Doran C.J.L. Geometric Algebra and Its Application To Mathematical Physics (PHD Thesis, Cambridge, 1994) (187s)
Doran C.J.L. Geometric Algebra and Its Application To Mathematical Physics (PHD Thesis, Cambridge, 1994) (187s)
Doran C.J.L. Geometric Algebra and Its Application To Mathematical Physics (PHD Thesis, Cambridge, 1994) (187s)
to Mathematical Physics
Chris J. L. Doran
Sidney Sussex College
Acknowledgements
Many people have given help and support over the last three years and I am grateful to
them all. I owe a great debt to my supervisor, Nick Manton, for allowing me the freedom
to pursue my own interests, and to my two principle collaborators, Anthony Lasenby and
Stephen Gull, whose ideas and inspiration were essential in shaping my research. I also
thank David Hestenes for his encouragement and his company on an arduous journey to
Poland. Above all, I thank Julie Cooke for the love and encouragement that sustained
me through to the completion of this work. Finally, I thank Stuart Rankin and Margaret
James for many happy hours in the Mill, Mike and Rachael, Tim and Imogen, Paul, Alan
and my other colleagues in DAMTP and MRAO.
I gratefully acknowledge
nancial support from the SERC, DAMTP and Sidney Sussex
College.
To my parents
Contents
1 Introduction 1
1.1 Some History and Recent Developments : : : : : : : : : : : : : : : : : : : 4
1.2 Axioms and De
nitions : : : : : : : : : : : : : : : : : : : : : : : : : : : : : 7
1.2.1 The Geometric Product : : : : : : : : : : : : : : : : : : : : : : : : 11
1.2.2 The Geometric Algebra of the Plane : : : : : : : : : : : : : : : : : 12
1.2.3 The Geometric Algebra of Space : : : : : : : : : : : : : : : : : : : : 15
1.2.4 Reections and Rotations : : : : : : : : : : : : : : : : : : : : : : : 17
1.2.5 The Geometric Algebra of Spacetime : : : : : : : : : : : : : : : : : 21
1.3 Linear Algebra : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : 23
1.3.1 Linear Functions and the Outermorphism : : : : : : : : : : : : : : 23
1.3.2 Non-Orthonormal Frames : : : : : : : : : : : : : : : : : : : : : : : 26
2 Grassmann Algebra and Berezin Calculus 29
2.1 Grassmann Algebra versus Cliord Algebra : : : : : : : : : : : : : : : : : : 29
2.2 The Geometrisation of Berezin Calculus : : : : : : : : : : : : : : : : : : : 30
2.2.1 Example I. The \Grauss" Integral : : : : : : : : : : : : : : : : : : : 33
2.2.2 Example II. The Grassmann Fourier Transform : : : : : : : : : : : 34
2.3 Some Further Developments : : : : : : : : : : : : : : : : : : : : : : : : : : 37
3 Lie Groups and Spin Groups 39
3.1 Spin Groups and their Generators : : : : : : : : : : : : : : : : : : : : : : : 39
3.2 The Unitary Group as a Spin Group : : : : : : : : : : : : : : : : : : : : : 44
3.3 The General Linear Group as a Spin Group : : : : : : : : : : : : : : : : : 48
3.3.1 Endomorphisms of <n : : : : : : : : : : : : : : : : : : : : : : : : : 54
3.4 The Remaining Classical Groups : : : : : : : : : : : : : : : : : : : : : : : 59
3.4.1 Complexi
cation | so(n,C) : : : : : : : : : : : : : : : : : : : : : : 59
3.4.2 Quaternionic Structures | sp(n) and so (2n) : : : : : : : : : : : : 61
3.4.3 The Complex and Quaternionic General Linear Groups : : : : : : : 64
3.4.4 The symplectic Groups Sp(n,R) and Sp(n,C) : : : : : : : : : : : : : 65
3.5 Summary : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : 66
4 Spinor Algebra 68
4.1 Pauli Spinors : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : 69
4.1.1 Pauli Operators : : : : : : : : : : : : : : : : : : : : : : : : : : : : : 73
4.2 Multiparticle Pauli States : : : : : : : : : : : : : : : : : : : : : : : : : : : 74
i
4.2.1 The Non-Relativistic Singlet State : : : : : : : : : : : : : : : : : : : 77
4.2.2 Non-Relativistic Multiparticle Observables : : : : : : : : : : : : : : 78
4.3 Dirac Spinors : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : 79
4.3.1 Changes of Representation | Weyl Spinors : : : : : : : : : : : : : 83
4.4 The Multiparticle Spacetime Algebra : : : : : : : : : : : : : : : : : : : : : 85
4.4.1 The Lorentz Singlet State : : : : : : : : : : : : : : : : : : : : : : : 89
4.5 2-Spinor Calculus : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : 91
4.5.1 2-Spinor Observables : : : : : : : : : : : : : : : : : : : : : : : : : : 93
4.5.2 The 2-spinor Inner Product : : : : : : : : : : : : : : : : : : : : : : 95
4.5.3 The Null Tetrad : : : : : : : : : : : : : : : : : : : : : : : : : : : : : 96
4.5.4 The rA A Operator : : : : : : : : : : : : : :
0
: : : : : : : : : : : : : 98
4.5.5 Applications : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : 99
5 Point-particle Lagrangians 102
5.1 The Multivector Derivative : : : : : : : : : : : : : : : : : : : : : : : : : : : 102
5.2 Scalar and Multivector Lagrangians : : : : : : : : : : : : : : : : : : : : : : 105
5.2.1 Noether's Theorem : : : : : : : : : : : : : : : : : : : : : : : : : : : 106
5.2.2 Scalar Parameterised Transformations : : : : : : : : : : : : : : : : 106
5.2.3 Multivector Parameterised Transformations : : : : : : : : : : : : : 107
5.3 Applications | Models for Spinning Point Particles : : : : : : : : : : : : : 108
6 Field Theory 121
6.1 The Field Equations and Noether's Theorem : : : : : : : : : : : : : : : : : 122
6.2 Spacetime Transformations and their Conjugate Tensors : : : : : : : : : : 124
6.3 Applications : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : 128
6.4 Multivector Techniques for Functional Dierentiation : : : : : : : : : : : : 135
7 Gravity as a Gauge Theory 137
7.1 Gauge Theories and Gravity : : : : : : : : : : : : : : : : : : : : : : : : : : 138
7.1.1 Local Poincare Invariance : : : : : : : : : : : : : : : : : : : : : : : 140
7.1.2 Gravitational Action and the Field Equations : : : : : : : : : : : : 143
7.1.3 The Matter-Field Equations : : : : : : : : : : : : : : : : : : : : : : 148
7.1.4 Comparison with Other Approaches : : : : : : : : : : : : : : : : : : 152
7.2 Point Source Solutions : : : : : : : : : : : : : : : : : : : : : : : : : : : : : 155
7.2.1 Radially-Symmetric Static Solutions : : : : : : : : : : : : : : : : : 156
7.2.2 Kerr-Type Solutions : : : : : : : : : : : : : : : : : : : : : : : : : : 166
7.3 Extended Matter Distributions : : : : : : : : : : : : : : : : : : : : : : : : 169
7.4 Conclusions : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : 174
ii
List of Tables
1.1 Some algebraic systems employed in modern physics : : : : : : : : : : : : : 6
3.1 Bivector Basis for so(p,q) : : : : : : : : : : : : : : : : : : : : : : : : : : : : 44
3.2 Bivector Basis for u(p,q) : : : : : : : : : : : : : : : : : : : : : : : : : : : : 48
3.3 Bivector Basis for su(p,q) : : : : : : : : : : : : : : : : : : : : : : : : : : : : 48
3.4 Bivector Basis for gl(n,R) : : : : : : : : : : : : : : : : : : : : : : : : : : : 51
3.5 Bivector Basis for sl(n,R) : : : : : : : : : : : : : : : : : : : : : : : : : : : 51
3.6 Bivector Basis for so(n,C) : : : : : : : : : : : : : : : : : : : : : : : : : : : 60
3.7 Bivector Basis for sp(n) : : : : : : : : : : : : : : : : : : : : : : : : : : : : 62
3.8 Bivector Basis for so (n) : : : : : : : : : : : : : : : : : : : : : : : : : : : : 63
3.9 Bivector Basis for gl(n,C) : : : : : : : : : : : : : : : : : : : : : : : : : : : 64
3.10 Bivector Basis for sl(n,C) : : : : : : : : : : : : : : : : : : : : : : : : : : : : 65
3.11 Bivector Basis for sp(n,R) : : : : : : : : : : : : : : : : : : : : : : : : : : : 66
3.12 The Classical Bilinear Forms and their Invariance Groups : : : : : : : : : : 67
3.13 The General Linear Groups : : : : : : : : : : : : : : : : : : : : : : : : : : 67
4.1 Spin Currents for 2-Particle Pauli States : : : : : : : : : : : : : : : : : : : 79
4.2 Two-Particle Relativistic Invariants : : : : : : : : : : : : : : : : : : : : : : 91
4.3 2-Spinor Manipulations : : : : : : : : : : : : : : : : : : : : : : : : : : : : : 95
iii
Chapter 1
Introduction
This thesis is an investigation into the properties and applications of Cliord's geometric
algebra. That there is much new to say on the subject of Cliord algebra may be a surprise
to some. After all, mathematicians have known how to associate a Cliord algebra with
a given quadratic form for many years 11] and, by the end of the sixties, their algebraic
properties had been thoroughly explored. The result of this work was the classi
cation of
all Cliord algebras as matrix algebras over one of the three associative division algebras
(the real, complex and quaternion algebras) 12]{16]. But there is much more to geometric
algebra than merely Cliord algebra. To paraphrase from the introduction to \Cli ord
Algebra to Geometric Calculus" 24], Cliord algebra provides the grammar from which
geometric algebra is constructed, but it is only when this grammar is augmented with a
number of secondary de
nitions and concepts that one arrives at a true geometric algebra.
In fact, the algebraic properties of a geometric algebra are very simple to understand, they
are those of Euclidean vectors, planes and higher-dimensional (hyper)surfaces. It is the
computational power brought to the manipulation of these objects that makes geometric
algebra interesting and worthy of study. This computational power does not rest on the
construction of explicit matrix representations, and very little attention is given to the
matrix representations of the algebras used. Hence there is little common ground between
the work in this thesis and earlier work on the classi
cation and study of Cliord algebras.
There are two themes running through this thesis: that geometric algebra is the nat-
ural language in which to formulate a wide range of subjects in modern mathematical
physics, and that the reformulation of known mathematics and physics in terms of geo-
metric algebra leads to new ideas and possibilities. The development of new mathematical
formulations has played an important role in the progress of physics. One need only con-
sider the bene
ts of Lagrange's and Hamilton's reformulations of classical mechanics, or
Feynman's path integral (re)formulation of quantum mechanics, to see how important the
process of reformulation can be. Reformulations are often interesting simply for the novel
and unusual insights they can provide. In other cases, a new mathematical approach can
lead to signi
cant computational advantages, as with the use of quaternions for combining
rotations in three dimensions. At the back of any programme of reformulation, however,
lies the hope that it will lead to new mathematics or physics. If this turns out to be
the case, then the new formalism will usually be adopted and employed by the wider
community. The new results and ideas contained in this thesis should support the claim
1
that geometric algebra oers distinct advantages over more conventional techniques, and
so deserves to be taught and used widely.
The work in this thesis falls broadly into the categories of formalism, reformulation
and results. Whilst the foundations of geometric algebra were laid over a hundred years
ago, gaps in the formalism still remain. To
ll some of these gaps, a number of new alge-
braic techniques are developed within the framework of geometric algebra. The process
of reformulation concentrates on the subjects of Grassmann calculus, Lie algebra theory,
spinor algebra and Lagrangian
eld theory. In each case it is argued that the geometric
algebra formulation is computationally more ecient than standard approaches, and that
it provides many novel insights. The new results obtained include a real approach to
relativistic multiparticle quantum mechanics, a new classical model for quantum spin-1/2
and an approach to gravity based on gauge
elds acting in a at spacetime. Through-
out, consistent use of geometric algebra is maintained and the bene
ts arising from this
approach are emphasised.
This thesis begins with a brief history of the development of geometric algebra and a
review of its present state. This leads, inevitably, to a discussion of the work of David
Hestenes 17]{34], who has done much to shape the modern form of the subject. A number
of the central themes running through his research are described, with particular emphasis
given to his ideas on mathematical design. Geometric algebra is then introduced, closely
following Hestenes' own approach to the subject. The central axioms and de
nitions
are presented, and a notation is introduced which is employed consistently throughout
this work. In order to avoid introducing too much formalism at once, the material in
this thesis has been split into two halves. The
rst half, Chapters 1 to 4, deals solely
with applications to various algebras employed in mathematical physics. Accordingly,
only the required algebraic concepts are introduced in Chapter 1. The second half of the
thesis deals with applications of geometric algebra to problems in mechanics and
eld
theory. The essential new concept required here is that of the dierential with respect to
variables de
ned in a geometric algebra. This topic is known as geometric calculus , and
is introduced in Chapter 5.
Chapters 2, 3 and 4 demonstrate how geometric algebra embraces a number of alge-
braic structures essential to modern mathematical physics. The
rst of these is Grass-
mann algebra, and particular attention is given to the Grassmann \calculus" introduced
by Berezin 35]. This is shown to have a simple formulation in terms of the properties
of non-orthonormal frames and examples are given of the algebraic advantages oered by
this new approach. Lie algebras and Lie groups are considered in Chapter 3. Lie groups
underpin many structures at the heart of modern particle physics, so it is important to
develop a framework for the study of their properties within geometric algebra. It is
shown that all (
nite dimensional) Lie algebras can be realised as bivector algebras and it
follows that all matrix Lie groups can be realised as spin groups. This has the interesting
consequence that every linear transformation can be represented as a monomial of (Clif-
ford) vectors. General methods for constructing bivector representations of Lie algebras
are given, and explicit constructions are found for a number of interesting cases.
The
nal algebraic structures studied are spinors. These are studied using the space-
time algebra | the (real) geometric algebra of Minkowski spacetime. Explicit maps are
constructed between Pauli and Dirac column spinors and spacetime multivectors, and
2
it is shown that the role of the scalar unit imaginary of quantum mechanics is played
by a
xed spacetime bivector. Changes of representation are discussed, and the Dirac
equation is presented in a form in which it can be analysed and solved without requiring
the construction of an explicit matrix representation. The concept of the multiparticle
spacetime algebra is then introduced and is used to construct both non-relativistic and
relativistic two-particle states. Some relativistic two-particle wave equations are consid-
ered and a new equation, based solely in the multiparticle spacetime algebra, is proposed.
In a
nal application, the multiparticle spacetime algebra is used to reformulate aspects
of the 2-spinor calculus developed by Penrose & Rindler 36, 37].
The second half of this thesis deals with applications of geometric calculus. The essen-
tial techniques are described in Chapter 5, which introduces the concept of the multivector
derivative 18, 24]. The multivector derivative is the natural extension of calculus for func-
tions mapping between geometric algebra elements (multivectors). Geometric calculus is
shown to be ideal for studying Lagrangian mechanics and two new ideas are developed |
multivector Lagrangians and multivector-parameterised transformations. These ideas are
illustrated by detailed application to two models for spinning point particles. The
rst,
due to Barut & Zanghi 38], models an electron by a classical spinor equation. This model
suers from a number of defects, including an incorrect prediction for the precession of
the spin axis in a magnetic
eld. An alternative model is proposed which removes many
of these defects and hints strongly that, at the classical level, spinors are the generators
of rotations. The second model is taken from pseudoclassical mechanics 39], and has the
interesting property that the Lagrangian is no longer a scalar but a bivector-valued func-
tion. The equations of motion are solved exactly and a number of conserved quantities
are derived.
Lagrangian
eld theory is considered in Chapter 6. A unifying framework for vectors,
tensors and spinors is developed and applied to problems in Maxwell and Dirac theory.
Of particular interest here is the construction of new conjugate currents in the Dirac
theory, based on continuous transformations of multivector spinors which have no simple
counterpart in the column spinor formalism. The chapter concludes with the development
of an extension of multivector calculus appropriate for multivector-valued linear functions.
The various techniques developed throughout this thesis are brought together in Chap-
ter 7, where a theory of gravity based on gauge transformations in a at spacetime is
presented. The motivation behind this approach is threefold: (1) to introduce gravity
through a similar route to the other interactions, (2) to eliminate passive transformations
and base physics solely in terms of active transformations and (3) to develop a theory
within the framework of the spacetime algebra. A number of consequences of this theory
are explored and are compared with the predictions of general relativity and spin-torsion
theories. One signi
cant consequence is the appearance of time-reversal asymmetry in
radially-symmetric (point source) solutions. Geometric algebra oers numerous advan-
tages over conventional tensor calculus, as is demonstrated by some remarkably compact
formulae for the Riemann tensor for various
eld con
gurations. Finally, it is suggested
that the consistent employment of geometric algebra opens up possibilities for a genuine
multiparticle theory of gravity.
3
1.1 Some History and Recent Developments
There can be few pieces of mathematics that have been re-discovered more often than
Cliord algebras 26]. The earliest steps towards what we now recognise as a geometric
algebra were taken by the pioneers of the use of complex numbers in physics. Wessel,
Argand and Gauss all realised the utility of complex numbers when studying 2-dimensional
problems and, in particular, they were aware that the exponential of an imaginary number
is a useful means of representing rotations. This is simply a special case of the more general
method for performing rotations in geometric algebra.
The next step was taken by Hamilton, whose attempts to generalise the complex num-
bers to three dimensions led him to his famous quaternion algebra (see 40] for a detailed
history of this subject). The quaternion algebra is the Cliord algebra of 2-dimensional
anti-Euclidean space, though the quaternions are better viewed as a subalgebra of the
Cliord algebra of 3-dimensional space. Hamilton's ideas exerted a strong inuence on
his contemporaries, as can be seen form the work of the two people whose names are most
closely associated with modern geometric algebra | Cliord and Grassmann.
Grassmann is best known for his algebra of extension. He de
ned hypernumbers
ei, which he identi
ed with unit directed line segments. An arbitrary vector was then
written as aiei, where the ai are scalar coecients. Two products were assigned to these
hypernumbers, an inner product
ei ej = ej ei = ij (1:1)
and an outer product
ei ^ ej = ;ej ^ ei: (1:2)
The result of the outer product was identi
ed as a directed plane segment and Grassmann
extended this concept to include higher-dimensional objects in arbitrary dimensions. A
fact overlooked by many historians of mathematics is that, in his later years, Grassmann
combined his interior and exterior products into a single, central product 41]. Thus he
wrote
ab = a b + a ^ b (1:3)
though he employed a dierent notation. The central product is precisely Cliord's prod-
uct of vectors, which Grassmann arrived at independently from (and slightly prior to)
Cliord. Grassmann's motivation for introducing this new product was to show that
Hamilton's quaternion algebra could be embedded within his own extension algebra. It
was through attempting to unify the quaternions and Grassmann's algebra into a sin-
gle mathematical system that Cliord was also led to his algebra. Indeed, the paper in
which Cliord introduced his algebra is entitled \Applications of Grassmann's extensive
algebra" 42].
Despite the eorts of these mathematicians to
nd a simple uni
ed geometric algebra
(Cliord's name for his algebra), physicists ultimately adopted a hybrid system, due
largely to Gibbs. Gibbs also introduced two products for vectors. His scalar (inner)
product was essentially that of Grassmann, and his vector (cross) product was abstracted
from the quaternions. The vector product of two vectors was a third, so his algebra
was closed and required no additional elements. Gibbs' algebra proved to be well suited
4
to problems in electromagnetism, and quickly became popular. This was despite the
clear de
ciencies of the vector product | it is not associative and cannot be generalised
to higher dimensions. Though special relativity was only a few years o, this lack of
generalisability did not appear to deter physicists and within a few years Gibbs' vector
algebra had become practically the exclusive language of vector analysis.
The end result of these events was that Cliord's algebra was lost amongst the wealth
of new algebras being created in the late 19th century 40]. Few realised its great promise
and, along with the quaternion algebra, it was relegated to the pages of pure algebra
texts. Twenty more years passed before Cliord algebras were re-discovered by Dirac in
his theory of the electron. Dirac arrived at a Cliord algebra through a very dierent
route to the mathematicians before him. He was attempting to
nd an operator whose
square was the Laplacian and he hit upon the matrix operator @ , where the -matrices
satisfy
+ = 2I : (1:4)
Sadly, the connection with vector geometry had been lost by this point, and ever since
the -matrices have been thought of as operating on an internal electron spin space.
There the subject remained, essentially, for a further 30 years. During the interim
period physicists adopted a wide number of new algebraic systems (coordinate geometry,
matrix algebra, tensor algebra, dierential forms, spinor calculus), whilst Cliord algebras
were thought to be solely the preserve of electron theory. Then, during the sixties, two
crucial developments dramatically altered the perspective. The
rst was made by Atiyah
and Singer 43], who realised the importance of Dirac's operator in studying manifolds
which admitted a global spin structure. This led them to their famous index theorems, and
opened new avenues in the subjects of geometry and topology. Ever since, Cliord algebras
have taken on an increasingly more fundamental role and a recent text proclaimed that
Cliord algebras \emerge repeatedly at the very core of an astonishing variety of problems
in geometry and topology" 15].
Whilst the impact of Atiyah's work was immediate, the second major step taken in
the sixties has been slower in coming to fruition. David Hestenes had an unusual training
as a physicist, having taken his bachelor's degree in philosophy. He has often stated that
this gave him a dierent perspective on the role of language in understanding 27]. Like
many theoretical physicists in the sixties, Hestenes worked on ways to incorporate larger
multiplets of particles into the known structures of
eld theory. During the course of these
investigations he was struck by the idea that the Dirac matrices could be interpreted as
vectors, and this led him to a number of new insights into the structure and meaning of
the Dirac equation and quantum mechanics in general 27].
The success of this idea led Hestenes to reconsider the wider applicability of Cliord
algebras. He realised that a Cliord algebra is no less than a system of directed numbers
and, as such, is the natural language in which to express a number of theorems and results
from algebra and geometry. Hestenes has spent many years developing Cliord algebra
into a complete language for physics, which he calls geometric algebra. The reason for
preferring this name is not only that it was Cliord's original choice, but also that it serves
to distinguish Hestenes' work from the strictly algebraic studies of many contemporary
texts.
During the course of this development, Hestenes identi
ed an issue which has been
5
coordinate geometry spinor calculus
complex analysis Grassmann algebra
vector analysis Berezin calculus
tensor analysis dierential forms
Lie algebras twistors
Cliord algebra
6
was not formulated as such. By adopting good design principles in the development of
mathematics, the bene
ts of these insights would be available to all. Some issues of what
constitutes good design are debated at various points in this introduction, though this
subject is only in its infancy.
