [k ]. From the context, it should be clear if a mapping denoted by [k ] acts on
A or its dual Ǎ.
Next, let us choose an element ϕ0 in Ǎ and consider the Hilbert space / of L2-
functions defined on 6ϕ = {ϕ : ϕ = [kˇ ]ϕ , k ∈ K }, the K-orbit of ϕ . The inner
0 0 0
product of / is defined in the usual manner,

( f1 , f2 ) = ∑∫6ϕ dμ (ϕ) f1* (ϕ , i ) f2 (ϕ , i ), (B.21)


i 0

B-6
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

where μ is a K-invariant measure on 6ϕ0 . The meaning of the multiplicity index i


will be clarified below. It is also possible to introduce generalized basis vectors
(distributions) ∣ ϕ, i 〉, defined as continuous antilinear functionals on sufficiently
well-behaved functions f ∈ / ,
ϕ , i : f → ϕ , i f := f (ϕ , i ) ∈  , (B.22)
such that

f= ∑∫6ϕ dμ ( ϕ ) ϕ , i ϕ , i f . (B.23)
i 0

From (B.18), we define a unitary representation of the Abelian subgroup A on


/ by
(U (a ) f )(ϕ , i ) = e iϕ(a ) f (ϕ , i ) (B.24)
or, equivalently, by the transformations of the generalized basis vectors ϕ, i ,
−1)
U ×(a ) ϕ , i = e−iϕ(a ϕ, i . (B.25)
The two representations U and U × are linked by the duality relation

( )
f U × g − 1 ϕ , i = U (g ) f ϕ , i . (B.26)

It remains to determine the action of the factor group K on / . To this end, note
that for any ϕ ∈ 6ϕ0 , there exists an element k ϕ ∈ K such that
k ϕ : ϕ0 → ⎡⎣ k ϕ ⎤⎦ ϕ0 = ϕ . (B.27)

The k ϕ ∈ K is clearly not defined uniquely by (B.27) because for any element γ of
Γϕ0 = {γ : γ ∈ K ; [γ ] ϕ0 = ϕ0 }, the stability group of ϕ0 , we have
⎡⎣ k ϕ γ ⎤⎦ ϕ = [k ]ϕ [γ ] ϕ = ϕ . (B.28)
0 0

This equation shows that for a given element ϕ of 6ϕ0 , the stability group Γϕ0
introduces, say, a right, coset decomposition of K. That is to say, for each auto-
morphism [k ϕ ] there corresponds an equivalent class of elements k ϕ Γϕ0 of K. Sup-
pose that from each class we choose a specific element k˜ ϕ . Then, from (B.27) and
(B.28), it follows
[k ] ϕ = [k ] ⎡⎣ k˜ ϕ ⎤⎦ ϕ0
⎡ −1 ⎤
= ⎡⎣ k˜[k ] ϕ ⎤⎦ ⎣ k˜[k ] ϕ kk˜ ϕ ⎦ ϕ0 . (B.29)
−1
We see that the automorphism [k˜[k ] ϕ kk˜ ϕ ] leaves ϕ0 invariant. Therefore, there must
exist an element γ (k , ϕ ) ∈ Γϕ0 , such that
⎡ −1 ⎤
[γ (k , ϕ)] = ⎣ k˜[k ] ϕ kk˜ ϕ ⎦ (B.30)

B-7
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

and

[k ] ϕ = ⎡⎣ k˜[k ] ϕ ⎤⎦ [γ (k , ϕ)] ϕ0 . (B.31)

Next, for each representative element k˜ ϕ of cosets of (B.27), let us define an


operator U (k˜ ϕ ) by demanding

( )
U × k˜ ϕ ϕ0 , i := ϕ , i . (B.32)

This equation is really a definition of how the multiplicity index i is related to various
points on the orbit 6ϕ .
Using the definition (B.32) and the identity (B.31), we can now define the operator
×
U (k ) for arbitrary k ∈ H :

U ×(k ) ϕ , i = U ×(k ) U × k˜ ϕ ϕ0 , i ( )
( )
= U × k˜[k ] ϕ U ×(γ (k , ϕ)) ϕ0 , i . (B.33)

Since [γ (k , ϕ )] ϕ0 = ϕ0 , we note that U ×(γ (k , ϕ )) ϕ0 , i can at most be a linear


combination of the vectors ϕ0 , i . Thus, let

U ×(γ (k , ϕ)) ϕ0 , i = ∑Δ ji (γ (k, ϕ)) ϕ0 , j , (B.34)


j

where Δij (γ (k , ϕ )) are an array of complex numbers that may depend on k, ϕ0 and ϕ.
Substituting this expression in (B.33) and making use of the definition (B.32) again,
we obtain
U ×(k ) ϕ , i = ∑Δij (γ (k, ϕ)) [k ] ϕ , j . (B.35)
j

This transformation formula defines a representation of K if and only if the


operators Δ furnish a representation of the stability group Γϕ0 . To see this, let us
apply (B.35) twice,

U ×( k 2 ) U ×( k1) ϕ , i = U ×( k 2 ) ∑Δ ji (γ (k , ϕ)) ⎡⎣ k1⎤⎦ ϕ , j


j

= ∑Δ ji γ ( k1, ϕ) Δlj γ k 2, ⎡⎣ k1⎤⎦ ϕ


( ) (( ) ) ⎡⎣ k2 ⎤⎦ ⎡⎣ k1⎤⎦ ϕ, l
jl

l
{ (
= ∑ Δ γ k 2, ⎡⎣ k1⎤⎦ ϕ Δ γ ( k1, ϕ)
( )) ( ) }li ⎡⎣ k 2 k1⎤⎦ ϕ , l (B.36)

and compare the result with the action of U ( k 2 k1 ):

U ×( k 2 k1) = ∑Δli ( γ ( k2 k1, ϕ) ) ⎡⎣ k2 k1⎤⎦ ϕ, l . (B.37)


l

B-8
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

Now, from (B.30), we obtain

γ k 2, ⎡⎣ k1⎤⎦ ϕ γ ( k1, ϕ) = γ ( k 2 k1, ϕ) .


( ) (B.38)

Therefore, the equality of (B.36) and (B.37) holds if and only if

Δ γ k 2, ⎡⎣ k1⎤⎦ ϕ Δ γ ( k1, ϕ) = Δ γ k 2, ⎡⎣ k1⎤⎦ ϕ γ ( k1, ϕ) ,


(( )) ( ) (( ) ) (B.39)

i.e., if and only if the operators Δ (γ (k , ϕ )) furnish a representation of the stability


group Γϕ0. In other words, the problem of finding representations of the factor group
K of the group G of (B.17) reduces to that of finding representations of the stability
group Γϕ0 , often an easier task. Furthermore, if these stability group representations
are unitary, so are the representations of K. If they are irreducible, so are the
resulting representations of K.
It is now possible to obtain a representation of the whole of G from the
representations of A (B.25) and H (B.35), provided we know the composition rule
for G = {(a, h )} in terms of those of A and K. If it is of the form (B.9), for instance,
so that
(a , k ) = (a , e )(0, k ), (B.40)
we obtain:

U ×(g ) ϕ , i = U ×(a , 0) U ×(0, k ) ϕ , i


= e−i([k ] ϕ )(−a ) ∑Δ ji (γ (k , ϕ)) [k ] ϕ , j
j

= e−iϕ([k ](−a )) ∑Δ ji (γ (k , ϕ)) [k ] ϕ , j , (B.41)


j

where we have used the + sign for the composition rule of the Abelian group A. The
third equality follows from the duality condition (B.20).
The dual representation of G on the L2-functions of / readily follows from (B.41)
and (B.26):
(U (g ) f )(ϕ , i ) = e iϕ(a ) ∑Δ*ij (γ (k , ϕ)) f ⎡⎣ k −1⎤⎦ ϕ , j ,
( ) (B.42)
j

where we have taken Δ, and therewith U, to be unitary operators so that


Δ (γ (k −1, ϕ )) = Δ(γ (k , ϕ ))†. As mentioned above, this unitary representation of G is
irreducible if the operators Δ define an irreducible representation of Γϕ0. As seen
from appendix A, the stability groups of any two points on a given orbit are iso-
morphic. Therefore, if we had chosen a point different from ϕ0 on the same orbit, we
would have obtained a representation of G completely equivalent to (B.42). A point
on a different orbit, on the other hand, would lead to a representation inequivalent
to (B.42). This concludes our general discussion on the method of induced repre-
sentations. In the remainder of this appendix, we apply it to construct representa-
tions of the Galilei group.

B-9
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

Galilei group. The projective representations of the Galilei group . arise from both
its topological and algebraic structures. In the topological case, the projective
representations are equivalent to true representations of the Galilei group extended
by Z2, its first homotopy group. Essentially, this is tantamount to replacing the
rotation subgroup SO (3) of the Galilei group by SU (2) to obtain its covering group.
In particular, as we see below, this means that the stability group that defines
induced representations of the Galilei group along the lines of (B.42) is SU (2).
The Galilei group also has projective representations corresponding to algebraic
central extensions. The phase factor eiω( g2, g1) of (B.1) is now given by the two-cocycle
c
ω ( g2, g1) = c ξ ( g2, g1) = ( a 2 · R 2 v1 − v2 · R 2 a1 + b1v2 · R 2 v1) , (B.43)
2
where c is a constant. As mentioned above, this two-cocycle is defined only up to a
two-coboundary. For instance,
⎛ 1 ⎞
ω′ ( g2, g1) = c ⎜ v2 · R 2 a1 + v22 b1⎟ (B.44)
⎝ 2 ⎠

differs from (B.43) by a two-coundary. As such, both (B.43) and (B.44) give rise to
equivalent projective representations or central extensions of the Galilei group.
However, different values of the constant c in either (B.43) or (B.44) lead to
inequivalent projective representations.
Applying the above analysis on Lie alegbras, the operator Lie algebra for a
unitary vector representation of the Galilei group is defined by the commutation
relations
⎡⎣ Ji , J j ⎤⎦ = i ϵijk Jk ⎡⎣ Ji , K j ⎤⎦ = i ϵijk K k ⎡⎣ Ji , Pj ⎤⎦ = i ϵijk Pk

⎡⎣ K i , K j ⎤⎦ = 0 ⎡⎣ Pi , Pj ⎤⎦ = 0 ⎡⎣ K i , Pj ⎤⎦ = 0

⎡⎣ Ji , H ⎤⎦ = 0 ⎡⎣ Pi , H ⎤⎦ = 0 ⎡⎣ K i , H ⎤⎦ = iPi . (B.45)

In order to obtain the commutation relations for the operator Lie algebra of a
unitary representation of the centrally extended Galilei group, let us compute the
additional structure constants from (B.43) and (B.15). The only non-vanishing
ones are:
cvφia j = −caφj vi = δ ij . (B.46)