In conclusion, the subject of geometric algebra is in a curious state. On the one
hand, the algebraic structures keeps reappearing in central ideas in physics, geometry
and topology, and most mathematicians are now aware of the importance of Cliord
algebras. On the other, there is far less support for Hestenes' contention that geometric
algebra, built on the framework of Cliord algebra, provides a uni
ed language for much
of modern mathematics. The work in this thesis is intended to oer support for Hestenes'
ideas.
7
4. The de
nition is completely useless for introducing geometric algebra to a physi-
cist or an engineer. It contains too many concepts that are the preserve of pure
mathematics.
Clearly, it is desirable to
nd an alternative axiomatic basis for geometric algebra which
does not share these de
ciencies. The axioms should be consistent with our ideas of what
constitutes good design. The above considerations lead us propose the following principle:
The axioms of an algebraic system should deal directly with the objects of
interest.
That is to say, the axioms should oer some intuitive feel of the properties of the system
they are de
ning.
The central properties of a geometric algebra are the grading, which separates objects
into dierent types, and the associative product between the elements of the algebra. With
these in mind, we adopt the following de
nition. A geometric algebra G is a graded linear
space, the elements of which are called multivectors. The grade-0 elements are called
scalars and are identi
ed with the
eld of real numbers (we will have no cause to consider
a geometric algebra over the complex
eld). The grade-1 elements are called vectors, and
can be thought of as directed line segments. The elements of G are de
ned to have an
addition, and each graded subspace is closed under this. A product is also de
ned which
is associative and distributive, though non-commutative (except for multiplication by a
scalar). The
nal axiom (which distinguishes a geometric algebra from other associative
algebras) is that the square of any vector is a scalar.
Given two vectors, a and b, we
nd that
(a + b)2 = (a + b)(a + b)
= a2 + (ab + ba) + b2: (1.7)
It follows that
ab + ba = (a + b)2 ; a2 ; b2 (1:8)
and hence that (ab + ba) is also a scalar. The geometric product of 2 vectors a b can
therefore be decomposed as
ab = a b + a ^ b (1:9)
where
a b 21 (ab + ba) (1:10)
is the standard scalar, or inner, product (a real scalar), and
a ^ b 12 (ab ; ba) (1:11)
is the antisymmetric outer product of two vectors, originally introduced by Grassmann.
The outer product of a and b anticommutes with both a and b,
a(a ^ b) = 21 (a2b ; aba)
= 12 (ba2 ; aba)
= ; 21 (ab ; ba)a
= ;(a ^ b)a (1.12)
8
so a ^ b cannot contain a scalar component. The axioms are also sucient to show that
a ^ b cannot contain a vector part. If we supposed that a ^ b contained a vector part c,
then the symmetrised product of a ^ b with c would necessarily contain a scalar part. But
c(a ^ b) + (a ^ b)c anticommutes with any vector d satisfying d a = d b = d c = 0, and
so cannot contain a scalar component. The result of the outer product of two vectors
is therefore a new object, which is dened to be grade-2 and is called a bivector. It can
be thought of as representing a directed plane segment containing the vectors a and b.
The bivectors form a linear space, though not all bivectors can be written as the exterior
product of two vectors.
The de
nition of the outer product is extended to give an inductive de
nition of the
grading for the entire algebra. The procedure is illustrated as follows. Introduce a third
vector c and write
c(a ^ b) = 12 c(ab ; ba)
= (a c)b ; (b c)a ; 12 (acb ; bca)
= 2(a c)b ; 2(b c)a + 12 (ab ; ba)c (1.13)
so that
c(a ^ b) ; (a ^ b)c = 2(a c)b ; 2(b c)a: (1:14)
The right-hand side of (1.14) is a vector, so one decomposes c(a ^ b) into
c(a ^ b) = c (a ^ b) + c ^ (a ^ b) (1:15)
where
c (a ^ b) 21 c(a ^ b) ; (a ^ b)c] (1:16)
and
c ^ (a ^ b) 12 c(a ^ b) + (a ^ b)c] : (1:17)
The de
nitions (1.16) and (1.17) extend the de
nitions of the inner and outer products
to the case where a vector is multiplying a bivector. Again, (1.17) results in a new object,
which is assigned grade-3 and is called a trivector. The axioms are sucient to prove that
the outer product of a vector with a bivector is associative:
c ^ (a ^ b) = 12 c(a ^ b) + (a ^ b)c]
= 41 cab ; cba + abc ; bac]
= 41 2(c ^ a)b + acb + abc + 2b(c ^ a) ; bca ; cba]
= 12 (c ^ a)b + b(c ^ a) + a(b c) ; (b c)a]
= (c ^ a) ^ b: (1.18)
The de
nitions of the inner and outer products are extended to the geometric product
of a vector with a grade-r multivector Ar as,
aAr = a Ar + a ^ Ar (1:19)
where the inner product
a Ar haArir;1 = 21 (aAr ; (;1)r Ar a) (1:20)
9
lowers the grade of Ar by one and the outer (exterior) product
a ^ Ar haArir+1 = 12 (aAr + (;1)r Ar a) (1:21)
raises the grade by one. We have used the notation hAir to denote the result of the
operation of taking the grade-r part of A (this is a projection operation). As a further
abbreviation we write the scalar (grade 0) part of A simply as hAi.
The entire multivector algebra can be built up by repeated multiplication of vectors.
Multivectors which contain elements of only one grade are termed homogeneous, and will
usually be written as Ar to show that A contains only a grade-r component. Homogeneous
multivectors which can be expressed purely as the outer product of a set of (independent)
vectors are termed blades.
The geometric product of two multivectors is (by de
nition) associative, and for two
homogeneous multivectors of grade r and s this product can be decomposed as follows:
ArBs = hAB ir+s + hAB ir+s;2 : : : + hAB ijr;sj : (1:22)
The \" and \^" symbols are retained for the lowest-grade and highest-grade terms of
this series, so that
Ar Bs hAB ijs;rj (1.23)
Ar ^ Bs hAB is+r (1.24)
which we call the interior and exterior products respectively. The exterior product is
associative, and satis
es the symmetry property
Ar ^ Bs = (;1)rs Bs ^ Ar: (1:25)
An important operation which can be performed on multivectors is reversion, which
reverses the order of vectors in any multivector. The result of reversing the multivector
A is written A~, and is called the reverse of A. The reverse of a vector is the vector itself,
and for a product of multivectors we have that
(AB )~= B~ A:
~ (1:26)
It can be checked that for homogeneous multivectors
A~r = (;1)r(r;1)=2Ar: (1:27)
It is useful to de
ne two further products from the geometric product. The
rst is the
scalar product
A B hAB i: (1:28)
This is commutative, and satis
es the useful cyclic-reordering property
hA : : : BC i = hCA : : :B i: (1:29)
In positive de
nite spaces the scalar product de
nes the modulus function
jAj (A A)1=2: (1:30)
10
The second new product is the commutator product, de
ned by
A B 12 (AB ; BA): (1:31)
The associativity of the geometric product ensures that the commutator product satis
es
the Jacobi identity
A (B C ) + B (C A) + C (A B ) = 0: (1:32)
Finally, we introduce an operator ordering convention. In the absence of brackets,
inner, outer and scalar products take precedence over geometric products. Thus a bc
means (a b)c and not a (bc). This convention helps to eliminate unruly numbers of
brackets. Summation convention is also used throughout this thesis.
One can now derive a vast number of properties of multivectors, as is done in Chapter 1
of 24]. But before proceeding, it is worthwhile stepping back and looking at the system
we have de
ned. In particular, we need to see that the axioms have produced a system
with sensible properties that match our intuitions about physical space and geometry in
general.
Calculations rarely have to be performed in this detail, but this exercise does serve to
illustrate how geometric algebras can be made intrinsic to a computer language. One can
even think of (1.46) as generalising Hamilton's concept of complex numbers as ordered
pairs of real numbers.
13
The square of the bivector 1^2 is ;1, so the even-grade elements z = x + y12 form
a natural subalgebra, equivalent to the complex numbers. Furthermore, 1 ^ 2 has the
geometric eect of rotating the vectors f1 2g in their own plane by 90 clockwise when
multiplying them on their left. It rotates vectors by 90 anticlockwise when multiplying
on their right. (This can be used to de
ne the orientation of 1 and 2).
The equivalence between the even subalgebra and complex numbers reveals a new
interpretation of the structure of the Argand diagram. From any vector r = x1 + y2 we
can form an even multivector z by
z 1r = x + Iy (1:47)
where
I 12: (1:48)
There is therefore a one-to-one correspondence between points in the Argand diagram
and vectors in two dimensions,
r = 1z (1:49)
where the vector 1 de
nes the real axis. Complex conjugation,
z z~ = r1 = x ; Iy (1:50)
now appears as the natural operation of reversion for the even multivector z. Taking the
complex conjugate of z results in a new vector r given by
r = 1z~
= (z1)~
= (1r1)~
= 1r1
= ;2r2: (1.51)
We will shortly see that equation (1.51) is the geometric algebra representation of a
reection in the 1 axis. This is precisely what one expects for complex conjugation.
This identi
cation of points on the Argand diagram with (Cliord) vectors gives ad-
ditional operational signi
cance to complex numbers of the form exp(i ). The even mul-
tivector equivalent of this is exp(I ), and applied to z gives
eI z = eI 1r
= 1e;I r: (1.52)
But we can now remove the 1, and work entirely in the (real) Euclidean plane. Thus
r0 = e;I r (1:53)
rotates the vector r anticlockwise through an angle . This can be veri
ed from the fact
that
e;I 1 = (cos ; sin I )1 = cos 1 + sin 2 (1:54)
14
and
e;I 2 = cos 2 ; sin 1: (1:55)
Viewed as even elements in the 2-dimensional geometric algebra, exponentials of \imag-
inaries" generate rotations of real vectors. Thinking of the unit imaginary as being a
directed plane segment removes much of the mystery behind the usage of complex num-
bers. Furthermore, exponentials of bivectors provide a very general method for handling
rotations in geometric algebra, as is shown in Chapter 3.
15
The algebra of 3-dimensional space is the Pauli algebra familiar from quantum me-
chanics. This can be seen by multiplying the pseudoscalar in turn by 3, 1 and 2 to
nd
(123)3 = 12 = i3
23 = i1 (1:60)
31 = i2
which is immediately identi
able as the algebra of Pauli spin matrices. But we have
arrived at this algebra from a totally dierent route, and the various elements in it have
very dierent meanings to those assigned in quantum mechanics. Since 3-dimensional
space is closest to our perception of the world, it is worth emphasising the geometry of
this algebra in greater detail. A general multivector M consists of the components
M = + a + ib + i
(1:61)
scalar vector bivector pseudoscalar
where a ak k and b bk k . The reason for writing spatial vectors in bold type is
to maintain a visible dierence between spatial vectors and spacetime 4-vectors. This
distinction will become clearer when we consider relativistic physics. The meaning of the
fk g is always unambiguous, so these are not written in bold type.
Each of the terms in (1.61) has a separate geometric signi
cance:
1. scalars are physical quantities with magnitude but no spatial extent. Examples are
mass, charge and the number of words in this thesis.
2. vectors have both a magnitude and a direction. Examples include relative positions,
displacements and velocities.
3. bivectors have a magnitude and an orientation. They do not have a shape. In Fig-
ure 1.1 the bivector a^b is represented as a parallelogram, but any other shape could
have been chosen. In many ways a circle is more appropriate, since it suggests the
idea of sweeping round from the a direction to the b direction. Examples of bivec-
tors include angular momentum and any other object that is usually represented as
an \axial" vector.
4. trivectors have simply a handedness and a magnitude. The handedness tells whether
the vectors in the product a^b^c form a left-handed or right-handed set. Examples
include the scalar triple product and, more generally, alternating tensors.
These four objects are represented pictorially in Figure 1.1. Further details and discussions
are contained in 25] and 44].
The space of even-grade elements of the Pauli algebra,
= + ib (1:62)
is closed under multiplication and forms a representation of the quarternion algebra.
Explicitly, identifying i, j , k with i1, ;i2, i3 respectively, the usual quarternion
relations are recovered, including the famous formula
i2 = j 2 = k2 = ijk = ;1: (1:63)
16
c
a a b a b
17
where
ak = a nn (1.68)
a? = nn ^ a: (1.69)
The result of reecting a in the hyperplane orthogonal to n is the vector a? ; ak, which
can be written as
a? ; ak = nn ^ a ; a nn
= ;n an ; n ^ an
= ;nan: (1.70)
This formula for a reection extends to arbitrary multivectors. For example, if the vectors
a and b are both reected in the hyperplane orthogonal to n, then the bivector a ^ b is
reected to
(;nan) ^ (;nbn) = 21 (nannbn ; nbnnan)
= na ^ bn: (1.71)
In three dimensions, the sign dierence between the formulae for vectors and bivectors
accounts for the dierent behaviour of \polar" and \axial" vectors under reections.
Rotations are built from pairs of reections. Taking a reection
rst in the hyperplane
orthogonal to n, and then in the hyperplane orthogonal to m, leads to the new vector
;m(;nan)m = mnanm
= RaR~ (1.72)
where
R mn: (1:73)
The multivector R is called a rotor. It contains only even-grade elements and satis
es the
identity
RR~ = RR~ = 1: (1:74)
Equation (1.74) ensures that the scalar product of two vectors is invariant under rotations,
(RaR~ ) (RbR~ ) = hRaRRb ~ R~ i
~ RR
= haRRb ~ i
= habi
= a b: (1.75)
As an example, consider rotating the unit vector a into another unit vector b, leaving
all vectors perpendicular to a and b unchanged. This is accomplished by a reection
perpendicular to the unit vector half-way between a and b (see Figure 1.2)
n (a + b)=ja + bj: (1:76)
18
b
-nan
19
is seen to be a pure vector by Taylor expanding in ,
b = a + B a + 2! B (B a) + :
2
(1:83)
The right-hand side of (1.83) is a vector since the inner product of a vector with a bivector
is always a vector (1.14). This method of representing rotations directly in terms of the
plane in which they take place is very powerful. Equations (1.54) and (1.55) illustrated this
in two dimensions, where the quantity exp(;I ) was seen to rotate vectors anticlockwise
through an angle . This works because in two dimensions we can always write
e;I=2reI=2 = e;I r: (1:84)
In higher dimensions the double-sided (bilinear) transformation law (1.78) is required.
This is much easier to use than a one-sided rotation matrix, because the latter becomes
more complicated as the number of dimensions increases. This becomes clearer in three
dimensions. The rotor
R exp(;ia=2) = cos(jaj=2) ; i jaaj sin(jaj=2) (1:85)
represents a rotation of jaj = (a2)1=2 radians about the axis along the direction of a.
This is already simpler to work with than 3 3 matrices. In fact, the representation of
a rotation by (1.85) is precisely how rotations are represented in the quaternion algebra,
which is well-known to be advantageous in three dimensions. In higher dimensions the
improvements are even more dramatic.
Having seen how individual rotors are used to represent rotations, we must look at
their composition law. Let the rotor R transform the unit vector a into a vector b,
b = RaR:~ (1:86)
Now rotate b into another vector b0, using a rotor R0. This requires
b0 = R0bR~ 0 = (R0R)a(R0 R)~ (1:87)
so that the transformation is characterised by
R 7! R0R (1:88)
which is the (left-sided) group combination rule for rotors. It is immediately clear that
the product of two rotors is a third rotor,
R0 R(R0R)~ = R0RR~ R~0 = R0R~0 = 1 (1:89)
so that the rotors do indeed form a (Lie) group.
The usefulness of rotors provides ample justi
cation for adding up terms of dierent
grades. The rotor R on its own has no geometric signi
cance, which is to say that no
meaning should be attached to the individual scalar, bivector, 4-vector : : : parts of R.
When R is written in the form R = eB=2, however, the bivector B has clear geometric
signi
cance, as does the vector formed from RaR~ . This illustrates a central feature of
geometric algebra, which is that both geometrically meaningful objects (vectors, planes
: : : ) and the elements that act on them (rotors, spinors : : : ) are represented in the same
algebra.
20
1.2.5 The Geometric Algebra of Spacetime
As a
nal example, we consider the geometric algebra of spacetime. This algebra is
suciently important to deserve its own name | spacetime algebra | which we will
usually abbreviate to STA. The square of a vector is no longer positive de
nite, and we
say that a vector x is timelike, lightlike or spacelike according to whether x2 > 0, x2 = 0
or x2 < 0 respectively. Spacetime consists of a single independent timelike direction, and
three independent spacelike directions. The spacetime algebra is then generated by a set
of orthonormal vectors f g, = 0 : : : 3, satisfying
= = diag(+ ; ; ;): (1:90)
(The signi
cance of the choice of metric signature will be discussed in Chapter 4.) The
full STA is 16-dimensional, and is spanned by the basis
1 f g fk ik g fi g i: (1:91)
The spacetime bivectors fk g, k = 1 : : : 3 are de
ned by
k k 0: (1:92)
They form an orthonormal frame of vectors in the space relative to the 0 direction. The
spacetime pseudoscalar i is de
ned by
i 0123 (1:93)
and, since we are in a space of even dimension, i anticommutes with all odd-grade elements
and commutes with all even-grade elements. It follows from (1.92) that
123 = 102030 = 0123 = i: (1:94)
The following geometric signi
cance is attached to these relations. An inertial system
is completely characterised by a future-pointing timelike (unit) vector. We take this to
be the 0 direction. This vector/observer determines a map between spacetime vectors
a = a and the even subalgebra of the full STA via
a0 = a0 + a (1:95)
where
a0 = a 0 (1.96)
a = a ^ 0: (1.97)
The even subalgebra of the STA is isomorphic to the Pauli algebra of space de
ned in
Section 1.2.3. This is seen from the fact that the k = k 0 all square to +1,
k 2 = k 0k 0 = ;k k 00 = +1 (1:98)
and anticommute,
j k = j 0k 0 = k j 00 = ;k 0j 0 = ;k j (j 6= k): (1:99)
21
There is more to this equivalence than simply a mathematical isomorphism. The way we
think of a vector is as a line segment existing for a period of time. It is therefore sensible
that what we perceive as a vector should be represented by a spacetime bivector. In this
way the algebraic properties of space are determined by those of spacetime.
As an example, if x is the spacetime (four)-vector specifying the position of some point
or event, then the \spacetime split" into the 0-frame gives
x0 = t + x (1:100)
which de
nes an observer time
t = x 0 (1:101)
and a relative position vector
x = x ^ 0: (1:102)
One useful feature of this approach is the way in which it handles Lorentz-scalar quantities.
The scalar x2 can be decomposed into
x2 = x00x
= (t + x)(t ; x)
= t2 ; x2 (1.103)
which must also be a scalar. The quantity t2 ; x2 is now seen to be automatically
Lorentz-invariant, without needing to consider a Lorentz transformation.
The split of the six spacetime bivectors into relative vectors and relative bivectors is
a frame/observer-dependent operation. This can be illustrated with the Faraday bivector
F = 21 F ^ , which is a full, 6-component spacetime bivector. The spacetime split
of F into the 0-system is achieved by separating F into parts which anticommute and
commute with 0. Thus
F = E + iB (1:104)
where
E = 21 (F ; 0F0) (1.105)
iB = 2 (F + 0F0):
1 (1.106)
Here, both E and B are spatial vectors, and iB is a spatial bivector. This decomposes F
into separate electric and magnetic
elds, and the explicit appearance of 0 in the formulae
for E and B shows that this split is observer-dependent. In fact, the identi
cation
of spatial vectors with spacetime bivectors has always been implicit in the physics of
electromagnetism through formulae like Ek = Fk0.
The decomposition (1.104) is useful for constructing relativistic invariants from the E
and B
elds. Since F 2 contains only scalar and pseudoscalar parts, the quantity
F 2 = (E + iB)(E + iB)
= E 2 ; B 2 + 2iE B (1.107)
is Lorentz-invariant. It follows that both E 2 ; B 2 and E B are observer-invariant
quantities.
22
Equation (1.94) is an important geometric identity, which shows that relative space
and spacetime share the same pseudoscalar i. It also exposes the weakness of the matrix-
based approach to Cliord algebras. The relation
123 = i = 0123 (1:108)
cannot be formulated in conventional matrix terms, since it would need to relate the
2 2 Pauli matrices to 4 4 Dirac matrices. Whilst we borrow the symbols for the
Dirac and Pauli matrices, it must be kept in mind that the symbols are being used in
a quite dierent context | they represent a frame of orthonormal vectors rather than
representing individual components of a single isospace vector.
The identi
cation of relative space with the even subalgebra of the STA necessitates
developing a set of conventions which articulate smoothly between the two algebras. This
problem will be dealt with in more detail in Chapter 4, though one convention has already
been introduced. Relative (or spatial) vectors in the 0-system are written in bold type to
record the fact that in the STA they are actually bivectors. This distinguishes them from
spacetime vectors, which are left in normal type. No problems can arise for the fk g,
which are unambiguously spacetime bivectors, so these are also left in normal type. The
STA will be returned to in Chapter 4 and will then be used throughout the remainder of
this thesis. We will encounter many further examples of its utility and power.
23
where B is the plane(s) of rotation. The outermorphism extension of this is simply
R(A) = eB=2Ae;B=2: (1:111)
An important property of the outermorphism is that the outermorphism of the product
of two functions in the product of the outermorphisms,
f g(a)] ^ f g(b)] : : :^ f g(c)] = f g(a) ^ g(b) : : :^ g(c)]
= f g(a ^ b ^ : : :^ c)]: (1.112)
To ease notation, the product of two functions will be written simply as f g(A), so that
(1.112) becomes
fg(a) ^ fg(b) : : : ^ fg(c) = f g(a ^ b ^ : : :^ c): (1:113)
The pseudoscalar of an algebra is unique up to a scale factor, and this is used to de
ne
the determinant of a linear function via
det(f ) f (I )I ;1 (1:114)
so that
f (I ) = det(f )I: (1:115)
This de
nition clearly illustrates the role of the determinant as the volume scale factor.
The de
nition also serves to give a very quick proof of one of the most important properties
of determinants. It follows from (1.113) that
f g(I ) = f (det(g)I )
= det(g)f (I )
= det(f ) det(g)I (1.116)
and hence that
det(fg) = det(f ) det(g): (1:117)
This proof of the product rule for determinants illustrates our third (and
nal) principle
of good design:
Denitions should be chosen so that the most important theorems can be proven
most economically.
The de
nition of the determinant clearly satis
es this criteria. Indeed, it is not hard to
see that all of the main properties of determinants follow quickly from (1.115).