Hence, the non-vanishing commutation relations for the operator Lie algebra of the
vector representation of .̃ are (omitting the tilde for notational simplicity):
⎡⎣ Ji , J j ⎤⎦ = i ϵijk Jk ⎡⎣ Ji , K j ⎤⎦ = i ϵijk K k ⎡⎣ Ji , Pj ⎤⎦ = i ϵijk Pk

⎡⎣ K i , K j ⎤⎦ = 0 ⎡⎣ Pi , Pj ⎤⎦ = 0 ⎡⎣ K i , Pj ⎤⎦ = iδ ij C = ic δ ij I

⎡⎣ Ji , H ⎤⎦ = 0 ⎡⎣ Pi , H ⎤⎦ = 0 ⎡⎣ K i , H ⎤⎦ = iPi . (B.47)

B-10
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

Note the appearance of the central generator C in the right-hand of the commu-
tation relation between boost and momentum operators. For c ≠ 0, a necessary
condition for the existence of the central extension .̃ , we may multiply these
commutation relations by the inverse operator C −1 to obtain
⎡ Ki ⎤
⎢⎣ , Pj ⎥⎦ = iδ ij . (B.48)
C

These are the familiar Heisenberg commutation relations. Note the critical fact that
they exist only in the algebra of the centrally extended Galilei group .̃ , not in that of
the Galilei group itself.
From the commutation relations (B.47), we can verify that the following operators
in the enveloping algebra of the Lie algebra are central:
C = cI
⎛ 1 ⎞2
S 2 := ⎜ J − K × P⎟ = s (s + 1) I , s = 0, 1/2, 1, 3/2, …
⎝ C ⎠
1 2
W := H − P = wI , w ∈  . (B.49)
2C
1
The last expression shows that C P 2 has units of energy. This suggests the interpretation
1
of C as mass and 2C P 2 as kinetic energy. (It is for this reason that we denoted the
constant c that appears the cocycle (B.43) by m when we introduced it in chapter 2.)
1
Then, W , the difference between total energy H and kinetic energy 2C P 2 has natural
interpretation as internal energy. Its Galilean invariance, derived here as a consequence
of the unitary representations of the centrally extended Galilei group, is a property with
which we are already familiar from classical Newtonian physics. With C being the mass
1
operator, we identify C K as the position operator, canonically conjugated to
1
momentum operator (B.48). Therewith, the term C K × P acquires interpretation as
orbital angular momentum. We then interpret the difference between total angular
1
momentum operator J and orbital angular momentum operator C K × P as the
intrinsic spin operator, the square of which is another Galilean invariant. Thus, the
properties by which we identify an elementary system in Galilean quantum theory, namely
mass, spin and internal energy, all follow from unitary representations of the centrally
extended Galilei group.

Construction of the representations of .̃ . The centrally extended Galilei group con-


sists of elements (φ , a, b, v, R ), with the composition rule

( φ2 , a 2, b2 , v2 , R 2 )( φ1, a1, v1, R1)


= ( φ2 + φ1 + ξ ( g2, g1), a 2 + R 2 a1 + v2 b1, b 2 + b1, v2 + R 2 v1, R 2 R1) , (B.50)

where ξ ( g2, g1 ) is given by (B.43).

B-11
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

The largest Abelian invariant subgroup of the extended Galilei group is


A = {(φ , a, b, 0, I )}. The factor group .̃/A consists of Galilean boosts and rota-
tions, K = {(0, 0, 0, v, R )} and it is isomorphic to E (3). For notational simplicity,
let us denote the elements of A by a = (φ , a, b ) and elements of K by k = (v, R ). The
action of K on A by conjugation can be readily computed from (B.50):

⎡ k −1⎤ a = ⎡ (v , R )−1⎤ (φ , a , b) = ⎛⎜ φ + v · a − 1 v2b , R−1 (a − vb), b⎞⎟ . (B.51)


⎣ ⎦ ⎣ ⎦ ⎝ ⎠
2

We can characterize the elements of the dual group of A by ϕ = (m, p, E ), where


m, E ∈  and p ∈ 3 so that the action ϕ : a → ϕ (a ) is given by
ϕ (a ) = (m , p , E )(φ , a , b) := mφ + p · a − Eb . (B.52)

Next, we must obtain the action of K on Aˇ = {(m , p , E )}. Suppose


[k ] ϕ = [(v , R )](m , p , E ) ≡ (m′ , p′ , E ′). (B.53)

Then, from (B.20), (B.51) and (B.52),

m′φ + p′ · a − E ′b = (m , p , E ) ⎡⎣ (v , R )−1⎤⎦ (φ , a , b)
( )
= m (φ + v · a ) + p · R−1 (a − vb) − Eb . (B.54)

Comparing coefficients of φ, a and b,


m′ = m
p′ = Rp + mv
1 2
E ′ = E + Rp · v + mv (B.55)
2
and so
⎛ 1 ⎞
[(v , R )](m , p , E ) = ⎜ m , Rp + mv , E + Rp · v + mv2⎟ . (B.56)
⎝ 2 ⎠

Eliminating v and R from the last two equalities and using the first of (B.55), we
obtain
1 2 1 2
E′ − p′ = E − p := w. (B.57)
2m 2m
The first equality of (B.55) and (B.57) should be compared with the first and third
expressions of (B.49). The invariance of m under the action of (v, R ) is to be
expected, since m is the parameter conjugated to the central variable φ. It is identical
to the constant c that appears in the two-cocycle (B.43) and has the same inter-
pretation as the eigenvalue of the mass operator. The equation (B.57) describes the

B-12
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

K-orbit of a point (m, p, E ) ∈ Aˇ . It tells us that E and p are not independent for
non-zero mass. Rather,
1 2
E= p +w (B.58)
2m

for a Galiliean invariant parameter w. Therefore, we may label an element of Ǎ by


(m, w, p ), or with the understanding that the invariant parameters m and w are
implicitly at play, simply by p .
In order to write down the representation of the centrally extended Galilei group,
we only have to determine the stability group of a certain point (m, p0 , E0 ), or
in our abbreviated notation p0, and choose a standard boost that carries p0
to an arbitrary point p on the K-orbit of p0. All the points on the K-orbit of
1
(m, p0 , E0 = w + 2m p02 ) are given by either (B.55) or (B.57). The stability group of
p0 can be obtained from (B.55) by the requirement p′ = p0 , or E ′ = E0:
p′ = Rp0 + mv = p0
1 2
E ′ = E 0 + Rp · v + mv = E 0. (B.59)
2
From (B.57) or (B.58), these two equations are clearly not independent constraints
on v and R. The solution of either gives the stability group of p0:
p
Γ( m,w,p 0 ) = { ( v ( p0 ), R ) : v ( p0 ) = (I − R) m0 } . (B.60)

Here, R is an arbitrary rotation. Once a specific R is chosen, the Galilean boost v( p0 )


is uniquely determined by it and the chosen point on the K-orbit. Hence, for m ≠ 0,
the stability group is isomorphic to the rotation group, or its universal covering
SU (2) if we deal with the central extensions of the covering group of the Galilei
group.
The transformation formula (B.55) also suggests the natural choice for the
standard boost ⎡⎣ k˜ p ⎤⎦ : p0 → p:
⎛1 ⎞ 1
(v ,˜R ) p = ⎜ ( p − p0 ), I ⎟ ≡ ( p − p0 ) (B.61)
⎝m ⎠ m

Therefore, the action of the element of the stability group denoted in the transfor-
mation formulas (B.39) and (B.42) by γ (k , ϕ ) becomes
⎡ −1 ⎤
[γ (k , ϕ)] := ⎣ k˜[k ] ϕ hk˜ ϕ ⎦

⎡ ⎛ (I − R ) p ⎞⎤
= ⎢⎜ , R⎟ ⎥ .
0
(B.62)
⎢⎣ ⎝ m ⎠ ⎥⎦

B-13
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

Now we can write down the representations of .̃ . For the Abelian invariant
group A = {(φ , a , b )}, with the use of (B.52), from (B.24) we obtain,

(U ( (φ , a , b) ) f ) (p , i ; m , w) = e i(mφ+p·a−Eb) f (p , i ; m , w) (B.63)

or, from (B.25),


U ×( (φ , a , b) ) p , i ; m , w = e i(mφ+p·a−Eb) p , i ; m , w . (B.64)

For the group K = {(v, R )} of Galilean boosts and rotations, the representation
follows from (B.62), (B.53) and (B.35):

U ×((v , R )) p , i ; m , s , w = ∑D jis ( (I − R) p0 /m, R )


j

× Rp + mv , j ; m , s , w , (B.65)
where, since the stability group is isomorphic to the rotation group, we have used
the standard symbol D s for its irreducible representations. The superscript s has
the same meaning as in the second expression of (B.49) and assumes one value from
the set s = 0, 1/2, 1, 3/2, …. For a given s, the i varies from −s to +s in integer steps.
We have included the invariant parameter s, along with m and w, in the labeling of
the basis vectors of the representation. Note that we use the more commonly used
notation of σ to indicate spin projection in chapter 2.
In order to obtain the representation of the full group .̃ by combining (B.63) and
(B.65), we must first consider the decomposition of an arbitrary element
(φ , a, b, v, R ) into an element of A and an element of K:
⎛ 1 ⎞
(φ , a , b , v , R ) = ⎜φ − v · a , a , b , 0, I ⎟ (0, 0, 0, v , R ). (B.66)
⎝ 2 ⎠

(This shows that .̃ is not the semi-direct product of A and K. Neither is it an


extension in the sense of (B.9), Cf (B.40).) Then,
⎛ 1 ⎞
U ×(φ , a , b , v , R ) p , j ; m , s , w = U ×⎜φ − v · a , a , b⎟ U ×((v , R ))
⎝ 2 ⎠
× p, j; m, s , w
⎛ 1 ⎞
= U ×⎜φ − v · a , a , b⎟ ∑D jis ( (1 − R ) p0 / m , R )
⎝ 2 ⎠
j

× p′ , j ; m , s , w
= e− 2 mv·ae i(mφ+p′·a−E ′b) ∑D jis ( (1 − R ) p0 / m , R )
i

× p′ , j ; m , s , w , (B.67)
1 2
where p′ and E′ are defined by (B.55) and E = 2m
p + w.