The adjoint to f , written as f , is de
ned by
f (a) eihf (ei )ai (1:118)
where feig is an arbitrary frame of vectors, with reciprocal frame feig. A frame-invariant
de
nition of the adjoint can be given using the vector derivative, but we have chosen not
to introduce multivector calculus until Chapter 5. The de
nition (1.118) ensures that
b f (a) = a (b eif (ei))
= a f (b): (1.119)
24
A symmetric function is one for which f = f .
The adjoint also extends via outermorphism and we
nd that, for example,
f (a ^ b) = f (a) ^ f (b)
= ei ^ ej a f (ei)b f (ej )
1 ei ^ ej a f (e )b f (e ) ; a f (e )b f (e )
= 2 i j j i
= 1 ei ^ ej (a ^ b) f (e ^ e ): (1.120)
2 j i
By using the same argument as in equation (1.119), it follows that
hf (A)B i = hAf (B )i (1:121)
for all multivectors A and B . An immediate consequence is that
det f = hI ;1f (I )i
= hf (I ;1)I i
= det f: (1.122)
Equation (1.121) turns out to be a special case of the more general formulae,
Ar f (Bs ) = f f (Ar) Bs] rs (1:123)
f (Ar ) Bs = f Ar f (Bs)] r s
which are derived in 24, Chapter 3].
As an example of the use of (1.123) we
nd that
f (f (AI )I ;1) = AIf (I ;1) = A det f (1:124)
which is used to construct the inverse functions,
f ;1(A) = det(f );1 f (AI )I ;1 (1:125)
f ;1(A) = det(f );1 I ;1f (IA):
These equations show how the inverse function is constructed from a double-duality op-
eration. They are also considerably more compact and ecient than any matrix-based
formula for the inverse.
Finally, the concept of an eigenvector is generalized to that of an eigenblade Ar, which
is an r-grade blade satisfying
f (Ar ) = Ar (1:126)
where is a real eigenvalue. Complex eigenvalues are in general not considered, since
these usually loose some important aspect of the geometry of the function f . As an
example, consider a function f satisfying
f (a) = b (1:127)
f (b) = ;a
25
for some pair of vectors a and b. Conventionally, one might write
f (a + jb) = ;j (a + jb) (1:128)
and say that a + bj is an eigenvector with eigenvalue ;j . But in geometric algebra one
can instead write
f (a ^ b) = b ^ (;a) = a ^ b (1:129)
which shows that a ^ b is an eigenblade with eigenvalue +1. This is a geometrically more
useful result, since it shows that the a^b plane is an invariant plane of f . The unit blade
in this plane generates its own complex structure, which is the more appropriate object
for considering the properties of f .
28
Chapter 2
Grassmann Algebra and Berezin
Calculus
This chapter outlines the basis of a translation between Grassmann calculus and geo-
metric algebra. It is shown that geometric algebra is sucient to formulate all of the
required concepts, thus integrating them into a single unifying framework. The transla-
tion is illustrated with two examples, the \Grauss integral" and the \Grassmann Fourier
transform". The latter demonstrates the full potential of the geometric algebra approach.
The chapter concludes with a discussion of some further developments and applications.
Some of the results presented in this chapter
rst appeared in the paper \Grassmann
calculus, pseudoclassical mechanics and geometric algebra" 1].
nd applications in other
elds. As a further comment, this translation also makes it clear
why no measure is associated with Grassmann integrals: nothing is being added up!
where the convention is adopted that terms where Cj ^ H or C~j ^ G have grade less than
n do not contribute. Since G and H only contain terms purely constructed from the fekg
and ff k g respectively, (2.57) can be written as
Xn
G(e) = (Cn;j ^hH (f )ij ) Fn
j =0
X n
H (f ) = (;1)j (hG(e)ij ^ Cn;j ) E n: (2.58)
j =0
So far we have only derived a formula analogous to (2.44), but we can now go much
further. By using
eJ = cosn + cosn;1 sin C1 + : : : + sinn I (2:59)
to decompose eJ (+=2) = eJ I in two ways, it can be seen that
Cn;r = (;1)r Cr I = (;1)r ICr (2:60)
and hence (using some simple duality relations) (2.58) become
Xn
G(e) = Cj Hj En
j =0
X n
H (f ) = (;1) Gj Cj F n:
n (2.61)
j =0
Finally, since G and H are pure in the fek g and ff k g respectively, the eect of dotting
with Ck is simply to interchange each ek for an ;fk and each fk for an ek . For vectors
36
this is achieved by dotting with J . But, from (2.53), the same result is achieved by a
rotation through =2 in the planes of J . Rotations extend simply via outermorphism, so
we can now write
Cj Hj = eJ=4Hj e;J=4
Gj Cj = e;J=4Gj eJ=4: (2.62)
We thus arrive at the following equivalent expressions for (2.57):
G(e) = eJ=4H (f )e;J=4En
H (f ) = (;1)ne;J=4 G(e)eJ=4F n: (2.63)
The Grassmann Fourier transformation has now been reduced to a rotation through =2
in the planes speci
ed by J , followed by a duality transformation. Proving the \inversion"
theorem (i.e. that the above expressions are consistent) amounts to no more than carrying
out a rotation, followed by its inverse,
G(e) = eJ=4 (;1)ne;J=4 G(e)eJ=4F n e;J=4 En
= G(e)E nEn = G(e): (2.64)
This proof is considerably simpler than any that can be carried out in the more restrictive
system of Grassmann algebra.
37
satisfy the Cliord algebra generating relations
fQ^ j Q^ k g = 2jk (2:66)
and this has been used by Sherry to provide an alternative approach to quantizing a
Grassmann system 55, 56]. The geometric algebra formalism oers a novel insight into
these relations. By utilising the
ducial tensor, we can write
Q^ k a( ) $ ek ^ A + ek A = h(k ) ^ A + h;1 (k ) A
= h(k ^ h;1(A)) + h(sk h;1(A))
= hk h;1(A)] (2.67)
where A is the multivector equivalent of a( ) and we have used (1.123). The operator Q^ k
thus becomes an orthogonal Cliord vector (now Cliord multiplied), sandwiched between
a symmetric distortion and its inverse. It is now simple to see that
fQ^ j Q^ k ga( ) $ h(2j k h;1(A)) = 2jk A: (2:68)
The above is an example of the ubiquity of the
ducial tensor in applications involving
non-orthonormal frames. In this regard it is quite surprising that the
ducial tensor is
not more prominent in standard expositions of linear algebra.
Berezin 35] de
nes dual operators to the Q^ k by
P^k = ;j (k ; @@ ) (2:69)
k
though a more useful structure is derived by dropping the j , and de
ning
P^k = k ; @@ : (2:70)
k
These satisfy
fP^j P^k g = ;2jk (2:71)
and
fP^j Q^ kg = 0 (2:72)
so that the P^k Q^ k span a 2n-dimensional balanced algebra (signature n n). The P^k can be
translated in the same manner as the Q^ k , this time giving (for a homogeneous multivector)
P^k a( ) $ ek ^ Ar ; ek Ar = (;1)r hh;1(Ar )k ]: (2:73)
The fk g frame now sits to the right of the multivector on which it operates. The factor
of (;1)r accounts for the minus sign in (2.71) and for the fact that the left and right
multiples anticommute in (2.72). The Q^ k and P^k can both be given right analogues
if desired, though this does not add anything new. The fQ^ kg and fP^k g operators are
discussed more fully in Chapter 4, where they are related to the theory of the general
linear group.
38
Chapter 3
Lie Groups and Spin Groups
This chapter demonstrates how geometric algebra provides a natural arena for the study of
Lie algebras and Lie groups. In particular, it is shown that every matrix Lie group can be
realised as a spin group. Spin groups consist of even products of unit magnitude vectors,
and arise naturally from the geometric algebra treatment of reections and rotations
(introduced in Section 1.2.4). The generators of a spin group are bivectors, and it is
shown that every Lie algebra can be represented by a bivector algebra. This brings the
computational power of geometric algebra to applications involving Lie groups and Lie
algebras. An advantage of this approach is that, since the rotors and bivectors are all
elements of the same algebra, the discussion can move freely between the group and its
algebra. The spin version of the general linear group is studied in detail, revealing some
novel links with the structures of Grassmann algebra studied in Chapter 2. An interesting
result that emerges from this work is that every linear transformation can be represented
as a (geometric) product of vectors. Some applications of this result are discussed. A
number of the ideas developed in this chapter appeared in the paper \Lie groups as spin
groups" 2].
Throughout this chapter, the geometric algebra generated by p independent vectors
of positive norm and q of negative norm is denoted as <pq. The grade-k subspace of
this algebra is written as <kpq and the space of vectors, <1pq , is abbreviated to <pq. The
Euclidean algebra <n0 is abbreviated to <n, and the vector space <1n is written as <n .
Lie groups and their algebras are labeled according to the conventions of J.F. Cornwell's
\Group Theory in Physics", Vol. 2 57]. (A useful table of these conventions is found on
page 392).
39
where we have recalled equation (1.70). (A convenient feature of the underbar/overbar
notation for linear functions is that a function can be written in terms of the multivector
that determines it.) The function n satis
es
n(a) n(b) = hnannbni = a b (3:2)
and so preserves the inner product. On combining n with a second reection m, where
m(a) = ;mam (3:3)
the function
m n(a) = mnanm (3:4)
is obtained. This function also preserves inner products, and in Section 1.2.4 was identi
ed
as a rotation in the m ^ n plane. The group of even products of unit vectors is denoted
spin(n). It consists of all even multivectors (rotors) satisfying
RR~ = 1 (3:5)
and such that the quantity RaR~ is a vector for all vectors a. The double-sided action of
a rotor R on a vector a is written as
R(a) = RaR~ (3:6)
and the R form the group of rotations on <n , denoted SO(n). The rotors aord a spin-1/2
description of rotations, hence rotor groups are referred to as spin groups.
In spaces with mixed signature the situation is slightly more complicated. In order to
take care of the fact that a unit vector can now have n2 = 1, equation (3.1) must be
modi
ed to
n(a) = ;nan;1: (3:7)
Taking even combinations of reections now leads to functions of the type
M (a) = MaM ;1 (3:8)
as opposed to MaM~ . Again, the spin group spin(p q) is de
ned as the group of even
products of unit vectors, but its elements now satisfy M M~ = 1. The term \rotor"
is retained for elements of spin(p q) satisfying RR~ = 1. The subgroup of spin(p q)
containing just the rotors is called the rotor group (this is sometimes written as spin+ (p q)
in the literature). The action of a rotor on a vector a is always de
ned by (3.6). Spin
groups and rotor groups are both Lie groups and, in a space with mixed signature, the spin
group diers from the rotor group only by a direct product with an additional subgroup
of discrete transformations.
The generators of a spin group are found by adapting the techniques found in any of
the standard texts of Lie group theory (see 57], for example). We are only interested
in the subgroup of elements connected to the identity, so only need to consider the rotor
group. We introduce a one-parameter set of rotors R(t), so that
R(t)aR~ (t) = hR(t)aR~ (t)i1 (3:9)
40
for all vectors a and for all values of the parameter t. On dierentiating with respect to
t, we
nd that the quantity
R0aR~ + RaR~0 = R0R~ (RaR~ ) + (RaR~ )RR~0
= R0R~ (RaR~ ) ; (RaR~)R0 R~ (3.10)
must be a vector, where we have used RR~ = 1 to deduce that
R0R~ = ;RR~0 : (3:11)
The commutator of R0 R~ with an arbitrary vector therefore results in a vector, so R0R~ can
only contain a bivector part. (R0R~ cannot contain a scalar part, since (R0R~ )~ = ;R0R~ .)
The generators of a rotor group are therefore a set of bivectors in the algebra containing
the rotors.
A simple application of the Jacobi identity gives, for vectors a, b, c, and d,
(a ^ b) (c ^ d) = (a ^ b) c] ^ d ; (a ^ b) d] ^ c (3:12)
so the commutator product of two bivector blades results in a third bivector. It follows
that the space of bivectors is closed under the commutator product, and hence that the
bivectors (together with the commutator product) form the Lie algebra of a spin group.
It should be noted that the commutator product, , in equation (3.12) diers from the
commutator bracket by a factor of 1=2. The commutator product is simpler to use, since
it is the bivector part of the full geometric product of two bivectors A and B :
AB = A B + A B + A ^ B (3:13)
where
A B + A ^ B = 21 (AB + BA) (3.14)
A B = 12 (AB ; BA): (3.15)
For this reason the commutator product will be used throughout this chapter.
Since the Lie algebra of a spin group is generated by the bivectors, it follows that all
rotors simply connected to the identity can be written in the form
R = eB=2 (3:16)
which ensures that
R~ = e;B=2 = R;1: (3:17)
The form of a rotor given by equation (3.16) was found in Section 1.2.4, where rotations
in a single Euclidean plane were considered. The factor of 1=2 is included because rotors
provide a half-angle description of rotations. In terms of the Lie algebra, the factor of
1=2 is absorbed into our use of the commutator product, as opposed to the commutator
bracket.
41
It can be shown that, in positive de
nite spaces, all rotors can be written in the form
of (3.16). The bivector B is not necessarily unique, however, as can be seen by considering
the power series de
nition of the logarithm,
H 3 H5
ln X = 2H + 3 + 5 + ] (3:18)
where
H=X X ; 1: (3:19)
+1
It is implicit in this formula that 1 + X is invertible, and the logarithm will not be well-
de
ned if this is not the case. For example, the pseudoscalar I in <40 is a rotor (I I~ = 1),
the geometric eect of which is to reverse the sign of all vectors. But 1+ I is not invertible,
since (1 + I )2 = 2(1 + I ). This manifests itself as a non-uniqueness in the logarithm of I
| given any bivector blade B satisfying B 2 = ;1, I can be written as
I = expfB (1 ; I ) 2 g: (3:20)
Further problems can arise in spaces with mixed signature. In the spacetime algebra, for
example, whilst the rotor
R = (0 + 1 ; 2)2 = 1 + (0 + 1)2 (3:21)
can be written as
R = expf(0 + 1)2g (3:22)
the rotor
; R = expf12 2 gR = ;1 ; (0 + 1)2 (3:23)
cannot be written as the exponential of a bivector. The problem here is that the series
for ln(;X ) is found by replacing H by H ;1 in equation (3.18) and, whilst 1 + R =
2 + (0 + 1)2 is invertible, 1 ; R = ;(0 + 1)2 is null and therefore not invertible.
Further examples of rotors with no logarithm can be constructed in spaces with other
signatures. Near the identity, however, the Baker-Campbell-Hausdor formula ensures
that, for suitably small bivectors, one can always write
eA=2eB=2 = eC=2: (3:24)
So, as is usual in Lie group theory, the bulk of the properties of the rotor (and spin)
groups are transferred to the properties of their bivector generators.
In the study of Lie groups and their algebras, the adjoint representation plays a partic-
ularly important role. The adjoint representation of a spin group is formed from functions
mapping the Lie algebra to itself,
AdM (B ) MBM ;1 = M (B ): (3:25)
The adjoint representation is therefore formed by the outermorphism action of the linear
functions M (a) = MaM ;1 . For the rotor subgroup, we have
AdR(B ) = R(B ) = RB R: ~ (3:26)
42
It is immediately seen that the adjoint representation satis
es
AdM1 AdM2 (B )] = AdM1M2 (B ): (3:27)
The adjoint representation of the Lie group induces a representation of the Lie algebra as
adA=2(B ) = A B (3:28)
or
adA (B ) = 2A B: (3:29)
The Jacobi identity ensures that
2 (adA adB ; adB adA )(C ) = 2A (B C ) ; B (A C )]
1
= 2(A B ) C
= adAB (C ): (3.30)
The Killing form is constructed by considering adA as a linear operator on the space of
bivectors, and de
ning
K (A B ) = Tr(adAadB ): (3:31)
For the case where the Lie algebra is the set of all bivectors, we can de
ne a basis set of
bivectors as BK = ei ^ ej (i < j ) with reciprocal basis B K = ej ^ ei . Here, the index K
is a simplicial index running from 1 to n(n ; 1)=2 over all combinations of i and j with
i < j . A matrix form of the adjoint representation is now given by
(adA)K J = 2(A BJ ) B K (3:32)
so that the Killing form becomes
n(nX
;1)=2
K (A B ) = 4 (A BJ ) B K (B BK ) B J
JK =1
= 2A (B (ei ^ ej ))] (ej ^ ei)
= hABei ^ ej ej ^ ei ; Aei ^ ej Bej ^ eii (3.33)
Now,
ei ^ ej ej ^ ei = eiej ej ^ ei
= n(n ; 1) (3.34)
and
ei ^ ej Bej ^ ei = eiej Bej ^ ei
= eiej Bej ei ; eiej ei ej B
= (n ; 4)2 ; n]B (3.35)
where we have used equations (1.139). On recombining (3.34) and (3.35), the Killing form
on a bivector algebra becomes
K (A B ) = 8(n ; 2)hAB i (3:36)
43
Eij = ei ^ ej (i < j i j = 1 : : : p)
Fij = fi ^ fj (i < j i j = 1 : : : q)
Gij = ei ^ fj (i = 1 : : : p, j = 1 : : : q):
46
unitary group. All orthogonal transformations leave a b invariant, but only a subset will
leave (a ^ b) J invariant as well. These transformations must satisfy
f (a) ^ f (b) J = (a ^ b) f (J ) = (a ^ b) J (3:59)
for all vectors a and b. The invariance group therefore consists of all orthogonal transfor-
mations whose outermorphism satis
es
f (J ) = J: (3:60)
This requirement excludes all discrete transformations, since a vector n will only generate
a symmetry if
n(J ) = nJn;1 = J
) n J = 0 (3.61)
and no such vector n exists. It follows that the symmetry group is constructed entirely
from the double sided action of the elements of the spin group which satisfy
MJ = JM: (3:62)
These elements aord a spin group representation of the unitary group.
Equation (3.62) requires that, for a rotor R simply connected to the identity, the
bivector generator of R commutes with J . The Lie algebra of a unitary group is therefore
realised by the set of bivectors commuting with J , which we have seen are also eigen-
bivectors of J . Given an arbitrary bivector B , therefore, the bivector
BJ = B + J (B ) (3:63)
is contained in the bivector algebra of u(p,q). This provides a quick method for writing
down a basis set of generators. It is convenient at this point to introduce an orthonormal
frame of vectors fei fig satisfying
ei ej = fi fj = ij (3.64)
ei fj = 0 (3.65)
where ij = ijk (no sum) and i is the metric indicator (= 1 or ;1). This frame is used
to write down a basis set of generators which are orthogonal with respect to the Killing
form. Such a basis for u(p,q) is contained in Table 3.2. This basis has dimension
2 n(n ; 1) + 2 n(n ; 1) + n = n : (3:66)
1 1 2
Of these, p2 + q2 bivectors have negative norm, and 2pq have positive norm.
The algebra of Table 3.2 contains the bivector J , which commutes with all other
elements of the algebra and generates a U(1) subgroup. This is factored out to give the
basis for su(p,q) contained in Table 3.3. The Hi are written in the form given to take care
of the metric signature of the vector space. When working in <2n one can simply write
Hi = Ji ; Ji+1: (3:67)
47
Eij = eiej + fi fj (i < j = 1 : : : n)
Fij = eifj ; fiej 00
Ji = eifi (i = 1 : : : n):
50
Eij = eiej ; e^ie^j (i < j = 1 : : : n)
Fij = eie^j ; e^iej 00
Ji = eie^i (i = 1 : : : n):
52
where M is a member of the spin group spin(n n) which commutes with K . We start by
considering the polar decomposition of an arbitrary matrix M . Assuming that det M 6= 0,
the matrix M M ' can be written (not necessarliy uniquely) as
M M' = S S' (3:106)
where S is an orthogonal transformation (which can be arranged to be a rotation matrix),
and is a diagonal matrix with positive entries. One can now write
M = S 1=2R (3:107)
where 1=2 is the diagonal matrix of positive square roots of the entries of and R is a
matrix de
ned by
R = ;1=2S' M : (3:108)
The matrix R satis
es
RR' = ;1=2S' M MS ' ;1=2
= ;1=2 ;1=2
= I (3.109)
and so is also orthogonal. It follows from (3.107) that an arbitrary non-singular matrix
can be written as a diagonal matrix with positive entries sandwiched between a pair of
orthogonal matrices. As a check, this gives n2 degrees of freedom. To prove the desired
result, we need only show that orthogonal transformations and positive dilations can be
written in the form of equation (3.105).
We
rst consider rotations. The Eij generators in Table 3.4 produce rotors of the form
R = expf(E ; E^ )=2g (3:110)
where
E = ij Eij (3:111)
and the ij are a set of scalar coecients. The eect of the rotor R on a+ generates
R(a+) = R(a + ^a)R~
= RaR~ + (RaR~ ) K
= eE=2ae;E=2 + (eE=2ae;E=2) K (3.112)
and so accounts for all rotations of the vector a in <n. To complete the set of orthogonal
transformations, a representation for reections must be found as well. A reection in
the hyperplane orthogonal to the vector n in <n is represented by the element nn^ in <nn .
Since nn^ n^ n = ;1, nn^ is not a rotor and belongs to the disconnected part of spin(n n).
That nn^ commutes with K , and so is contained in spin(n n), is veri
ed as follows,
nn^ K = 2nn^ K + nK n^
= 2n2 + 2n K n^ + Knn^
= 2(n2 + n^ 2) + Knn^
= Knn^ : (3.113)
53
The action of nn^ on a vector is determined by (3.8), and gives
nn^ a+ nn^ = ;nn^ an^n ; (nn^ an^n) K
= ;nan ; (nan) K (3.114)
as required.
Finally, we need to see how positive dilations are given a rotor description. A dilation
in the n direction by an amount e is generated by the rotor
R = e;nn^=2 (3:115)
where the generator ;nn^ =2 is built from the Ki in Table 3.4. Acting on the null vector
n+ = n + n^ , the rotor (3.115) gives
Rn+ R~ = e;nn^=2n+ e+nn^=2
= e;nn^ (n + n^)
= (cosh ; nn^ sinh )(n + n^ )
= (cosh + sinh )(n + n^ )
= en+: (3.116)
In addition, for vectors perpendicular to n in <n, the action of R on their null vector
equivalents has no eect. These are precisely the required properties for a dilation in
the n direction. This concludes the proof that the general linear group is represented by
the subgroup of spin(n n) consisting of elements commuting with K . As an aside, this
construction has led us to the Eij and Ki generators in Table (3.4). Commutators of the
Eij and Ki give the remaining Fij generators, which are sucient to close the algebra.
The determinant of a linear function on <n is easily represented in <nn since
f (e1) ^ f (e2) ^ : : :^ f (en) = det f En (3:117)
becomes
MWn M ;1 = det fWn (3:118)
in the null space of V n. Here M is the spin group element representing the linear function
f . From the de
nitions of Wn and Wn (3.102), we can write
det f = 2n hW~ n MWn M ;1i (3:119)
from which many of the standard properties of determinants can be derived.
55
A family of commuting idempotents are now de
ned by
Ii 21 (1 + Ki) = wi wi (3:130)
and have the following properties:
Ii2 = Ii (3.131)
IiIj = Ij Ii (3.132)
eiIi = wi = e^iIi (3.133)
Iiei = wi = ;Iie^i (3.134)
KiIi = Ii: (3.135)
From the Ii the idempotent I is de
ned by
Yn
I Ii = I1I2 : : :In = w1 w1w2 w2 : : : wnwn = Wn W~ n: (3:136)
i=1
I has the following properties:
I2 = I (3.137)
eiI = e^iI (3.138)
and
EnI = E^nI = Wn I = Wn (3:139)
where En is the pseudoscalar for the Euclidean algebra <n and E^n is the pseudoscalar
for the anti-Euclidean algebra <0n. The relationships in (3.139) establish an equivalence
between the <n, <0n and V n vector spaces.
Whilst the construction of I has made use of an orthonormal frame, the form of I is
actually independent of this choice. This can be seen by writing I in the form
1 K ^ K K ^ K ^ : : : ^ K
I = 2n 1 + K + 2! + : : : + n! (3:140)
and recalling that K is frame-independent. It is interesting to note that the bracketed term
in (3.140) is of the same form as the Grassmann exponential considered in Section 2.2.1.
The full 2n -dimensional space I n is generated by left multiplication of I by the entire
algebra <nn ,
I n = <nnI: (3:141)
Since multiplication of I by ei and e^i are equivalent, every occurrence of an e^i in a multi-
vector in <nn can be replaced by an ei, so that there is a simple 1 $ 1 equivalence between
elements of <n and I n. The action of an element of end(<n) can now be represented in
<nn by left multiplication of I n by the appropriate multivector. For a multivector Ar in
<n the equivalence between the basic operators (3.122) is seen from
eiArI $ eiAr (3:142)
56
and
e^iArI $ A^r ei: (3:143)
The parity operation on the right-hand side of (3.143) arises because the e^i vector must
be anticommuted through each of the vectors making up the Ar multivector. This is
the reason for the dierent uses of the overhat notation for the <n and <nn algebras.
Symbolically, we can now write
eiI n $ ei<n (3.144)
e^iI $ <^ nei:
n (3.145)
Also, from the de
nitions of wi and wi (3.99), we
nd the equivalences
wiI n $ ei ^<n (3.146)
wi I $ ei <n
n (3.147)
which establishes contact with the formalism of Grassmann/Berezin calculus given in
Chapter 2. We can now move easily between the formalism with dot and wedge products
used in Chapter 2 and the null-vector formalism adopted here. The chosen application
should dictate which is the more useful.
We next consider the quantity nn^ , where n is a unit vector. The action of this on I n
gives
nn^ I n $ n<^ nn: (3:148)
The operation on the right-hand side is the outermorphism action of a reection in the
hyperplane perpendicular to n. In the previous section we used a double-sided application
of nn^ on null vectors to represent reections in <n . We now see that the same object can
be applied single-sidedly in conjunction with the idempotent I to also produce reections.
The same is true of products of reections. For example, the rotor (3.110) gives
e(E;E^)=2MI = eE=2Me;E= ^ 2
I $ eE=2Me;E=2 (3:149)
demonstrating how the two-bladed structure of the Eij generators is used to represent
concurrent left and right multiplication in <n .
The operation <n 7! <^ n is performed by successive reections in each of the ei direc-
tions. We therefore
nd the equivalence
e1e^1e2e^2 : : : ene^nI n $ <^ n : (3:150)
But
e1e^1e2e^2 : : : ene^n = en : : : e2e1e^1e^2 : : : e^n = E~nE^n = Enn (3:151)
is the unit pseudoscalar in <nn , so multiplication of an element of I n by Enn corresponds
to the parity operation in <n. As a check, (Enn )2 is always +1, so the result of two parity
operations is always the identity.
The correspondence between the single-sided and double-sided forms for a dilation are
not quite so simple. If we consider the rotor expf;nn^ =2g again, we
nd that, for the
vector n,
e;nn^=2nI = e=2nI $ e=2n (3:152)
57
For vectors perpendicular to n, however, we
nd that
e;nn^=2n? I = n?e;=2nn^ I $ e;=2n? (3:153)
so the single-sided formulation gives a stretch along the n direction of expfg, but now
combined with an overall dilation of expf;=2g. This overall factor can be removed by
an additional boost with the exponential of a suitable multiple of K . It is clear, however,
that both single-sided and double-sided application of elements of the spin group which
commute with K can be used to give representations of the general linear group.
Finally, we consider even products of the null vectors wi and wi . These generate the
operations
wiwi I n $ ei (ei ^<n)
wi wiI n $ ei ^ (ei <n) (3.154)
which are rejection and projection operations in <n respectively. For a vector a in <n ,
the operation of projecting a onto the ei direction is performed by
Pi (a) = eiei a (3:155)
and for a general multivector,
Pi(A) = ei ^ (ei A): (3:156)
This projects out the components of A which contain a vector in the ei direction. The
projection onto the orthogonal complement of ei (the rejection) is given by
Pi? (A) = ei (ei ^ A): (3:157)
Projection operations correspond to singular transformations, and we now see that these
are represented by products of null multivectors in <nn . This is sucient to ensure that
singular transformations can also be represented by an even product of vectors, some of
which may now be null.
Two results follow from these considerations. Firstly, every matrix Lie group can be
represented by a spin group | every matrix Lie group can be de
ned as a subgroup
of GL(n,R) and we have shown how GL(n,R) can be represented as a spin group. It
follows that every Lie algebra can be represented by a bivector algebra, since all Lie
algebras have a matrix representation via the adjoint representation. The discussion of
the unitary group has shown, however, that subgroups of GL(n,R) are not, in general,
the best way to construct spin-group representations. Other, more useful, constructions
are given in the following Sections. Secondly, every linear transformation on <n can be
represented in <nn as an even product of vectors, the result of which commutes with
K . It is well known that quaternions are better suited to rotations in three dimensions
than 3 3 matrices. It should now be possible to extend these advantages to arbitrary
linear functions. A number of other applications for these results can be envisaged. For
example, consider the equation
u0(s) = M (s)u(s) (3:158)
58
where u(s) and M (s) are vector and matrix functions of the parameter s and the prime
denotes the derivative with respect to s. By replacing the vector u by the null vector u
in <nn, equation (3.158) can be written in the form
u0 = B (s) u (3:159)
where B (s) is a bivector. If we now write u = Ru0R~ , where u0 is a constant vector, then
equation (3.158) reduces to the rotor equation
R0 = 21 BR (3:160)
which may well be easier to analyse (a similar rotor reformulation of the Lorentz force
law is discussed in 20]).
3.4.1 Complexi
cation | so(n,C)
Complexi
cation of the Orthogonal groups O(p q) leads to a single, non-compact, Lie
group in which all reference to the underlying metric is lost. With the uk and vk de
ned
as in Equation (3.55), the invariant bilinear form is
(u v) = uk vk = xk rk ; yk sk + j (xk sk + yk rk ): (3:161)
This is symmetric, and the real part contains equal numbers of positive and negative norm
terms. The Lie group O(n,C) will therefore be realised in the \balanced" algebra <nn . To
construct the imaginary part of (3.161), however, we need to
nd a symmetric function
which squares to give minus the identity. This is in contrast to the K function, which is
antisymmetric, and squares to +1. The solution is to introduce the \star" function
a (;1)n+1 EnaEn;1 (3:162)
so that
ei = ei (3:163)
e^ = ;e^ :
i i
The use of the notation is consistent with the de
nitions of fwig and fwi g bases (3.99).
The star operator is used to de
ne projections into the Euclidean and anti-Euclidean
subspaces of <nn :
E n (a) = 21 (a + a ) = a EnEn;1 (3:164)
E^ (a) = 1 (a ; a ) = a ^ E E ;1:
n 2 n n
59
Eij = eiej ; e^ie^j (i < j = 1 : : : n)
Fij = eie^j + e^iej : 00
which de
ne the three functions
J i(a) = a Ji: (3:175)
(The introduction of an orthonormal frame is not essential since each of the Ji are inde-
pendent of the intial choice of frame. Orthonormal frames do ease the discussion of the
properties of the Ji, however, so will be used frequently in this and the following sections).
The combined eect of J 1 and J 2 on a vector a produces
J 1J 2(a) = J 1(a eie2i ; a e2i ei + a e3i e1i ; a e1i e3i )
= a eie3i ; a e2i e1i ; a e3i ei + a e1i e2i
= J 3(a): (3.176)
The J i functions therefore generate the quaternionic structure
J 21 = J 22 = J 23 = J 1J 2J 3 = ;1: (3:177)
The Hermitian-symmetric quaternion inner product can be realised in <4n by
(a b) = a b + a J 1(b)i + a J 2(b)j + a J 3(b)j (3:178)
where fi j kg are a basis set of quaterions (see Section 1.2.3). The inner product (3.178)
contains four separate terms, each of which must be preserved by the invariance group.
This group therefore consists of orthogonal transformations satisfying
f (Ji) = Ji i = 1:::3 (3:179)
and the spin group representation consists of the elements of spin(4n) which commute
with all of the Ji. The bivector generators of the invariance group therefore also commute
with the Ji. The results established in Section 3.2 apply for each of the Ji in turn, so an
arbitrary bivector in the Lie algebra of Sp(n) must be of the form
BHU = B + J 1(B ) + J 2(B ) + J 3(B ): (3:180)
This result is used to write down the orthogonal basis set in Table 3.7. The algebra has
dimension 2n2 + n and rank n.
61
Eij = eiej + e1i e1j + e2i e2j + e3i e3j (i < j = 1 : : : n)
Fij = eie1j ; e1i ej ; e2i e3j + e3i e2j 00
Hi = eifi + e^if^i (i = 1 : : : n)
63
Eij = eiej + fifj ; e^ie^j ; f^if^j (i < j = 1 : : : n)
Fij = eifj ; fiej ; e^if^j + f^ie^j 00
Ji = eifi ; e^if^i (i = 1 : : : n)
Ki = eie^i + fif^i 00
Gi = Ji ; Ji+1 (i = 1 : : : n ; 1)
Hi = Ki ; Ki+1 00
65
Eij = eiej + fifj ; e^ie^j ; f^if^j (i < j = 1 : : : n)
Fij = eifj ; fiej ; e^if^j + f^ie^j 00
Fi = eifi ; e^if^i (i = 1 : : : n)
Gi = eie^i ; fif^i 00
Hi = eif^i + fie^i 00
3.5 Summary
In the preceding sections we have seen how many matrix Lie groups can be represented
as spin groups, and how all (
nite dimensional) Lie algebras can be realised as bivector
algebras. These results are summarised in Tables 3.12 and 3.13. Table 3.12 lists the
classical bilinear forms, their invariance groups, the base space in which the spin group
representation is constructed and the general form of the bivector generators. The re-
maining general linear groups are listed in Table 3.13. Again, their invariant bivectors
and the general form of the generators are listed. For both tables, the conventions for
the various functions and bivectors used are those of the section where the group was
discussed.
A number of extensions to this work can be considered. It is well known, for example,
that the Lie group G2 can be constructed in <07 as the invariance group of a particular
trivector (which is given in 46]). This suggests that the techniques explored in this chapter
can be applied to the exceptional groups. A geometric algebra is a graded space and in
Chapter 5 we will see how this can be used to de
ne a multivector bracket which satis
es
66
Base Form of Bivector
Type Form of (a b) Group Space Generators
R-symmetric ab SO(p q) <pq B
R-skew a J (b) Sp(n,R) <2n2n B + J S (B ) ; K (B + J S (B ))
C-symmetric a b + ja K (b) SO(n,C) <nn B ; K (B )
C-skew a J 1(b) + ja J 2(b) Sp(n,C) < 4n4 n B + J (B ) + J S (B ) + J J S (B )
;K ( 00 )
C-Hermitian a b + ja J (b) U(p q) <2p2q B + J (B )
H-Hermitian a b + a J 1(b)i + Sp(n) <4n B + J 1(B ) + J 2(B ) + J 3(B )
a J 2(b)j + a J 3(b)j
H-Skew a J (b) + a K 1(b)+ SO (2n) <2n2n B + J (B ) ; K 1(B ) ; K 2(B )
a bi + a K 2(b)k
Table 3.12: The Classical Bilinear Forms and their Invariance Groups
the super-Jacobi identities. This opens up the possibility of further extending the work of
this chapter to include super-Lie algebras. Furthermore, we shall see in Chapter 4 that the
techniques developed for doubling spaces are ideally suited to the study of multiparticle
quantum theory. Whether some practical bene
ts await the idea that all general linear
transformations can be represented as even products of vectors remains to be seen.
67
Chapter 4
Spinor Algebra
This chapter describes a translation between conventional matrix-based spinor algebra
in three and four dimensions 59, 60], and an approach based entirely in the (real) geo-
metric algebra of spacetime. The geometric algebra of Minkowski spacetime is called the
spacetime algebra or, more simply, the STA. The STA was introduced in Section 1.2.5 as
the geometric algebra generated by a set of four orthonormal vectors f g, = 0 : : : 3,
satisfying
= = diag(+ ; ; ;): (4:1)
Whilst the f g satisfy the Dirac algebra generating relations, they are to be thought
of as an orthonormal frame of independent vectors and not as components of a single
\isospace" vector. The full STA is spanned by the basis
1 f g fk ik g fi g i (4:2)
where
i 0123 (4:3)
and
k k 0: (4:4)
The meaning of these equation was discussed in Section 1.2.5.
The aim of this chapter is to express both spinors and matrix operators within the
real STA. This results in a very powerful language in which all algebraic manipulations
can be performed without ever introducing a matrix representation. The Pauli matrix
algebra is studied
rst, and an extension to multiparticle systems is introduced. The Dirac
algebra and Dirac spinors are then considered. The translation into the STA quickly yields
the Dirac equation in the form
rst found by Hestenes 17, 19, 21, 27]. The concept of
the multiparticle STA is introduced, and is used to formulate a number of two-particle
relativistic wave equations. Some problems with these are discussed and a new equation,
which has no spinorial counterpart, is proposed. The chapter concludes with a discussion
of the 2-spinor calculus of Penrose & Rindler 36]. Again, it is shown how a scalar unit
imaginary is eliminated by the use of the real multiparticle STA. Some sections of this
chapter appeared in the papers \States and operators in the spacetime algebra" 6] and
\2-Spinors, twistors and supersymmetry in the spacetime algebra 4].
68
4.1 Pauli Spinors
This section establishes a framework for the study of the Pauli operator algebra and Pauli
spinors within the geometric algebra of 3-dimensional space. The geometric algebra of
space was introduced in Section 1.2.3 and is spanned by
1 fk g fik g i: (4:5)
Here the fk g are a set of three relative vectors (spacetime bivectors) in the 0-system.
Vectors in this system are written in bold type to distinguish them from spacetime vectors.
There is no possible confusion with the fk g symbols, so these are left in normal type.
When working non-relativistically within the even subalgebra of the full STA some nota-
tional modi
cations are necessary. Relative vectors fk g and relative bivectors fikg are
both bivectors in the full STA, so spatial reversion and spacetime reversion have dierent
eects. To distinguish these, we de
ne the operation
~ 0
Ay = 0A (4:6)
which de
nes reversion in the Pauli algebra. The presence of the 0 vector in the de
nition
of Pauli reversion shows that this operation is dependent on the choice of spacetime frame.
The dot and wedge symbols also carry dierent meanings dependent on whether their
arguments are treated as spatial vectors or spacetime bivectors. The convention adopted
here is that the meaning is determined by whether their arguments are written in bold
type or not. Bold-type objects are treated as three-dimensional multivectors, whereas
normal-type objects are treated as belonging to the full STA. This is the one potentially
confusing aspect of our conventions, though in practice the meaning of all the symbols
used is quite unambiguous.
The Pauli operator algebra 59] is generated by the 2 2 matrices
! ! !
0 1 0 ; j
^1 = 1 0 ^2 = j 0 ^3 = 0 ;1 : 1 0 (4:7)
These operators act on 2-component complex spinors
!
ji = 12 (4:8)
where 1 and 2 are complex numbers. We have adopted a convention by which standard
quantum operators appear with carets, and quantum states are written as kets and bras.
We continue to write the unit scalar imaginary of conventional quantum mechanics as j ,
which distinguishes it from the geometric pseudoscalar i.
To realise the Pauli operator algebra within the algebra of space, the column Pauli
spinor ji is placed in one-to-one correspondence with the even multivector (which
satis
es = 00) through the identi
cation1
0 3 !
a + ja
ji = ;a2 + ja1 $ = a0 + ak ik : (4:9)
1 This mapping was rst found by Anthony Lasenby.
69
In particular, the basis spin-up and spin-down states become
!
1 $ 1 (4:10)
0
and !
0 $ ;i2: (4:11)
1
The action of the four quantum operators f^k j g can now be replaced by the operations
^k ji $ k 3 (k = 1 2 3) (4:12)
and
j ji $ i3: (4:13)
Verifying these relations is a matter of routine computation, for example
2 1 !
; a +
^1 ji = a0 + ja3 ja $ ;;aa0i+2a+ia33i1 = 1 a0 + akik 3:
2 1
(4:14)
With these de
nitions, the action of complex conjugation of a Pauli spinor translates to
ji $ 22: (4:15)
The presence of a
xed spatial vector on the left-hand side of shows that complex
conjugation is a frame-dependent concept.
As an illustration, the Pauli equation (in natural units),
j@t ji = 21m (;j r ; eA)2 ; e^k B k ji + eV ji (4:16)
can be written (in the Coulomb gauge) as 22]
@ti3 = 21m (;r2 + 2eA ri3 + e2A2) ; 2em B 3 + eV (4:17)
where B is the magnetic
eld vector B k k . This translation achieves two important
goals. The scalar unit imaginary is eliminated in favour of right-multiplication by i3,
and all terms (both operators and states) are now real-space multivectors. Removal of
the distinction between states and operators is an important conceptual simpli
cation.
We next need to
nd a geometric algebra equivalent of the spinor inner product h ji.
In order to see how to handle this, we need only consider its real part. This is given by
<h ji $ hyi (4:18)
so that, for example,
h ji $ hyi = h(a0 ; iaj j )(a0 + iakk )i
= (a0)2 + ak ak: (4.19)
70
Since
hji = <hji ; j <hjji (4:20)
the full inner product becomes
h ji $ ( )S hyi ; hyi3ii3: (4:21)
The right hand side projects out the f1 i3g components from the geometric product
y. The result of this projection on a multivector A is written hAiS . For Pauli-even
multivectors this projection has the simple form
hAiS = 21 (A ; i3Ai3): (4:22)
As an example of (4.21), consider the expectation value
hj^k ji $ hyk 3i ; hyk iii3 = k h3yi1 (4:23)
which gives the mean value of spin measurements in the k direction. The STA form
indicates that this is the component of the spin vector s = 3y in the k direction,
so that s is the coordinate-free form of this vector. Since 3y is both Pauli-odd and
Hermitian-symmetric (reverse-symmetric in the Pauli algebra), s contains only a vector
part. (In fact, both spin and angular momentum are better viewed as bivector quantities,
so it is usually more convenient to work with is instead of s.)
Under an active rotation, the spinor transforms as
7! 0 = R0 (4:24)
where R0 is a constant rotor. The quantity 0 is even, and so is a second spinor. (The
term \spinor" is used in this chapter to denote any member of a linear space which is
closed under left-multiplication by a rotor R0.) The corresponding transformation law for
s is
s 7! s0 = R0sRy0 (4:25)
which is the standard double-sided rotor description for a rotation, introduced in Sec-
tion 1.2.4.
The de
nitions (4.9), (4.12) and (4.13) have established a simple translation from the
language of Pauli operators and spinors into the geometric algebra of space. But the STA
formulation can be taken further to aord new insights into the role of spinors in the
Pauli theory. By de
ning
= y (4:26)
the spinor can be written
= 1=2R (4:27)
where R is de
ned as
R = ;1=2: (4:28)
R satis
es
RRy = 1 (4:29)
71
and is therefore a spatial rotor. The spin vector can now be written
s = R3Ry (4:30)
which demonstrates that the double-sided construction of the expectation value (4.23)
contains an instruction to rotate and dilate the
xed 3 axis into the spin direction. The
original states of quantum mechanics have now become operators in the STA, acting on
vectors. The decomposition of the spinor into a density term and a rotor R suggests
that a deeper substructure underlies the Pauli theory. This is a subject which has been
frequently discussed by David Hestenes 19, 22, 23, 27]. As an example of the insights
aorded by this decomposition, it is now clear \why" spinors transform single-sidedly
under active rotations of
elds in space. If the vector s is to be rotated to a new vector
R0sRy0 then, according to the rotor group combination law, R must transform to R0R.
This produces the spinor transformation law (4.24).
We should now consider the status of the
xed fk g frame. The form of the Pauli
equation (4.17) illustrates the fact that, when forming covariant expressions, the fkg
only appear explicitly on the right-hand side of . In an expression like
Ak ^k ji $ A3 (4:31)
for example, the quantity A is a spatial vector and transforms as
A 7! A0 = R0ARy0: (4:32)
The entire quantity therefore transforms as
A3 7! R0ARy0R03 = R0A3 (4:33)
so that A3 is another spinor, as required. Throughout this derivation, the 3 sits on
the right-hand side of and does not transform | it is part of a
xed frame in space.
A useful analogy is provided by rigid-body dynamics, in which a rotating frame fek g,
aligned with the principal axes of the body, can be related to a
xed laboratory frame
fk g by
ek = Rk Ry: (4:34)
The dynamics is now completely contained in the rotor R. The rotating frame fek g is
unaected by the choice of laboratory frame. A dierent
xed laboratory frame,
k0 = R1k Ry1 (4:35)
simply requires the new rotor
R0 = RRy1 (4:36)
to produce the same rotating frame. Under an active rotation, the rigid body is rotated
about its centre of mass, whilst the laboratory frame is
xed. Such a rotation takes
ek 7! e0k = R0ek Ry0 (4:37)
which is enforced by the rotor transformation R 7! R0R. The
xed frame is shielded from
this rotation, and so is unaected by the active transformation. This is precisely what
72
happens in the Pauli theory. The spinor contains a rotor, which shields vectors on the
right-hand side of the spinor from active rotations of spatial vectors.
Since multiplication of a column spinor by j is performed in the STA by right-sided
multiplication by i3, a U(1) gauge transformation is performed by
7! 0 = ei 3 : (4:38)
This right-sided multiplication by the rotor R = expfi3g is equivalent to a rotation
of the initial (
xed) frame to the new frame fRk Ryg. Gauge invariance can therefore
now be interpreted as the requirement that physics is unaected by the position of the 1
and 2 axes in the i3 plane. In terms of rigid-body dynamics, this means that the body
behaves as a symmetric top. These analogies between rigid-body dynamics and the STA
form of the Pauli theory are quite suggestive. We shall shortly see how these analogies
extend to the Dirac theory.