B-14
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

As usual, the transformation formula for the L2-functions on the K-orbit of p0 can
be obtained from (B.67) and the duality formula (B.26):
i
U ((φ , a , b , v , R ) f ) (p , j ; m , s , w) = e− 2 mv·ae i(mφ+p·a−Eb)
× ∑Dijs ( (I − R ) p0 / m , R ) f ( p˜ , j ; m , s , w) ,
j

(B.68)
where
p˜ = ⎡⎣ (v , R )−1⎤⎦ p = R−1p − mR−1v . (B.69)

As a final remark, note that (B.61) shows that the point p = 0 is always on
any orbit of m ≠ 0. Since the stability groups of all points on a given orbit are
isomorphic, without loss of generality, we may set p0 = 0. This reduces (B.62) to
⎡ −1 ⎤
[γ (k , ϕ)] := ⎣ k˜[k ] ϕ kk˜ ϕ ⎦ = [R ] (B.70)

and simplifies the representation:


i
U ×(φ , a , b , v , R ) p , j ; m , s , w = e− 2 mv·ae i(mφ+p′·a−E ′b)
× ∑D jis (R ) p′ , j ; m , s , w
j
i
U ((φ , a , b , v , R ) f ) ( p , j ; m , s , w) = e− 2 mv·ae i(mφ+p·a−Eb)
× ∑Dijs (R ) f ( p˜ , j ; m , s , w) . (B.71)
j

Note that we have replaced F with φ in chapter 2 to make the notation more in line
with the notation commonly used in quantum mechanics.
This concludes our discussion of central extensions of the Galilei group and its
unitary, irreducible representations. The same method can be used to construct
unitary, irreducible representations of the Poincaré group. It is remarkable that
all unitary irreducible representations of both Poincaré and (centrally extended)
Galilei groups can be obtained this way.

Bibliography
[1] Wigner E P 1939 On the unitary representations of the inhomogeneous Lorentz group Ann.
Math. 40
[2] Mackey G W 1952 Induced representations of locally compact groups I Ann. Math. 55 101;
1953 Induced representations of locally compact groups II The Frobenius reciprocity theorem
Ann. Math. 58 193
[3] Bargmann V 1954 On unitary ray representations of continuous groups Ann. Math. 59 1
[4] Taylor M E 1986 Noncommutative Harmonic Analysis (Mathematical Surveys and Mono-
graphs, no. 22) (Providence, RI: American Mathematical Society)

B-15
IOP Concise Physics

Relativity, Symmetry and the Structure of Quantum Theory I:


Galilean quantum theory
William H Klink and Sujeev Wickramasekara

Appendix C
Decomposition of n-fold tensor products
and Clebsch–Gordan coefficients

Given the irreducible representations of a group, it is always possible to generate


new representations by taking tensor products. For the groups considered in this
book, this corresponds to passing from single particle states to multiparticle states.
Representations generated by tensor products are generally reducible and, for the
groups under consideration here, can be decomposed into direct sums (integrals) of
irreducible representations. This appendix concerns the transformation coefficients
(traditionally called Clebsch–Gordan coefficients) connecting the reducible tensor
product representations with the direct sum (integral) of irreducible representations.
To see how these coefficients arise, consider a basis state χ , x for the irreducible
representation χ of a group G. x denotes the eigenvalues of a complete set of
commuting operators from the Lie algebra of G. Then an n-fold tensor product can
be written as ∣ χ1 , x1〉 ⊗ … ⊗ ∣ χn , xn〉, and provides a basis for the reducible repre-
sentation of G.
Assume now that this reducible representation of G has been decomposed into a
direct sum (integral) of irreducible representations of G, with a basis written
χ , y, η , where χ is again an irreducible representation (irrep) label. Also, y again
labels the eigenvalues of a compete set of commuting operators from the Lie algebra
of G, but not necessarily the same set as occurred in the tensor product.
In the decomposition to irreducibles, it is generally the case that a given irre-
ducible representation may occur more than once. This is the so-called multiplicity
problem, wherein it is necessary to find some way to distinguish between repre-
sentations that occur more than once in the decomposition. The label η distinguishes
between these equivalent representations.

doi:10.1088/978-1-6270-5624-3ch9 C-1 ª Morgan & Claypool Publishers 2015


Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

Given these two bases in the n-fold tensor product space, it is possible to write

χ1 , x1 ⊗ … ⊗ χn , xn = ∑∫ χ,y,η χ , y , η χ1 , x1; … ; χn , xn χ , y , η , (C.1)

χ, y, η = ∑∫x .. x χ1 , x1; … ; χn , xn χ , y , η
1 n

χ1 , x1 ⊗ … ⊗ χn , xn ; (C.2)

here 〈∣〉 and inverse are the Clebsch–Gordan coefficients. If the irreps of G are
unitary (as they are always assumed to be in this book), the Clebsch–Gordan
coefficients will be unitary matrices (kernels).
There are number of different ways of generating multiplicity labels η. The best
known involves a method called stepwise coupling. If there is no multiplicity
appearing in the two-fold tensor product (that is, if every irrep of G occurs either
once or not at all), then it is possible to keep decomposing two-fold pieces until the
full n-fold tensor product has been decomposed. The best-known example of this is
the group SU (2), discussed below. For stepwise coupling the multiplicity label η is
given by the set of irrep labels that occur in the intermediate coupling. For example,
for a three-fold tensor product, if 1 is coupled to 2 to form an intermediate state with
irrep χ1−2 , which is then coupled to 3, the multiplicity label η will be given by χ1−2 .
The disadvantage of stepwise coupling is that it singles out one stepwise scheme
over many other possibilities. In the above three-fold example, 1 could be coupled to
3 instead of 2, in which case the multiplicity label would be χ1−3.
To get around this problem it is possible to introduce the notion of simultaneous
coupling, wherein all n irreps are coupled together at once, resulting in a new set of
multiplicity labels. For simultaneous coupling, each group must be treated sepa-
rately, as seen in the following examples.
The analysis of tensor product representations discussed here was first developed
for the angular momentum group SU (2) (SO(3)). As is shown in most books on
quantum theory, the label χ becomes the total angular momentum j, and every
irreducible representation is given by a value of j = 0, 1 2, 1…. The angular
momentum commutation relations show that the Lie algebra elements, Ji, i = 1,2,3,
do not commute with one another, so that one operator, usually chosen to be J3,
constitutes a complete set of commuting operators, with eigenvalues running
between −j and +j ; in the notation used above, these are the eigenvalues x. As is also
shown in quantum mechanics texts, no multiplicity occurs with two-fold tensor
products; rather, if the irrep j1 is tensored with the irrep j2, the only irreps to appear
in the decomposition of the two-fold tensor product are those between j1 + j2 and
∣ j1 − j2 ∣.
What we wish to do in this appendix is generalize these results for the unitary
groups; as explained in chapter 5, these groups seem to play the most important role
in internal symmetries. We will first present a method for obtaining all of the unitary
irreducible representations of the unitary groups and then discuss how to deal with
tensor products.

C-2
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

It is possible to obtain all the unitary irreducible representations of the unitary


groups by a method called holomorphic induction [1, 2]. Just as all the irreducible
representations of the Galilei group and the Galilean line group were obtained by
the method of induced representations, so too can all the irreducible representations
of the unitary groups, but on different representation spaces. The representation
spaces are holomorphic functions over complex N × N matrices, with certain
transformation properties; for the purposes of this appendix it suffices to consider
only polynomial functions. Thus, consider polynomial complex functions over N × N
complex matrices satisfying the left covariance conditions:

V (m ) = { f :  N ×N →  f (dz ) = d1m1…d Nm N f (z ); f (bz ) = f (z ); f < ∞} , (C.3)

T (g )f (z ) = f (zg ), (C.4)


L i , j f (z ) = ∑zi,k ∂z f (z ); (C.5)
k j ,k

here z is an N × N complex matrix with indices i , j ranging from 1 to N and g is an


element of GL(N , ), the group of N × N non-singular matrices over the complex
numbers. (m ) is an array of N non-negative integers satisfying m1 ⩾ m 2 ⩾ …. ⩾ m N ,
d is the diagonal subgroup of GL(N , ), and b is an element of the ‘Borel’ subgroup
of GL(N , ), consisting of lower triangular matrices. There are two inner products
which can be shown to be equivalent, one an integration inner product, the other a
differentiation inner product. The differentiation inner product is more useful from
a computational point of view and is given by

⎛ ∂⎞
f 2
:= f ⎜ ⎟ f * z*
⎝ ∂z ⎠ ( ) z=0 . (C.6)

That is, the length squared of a polynomial f is given by replacing z by its partial
derivatives in f, which then act on complex conjugated functions, all evaluated at the
point zero.
If the elements of GL(N , ) are restricted to the unitary subgroup, it can be shown
that the representation, (C.4), acting on the representation space (C.3), gives unitary
irreducible representations for all values of (m ) with respect to the inner product
(C.6). Further, the representation of the Lie algebra, given in (C.5), agrees with the
commutation relations given in (5.1), and thus furnishes a representation of the Lie
algebra of the unitary groups by differential operators acting on the representation
spaces (C.3).
Consider as a simple example the so-called ‘symmetric’ or ‘bosonic’ representa-
tions, in which (m ) = (m, 0, …, 0). In this case the representation space can be
chosen to be polynomials in 1 × N complex variables so that f (dz ) = f (d (z1…zN ) =
d mf (z ). In particular, if N = 2 this provides a representation for the group SU (2),
m
with the usual irrep label j = 2 .

C-3
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

A main advantage of using holomorphic induction to get the unitary irreducible


representations of the unitary groups is that it is basis independent. However, to be
of use in physical applications a basis in the vector space must be chosen. For the
unitary groups a well-known basis is the Gelfand–Zetlin basis, which is obtained by
making use of the following properties of representations of unitary subgroups of the
unitary groups. If U (N − 1) is a subgroup of U(N), then the only representations of
U (N − 1) that occur in a given representation of U(N) are those satisfying certain
betweeness relations with respect to the integers (m ). Since there is no multiplicity,
the chain from U(N) down to U (1) uniquely fixes a basis in the representation space
for the representation (m ) [1, 2].
For the group SU (2) the usual basis is one in which the subgroup U (1) is
embedded in SU (2) in such a way that the irrep k = σ + j of U (1) is between m = 2j
and zero; that is, j ⩾ σ ⩾ −j , which is the usual condition for the spin projection.
Further, a basis in V m can be chosen such that hσj = z1j +σ z2j −σ , with Lie algebra basis
elements given by

J+ = L1,2 = z1 , (C.7)
∂z2

J − = L 2,1 = z2 , (C.8)
∂z1

1 1⎛ ∂ ∂ ⎞
J3 =
2
( L1,1 − L 2,2 ) = ⎜ z1
2 ⎝ ∂z1
− z2 ⎟,
∂z2 ⎠
(C.9)

with commutation relations agreeing with those in (5.4). Using the differentiation
inner product, the basis elements can be normalized so that
z1j +σ z2j −σ
z jσ = ; (C.10)
( j + σ) ! ( j − σ)!

∂ z1j +σ z2j −σ
z J+ jσ = z1
∂z2 ( j + σ) ! ( j − σ)!