73
The adjoint of a multilinear function is de
ned in the same way as that of a linear function
(Section 1.3), so that
hM' ()i = hM ()i: (4:43)
The Pauli operator adjoint is therefore given by the combination of a reversion, the
geometric adjoint, and a second reversion,
MHA() = M' y(y): (4:44)
For example, if M () = AB , then
M' () = BA (4:45)
and
MHA() = (ByA)y
= AyB y (4.46)
Since the STA action of the ^k operators takes into k 3, it follows that these operators
are, properly, Hermitian. Through this approach, the Pauli operator algebra can now be
fully integrated into the wider subject of multilinear function theory.
75
The STA representation of a 2-particle Pauli spinor is now given by 12E , where 1
and 2 are spinors (even multivectors) in their own spaces. A complete basis for 2-particle
spin states is provided by
! !
1 1 $ E
0! 0!
0 1 $ ;i21E
1
! ! 0 (4:58)
1 0 $ ;i2E 2
0! 1!
0 0 $ i21i22E:
1 1
This procedure extends simply to higher multiplicities. All that is required is to
nd
the \quantum correlator" En satisfying
En i3j = Eni3k = Jn for all j , k: (4:59)
En can be constructed by picking out the j = 1 space, say, and correlating all the other
spaces to this, so that
Yn
En = 21 (1 ; i31i3j ): (4:60)
j =2
The form of En is independent of which of the n spaces is singled out and correlated to.
The complex structure is de
ned by
Jn = Eni3j (4:61)
where i3j can be chosen from any of the n spaces. To illustrate this consider the case of
n = 3, where
E3 = 41 (1 ; i31i32)(1 ; i31i33) (4.62)
= 41 (1 ; i31i32 ; i31i33 ; i32i33) (4.63)
and
J3 = 14 (i31 + i32 + i33 ; i31i32i33): (4:64)
Both E3 and J3 are symmetric under permutations of their indices.
A signi
cant feature of this approach is that all the operations de
ned for the single-
particle STA extend naturally to the multiparticle algebra. The reversion operation, for
example, still has precisely the same de
nition | it simply reverses the order of vectors
in any given multivector. The spinor inner product (4.21) also generalises immediately,
to
( )S = hEn i;1hyEn iEn ; hyJn iJn]: (4:65)
The factor of hEn i;1 is included so that the operation
P (M ) = hEn i;1hMEn iEn ; hMJn iJn ] (4:66)
76
is a projection operation (i.e. P (M ) satis
es P 2(M ) = P (M )). The fact that P (M ) is a
projection operation follows from the results
P (En ) = hEn i;1hEn EniEn ; hEn Jn iJn]
= hEn i;1hEn iEn ; hEn i3j iJn ]
= En (4.67)
and
P (Jn) = hEn i;1 hJnEn iEn ; hJnJn iJn ]
= Jn : (4.68)
79
A Dirac column spinor ji is placed in one-to-one correspondence with an 8-component
even element of the STA via 4, 61]
0 a0 + ja3 1
B 2 1 C
ji = BB@ ;;ab3 ++ ja
jb0 C
C $ = a0 + ak ik + i(b0 + bk ik ):
A (4:87)
;b1 ; jb2
With the spinor ji now replaced by an even multivector, the action of the operators
f^ ^5 j g (where ^5 = ^5 = ;j ^0^1^2^3) becomes
^ ji $ 0 ( = 0 : : : 3)
j ji $ i3 (4:88)
^5 ji $ 3
which are veri
ed by simple computation! for example
0 ;b3 + jb0 1
B ;b1 ; jb2 C ;b3 + b03 + b1i2 ; b2i1 = 3:
^5 ji = B
B@ a0 + ja3 C C
A $ +a03 + a3i ; a21 + a12 (4:89)
;a2 + ja1
Complex conjugation in this representation becomes
ji $ ;22 (4:90)
which picks out a preferred direction on the left-hand side of and so is not a Lorentz-
invariant operation.
As a simple application of (4.87) and (4.88), the Dirac equation
^ (j@ ; eA ) ji = m ji (4:91)
becomes, upon postmultiplying by 0,
ri3 ; eA = m0 (4:92)
which is the form
rst discovered by Hestenes 17]. Here r = @ is the vector derivative
in spacetime. The properties of r will be discussed more fully in Chapter 6. This
translation is direct and unambiguous, leading to an equation which is not only coordinate-
free (since the vectors r = @ and A = A no longer refer to any frame) but is
also representation-free. In manipulating (4.92) one needs only the algebraic rules for
multiplying spacetime multivectors, and the equation can be solved completely without
ever having to introduce a matrix representation. Stripped of the dependence on a matrix
representation, equation (4.92) expresses the intrinsic geometric content of the Dirac
equation.
To discuss the spinor inner product, it is necessary to distinguish between the Hermi-
tian and Dirac adjoint. These are written as
h'j ; Dirac adjoint (4:93)
hj ; Hermitian adjoint
80
which translate as follows,
h'j $ ~ (4:94)
hj $ ~ 0:
y = 0
This makes it clear that the Dirac adjoint is the natural frame-invariant choice. The inner
product is handled in the same manner as in equation (4.21), so that
h' ji $ h ~ i ; hi
~ 3ii3 = h~ iS (4:95)
which is also easily veri
ed by direct calculation. In Chapters 6 and 7 we will be interested
in the STA form of the Lagrangian for the Dirac equation so, as an illustration of (4.95),
this is given here:
L = h'j(^ (j@ ; eA ) ; m)ji $ hri3~ ; eA0~ ; m~i: (4:96)
By utilising (4.95) the STA forms of the Dirac spinor bilinear covariants 60] are readily
found. For example,
h'j^ ji $ h ~ 0i ; h ~ i3ii3 = h0~i1 (4:97)
identi
es the vector 0~ as the coordinate-free representation of the Dirac current. Since
~ is even and reverses to give itself, it contains only scalar and pseudoscalar terms. We
can therefore de
ne
ei
:~ (4:98)
Assuming 6= 0, can now be written as
= 1=2ei
=2R (4:99)
where
R = (ei
);1=2: (4:100)
The even multivector R satis
es RR~ = 1 and is therefore a spacetime rotor. Double-sided
application of R on a vector a produces a Lorentz transformation. The STA equivalents
of the full set of bilinear covariants 33] can now be written as
Scalar h' ji $ h~i = cos
Vector h'j^ ji $ 0~ = v
Bivector h'jj ^ ji $ i3~ = ei
S (4:101)
'
Pseudovector hj^ ^5 ji $ 3 = s ~
Pseudoscalar h'jj ^5 ji $ hi ~ i = ; sin
where
v = R0R~ (4:102)
s = R3R~
and
S = isv: (4:103)
81
These are summarised neatly by the equation
(1 + 0)(1 + i3)~ = cos
+ v + ei
S + is + i sin
: (4:104)
The full Dirac spinor contains (in the rotor R) an instruction to carry out a rotation
of the
xed f g frame into the frame of observables. The analogy with rigid-body
dynamics discussed in Section 4.1 therefore extends immediately to the relativistic theory.
The single-sided transformation law for the spinor is also \understood" in the same way
that it was for Pauli spinors.
Once the spinor bilinear covariants are written in STA form (4.101) they can be manip-
ulated far more easily than in conventional treatments. For example the Fierz identities,
which relate the various observables (4.101), are simple to derive 33]. Furthermore, recon-
stituting from the observables (up to a gauge transformation) is now a routine exercise,
carried out by writing
hiS = 41 ( + 00 ; i3( + 00)i3)
= 41 ( + 00 + 33 + 33) (4.105)
so that
h~iS = 41 (ei
+ v0 ; ei
Si3 + s3): (4:106)
The right-hand side of (4.106) can be found directly from the observables, and the left-
hand side gives to within a complex multiple. On de
ning
Z = 41 (ei
+ v0 ; ei
Si3 + s3) (4:107)
we
nd that, up to an arbitrary phase factor,
= (ei
)1=2Z (Z Z~ );1=2: (4:108)
An arbitrary Dirac operator M^ ji is replaced in the STA by a multilinear function
M (), which acts linearly on the entire even subalgebra of the STA. The 64 real dimen-
sions of this space of linear operators are reduced to 32 by the constraint (4.39)
M (i3) = M ()i3: (4:109)
Proceeding as at (4.44), the formula for the Dirac adjoint is
MDA () = M~' (~): (4:110)
Self-adjoint Dirac operators satisfy M~ () = M' (~) and include the ^ . The Hermitian
adjoint, MHA , is derived in the same way:
MHA() = M' y(y) (4:111)
in agreement with the non-relativistic equation (4.44).
Two important operator classes of linear operators on are projection and symmetry
operators. The particle/antiparticle projection operators are replaced by
1 (m ^ p )ji $ 1 (m p ) (4:112)
0
2m 2m
82
and the spin-projection operators become
1 (1 ^ s ^ )j i $ 1 ( s ): (4:113)
2 5 2 3
Provided that p s = 0, the spin and particle projection operators commute.
The three discrete symmetries C , P and T translate equally simply (following the
convention of Bjorken & Drell 59]):
P^ ji $ 0('x)0
C^ ji $ 1 (4:114)
^T ji $ i0(;x')1
where x' = 0x0 is (minus) a reection of x in the time-like 0 axis.
The STA representation of the Dirac matrix algebra will be used frequently throughout
the remainder of this thesis. In particular, it underlies much of the gauge-theory treatment
of gravity discussed in Chapter 7.
The 2-spinors ji and j'i can therefore be given the STA equivalents
ji $ p12 (1 + 3) (4:173)
j'i $ ;' p12 (1 ; 3):
These dier from the representation of Pauli spinors, and are closer to the \minimal left
ideal" de
nition of a spinor given by some authors (see Chapter 2 of 13], for example).
Algebraically, the (1 3) projectors ensure that the 4-dimensional spaces spanned by
elements of the type p12 (1 + 3) and ' p12 (1 ; 3) are closed under left multiplication by
a relativistic rotor. The signi
cance of the (1 3) projectors lies not so much in their
algebraic properties, however, but in the fact that they are the 0-space projections of
the null vectors 0 3 . This will become apparent when we construct some 2-spinor
\observables".
Under a Lorentz transformation the spinor transforms to R, where R is a relativistic
rotor. If we separate the rotor R into Pauli-even and Pauli-odd terms,
R = R+ + R; (4:174)
where
R+ = 21 (R + 0R0 ) (4.175)
R; = 21 (R ; 0R0) (4.176)
91
then we can write
R p12 (1 + 3) = R+ p12 (1 + 3) + R; 3 p12 (1 + 3)
R' p12 (1 ; 3) = R+ ' p12 (1 ; 3) ; R; '3 p12 (1 ; 3): (4:177)
The transformation laws for the Pauli-even elements and ' are therefore
!7 R+ + R; 3 (4.178)
' !7 R+' ; R; '3 (4.179)
which con
rms that ji transforms under the operator equivalent of R, but that j'i
transforms under the equivalent of
~ 0)~= (R;1)y:
R+ ; R; = 0R0 = (0R (4:180)
This split of a Lorentz transformations into two distinct operations is an unattractive
feature of the 2-spinor formalism, but it is an unavoidable consequence of attempting to
perform relativistic calculations within the Pauli algebra of 2 2 matrices. The problem
is that the natural anti-involution operation is Hermitian conjugation. This operation is
dependent on the choice of a relativistic timelike vector, which breaks up expressions in
a way that disguises their frame-independent meaning.
The 2-spinor calculus attempts to circumvent the above problem by augmenting the
basic 2-component spinor with a number of auxilliary concepts. The result is a language
which has proved to be well-suited to the study of spinors in a wide class of problems and it
is instructive to see how some features of the 2-spinor are absorbed into the STA formalism.
The central idea behind the 2-spinor calculus is that a two-component complex spinor ji,
derived form the Weyl representation (4.121), is replaced by the complex \vector" A .
Here the A is an abstract index labeling the fact that A is a single spinor belonging to
some complex, two-dimensional linear space. We represent this object in the STA as
A $ 21 (1 + 3): (4:181)
p
(The factor of 1=2 replaces 1= 2 simply for convenience in some of the manipulations
that follow.) The only dierence now is that, until a frame is chosen in spin-space, we
have no direct mapping between the components of A and . Secifying a frame in
spin space also picks out a frame in spacetime (determined by the null tetrad). If this
spacetime frame is identi
ed with the f g frame, then the components A of A specify
the Pauli-even multivector via the identi
cation of equation (4.9). A second frame in
spin-space produces dierent components A, and will require a dierent identi
cation
to equation (4.9), but will still lead to the same multivector 21 (1 + 3). 2-Spinors are
equipped with a Lorentz-invariant inner product derived from a metric tensor AB . This
is used to lower indices so, for every 2-spinor A , there is a corresponding A. Both of
these must have the same multivector equivalent, however, in the same way that a and
a both have the STA equivalent a.
To account for the second type of relativistic 2-spinor, j'i (4.121), a second linear
space (or module) is introduced and elements of this space are labeled with bars and
92
primed indices. Thus an abstract element of this space is written as !' A . In a given basis,
0
!' 0 = !0
0
!' 1 = !1:
0
(4:182)
To construct the STA equivalent of !' A we need a suitable equivalent for this operation.
0
!' A 0
$ ;! 12 (1 + 3)1 = ;!i2 21 (1 ; 3): (4:183)
Again, once a basis is chosen, ! is constructed using the identi
cation of equation (4.9)
with the components !0 = !' 0 and !1 = !' 1 .
0 0
A ' A
0
$ ;1 21 (1 + 31)2i22 12 (1 ; 32) 12 (1 ; i31i32): (4:184)
To see how to manipulate the right-hand side of (4.184) we return to the relativistic
two-particle singlet (4.159). The essential property of under multiplication by even
elements was equation (4.162). This relation is unaected by further multiplication of
on the right-hand side by an element that commutes with E . We can therefore form the
object
= 21 (1 + 31) (4:185)
(not to be confused with the non-relativistic Pauli singlet state) which will still satisfy
M 1 = M~ 2 (4:186)
93
for all even multivectors M . The 2-particle state is still a relativistic singlet in the sense
of equation (4.163). From (4.185) we see that contains
2 (1 ; i i ) 2 (1 + 3 ) 2 (1 ; i3 i3 ) = 2 (1 ; i3 i ) 2 (1 + 3 )E
1 12 1 1 1 1 2 1 12 1 1
= 12 (1 ; i32i2) 21 (1 + 31)E
= 12 (1 + 32) 21 (1 + 31)E (4.187)
so we can write
= p12 (i21 ; i22) 21 (1 + 32) 12 (1 + 31)E: (4:188)
A second invariant is formed by right-sided multiplication by (1 ; 31)=2, and we de
ne
' = 21 (1 ; 31): (4:189)
Proceeding as above, we
nd that
' = p12 (i21 ; i22) 21 (1 ; 32) 12 (1 ; 31)E: (4:190)
This split of the full relativistic invariant into and ' lies at the heart of much of the
2-spinor calculus. To see why, we return to equation (4.184) and from this we extract the
quantity 21 (1 + 31) 12 (1 ; 32) 12 (1 ; i31i32). This can be manipulated as follows:
1 (1 + 1 ) 1 (1 ; 2 )E = 1 1 (1 ; 1) 1 (1 ; 1) 1 (1 ; 2)E 1
2 3 2 3 02 3 2 3 2 3 0
= 0 i2 2 (1 ; 3 )(;i2 ) 2 (1 ; 3 ) 2 (1 ; 32)E01
1 2 1 1 2 1 1 1
= 01i22 21 (1 ; 31)(i21 ; i22) 12 (1 ; 31) 12 (1 ; 32)E01
= 01i22 p12 (1 ; 31)'01
= ; p12 (01 + 31)i21'01 (4.191)
which shows how an ' arises naturally in the 2-spinor product. This ' is then used to
project everything to its left back down to a single-particle space. We continue to refer
to each space as a \particle space" partly to stress the analogy with relativistic quantum
states, but also simply as a matter of convenience. In 2-spinor calculus there is no actual
notion of a particle associated with each copy of spacetime.
Returning to the example of A ' A (4.184), we can now write
0
na = !A !' A 0
$ ;!1!2 21 (1 + 31)i2 12 (1 ; 32)E
= p12 !(0 + 3)~! ]1'01
= p12 (0 ; 3)~]1'01 (4.199)
ma = A !' A 0
$ ;1!2 21 (1 + 31)i2 21 (1 ; 32)E
= p12 (0 + 3)~!]1'01
= p12 (1 + i2)~]1'01 (4.200)
and
m' a = !A 'A 0
$ ;!12 21 (1 + 31)i2 21 (1 ; 32)E
= p12 !(0 + 3)~]1'01
= p12 (1 ; i2)~]1'01: (4.201)
96
The key identity used to arrive at the
nal two expression is
(1 + i2)~ = (1 + 3)1~
= (1 + 3)1~
= ;1(1 + 3)i2!~
= ;1(1 + 3)1!~
= (0 + 3)~!: (4.202)
The simplest spin frame is formed when = 1. In this case we arrive at the following
comparison with page 120 of Penrose & Rindler 36]!
la = p12 (ta + za) $ p1
2 (0 + 3 )
na = p12 (ta ; za) $ p (0
1 ; 3)
ma = p12 (xa ; jya) $
2
p (1 + i2 )
1 (4:203)
2
m' a = p12 (xa + jya) $ p1 (1
2 ; i2):
The signi
cant feature of this translation is that the \complex vectors" ma and m' a have
been replaced by vector + trivector combinations. This agrees with the observation that
the imaginary scalar in the 2-spinor calculus plays the role of the spacetime pseudoscalar.
We can solve (4.203) for the Minkowski frame fta xa ya zag (note how the abstract
indices here simply record the fact that the t : : :z are vectors). The only subtlety is
that, in recovering the vector ya from our expression for jya, we must post-multiply our
2-particle expression by i31. The factor of (1 + 31) means that at the one-particle level
this operation reduces to right-multiplication by i. We therefore
nd that
ta $ 0 ya $ ;2 (4:204)
xa $ 1 za $ 3:
The only surprise here is the sign of the y-vector 2. This sign can be traced back
to the fact that Penrose & Rindler adopt an usual convention for the 2 Pauli matrix
(page 16). This is also reected in the fact that they identify the quaternions with vectors
(page 22), and we saw in Section 1.2.3 that the quaternion algebra is generated by the
spatial bivectors fi1 ;i2 i3g.
An arbitrary spin-frame, encoded in the rotor R, produces a new null tetrad simply
by Lorentz rotating the vectors in (4.203), yielding
l = R p12 (0 + 3)R ~ m = R p12 (1 + i2)R ~
n = R p2 (0 ; 3)R
1 ~ m' = R p2 (1 ; i2)R:
1 ~ (4:205)
In this manner, the (abstract) null tetrad becomes a set of four arbitrary vector/trivector
combinations in (4.205), satisfying the anticommutation relations 4]
1
2 fl ng = 1 1
2 fm m' g = 1 all others = 0: (4:206)
97
4.5.4 The rA A Operator 0
The
nal 2-spinor object that we need a translation of is the dierential operator rA A . 0
The translation of rA A will clearly involve the vector derivative r = @x and this must
0
appear in such a way that it picks up the correct transformation law under a rotation in
two-particle space. These observations lead us to the object
rA A $ r101
0
(4:207)
so that, under a rotation,
r101 7! R1R2r101 = R1rR201
= (RrR~ )101 (4.208)
and the r does indeed inherit the correct vector transformation law. In this chapter
we are only concerned with the \at-space" vector derivative r! a suitable formulation
for \curved-space" derivatives will emerge in Chapter 7. A feature of the way that the
multiparticle STA is employed here is that each spinor (1+ 3)=2 is a function of position
in its own spacetime,
j 21 (1 + 3j ) = j (xj ) 12 (1 + 3j ): (4:209)
When such an object is projected into a dierent copy of spacetime, the position depen-
dence must be projected as well. In this manner, spinors can be \pulled back" into the
same spacetime as the dierential operator r.
We are now in a position to form the contraction rA AB AB . We know that the role
0
of the AB is to antisymmetrise on the relevant p particle spaces (in this case the 2 and 3
spaces), together with introducing a factor of 2. Borrowing from the 2-spinor notation,
we denote this operation as 23. We can now write
rA A A = rA AB AB $ r112013 21 (1 + 33)E323
0 0
(4:210)
where we have introduced the notation ij for the invariant (singlet state) under joint
rotations in the ith and j th copies of spacetime. Equation (4.210) is manipulated to give
r112013 21 (1 + 33)E323
= r1 p12 (i21 ; i22)3 12 (1 + 31) 12 (1 + 32) 12 (1 + 33)E32301
= r1 p12 ;i21h(1 + i2)~i204 + hi2(1 + i2)~i204 12 (1 + 31)2301E3(4.211)
and projecting down into particle-one space, the quantity that remains is
rA A A $ r p12 i2h(1 + i2)i04 + h(1 + i2)i2i04] 12 (1 + 3)0: (4:212)
0
nd the equivalent to the above for the expression rAA !'A . The translation for rAA is
0
0
0
4.5.5 Applications
The above constitutes the necessary ingredients for a complete translation of the 2-spinor
calculus into the STA. We close this chapter by considering two important applications.
It should be clear from both that, whilst the steps to achieve the STA form are often
quite complicated, the end result is nearly always more compact and easier to understand
than the original 2-spinor form. An objective of future research in this subject is to
extract from 2-spinor calculus the techniques which are genuinely powerful and useful.
These can then be imported into the STA, which will suitably enriched by so doing. The
intuitive geometric nature of the STA should then make these techniques available to a
wider audience of physicists than currently employ the 2-spinor calculus.
The Dirac Equation
The Dirac equation in 2-spinor form is given by the pair of equations 36, page 222]
rA A A = !' A
0 0
(4:217)
rAA !'A = A :
0
0
p p
The quantity is de
ned to be m= 2, where m is the electron mass. The factor of 1= 2
demonstrates that such factors are intrinsic to the way that the rA A symbol encodes the
0
99
we
nd that
r0 = m!2 21 (1 ; 3) ; m 12 (1 + 3)]i
= ;mi3: (4.220)
We thus recover the STA version of the Dirac equation (4.92)
ri3 = m0: (4:221)
Of the pair of equations (4.217), Penrose & Rindler write \an advantage of the 2-spinor
description is that the -matrices disappear completely { and complicated -matrix identi-
ties simply evaporate! " 36, page 221]. Whilst this is true, the comment applies even more
strongly to the STA form of the Dirac equation (4.221), in which complicated 2-spinor
identities are also eliminated!