= ( j + σ + 1)( j − σ ) z j , σ + 1 . (C.11)
Using the group action, (C.4), it is also possible to compute matrix elements in a
j, σ basis, namely
D σj′σ (g ) := j , σ ′ T (g ) j , σ
∂ j +σ′ ∂ j −σ′
∂z1 ∂z2 z1j +σ z2j −σ
= T (g )
( j + σ ′) ! ( j − σ ′) ! ( j + σ) ! ( j − σ)!
z=0

∂ j +σ′ ∂ j −σ′
∂z1 ∂z2 ( z1g11 + z2 g21) j +σ ( z1g12 + z2 g22 ) j −σ
= , (C.12)
( j + σ ′) ! ( j − σ ′) ! ( j + σ) ! ( j − σ)!
z=0

C-4
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

and if the indicated differentiation is carried out after using the binomial expansion,
a polynomial in group elements results. In particular, if the group element is chosen
β β
so that g11 = g22 = cos 2 , g12 = g21 = sin 2 , where β is the middle Euler angle, then
the result agrees with the ‘little’ d function given in quantum texts [3], namely

( j + σ ′) ! ( j − σ ′) ! ( j + σ ) ! ( j − σ ) !
d σj′σ (β ) = ( −1)σ′−σ ∑ ( −1) μ
μ
( j + σ − μ) ! (σ ′ + μ − σ ) ! ( j − μ − σ ′) !
⎛ β ⎞⎟2j −2μ+σ−σ′ ⎛⎜ β ⎞2μ+σ′−σ
⎜ cos sin ⎟ . (C.13)
⎝ 2⎠ ⎝ 2⎠

Given these results for the one-particle representation spaces of the unitary
groups, we now proceed to tensor products of such representations. The tensor
product representation spaces are given by V (m1) ⊗ … ⊗ V (m n ) , that is, polynomials
over complex variables with left covariance conditions. The goal is to find maps
carrying an irreducible representation (M ) into this tensor product space; in general,
this involves multiplicity, in that the representation (M ) will occur more than once in
the tensor product space. What is needed are operators that act on the tensor pro-
duct space and commute with the group action. Such operators are called general-
ized Casimir operators, and their eigenvalues can be used to break the multiplicity.
The details for using generalized Casimir operators for computing Clebsch–Gordan
coefficients for the unitary groups is complicated and will not be given here. For
works that show how to carry out such a program and how using the differentiation
inner product provides a way to compute the relevant Clebsch–Gordan coefficients,
see [1].
However, for the purposes of this book, there is a method for dealing with tensor
products of the simpler unitary groups, such as SU (2), that does not involve the
complicated machinery mentioned above. Consider an n-fold tensor product
∣ j1 σ1 〉 ⊗ …∣ jn σn 〉 for which we want to find the Clebsch–Gordan coefficients for the
irrep ∣ jσ 〉; now j may occur more then once in the tensor product decomposition, so
it will be necessary to add a multiplicity label η in order to write

jσ , η = ∑ j1 σ1…ji σi…jn σn jσ , η ( j1 σ1 ⊗ .. ⊗ jn σn ), (C.14)


i

where are the coefficients to be calculated. Acting on both sides of this equation
with the operator J3 shows that for a given σ, the only possible spin projections in the
tensor product are those satisfying σ = σ1 + … + σn . Next set σ = j and apply the
raising operator J+ to both sides; the left side gives zero, while the right-hand-side is
evaluated using (C.11). This results in a set of equations for the Clebsch–Gordan
coefficients, for which there are three possible solutions. First, there may be no
solution to the equations, indicating that the desired irrep does not occur in the
tensor product decomposition. Second, there may be a unique solution (up to a
normalization), which indicates that the desired representation occurs once and
the multiplicity is one. Finally, there may be many solutions, indicating that the

C-5
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

multiplicity is greater than one, and other operators must be introduced whose
eigenvalues break the multiplicity.
In the following we give several examples from SU (2) to indicate how this works.
For a two-fold tensor product there is no multiplicity and the Clebsch–Gordan
coefficients can be calculated in closed form. In fact, there are computer programs
for computing any desired two-fold tensor product [2]. However, starting with three-
fold tensor products, multiplicity does occur. For example, if j = 1 is tensored with
itself three times, the irrep j = 1 occurs three times, while the irrep j = 2 occurs twice.
If stepwise coupling is used to generate multiplicity labels, then one possibility is to
couple 1 to 2, resulting in an intermediate coupling label j1−2 with possible values 2,
1, or 0. Using stepwise coupling is one way of using generalized Casimir operators;
that is, the eigenvalues that label stepwise coupling schemes are eigenvalues of
operators that commute with the group action. But in the example given of tensoring
j = 1 with itself three times, there is a permutation group symmetry which can be
used to break the multiplicity using simultaneous rather than stepwise coupling [4].
To begin, consider the j = 0 representation which occurs only once in the
decomposition. Since there is no multiplicity, we can write
j = 0, σ = 0 = a1 1, 0, −1 + a 2 1, −1, 0 + a3 0, 1, −1 + a 4 0, −1, 1
+ a5 −1, 1, 0 + a6 −1, 0, 1 + a 7 0, 0, 0 ; (C.15)