Maxwell's Equations
In the 2-spinor calculus the real, antisymmetric tensor F ab is written as
F ab = AB A B + AB A B
0 0 0 0
(4:222)
where AB is symmetric on its two indices. We
rst need the STA equivalent of AB .
Assuming initially that AB is arbitrary, we can write
AB $ 12 (1 + 31) 12 (1 + 32)E = 21 (1 + 31)11 (4:223)
where is an arbitrary element of the product space of the two single-particle Pauli-
even algebras. A complete basis for is formed by all combinations of the 7 elements
f1 ik1 ik2g. The presence of the singlet allows all elements of second space to be
projected down into the
rst space, and it is not hard to see that this accounts for all
possible even elements in the one-particle STA. We can therefore write
21 (1 + 31)11 = M 1 (4:224)
where M is an arbitrary even element. The condition that AB is symmetric on its two
indices now becomes (recalling that is antisymmetric on its two particle indices)
M 1 = ;M 2 = ;M~ 1 (4:225)
~
) M = ;M: (4:226)
This condition projects out from M the components that are bivectors in particle-one
space, so we can write
AB $ F 1 (4:227)
where F is now a bivector. For the case of electromagnetism, F is the Faraday bivector,
introduced in Section (1.2.5). The complete translation of F ab is therefore
F ab $ F 1 + F 11112 = F 1 (4:228)
100
where is the full relativistic invariant.
The 2-spinor form of the Maxwell equations an be written
rA B AC BC = ;J AA
0 0
(4:229)
where J AA is a \real" vector (i.e. it has no trivector components). Recalling the con-
0
vention that ij denotes the singlet state in coupled fi j g-space, the STA version of
equation (4.229) is
r11201F 33424 = ;J 11301: (4:230)
This is simpli
ed by the identity
123424 = 13 (4:231)
which is proved by expanding the left-hand side and then performing the antisymmetri-
sation. The resultant equation is
r1F 313 = ;J 113 (4:232)
which has a one-particle reduction to
rF = J: (4:233)
This recovers the STA form of the Maxwell equations 17]. The STA form is remarkably
compact, makes use solely of spacetime quantities and has a number of computational
advantages over second-order wave equations 8]. The 2-spinor calculus also achieves a
101
Chapter 5
Point-particle Lagrangians
In this chapter we develop a multivector calculus as the natural extension of the calculus
of functions of a single parameter. The essential new tool required for such a calculus is
the multivector derivative, and this is described
rst. It is shown how the multivector
derivative provides a coordinate-free language for manipulating linear functions (forming
contractions etc.). This supersedes the approach used in earlier chapters, where such
manipulations were performed by introducing a frame.
The remainder of this chapter then applies the techniques of multivector calculus
to the analysis of point-particle Lagrangians. These provide a useful introduction to the
techniques that will be employed in the study of
eld Lagrangians in the
nal two chapters.
A novel idea discussed here is that of a multivector-valued Lagrangian. Such objects are
motivated by the pseudoclassical mechanics of Berezin & Marinov 39], but can only
be fully developed within geometric algebra. Forms of Noether's theorem are given for
both scalar and multivector-valued Lagrangians, and for transformations parameterised
by both scalars and multivectors. This work is applied to the study of two semi-classical
models of electron spin. Some aspects of the work presented in this chapter appeared
in the papers \Grassmann mechanics, multivector derivatives and geometric algebra" 3]
and \Grassmann calculus, pseudoclassical mechanics and geometric algebra" 1].
104
5.2 Scalar and Multivector Lagrangians
As an application of the multivector derivative formalism just outlined, we consider
Lagrangian mechanics. We start with a scalar-valued Lagrangian L = L(Xi X_ i ), where
the Xi are general multivectors, and X_ i denotes dierentiation with respect to time. We
wish to
nd the Xi (t) which extremise the action
Z t2
S = dt L(Xi X_ i ): (5:16)
t1
The solution to this problem can be found in many texts (see e.g. 71]). We write
Xi (t) = Xi0(t) + Yi(t) (5:17)
where Yi is a multivector containing the same grades as Xi and which vanishes at the
endpoints, is a scalar and Xi0 represents the extremal path. The action must now
satisfy @
S = 0 when = 0, since = 0 corresponds to Xi (t) taking the extremal values.
By applying the chain rule and integrating by parts, we
nd that
Z t2
@
S = dt (@
Xi) @Xi L + (@
X_ i ) @X_ i L
Zt1t2
= t dt Yi @Xi L + Y_i @X_ i L
Z 1t2
= dt Yi @Xi L ; @t(@X_ i L) : (5.18)
t1
Setting to zero now just says that Xi is the extremal path, so the extremal path is
de
ned by the solutions to the Euler-Lagrange equations
@Xi L ; @t(@X_ i L) = 0: (5:19)
The essential advantage of this derivation is that it employs genuine derivatives in place
of the less clear concept of an in
nitessimal. This will be exempli
ed when we study
Lagrangians containing spinor variables.
We now wish to extend the above argument to a multivector-valued Lagrangian L.
Taking the scalar product of L with an arbitrary constant multivector A produces a scalar
Lagrangian hLAi. This generates its own Euler-Lagrange equations,
@Xi hLAi ; @t(@X_ i hLAi) = 0: (5:20)
A \permitted" multivector Lagrangian is one for which the equations from each A are
mutually consistent, so that each component of the full L is capable of simultaneous
extremisation.
By contracting equation (5.20) on the right-hand side by @A , we
nd that a necessary
condition on the dynamical variables is
@Xi L ; @t(@X_ i L) = 0: (5:21)
For a permitted multivector Lagrangian, equation (5.21) is also su
cient to ensure that
equation (5.20) is satis
ed for all A. This is taken as part of the de
nition of a multivector
Lagrangian. We will see an example of how these criteria can be met in Section 5.3.
105
5.2.1 Noether's Theorem
An important technique for deriving consequences of the equations of motion resulting
from a given Lagrangian is the study of the symmetry properties of the Lagrangian itself.
The general result needed for this study is Noether's theorem. We seek a form of this
theorem which is applicable to both scalar-valued and multivector-valued Lagrangians.
There are two types of symmetry to consider, depending on whether the transformation
of variables is governed by a scalar or by a multivector parameter. We will look at these
separately.
It is important to recall at this point that all the results obtained here are derived in
the coordinate-free language of geometric algebra. Hence all the symmetry transforma-
tions considered are active. Passive transformations have no place in this scheme, as the
introduction of an arbitrary coordinate system is an unnecessary distraction.
The de
nition of L0 ensures that it has the same functional form of L, so the quantity
@Xi hL0Ai ; @t(@X_ i hL0 Ai)L0
0 0 (5:25)
is obtained by taking the Euler-Lagrange equations in the form (5.20) and replacing the
Xi by Xi0. If we now assume that the Xi0 satisfy the same equations of motion (which
must be checked for any given case), we
nd that
@L0 = @t (@aXi0) @X_ i L00 (5:26)
and, if L0 is independent of , the corresponding quantity (@aXi0) @X_ i L0 is conserved.
0
= @t f M (Xi A) @X_ i L0
0 (5.35)
where again it is necessary to assume that the equations of motion are satis
ed for the
transformed variables. We can remove the A-dependence by dierentiating, which yields
@M L0 = @t @Af M (Xi A) @X_ i L0 0 (5:36)
and, if L0 is independent of M , the corresponding conserved quantity is
@Af M (Xi A) @X_ i L0 =@M f (Xi M ) @X_ i L0
0 0 (5:37)
107
where the overstar on M denote the argument of @M .
It is not usually possible to set M to zero in (5.35), but it is interesting to see that
conserved quantities can be found regardless. This shows that standard treatments of
Lagrangian symmetries 71] are unnecessarily restrictive in only considering in
nitesimal
transformations. The subject is richer than this suggests, though without multivector
calculus the necessary formulae are hard to
nd.
In order to illustrate (5.37), consider reection symmetry applied to the harmonic
oscillator Lagrangian
L(x x_ ) = 21 (x_ 2 ; !2x2): (5:38)
The equations of motion are
x& = ;!2x (5:39)
and it is immediately seen that, if x is a solution, then so to is x0, where
x0 = ;nxn;1: (5:40)
Here n is an arbitrary vector, so x0 is obtained from x by a reection in the hyperplance
orthogonal to n. Under the reection (5.40) the Lagrangian is unchanged, so we can
nd
a conserved quantity from equation (5.37). With f (x n) de
ned by
f (x n) = ;nxn;1 (5:41)
we
nd that
f n(x a) = ;axn;1 + nxn;1an;1: (5:42)
Equation (5.37) now yields the conserved quantity
@a(;axn;1 + nxn;1 an;1) (;nxn
_ ;1 ) = @ahaxxn _ ;1 ; axxn
_ ;1i
= hxxn
_ ;1 ; xxn _ ;1i1
= 2(x ^ x_ ) n;1: (5.43)
This is conserved for all n, from which it follows that the angular momentum x ^ x_
is conserved. This is not a surprise, since rotations can be built out of reections and
dierent reections are related by rotations. It is therefore natural to expect the same
conserved quantity from both rotations and reections. But the derivation does show
that the multivector derivative technique works and, to my knowledge, this is the
rst
time that a classical conserved quantity has been derived conjugate to transformations
that are not simply connected to the identity.
108
variables include spinor variables. The STA formalism of Chapter 4 is applied to this
Lagrangian and used to analyse the equations of motion. Some problems with the model
are discussed, and a more promising model is proposed. The second example is drawn
from pseudoclassical mechanics. There the dynamical variables are Grassmann-valued
entities, and the formalism of Chapter 2 is used to represent these by geometric vectors.
The resulting Lagrangian is multivector-valued, and is studied using the techniques just
developed. The equations of motion are found and solved, and again it is argued that the
model fails to provide an acceptable picture of a classical spin-half particle.
112
Thus the variation now leads directly to the precession equation for the spin. The
complete set of equations is now
S_ = 2p ^ x_ (5.78)
x_ = R0 R ~ (5.79)
p_ = qF x_ (5.80)
which are manifestly Lorentz covariant and gauge invariant. The Hamiltonian is now p x_
and the free-particle angular momentum is still de
ned by J (5.55), though now the spin
bivector S is always of unit magnitude.
A
nal problem remains, however, which is that we have still not succeeded in con-
structing a model which predicts the correct gyromagnetic moment. In order to achieve
the correct coupling between the spin and the Faraday bivector, the Lagrangian (5.69)
must be modi
ed to
_ 3;1 + p(x_ ; 0=
L = hi ~ ) + qx_ A ; q Fi3;1i: (5:81)
2m
The equations of motion are now
S_ = 2p ^ x_ + mq F S
x_ = R0R~ (5.82)
q
p_ = qF x_ ; 2m rF (x) S
which recover the correct precession formulae in a constant magnetic
eld. When p is set
equal to mx_ , the equations (5.82) reduce to the pair of equations studied in 72].
xed B , equation (5.88) does not lead to the full equations of motion (5.86). It is only
by allowing B to vary that we arrive at (5.86). It is therefore an essential feature of the
formalism that L is a multivector, and that (5.88) holds for all B .
The equations of motion (5.86) can be written out in full to give
e_1 = ;!2e3 + !3e2
e_2 = ;!3e1 + !1e3 (5:89)
e_3 = ;!1e2 + !2e1
which are a set of three coupled
rst-order vector equations. In terms of components,
this gives nine scalar equations for nine unknowns, which illustrates how multivector
Lagrangians have the potential to package up large numbers of equations into a single,
highly compact entity. The equations (5.89) can be neatly combined into a single equation
by introducing the reciprocal frame feig (1.132),
e1 = e2 ^ e3En;1 etc. (5:90)
114
where
En e1 ^ e2 ^ e3: (5:91)
With this, the equations (5.89) become
e_i = ei ! (5:92)
which shows that potentially interesting geometry underlies this system, relating the
equations of motion of a frame to its reciprocal.
We now proceed to solve equation (5.92). On feeding (5.92) into (5.85), we
nd that
!_ = 0 (5:93)
so that the ! plane is constant. We next observe that (5.89) also imply
E_ n = 0 (5:94)
which is important as it shows that, if the feig frame initially spans 3-dimensional space,
then it will do so for all time. The constancy of En means that the reciprocal frame (5.90)
satis
es
e_1 = ;!2e3 + !3e2 etc. (5:95)
We now introduce the symmetric metric tensor g, de
ned by
g(ei) = ei: (5:96)
This de
nes the reciprocal bivector
! g;1 (!)
= !1(e2 ^ e3) + !2(e3 ^ e1) + !3(e1 ^ e2) (5.97)
so that the reciprocal frame satis
es the equations
e_i = ei ! : (5:98)
But, from (1.123), we have that
ei ! = ei g;1 (!) = g;1 (ei !): (5:99)
Now, using (5.92), (5.98) and (5.99), we
nd that
g(e_i) = ei ! = e_i = @tg(ei) (5:100)
) g_ = 0: (5:101)
Hence the metric tensor is constant, even though its matrix coecients are varying. The
variation of the coecients of the metric tensor is therefore purely the result of the time
variation of the frame, and is not a property of the frame-independent tensor. It follows
115
that the
ducial tensor (1.144) is also constant, and suggests that we should look at the
equations of motion for the
ducial frame i = h;1 (ei). For the fig frame we
nd that
_ i = h;1(_ei)
= h;1(h;1(i) !)
= i h;1(!): (5.102)
If we de
ne the bivector
+ = h;1(!) = !123 + !231 + !312 (5:103)
(which must be constant, since both h and ! are), we see that the
ducial frame satis
es
the equation
_ i = i +: (5:104)
The underlying
ducial frame simply rotates at a constant frequency in the + plane. If
i(0) denotes the
ducial frame speci
ed by the initial setup of the feig frame, then the
solution to (5.104) is
i(t) = e;t=2i(0)et=2 (5:105)
and the solution for the feig frame is
ei(t) = h(e;t=2i(0)et=2) (5:106)
ei(t) = h;1(e;t=2i(0)et=2):
Ultimately, the motion is that of an orthonormal frame viewed through a constant (sym-
metric) distortion. The feig frame and its reciprocal representing the same thing viewed
through the distortion and its inverse. The system is perhaps not quite as interesting as
one might have hoped, and it has not proved possible to identify the motion of (5.106)
with any physical system, except in the simple case where h = I . On the other hand, we
did start with a very simple Lagrangian and it is reassuring to recover a rotating frame
from an action that was motivated by the pseudoclassical mechanics of spin.
Some simple consequences follow immediately from the solution (5.106). Firstly, there
is only one frequency in the system, say, which is found via
2 = ;+2
= !12 + !22 + !32: (5.107)
Secondly, since
+ = i(!11 + !22 + !33) (5:108)
the vectors
u !1e1 + !2e2 + !3e3 (5:109)
and
u = g;1 (u) (5:110)
are conserved. This also follows from
u = ;En! (5.111)
u = E !: n (5.112)
116
Furthermore,
eiei = h(i)h(i)
= ig(i)
= Tr(g) (5.113)
must also be time-independent (as can be veri
ed directly from the equations of motion).
The reciprocal quantity eiei = Tr(g;1 ) is also conserved. We thus have the set of four
standard rotational invariants, ii, the axis, the plane of rotation and the volume scale-
factor, each viewed through the pair of distortions h, h;1. This gives the following set of
8 related conserved quantities:
feiei, eiei, u, u , !, ! , En , E ng: (5:114)
Lagrangian Symmetries and Conserved Quantities
We now turn to a discussion of the symmetries of (5.84). Although we have solved the
equations of motion exactly, it is instructive to derive some of their consequences directly
from the Lagrangian. We only consider continuous symmetries parameterised by a single
scalar, so the appropriate form of Noether's theorem is equation (5.28), which takes the
form
@aL0j = @t 1 ei ^ (@ae0 ) :
=0 2 i =0
(5:115)
In writing this we are explicitly making use of the equations of motion and so are
nding
\on-shell" symmetries. The Lagrangian could be modi
ed to extend these symmetries
o-shell, but this will not be considered here.
We start with time translation. From (5.32), the Hamiltonian is
H = 21 ei ^ e_i ; L = ! (5:116)
which is a constant bivector, as expected. The next symmetry to consider is a dilation,
e0i = eei: (5:117)
For this transformation, equation (5.115) gives
2L = @t 12 ei ^ ei = 0 (5:118)
so dilation symmetry shows that the Lagrangian vanishes along the classical path. This
is quite common for
rst-order systems (the same is true of the Dirac Lagrangian), and
is important in deriving other conserved quantities.
The
nal \classical" symmetry to consider is a rotation,
e0i = eB=2eie;B=2: (5:119)
Equation (5.115) now gives
B L = @t 21 ei ^ (B ei) (5:120)
117
but, since L = 0 when the equations of motion are satis
ed, the left hand side of (5.120)
vanishes, and we
nd that the bivector ei ^ (B ei) in conserved. Had our Lagrangian
been a scalar, we would have derived a scalar-valued function of B at this point, from
which a single conserved bivector | the angular momentum | could be found. Here our
Lagrangian is a bivector, so we
nd a conserved bivector-valued function of a bivector |
a set of 3 3 = 9 conserved scalars. The quantity ei ^ (B ei) is a symmetric function of
B , however, so this reduces to 6 independent conserved scalars. To see what these are
we introduce the dual vector b = iB and replace the conserved bivector ei ^(B ei) by the
equivalent vector-valued function,
f (b) = ei (b ^ ei) = ei bei ; beiei = g(b) ; bTr(g): (5:121)
This is conserved for all b, so we can contract with @b and observe that ;2Tr(g) is constant.
It follows that g (b) is constant for all b, so rotational symmetry implies conservation of
the metric tensor | a total of 6 quantities, as expected.
Now that we have derived conservation of g and !, the remaining conserved quantities
can be found. For example, En = det(g)1=2i shows that En is constant. One interesting
scalar-controlled symmetry remains, however, namely
e0i = ei + !i a (5:122)
where a is an arbitrary constant vector. For this symmetry (5.115) gives
1 a ^ u_ = @ 1 e ^ (! a)
2 t 2 i i (5:123)
) a ^ u_ = 0 (5:124)
which holds for all a. Conservation of u therefore follows directly from the symmetry
transformation (5.122). This symmetry transformation bears a striking resemblance to
the transformation law for the fermionic sector of a supersymmetric theory 75]. Although
the geometry behind (5.122) is not clear, it is interesting to note that the pseudoscalar
transforms as
En0 = En + a ^ ! (5:125)
and is therefore not invariant.
Poisson Brackets and the Hamiltonian Formalism
Many of the preceding results can also be derived from a Hamiltonian approach. As a
by-product, this reveals a new and remarkably compact formula for a super-Lie bracket.
We have already seen that the Hamiltonian for (5.84) is !, so we start by looking at
how the Poisson bracket is de
ned in pseudoclassical mechanics 39]. Dropping the j and
adjusting a sign, the Poisson bracket is de
ned by
@; @
fa( ) b( )gPB = a @ @ b: (5:126)
k k
The geometric algebra form of this is
fA B gPB = (A ek) ^ (ek B ) (5:127)
118
where A and B are arbitrary multivectors. We will consider the consequences of this
de
nition in arbitrary dimensions initially, before returning to the Lagrangian (5.84).
Equation (5.127) can be simpli
ed by utilising the
ducial tensor,
(A h;1(k )) ^ (h;1(k ) B ) = hh;1(A) k] ^ hk h;1(B )]
= h(h;1(A) k) ^ (k h;1(B ))]: (5.128)
If we now assume that A and B are homogeneous, we can employ the rearrangement
(Ar k ) ^ (k Bs ) = 41 h(Ar k ; (;1)r k Ar )(k Bs ; (;1)sBs k )ir+s;2
= 41 hnAr Bs ; (n ; 2r)Ar Bs ; (n ; 2s)Ar Bs
+n ; 2(r + s ; 2)]ArBsir+s;2
= hAr Bs ir+s;2 (5.129)
to write the Poisson bracket as
fAr Bs gPB = hhh;1(Ar )h;1(Bs)ir+s;2 : (5:130)
This is a very neat representation of the super-Poisson bracket. The combination rule is
simple, since the h always sits outside everything:
D E
fAr fBs CtgPB gPB = h h;1(Ar)hh;1(Bs)h;1(Ct)is+t;2 r+s+t;4 : (5:131)
Cliord multiplication is associative and
hAr Bsir+s;2 = ;(;1)rshBs Arir+s;2 (5:132)
so the bracket (5.130) generates a super-Lie algebra. This follows from the well-known
result 76] that a graded associative algebra satisfying the graded commutator rela-
tion (5.132) automatically satis
es the super-Jacobi identity. The bracket (5.130) there-
fore provides a wonderfully compact realisation of a super-Lie algebra. We saw in Chap-
ter 3 that any Lie algebra can be represented by a bivector algebra under the commutator
product. We now see that this is a special case of the more general class of algebras closed
under the product (5.130). A subject for future research will be to use (5.130) to extend
the techniques of Chapter 3 to include super-Lie algebras.
Returning to the system de
ned by the Lagrangian (5.84), we can now derive the
equations of motion from the Poisson bracket as follows,
e_i = fei H gPB
= h(i +)
= ei !: (5.133)
It is intersting to note that, in the case where h = I , time derivatives are determined
by (one-half) the commutator with the (bivector) Hamiltonian. This suggests an inter-
esting comparison with quantum mechanics, which has been developed in more detail
elsewhere 1].
119
Similarly, some conservation laws can be derived, for example
fEn H gPB = hhi+i3 = 0 (5:134)
and
f! H gPB = hh++i2 = 0 (5:135)
show that En and ! are conserved respectively. The bracket (5.130) gives zero for any
scalar-valued functions, however, so is no help in deriving conservation of eiei. Further-
more, the bracket only gives the correct equations of motion for the feig frame, since
these are the genuine dynamical variables.
This concludes our discussion of pseudoclassical mechanics and multivector Lagran-
gians in general. Multivector Lagrangians have been shown to possess the capability to
package up large numbers of variables in a single action principle, and it is to be hoped
that further, more interesting applications can be found. Elsewhere 1], the concept of a
bivector-valued action has been used to give a new formulation of the path integral for
pseudoclassical mechanics. The path integrals considered involved genuine Riemann inte-
grals in parameter space, though it has not yet proved possible to extend these integrals
beyond two dimensions.
120
Chapter 6
Field Theory
We now extend the multivector derivative formalism of Chapter 5 to encompass
eld
theory. The multivector derivative is seen to provide great formal clarity by allowing
spinors and tensors to be treated in a uni
ed way. The relevant form of Noether's theorem
is derived and is used to
nd new conjugate currents in Dirac theory. The computational
advantages of the multivector derivative formalism are further exempli
ed by derivations
of the stress-energy and angular-momentum tensors for Maxwell and coupled Maxwell-
Dirac theory. This approach provides a clear understanding of the role of antisymmetric
terms in the stress-energy tensor, and the relation of these terms to spin. This chapter
concludes with a discussion of how the formalism of multivector calculus is extended to
incorporate dierentiation with respect to a multilinear function. The results derived
in this section are crucial to the development of an STA-based theory of gravity, given
in Chapter 7. Many of the results obtained in this chapter appeared in the paper \A
multivector derivative approach to Lagrangian eld theory" 7].
Some additional notation is useful for expressions involving the vector derivative r.
The left equivalent of r is written as r and acts on multivectors to its immediateleft.
(It is not always necessary to use r, as the overdot notation can be used to write A r as
_Ar_ .) The operator r
$
acts both to its left and right, and is taken as acting on everything
within a given expression, for example
A r B = A_ r_ B + Ar_ B: _
$
(6:1)
Transformations of spacetime position are written as
x0 = f (x): (6:2)
The dierential of this is the linear function
f (a) = a rf (x) = f x(a) (6:3)
where the subscript labels the position dependence. A useful result for vector derivatives
is that
rx = @aa rx
= @a(a rxx0) rx 0
= f x(rx ):
0 (6.4)
121
6.1 The Field Equations and Noether's Theorem
In what follows, we restrict attention to the application of multivector calculus to rela-
tivistic
eld theory. The results are easily extended to include the non-relativistic case.
Furthermore, we are only concerned with scalar-valued Lagrangian densities. It has not
yet proved possible to construct a multivector-valued
eld Lagrangian with interesting
properties.
We start with a scalar-valued Lagrangian density
L = L(i a ri) (6:5)
where fig are a set of multivector
elds. The Lagrangian (6.5) is a functional of i and
the directional derivatives of i. In many cases it is possible to write L as a functional of
and r, and this approach was adopted in 7]. Our main application will be to gravity,
however, and there we need the more general form of (6.5).
The action is de
ned as Z
S = jd4xjL (6:6)
where jd4xj is the invariant measure. Proceeding as in Chapter 5, we write
i(x) = i0(x) + i(x) (6:7)
where i contains the same grades as i, and i0 is the extremal path. Dierentiating,
and using the chain rule, we
nd that
Z
@
S = jd4xj (@
i) @i L + (@
i ) @i L]
Z
= jd4xj i @i L + (i ) @i L]: (6.8)
Here, a
xed frame f g has been introduced, determining a set of coordinates x x.
The derivative of i with respect to x is denoted as i . The multivector derivative
@i is de
ned in the same way as @i . The frame can be eliminated in favour of the
multivector derivative by de
ning
@ia a @i (6:9)
where a = a, and writing
Z
@
S = jd4xj i @i L + (@a ri) @ia L]: (6:10)
It is now possible to perform all manipulations without introducing a frame. This ensures
that Lorentz invariance is manifest throughout the derivation of the
eld equations.
Assuming that the boundary term vanishes, we obtain
Z
@
S = jd4xj i @i L ; @a r(@ia L)]: (6:11)
122
Setting = 0, so that the i takes their extremal values, we
nd that the extremal path
is de
ned by the solutions of the Euler-Lagrange equations
@i L ; @a r(@ia L) = 0: (6:12)
The multivector derivative allows for vectors, tensors and spinor variables to be handled
in a single equation | a useful and powerful uni
cation.
Noether's theorem for
eld Lagrangians is also be derived in the same manner as in
Chapter 5. We begin by considering a general multivector-parameterised transformation,
i0 = f (i M ) (6:13)
where f and M are position-independent functions and multivectors respectively. With
L0 L(i0 a ri0), we
nd that
A @M L0 = f M (i A) @i L0 + f M (@a ri A) @ia L0
0 0
123
6.2 Spacetime Transformations and their Conjugate
Tensors
In this section we use Noether's theorem in the form (6.17) to analyse the consequences of
Poincare and conformal invariance. These symmetries lead to the identi
cation of stress-
energy and angular-momentum tensors, which are used in the applications to follow.
1. Translations
A translation of the spacetime
elds i is achieved by de
ning new spacetime
elds i0 by
i0(x) = i(x0) (6:21)
where
x0 = x + n: (6:22)
Assuming that L is only x-dependent through the
elds i(x), equation (6.17) gives
0
124
2. Rotations
If we assume initially that all
elds i transform as spacetime vectors, then a rotation of
these
elds from one spacetime point to another is performed by
i0(x) = eB=2i(x0)e;B=2 (6:31)
where
x0 = e;B=2xeB=2: (6:32)
This diers from the point-particle transformation law (5.53) in the relative directions of
the rotations for the position vector x and the
elds i. The result of this dierence is a
change in the relative sign of the spin contribution to the total angular momentum. In
order to apply Noether's theorem (6.17), we use
@i0j=0 = B i ; (B x) ri (6:33)
and
@L0j=0 = ;(B x) rL = r (x B L): (6:34)
Together, these yield the conjugate vector
J (B ) = @aB i ; (B x) ri] @ia L + B xL (6:35)
which satis
es
r J (B ) = 0: (6:36)
The adjoint to the conservation equation (6.36) is
J_ (r_ ) B = 0 for all B
) J_ (r_ ) = 0: (6.37)
The adjoint function J (n) is a position-dependent bivector-valued linear function of the
vector n. We identify this as the canonical angular-momentum tensor. A conserved
bivector in the 0-system is constructed in the same manner as for T (n) (6.28). The
calculation of the adjoint function J (n) is performed as follows:
J (n) = @B hJ (B )ni
= @B h(B i ; B (x ^r)i) @in L + B xLni
= ;x ^ r_ _ i @in L ; nL] + hi @in Li2
= T (n) ^ x + hi @in Li2: (6.38)
If one of the
elds , say, transforms single-sidedly (as a spinor), then J (n) contains the
term h 21 @n Li2 .
The
rst term in J (n) (6.38) is the routine p ^ x component, and the second term is
due to the spin of the
eld. The general form of J (n) is therefore
J (n) = T (n) ^ x + S (n): (6:39)
125
By applying (6.37) to (6.39) and using (6.27), we
nd that
T (r_ ) ^ x_ + S_ (r_ ) = 0: (6:40)
The
rst term in (6.40) returns (twice) the characteristic bivector of T (n). Since the
antisymmetric part of T (n) can be written in terms of the characteristic bivector B as
T ; (a) = 21 B a (6:41)
equation (6.40) becomes
B = ;S_ (r_ ): (6:42)
It follows that, in any Poincare-invariant theory, the antisymmetric part of the stress-
energy tensor is a total divergence. But, whilst T; (n) is a total divergence, x ^ T;(n)
certainly is not. So, in order for (6.37) to hold, the antisymmetric part of T (n) must be
retained since it cancels the divergence of the spin term.
3. Dilations
While all fundamental theories should be Poincare-invariant, an interesting class go be-
yond this and are invariant under conformal transformations. The conformal group con-
tains two further symmetry transformations, dilations and special conformal transfor-
mations. Dilations are considered here, and the results below are appropriate to any
scale-invariant theory.
A dilation of the spacetime
elds is achieved by de
ning
i0(x) = edii(x0) (6:43)
where
x0 = ex (6:44)
) ri(x) = e
0 (d i +1)
0
rx i(x ):
0
(6:45)
If the theory is scale-invariant, it is possible to assign the \conformal weights" di in such
a way that the left-hand side of (6.17) becomes
@L0j=0 = r (xL): (6:46)
In this case, equation (6.17) takes the form
r (xL) = r @a(di i + x ri) @ia L] (6:47)
from which the conserved current
j = di@ai @ia L + T (x) (6:48)
is read o. Conservation of j (6.48) implies that
r T (x) = @aT (a) = ;r (di@ai @ia L) (6:49)
so, in a scale-invariant theory, the trace of the canonical stress-energy tensor is a total
divergence. By using the equations of motion, equation (6.49) can be written, in four
dimensions, as
di hi@i Li + (di + 1)(@a ri) @ia L = 4L (6:50)
which can be taken as an alternative de
nition for a scale-invariant theory.
126
4. Inversions
The remaining generator of the conformal group is inversion,
x0 = x;1: (6:51)
As it stands, this is not parameterised by a scalar and so cannot be applied to (6.17). In
order to derive a conserved tensor, the inversion (6.51) is combined with a translation to
de
ne the special conformal transformation 77]
x0 = h(x) (x;1 + n);1 = x(1 + nx);1: (6:52)
From this de
nition, the dierential of h(x) is given by
h(a) = a rh(x) = (1 + xn);1a(1 + nx);1 (6:53)
so that h de
nes a spacetime-dependent rotation/dilation. It follows that h satis
es
h(a) h(b) = (x)a b (6:54)
where
(x) = (1 + 2n x + 2x2n2);2: (6:55)
That the function h(a) satis
es equation (6.54) demonstrates that it belongs to the con-
formal group.
The form of h(a) (6.53) is used to postulate transformation laws for all
elds (including
spinors, which transform single-sidedly) such that
L0 = (det h)L(i (x0) h(a) rx i(x0))
0 (6:56)
which implies that
@L0j=0 = @ det hj=0 L + ( @x0j=0) rL: (6:57)
Since
det h = (1 + 2n x + 2x2n2);4 (6:58)
it follows that
@ det hj=0 = ;8x n: (6:59)
We also
nd that
@x0j=0 = ;(xnx) (6:60)
and these results combine to give
@L0j=0 = ;8x nL ; (xnx) rL = ;r (xnxL): (6:61)
Special conformal transformations therefore lead to a conserved tensor of the form
T SC (n) = @ah(;(xnx) ri + @i0(x)) @ia L + xnxLi=0
= ;T (xnx) + @ah(@i0(x)) @ia Li=0 : (6.62)
127
The essential quantity here is the vector ;xnx, which is obtained by taking the constant
vector n and reecting it in the hyperplane perpendicular to the chord joining the point
where n is evaluated to the origin. The resultant vector is then scaled by a factor of x2.
In a conformally-invariant theory, both the antisymmetric part of T (n) and its trace
are total divergences. These can therefore be removed to leave a new tensor T 0(n) which
is symmetric and traceless. The complete set of divergenceless tensors is then given by
fT 0(x), T 0(n), xT 0(n)x, J 0(n) T 0(n) ^ xg (6:63)
This yields a set of 1 + 4 + 4 + 6 = 15 conserved quantities | the dimension of the
conformal group. All this is well known, of course, but it is the
rst time that geometric
algebra has been systematically applied to this problem. It is therefore instructive to
see how geometric algebra is able to simplify many of the derivations, and to generate a
clearer understanding of the results.
6.3 Applications
We now look at a number of applications of the formalism established in the preceding
sections. We start by illustrating the techniques with the example of electromagnetism.
This does not produce any new results, but does lead directly to the STA forms of the
Maxwell equations and the electromagnetic stress-energy tensor. We next consider Dirac
theory, and a number of new conjugate currents will be identi
ed. A study of coupled
Maxwell-Dirac theory then provides a useful analogue for the discussion of certain aspects
of a gauge theory of gravity, as described in Chapter 7. The
nal application is to a two-
particle action which recovers the
eld equations discussed in Section 4.4.
The essential result needed for what follows is
@a hb rM i = a @ h(b M i
= a bP(M ) (6.64)
where P (M ) is the projection of M onto the grades contained in . It is the result (6.64)
that enables all calculations to be performed without the introduction of a frame. It
is often the case that the Lagrangian can be written in the form L(i ri), when the
following result is useful:
@a hrM i = @a hb rM@bi
= a bP (M@b)
= P (Ma): (6.65)
1. Electromagnetism
The electromagnetic Lagrangian density is given by
L = ;A J + 21 F F (6:66)
128
where A is the vector potential, F = r^A, and A couples to an external current J which
is not varied. To
nd the equations of motion we
rst write F F as a function of rA,
F F = 41 h(rA ; (rA)~)2i
= 12 hrArA ; rA(rA)~i: (6.67)
The
eld equations therefore take the form
;J ; @b r 21 hrAb ; (rA)~bi1 = 0
) ;J ; @b rF b = 0
) r F = J: (6.68)
This is combined with the identity r^F = 0 to yield the full set of Maxwell's equations,
rF = J .
To calculate the free-
eld stress-energy tensor, we set J = 0 in (6.66) and work with
L0 = 12 hF 2i: (6:69)
Equation (6.26) now gives the stress-energy tensor in the form
_ AF
T (n) = rh _ ni ; 21 nhF 2i: (6:70)
This expression is physically unsatisfactory as is stands, because it is not gauge-invariant.
In order to
nd a gauge-invariant form of (6.70), we write 60]
_ AF
rh _ ni = (r^ A) (F n) + (F n) rA
= F (F n) ; (F r_ ) nA_ (6.71)
and observe that, since rF = 0, the second term is a total divergence and can therefore
be ignored. What remains is
T em(n) = F (F n) ; 21 nF F
= 21 FnF ~ (6.72)
which is the form of the electromagnetic stress-energy tensor obtained by Hestenes 17].
The tensor (6.72) is gauge-invariant, traceless and symmetric. The latter two properties
follow simultaneously from the identity
@aT em(a) = @a 21 FaF~ = 0: (6:73)
The angular momentum is obtained from (6.38), which yields
J (n) = (rh _ AFn
_ i ; 21 nhF 2i) ^ x + A ^ (F n) (6:74)
where we have used the stress-energy tensor in the form (6.70). This expression suers
from the same lack of gauge invariance, and is
xed up in the same way, using (6.71) and
; (F n) ^ A + x ^ (F r_ ) nA_ ] = x ^ (F r
$
) nA] (6:75)
129
which is a total divergence. This leaves simply
J (n) = T em(n) ^ x: (6:76)
By rede
ning the stress-energy tensor to be symmetric, the spin contribution to the
angular momentum is absorbed into (6.72). For the case of electromagnetism this has the
advantage that gauge invariance is manifest, but it also suppresses the spin-1 nature of
the
eld. Suppressing the spin term in this manner is not always desirable, as we shall
see with the Dirac equation.
The free-
eld Lagrangian (6.69) is not only Poincare-invariant! it is invariant under
the full conformal group of spacetime 7, 77]. The full set of divergenceless tensors for
free-
eld electromagnetism is therefore T em(x), T em(n), xT em(n)x, and T em(n) ^ x. It is a
simple matter to calculate the modi
ed conservation equations when a current is present.
2. Dirac Theory1
The multivector derivative is particularly powerful when applied to the STA form of the
Dirac Lagrangian. We recall from Chapter 5 that the Lagrangian for the Dirac equation
can be written as (4.96)
L = hri3~ ; eA0~ ; m~i (6:77)
where is an even multivector and A is an external gauge
eld (which is not varied). To
verify that (6.77) does give the Dirac equation we use the Euler-Lagrange equations in
the form
@ L = @a r(@a L) (6:78)
to obtain
(ri3)~ ; 2e0A~ ; 2m~ = @a r(i3a ~)
= i3~ r :
(6.79)
Reversing this equation, and postmultiplying by 0, we obtain
ri3 ; eA = m0 (6:80)
as found in Chapter 4 (4.92). Again, it is worth stressing that this derivation employs a
genuine calculus, and does not resort to treating and ~ as independent variables.
We now analyse the Dirac equation from the viewpoint of the Lagrangian (6.77).
In this Section we only consider position-independent transformations of the spinor .
Spacetime transformations are studied in the following section. The transformations we
are interested in are of the type
0 = eM (6:81)
where M is a general multivector and and M are independent of position. Operations
on the right of arise naturally in the STA formulation of Dirac theory, and can be
1 The basic idea developed in this section was provided by Anthony Lasenby.
130
thought of as generalised gauge transformations. They have no such simple analogue in
the standard column-spinor treatment. Applying (6.17) to (6.81), we obtain
rhMi3~i1 = @L0j=0 (6:82)
which is a result that we shall exploit by substituting various quantities for M . If M is
odd both sides vanish identically, so useful information is only obtained when M is even.
The
rst even M to consider is a scalar, , so that hMi3~i1 is zero. It follows that
@ e2L =0 = 0
) L = 0 (6.83)
and hence that, when the equations of motion are satis
ed, the Dirac Lagrangian vanishes.
Next, setting M = i, equation (6.82) gives
;r (s) = ;m@he2iei
i=0
) r (s) = ;2m sin
(6.84)
where s = 3~ is the spin current. This equation is well-known 33], though it is not
usually observed that the spin current is the current conjugate to duality rotations. In
conventional versions, these would be called \axial rotations", with the role of i taken by
5. In the STA approach, however, these rotations are identical to duality transformations
for the electromagnetic
eld. The duality transformation generated by ei is also the
continuous analogue of the discrete symmetry of mass conjugation, since 7! i changes
the sign of the mass term in L. It is no surprise, therefore, that the conjugate current,
s, is conserved for massless particles.
Finally, taking M to be an arbitrary bivector B yields
r (B (i3)~) = D2hriB 3~ ; eAB E0~i
= eA(3B3 ; B )0~ (6.85)
where the Dirac equation (6.80) has beed used. Both sides of(6.85) vanish for B = i1 i2
and 3, with useful equations arising on taking B = 1 2 and i3. The last of these,
B = i3, corresponds to the usual U (1) gauge transformation of the spinor
eld, and gives
r J = 0 (6:86)
where J = 0~ is the current conjugate to phase transformations, and is strictly con-
served. The remaining transformations generated by e 1 and e 2 give
r (e1) = 2eA e2 (6:87)
r (e2) = ;2eA e1
where e = ~. Although these equations have been found before 33], the role of
e1 and e2 as currents conjugate to right-sided e 2 and e 1 transformations has not
been noted. Right multiplication by 1 and 2 generates charge conjugation, since the
transformation 7! 0 1 takes (6.80) into
r0i3 + eA0 = m00 : (6:88)
It follows that the conjugate currents are conserved exactly if the external potential van-
ishes, or the particle has zero charge.
131
3. Spacetime Transformations in Maxwell-Dirac Theory
The canonical stress-energy and angular-momentum tensors are derived from spacetime
symmetries. In considering these it is useful to work with the full coupled Maxwell-Dirac
Lagrangian, in which the free-
eld term for the electromagnetic
eld is also included. This
ensures that the Lagrangian is Poincare-invariant. The full Lagrangian is therefore
L = hri3~ ; eA0~ ; m~ + 21 F 2i (6:89)
in which both and A are dynamical variables.
From the de
nition of the stress-energy tensor (6.26) and the fact that the Dirac part
of the Lagrangian vanishes when the
eld equations are satis
ed (6.83), T (n) is given by
T (n) = rh _ 3n
_ i ~ i + rh
_ AFn
_ i ; 21 nF F: (6:90)
Again, this is not gauge-invariant and a total divergence must be removed to recover a
gauge-invariant tensor. The manipulations are as at (6.71), and now yield
T md(n) = rh _ 3n
_ i ~ i ; n JA + 21 FnF
~ (6:91)
where J = 0~. The tensor (6.91) is now gauge-invariant, and conservation can be
checked simply from the
eld equations. The
rst and last terms are the free-
eld stress-
energy tensors and the middle term, ;nJA, arises from the coupling. The stress-energy
tensor for the Dirac theory in the presence of an external
eld A is conventionally de
ned
by the
rst two terms of (6.91), since the combination of these is gauge-invariant.
Only the free-
eld electromagnetic contribution to T md (6.91) is symmetric! the other
terms each contain antisymmetric parts. The overall antisymmetric contribution is
T;(n) = ;
1 T (n) T (n)]
2 md md
= ^ ; r^h
1 n A J
2
_ _ 3~ 1]
i i
= h ; r
1 n AJ
2 i3~ + _ i
rh_ 3~ 3 2
ii
= r
n ( ( 41 is))
= ; r^
n ( i ( 41 s)) (6.92)
and is therefore completely determined by the exterior derivative of the spin current 78].
The angular momentum is found from (6.39) and, once the total divergence is removed,
the gauge-invariant form is
J (n) = T md(n) ^ x + 12 is ^ n: (6:93)
The ease of derivation of J (n) (6.93) compares favourably with traditional operator-based
approaches 60]. It is crucial to the identi
cation of the spin contribution to the angular
momentum that the antisymmetric component of T md(n) is retained. In (6.93) the spin
term is determined by the trivector is, and the fact that this trivector can be dualised to
the vector s is a unique property of four-dimensional spacetime.
132
The sole term breaking conformal invariance in (6.89) is the mass term hm~i, and
it is useful to consider the currents conjugate to dilations and special conformal transfor-
mations, and to see how their non-conservation arises from this term. For dilations, the
conformal weight of a spinor
eld is 32 , and equation (6.48) yields the current
jd = T md(x) (6:94)
(after subtracting out a total divergence). The conservation equation is
r jd = @a T md(a)
= hm~i: (6.95)
Under a spacetime transformation the A
eld transforms as
A(x) 7! A0(x) f A(x0)] (6:96)
where x0 = f (x). For a special conformal transformation, therefore, we have that
A0(x) = (1 + nx);1 A(x0)(1 + xn);1: (6:97)
Since this is a rotation/dilation, we postulate for the single-sided transformation
0(x) = (1 + nx);2 (1 + xn);1(x0): (6:98)
In order to verify that the condition (6.56) can be satis
ed, we need the neat result that
r (1 + nx);2 (1 + xn);1 = 0: (6:99)
This holds for all vectors n, and the bracketed term is immediately a solution of the
massless Dirac equation (it is a monogenic function on spacetime). It follows from (6.56)
that the conserved tensor is
T SC (n) = ;T md(xnx) ; n (ix ^ (s)): (6:100)
and the conservation equation is
r T SC (xnx) = ;2hm~in x: (6:101)
In both (6.95) and (6.101) the conjugate tensors are conserved as the mass goes to zero,
as expected.
133
where is a function of position in the 8-dimensional con
guration space, and r1 and
r2 are the vector derivatives in their respective spaces. The action is
Z
S = jd8xjL: (6:103)
If we de
ne the function h by
h(a) = i1 am i1 + i2 am i2
1 2
(6:104)
where i1 and i2 are the pseudoscalars for particle-one and particle-two spaces respectively,
then we can write the action as
Z
S = jd8xjhh(@b)b rJ (01 + 02)~ ; 2~i: (6:105)
Here r = r1 + r2 is the vector derivative in 8-dimensional con
guration space. The
eld
equation is
@ L = @a r(@a L) (6:106)
which leads to
h(@a)a rJ (01 + 02)]~ ; 4~ = @a rJ (01 + 02)h
~ (a)]: (6:107)
The reverse of this equation is
h(@a)a rJ (01 + 02) = 2 (6:108)
and post-multiplying by (01 + 02) obtains
r + r )J = ( 1 + 2)
(m
1 2
(6:109)
1 m2 0 0
as used in Section 4.4.