J+ j = 0, σ = 0 = 2 ( 1, 1, −1 ( a1 + a3 ) + 1, −1, 1 ( a 2 + a 4 ) + −1, 1, 1
(a5 + a6 ) + 1, 0, 0 ( a1 + a 2 + a 7 ) + 0, 1, 0 ( a3 + a5 + a 7 )
+ 0, 0, 1 ( a 4 + a6 + a 7 )
= 0. (C.16)
Here ai is shorthand for the Clebsch–Gordan coefficients in (C.15) that are to be
calculated, and ∣ σ1, σ2, σ3 〉 is shorthand for ∣ 1, σ1 〉 ⊗ ∣ 1, σ2 〉 ⊗ ∣ 1, σ3 〉. Since the
coefficients multiplying each tensor product state must be zero, there are six equa-
tions in seven unknowns. Using these equations all the coefficients can be solved in
terms of one coefficient, say a1, which will act like an overall coefficient. Since we
want the j = 0 state to be normalized to one, a1 is fixed by normalization. The final
result is [5]
1
j = 0, σ = 0 = ( 1, 0, −1 − 1, −1, 0 − 0, 1, −1
6
+ 0, −1, 1 + −1, 1, 0 − −1, 0, 1 ). (C.17)
As a second example consider j = 2, which occurs with multiplicity 2. The state with
maximum value of spin projection can be written as
j = 2, σ = 2, η = a1 1, 1, 0 + a 2 1, 0, 1 + a3 0, 1, 1 (C.18)

J+ 2, 2 = 2 ( a1 + a 2 + a3 ) 1, 1, 1 = 0 (C.19)

a1 + a 2 + a3 = 0. (C.20)

C-6
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

Equation (C.20) says that the three coefficients must sum to zero, which is insuffi-
cient to determine the coefficients. However, the tensor product is invariant under
permutations of the three spin projections and hence must carry a representation of
the permutation group on three numbers, S3. This is also the case for the previous
example with j = 0. Though there is no multiplicity, the Clebsch–Gordan coefficients
carry the antisymmetric representation of S3, as can be seen by permuting the
allowed spin projections. For the case j = 2 there is multiplicity and the irreps of S3
can be used to break the multiplicity.
The first step is to determine which irreps of S3 occur in the j = 2 Clebsch–Gordan
coefficients. (For definitions of terms such as characters, see chapter 5 of [6].) This
is done by constructing the matrix representation of S3 which is generated by
permuting the spin projections and seeing how the ai then get permuted. Taking
the trace of these matrices gives the characters, from which one deduces that the
symmetric and mixed (two-dimensional) irreps each occur once. But the symmetric
irrep applied to the constraint (C.20) implies that all the ai are zero. Thus the
degeneracy is broken via the two-dimensional mixed representation of S3.
While there are many ways of using representations of S3 to break the multiplicity
(which must finally be determined by the application), here we present a simple way
of breaking the multiplicity by making use of the fact that the two-dimensional
representation of S3 contains the symmetric and antisymmetric representation of its
subgroup S2. Choosing the two elements of the S2 subgroup as the identity
e = (1)(2)(3) and the element (12)(3), which permutes the first and second spin
projection, gives two conditions on the ai:
a 2 = a3; a1 = −2a 2 , (C.21)
a 2 = −a3; a1 = 0. (C.22)
Call the first multiplicity condition the S = symmetric condition and the second the
A = antisymmetric condition. Then the two j = 2 states can be written as
1
j = 2, σ = 2, S = ( −2 1, 1, 0 + 1, 0, 1 + 0, 1, 1 ) (C.23)
6
1
j = 2, σ = 2, A = ( 1, 0, 1 − 0, 1, 1 ; (C.24)
2
once the Clebsch–Gordan coefficients for the S state, (C.23), and A state, (C.24), are
known, all other coefficients can be obtained by application of the lowering operator
J − to these states. Note that the S and A states are normalized to one and are
orthogonal to one another.
Turning to the Galilei group, Clebsch–Gordan coefficients for the two-fold tensor
product decomposition were worked out in chapter 4, (4.3). In this case the irrep labels
are given by mass, spin and internal energy, so that χ → M = m1 + m 2 , j , and q.
Now x includes the overall velocity Q⃗ and spin projection mj. Unlike SU (2), the
two-fold tensor product reduction of the Galilei group is not multiplicity free.
The multiplicity label η includes l and s, the orbital and spin angular momentum.

C-7
Relativity, Symmetry and the Structure of Quantum Theory I: Galilean quantum theory

Finally, it is possible to carry out the n-fold tensor product reduction of Galilei
group irreps. This can be done in a stepwise fashion, by repeatedly using the two-
fold reduction; but this is quite complicated, particularly for identical particles.
Much more straightforward is the simultaneous reduction, in which the orbital and
spin angular momentum are degeneracy parameters for the full n-particle system.
The irrep labels are again given by mass, spin and internal energy. Details are given
in reference [7].

Bibliography
[1] Klink W H and Ton-That T 2009 Invariant Theory of tensor product decompositions of
U(N) and generalized Casimir operators Notices Amer. Math. Soc. 56 931
[2] Gliske S, Klink W H and Ton-That T 2005 Algorithms for computing generalized U(N)
Racah coefficients Acta Appl. Math. 88 229
Gliske S, Klink W H and Ton-That T 2007 Algorithms for computing U(N) Clebsch–Gordan
coefficients Acta Appl. Math. 95 51
[3] Sakurai J J and Napolitano J J 2014 Modern Quantum Mechanics 2nd edn (Harlow: Pearson)
[4] Klink W H and Ton-That T 1990 Representations of Sn × U(N) in repeated tensor products of
the unitary group J. Phys. A: Math. Gen. 23 2751
[5] Klink W H and Wickramasekara S 2010 A simple method for calculating Clebsch–Gordan
coefficients Eur. J. Phys. 31 1021
[6] Tung W-K 2003 Group Theory in Physics (Singapore: World Scientific)
[7] Warren R E and Klink W H 1970 Model independent analysis of nonrelativistic multiparticle
reactions J. Math. Phys. 11 1155

C-8

You might also like