The action (6.102) is invariant under phase rotations in two-particle space,
7! 0 e;J (6:110)
and the conserved current conjugate to this symmetry is
j = @a(;J ) @aL
~ (a)i
= @ahE (01 + 02)h
= hhE (01 + 02)~i1: (6.111)
This current satis
es the conservation equation
r j = 0 (6:112)
or, absorbing the factor of E into ,
r + r ) h( 1 + 2)~i = 0:
(m
1 2
(6:113)
1 m2 0 0 1
135
For a
xed grade-r multivector Ar, we can now write
@h(a)h(Ar) = @h(a)hh(Ar)Xr i@Xr
= hh(a Ar)Xr i1@Xr
= (n ; r + 1)h(a Ar) (6.122)
where n is the dimension of the space and a result from page 58 of 24] has been used.
Equation (6.121) can be used to derive formulae for the functional derivative of the
adjoint. The general result is
@h(a)h(Ar ) = @h(a)hh(Xr )Ar i@Xr
= hh(a X_ r )Ari1@_Xr : (6.123)
When A is a vector, this admits the simpler form
@h(a)h(b) = ba: (6:124)
If h is a symmetric function then h = h, but this cannot be exploited for functional
dierentiation, since h and h are independent for the purposes of calculus.
As two
nal applications, we derive results for the functional derivative of the deter-
minant (1.115) and the inverse function (1.125). For the determinant, we
nd that
@h(a)h(I ) = h(a I )
) @h(a) det(h) = h(a I )I ;1
= det(h)h;1(a) (6.125)
where we have recalled the de
nition of the inverse (1.125). This coincides with standard
formulae for the functional derivative of the determinant by its corresponding tensor. The
proof given here, which follows directly from the de
nitions of the determinant and the
inverse, is considerably more concise than any available to conventional matrix/tensor
methods. The result for the inverse function is found from
;1
@h(a)hh(Br )h;1(Ar )i = hh(a Br)h;1(Ar )i1 + @_h(a)hh_ (Ar)h(Br )i = 0 (6:126)
from which it follows that
@h(a)hh;1(Ar )Br i = ;hha h;1(Br )]h;1(Ar)i1
= ;hh;1 (a) Brh;1(Ar)i1 (6.127)
where use has been made of results for the adjoint (1.123).
We have now assembled most of the necessary formalism and results for the application
of geometric calculus to
eld theory. In the
nal chapter we apply this formalism to
develop a gauge theory of gravity.
136
Chapter 7
Gravity as a Gauge Theory
In this chapter the formalism described throughout the earlier chapters of this thesis is
employed to develop a gauge theory of gravity. Our starting point is the Dirac action, and
we begin by recalling how electromagnetic interactions arise through right-sided transfor-
mations of the spinor
eld . We then turn to a discussion of Poincare invariance, and
attempt to introduce gravitational interactions in a way that closely mirrors the intro-
duction of the electromagnetic sector. The new dynamical
elds are identi
ed, and an
action is constructed for these. The
eld equations are then found and the derivation
of these is shown to introduce an important consistency requirement. In order that the
correct minimally-coupled Dirac equation is obtained, one is forced to use the simplest
action for the gravitational
elds | the Ricci scalar. Some free-
eld solutions are ob-
tained and are compared with those of general relativity. Aspects of the manner in which
the theory employs only active transformations are then illustrated with a discussion of
extended-matter distributions.
By treating gravity as a gauge theory of active transformations in the (at) spacetime
algebra, some important dierences from general relativity emerge. Firstly, coordinates
are unnecessary and play an entirely secondary role. Points are represented by vectors,
and all formulae given are coordinate-free. The result is a theory in which spacetime does
not play an active role, and it is meaningless to assign physical properties to spacetime.
The theory is one of forces, not geometry. Secondly, the gauge-theory approach leads
to a
rst-order set of equations. Despite the fact that the introduction of a set of coor-
dinates reproduces the matter-free
eld equations of general relativity, the requirement
that the
rst-order variables should exist globally has important consequences. These are
illustrated by a discussion of point-source solutions.
There has, of course, been a considerable discussion of whether and how gravity can
be formulated as a gauge theory. The earliest attempt was by Utiyama 79], and his
ideas were later re
ned by Kibble 80]. This led to the development of what is now
known as the Einstein-Cartan-Kibble-Sciama (ECKS) theory of gravity. A detailed review
of this subject was given in 1976 by Hehl et al. 81]. More recently, the
bre-bundle
approach to gauge theories has been used to study general relativity 82]. All these
developments share the idea that, at its most fundamental level, gravity is the result
of spacetime curvature (and, more generally, of torsion). Furthermore, many of these
treatments rely on an uncomfortable mixture of passive coordinate transformations and
137
active tetrad transformations. Even when active transformations are emphasised, as by
Hehl et al., the transformations are still viewed as taking place on an initially curved
spacetime manifold. Such ideas are rejected here, as is inevitable if one only discusses the
properties of spacetime
elds, and the interactions between them.
elds. We
rst consider translations
(x) 7! 0(x) (x0) (7:18)
where
x0 = x + a (7:19)
and a is a constant vector. To make these translations local, the vector a must become a
function of position. This is achieved by replacing (7.19) with
x0 = f (x) (7:20)
where f (x) is now an arbitrary mapping between spacetime positions. We continue to
refer to (7.20) as a translation, as this avoids overuse of the word \transformation". It
is implicit in what follows that all translations are local and are therefore determined by
completely arbitrary mappings. The translation (7.20) has the interpretation that the
eld has been physically moved from the old position x0 to the new position x. The
same holds for the observables formed from , for example the current J (x) = 0~ is
transformed to J 0(x) = J (x0).
As it stands, the translation de
ned by (7.20) is not a symmetry of the action, since
r0(x) = r(f (x))
= f (rx )(x0)0 (7.21)
and the action becomes
Z
S 0 = jd4x0j (det f );1hf (rx )0i3~0 ; m0~0i:
0 (7:22)
To recover a symmetry from (7.20), one must introduce an arbitrary, position-dependent
linear function h. The new action is then written as
Z
Sh = jd4xj(det h);1 hh(r)i3~ ; m~i: (7:23)
140
Under the translation
(x) 7! 0(x) (f (x)) (7:24)
the action Sh transforms to
Z
Sh = jd4x0j(det f );1(det h0);1hh0 f (rx )0i3~0 ; m0~0i
0
0 (7:25)
and the original action is recovered provided that h has the transformation law
hx 7! h0x hx f ;x 1
0 where x0 = f (x): (7:26)
This is usually the most useful form for explicit calculations, though alternatively we can
write
hx 7! h0x hx f x
0 0 where x = f (x0) (7:27)
which is sometimes more convenient to work with.
In arriving at (7.23) we have only taken local translations into account | the currents
are being moved from one spacetime position to another. To arrive at a gauge theory
of the full Poincare group we must consider rotations as well. (As always, the term
\rotation" is meant in the sense of Lorentz transformation.) In Chapter 6, rotational
invariance of the Dirac action (7.1) was seen by considering transformations of the type
(x) 7! R0(R~ 0xR0), where R0 is a constant rotor. By writing the action in the form
of (7.23), however, we have already allowed for the most general type of transformation
of position dependence. We can therefore translate back to x, so that the rotation takes
place at a point. In doing so, we completely decouple rotations and translations. This is
illustrated by thinking in terms of the frame of observables fe g. Given this frame at a
point x, there are two transformations that we can perform on it. We can either move
it somewhere else (keeping it in the same orientation with respect to the f g frame),
or we can rotate it at a point. These two operations correspond to dierent symmetries.
A suitable physical picture might be to think of \experiments" in place of frames. We
expect that the physics of the experiment will be unaected by moving the experiment
to another point, or by changing its orientation in space.
Active rotations of the spinor observables are driven by rotations of the spinor
eld,
7! 0 R0: (7:28)
Since h(a) is a spacetime vector
eld, the corresponding law for h must be
h(a) 7! h0(a) R0h(a)R~ 0: (7.29)
By writing the action (7.23) in the form
Z
Sh = jd4xj(det h);1 hh(@a)a ri3~ ; m~i (7:30)
we observe that it is now invariant under the rotations de
ned by (7.28) and (7.29). All
rotations now take place at a point, and the presence of the h
eld ensures that a rotation
141
at a point has become a global symmetry. To make this symmetry local, we need only
replace the directional derivative of by a covariant derivative, with the property that
Da0 (R) = RDa (7:31)
where R is a position-dependent rotor. But this is precisely the problem that was tackled
at the start of this section, the only dierence being that the rotor R now sits to the left
of . Following the same arguments, we immediately arrive at the de
nition:
Da (a r + 21 +(a)) (7:32)
where +(a) is a (position-dependent) bivector-valued linear function of the vector a. Under
local rotations 7! R, +(a) transforms to
+(a) 7! +(a)0 R+(a)R~ ; 2a rRR: ~ (7:33)
Under local translations, +(a) must transform in the same manner as the a rRR~ term,
so
+x(a) 7! +x f x(a)
0 if x0 = f (x) (7:34)
+ (a) 7! + f (a) if x = f (x0):
x x x
0
;1
0
144
From (7.60) we write the Lagrangian density as
LG = 21 R det h;1 = LG (h(a) +(a) b r+(a)): (7:61)
The action integral (7.60) is over a region of at spacetime, so all the variational princi-
ple techniques developed in Chapter 6 hold without modi
cation. The only elaboration
needed is to de
ne a calculus for +(a). Such a calculus can be de
ned in precisely the
same way as the derivative @h(a) was de
ned (6.117). The essential results are:
@(a)h+(b)B i = a bB (7.62)
@(b)a hc r+(d)B i = a cb dB (7.63)
where B is an arbitrary bivector.
We assume that the overall action is of the form
L = LG ; LM (7:64)
where LM describes the matter content and = 8G. The
rst of the
eld equations is
found by varying with respect to h, producing
@h(a)LM = 21 @h(a)(hh(@b ^ @c)R(c ^ b)i det h;1 )
= (R(a) ; 21 h;1(a)R) det h;1: (7.65)
The functional derivative with respect to h(a) of the matter Lagrangian is taken to de
ne
the stress-energy tensor of the matter
eld through
T h;1(a) det h;1 @h(a)LM (7:66)
so that we arrive at the
eld equations in the form
R(a) ; 12 h;1(a)R = T h;1 (a): (7:67)
It is now appropriate to de
ne the functions
R(a ^ b) Rh(a ^ b) (7.68)
R(a) Rh(a) = @a R(a ^ b) (7.69)
G R(a) ; 12 aR: (7.70)
These are covariant under translations (they simply change their position dependence),
and under rotations they transform as e.g.
R(B ) 7! R0R(R~ 0BR0)R~0: (7:71)
Equation (7.71) is the de
ning rule for the transformation properties of a tensor, and
we hereafter refer to (7.68) through to (7.70) as the Riemann, Ricci and Einstein tensors
respectively. We can now write (7.67) in the form
G (a) = T (a) (7:72)
145
which is the (at-space) gauge-theory equivalent of Einstein's
eld equations.
In the limit of vanishing gravitational
elds (h(a) 7! a and +(a) 7! 0) the stress-
energy tensor de
ned by (7.66) agrees with the canonical stress-energy tensor (6.24), up
to a total divergence. When +(a) vanishes, the matter action is obtained from the free-
= ; 12 ih(a) ^ s: (7.83)
It follows immediately that the contraction in (7.82) vanishes, since
h;1 (@a) (ih(a) ^ s) = ;i@a ^ a ^ s = 0: (7:84)
We de
ne S by
S 21 i3~ (7:85)
so that we can now write
S (a) = h(a) S : (7:86)
Given that (7.81) does hold, we can now write (7.76) in the form
S (a) = h_ (r) ^ h_ (a) + +(b) h(@b) ^ h(a) ; +(b) h(@b) ^ h(a)
= h(@b) ^ b rh(a) + +(b) h(a)
= D^ h(a): (7.87)
The right-hand side of this equation could be viewed as the torsion though, since we are
working in a at spacetime, it is preferable to avoid terminology borrowed from Riemann-
Cartan geometry. When the left-hand side of (7.87) vanishes, we arrive at the simple
equation
D^ h(a) = 0 (7:88)
valid for all constant vectors a. All dierential functions f (a) = arf (x) satisfy r^f (a) =
0, and (7.88) can be seen as the covariant generalisation of this result. Our gravitational
148
and the
eld equations are
G (a) = ha Di3~i1 (7.97)
D^ h(a) = h(a) ( 12 i3~) = h(a) ( 21 is) (7.98)
Di3 = m0: (7.99)
It is not clear that self-consistent solutions to these equations could correspond to any
physical situation, as such a solution would describe a self-gravitating Dirac uid. Self-
consistent solutions have been found in the context of cosmology, however, and the solu-
tions have the interesting property of forcing the universe to be at critical density 10].
The Electromagnetic Field Equations
We now return to the introduction of the electromagnetic
eld. From the action (7.91),
and following the procedure of the start of this chapter, we arrive at the action
Z
SD+EM = jd4xj(det h);1 hh(@a)(Dai3~ ; ea A0~) ; m~i: (7:100)
The
eld equation from this action is
Di3 ; eA = m0 (7:101)
where we have introduced the notation
A = h(A): (7:102)
It is to be expected that A should appear in the
nal equation, rather than A. The
vector potential A originated as the generalisation of the quantity r. If we examine
what happens to this under the translation (x) 7! (x0), with x0 = f (x), we
nd that
r 7! f (rx (x0)):
0 (7.103)
It follows that A must also pick up a factor of f as it is moved from x0 to x,
A(x) 7! f (A(x0)) (7.104)
so it is the quantity A that is Poincare-covariant, as are all the other quantities in equa-
tion (7.101). However, A is not invariant under local U (1) transformations. Instead, we
must construct the Faraday bivector
F = r^ A: (7:105)
It could be considered a weakness of conventional spin-torsion theory that, in order to
construct the gauge-invariant quantity F , one has to resort to the use of the at-space
vector derivative. Of course, in our theory background spacetime has never gone away,
and we are free to exploit the vector derivative to the full.
The conventional approach to gauge theories of gravity (as discussed in 81], for ex-
ample) attempts to de
ne a minimal coupling procedure for all matter
elds, preparing
149
the way for a true curved-space theory. The approach here has been rather dierent, in
that everything is derived from the Dirac equation, and we are attempting to put electro-
magnetic and gravitational interactions on as similar a footing as possible. Consequently,
there is no reason to expect that the gravitational
eld should \minimally couple" into the
electromagnetic
eld. Instead, we must look at how F behaves under local translations.
We
nd that
F (x) 7! r^ fA(x0) = f (rx ^ A(x0))
0 (7.106)
= fF (x0) (7.107)
so the covariant form of F is
F h(F ): (7:108)
F is covariant under local Poincare transformations, and invariant under U (1) transfor-
mations. The appropriate action for the electromagnetic
eld is therefore
Z
SEM = jd4xj(det h);1 h 12 FF ; A J i (7:109)
which reduces to the standard electromagnetic action integral in the limit where h is the
identity. To
nd the electromagnetic
eld equations, we write
LEM = det h;1h 12 FF ; A J i = L(A a rA) (7:110)
and treat the h and J
elds as external sources. There is no +-dependence in (7.109), so
LEM satis
es the criteria of equation (7.82).
Variation of LEM with respect to A leads to the equation
@a r(a h(F ) det h;1) = r hh(r^ A) det h;1 = det h;1 J (7:111)
which combines with the identity
r^ F = 0 (7:112)
to form the Maxwell equations in a gravitational background. Equation (7.111) corre-
sponds to the standard second-order wave equation for the vector potential A used in
general relativity. It contains only the functions hh g;1 and det h;1 = (det g)1=2,
where g is the symmetric \metric" tensor. The fact that equation (7.111) only involves h
through the metric tensor is usually taken as evidence that the electromagnetic
eld does
not couple to torsion.
So far, we only have the Maxwell equations as two separate equations (7.111) and
(7.112). It would be very disappointing if our STA approach did not enable us to do better
since one of the many advantages of the STA is that, in the absence of a gravitational
150
This is more than a mere notational convenience. The r operator is invertible, and can be
used to develop a
rst-order propagator theory for the F -
eld 8]. This has the advantages
of working directly with the physical
eld, and of correctly predicting the obliquity factors
that have to be put in by hand in the second-order approach (based on a wave equation
for A). It would undermine much of the motivation for pursuing
rst-order theories if
this approach cannot be generalised to include gravitational eects. Furthermore, if we
construct the stress-energy tensor, we
nd that
TEM h;1(a) det h;1 = 21 @h(a)hh(F )h(F ) det h;1i
= h(a F ) F ; 12 h;1(a)F F det h;1 (7.115)
which yields
TEM (a) = ;(F a) F ; 21 aF F
= ; 12 F aF : (7.116)
This is the covariant form of the tensor found in Section (6.2). It is intersting to see how
the de
nition of TEM as the functional derivative of L with respect to @h(a) automatically
preserves gauge invariance. For electromagnetism this has the eect of forcing TEM to
be symmetric. The form of TEM (7.116) makes it clear that it is F which is the genuine
physical
eld, so we should seek to express the
eld equations in terms of this object. To
achieve this, we
rst write the second of the
eld equations (7.90) in the form
D^ h(a) = h(r^ a) + h(a) S (7:117)
which holds for all a. If we now de
ne the bivector B = a ^ b, we
nd that
D^ h(B ) = D^ h(a)] ^ h(b) ; h(a) ^D^ h(b)
= h(r^ a) ^ h(b) ; h(a) ^ h(r^ b) + (h(a) S ) ^ h(b)
;h(a) ^ (h(b) S )
= h(r^ B ) ; h(B ) S (7.118)
which is used to write equation (7.112) in the form
D^F ; SF = h(r^ F ) = 0: (7:119)
Next, we use a double-duality transformation on (7.111) to write the left-hand side as
r (h(F ) det h;1) = ir^ (ih(F ) det h;1)
= ir^ (h;1(iF ))
= ih;1 (D^ (iF ) + (iF ) S ) (7.120)
so that (7.111) becomes
DF ; SF = ih(Ji) det h;1 = h;1(J ): (7:121)
151
Writing
J = h;1 (J ) (7:122)
we can now combine (7.119) and (7.121) into the single equation
DF ; SF = J (7:123)
which achieves our objective. The gravitational background has led to the vector deriva-
tive r being generalised to D ; S . Equation (7.123) surely deserves considerable study.
In particular, there is a clear need for a detailed study of the Green's functions of the
D ; S operator. Furthermore, (7.123) makes it clear that, even if the A equation does
not contain any torsion term, the F equation certainly does. This may be of importance
in studying how F propagates from the surface of an object with a large spin current.
152
and under an active rotation g(a) is unchanged. The fact that g(a) is unaected by active
rotations limits its usefulness, and this is a strong reason for not using the metric tensor
as the foundation of our theory.
The comparison with general relativity is clari
ed by the introduction of a set of 4
coordinate functions over spacetime, x = x (x). From these a coordinate frame is de
ned
by
e = @ x (7:129)
where @ = @x . The reciprocal frame is de
ned as
e = rx (7:130)
and satis
es
e e = (@ x) rx = @x x = : (7:131)
From these we de
ne a frame of \contravariant" vectors
g = h;1(e ) (7:132)
and a dual frame of \covariant" vectors
g = h(e ): (7:133)
These satisfy (no torsion)
g g = (7.134)
D^ g = 0 (7.135)
and
g Dg ; g Dg = 0: (7:136)
The third of these identities is the at-space equivalent of the vanishing of the Lie bracket
for a coordinate frame in Riemannian geometry.
From the fg g frame the metric coecients are de
ned by
g = g g (7:137)
which enables us to now make contact with Riemannian geometry. Writing + for +(e ),
we
nd from (7.125) that
2+ = g ^ (@ g) + g ^ g
@
g : (7:138)
The connection is de
ned by
; = g (D g) (7:139)
so that, with a a g,
@ a ; ; a = @ (a g) ; a (D g)
= g (D a) (7.140)
153
as required | the connection records the fact that, by writing a = a g , additional
x-dependence is introduced through the g.
By using (7.138) in (7.139), ; is given by
; = 21 g (@ g + @g ; @g ) (7:141)
which is the conventional expression for the Christoel connection. In the absence of
spin, the introduction of a coordinate frame unpackages our equations to the set of scalar
equations used in general relativity. The essential dierence is that in GR the quantity g
is fundamental, and can only be de
ned locally, whereas in our theory the fundamental
variables are the h and +
elds, which are de
ned globally throughout spacetime. One
might expect that the only dierences that could show up from this shift would be due
to global, topological considerations. In fact, this is not the case, as is shown in the
following sections. The reasons for these dierences can be either physical, due to the
dierent understanding attached to the variables in the theory, or mathematical, due often
to the constraint that the metric must be generated from a suitable h function. It is not
always the case that such an h function can be found, as is demonstrated in Section 7.2.1.
The ability to develop a coordinate-free theory of gravity oers a number of advantages
over approaches using tensor calculus. In particular, the physical content of the theory
is separated from the artefacts of the chosen frame. Thus the h and +
elds only dier
from the identity and zero in the presence of matter. This clari
es much of the physics
involved, as well as making many equations easier to manipulate.
Many of the standard results of classical Riemannian geometry have particularly simple
expressions in this STA-based theory. Similar expressions can be found in Chapter 5 of
Hestenes & Sobczyk 24], who have developed Riemannian geometry from the viewpoint
of geometric calculus. All the symmetries of the Riemann tensor are summarised in the
single equation
@a ^R(a ^ b) = 0: (7:142)
This says that the trivector @a ^R(a ^ b) vanishes for all values of the vector b, and so
represents a set of 16 scalar equations. These reduce the 36-component tensor R(B ) to a
function with only 20 degrees of freedom | the correct number for Riemannian geometry.
Equation (7.142) can be contracted with @b to yield
@a ^R(a) = 0 (7:143)
which says that the Ricci tensor is symmetric. The Bianchi identity is also compactly
written:
_ R_ (B ) = 0
D^ (7:144)
where the overdot notation is de
ned via
D_ T_ (M ) DT (M ) ; @aT (a DM ): (7:145)
Equation (7.144) can be contracted with @b ^ @a to yield
_ R_ (a ^ b) = @b R_ (D^ _ b) ; D^
_ R_ (b)
(@b ^ @a) D^
= ;2R_ (D_ ) + DR = 0: (7.146)
154
It follows that
G_ (D_ ) = 0 (7:147)
which, in conventional terms, represents conservation of the Einstein tensor. Many other
results can be written equally compactly.
The inclusion of torsion leads us to a comparison with the ECKS theory, which is
certainly closest to the approach adopted here. The ECKS theory arose from attempts
to develop gravity as a gauge theory, and modern treatments do indeed emphasise active
transformations 81]. However, the spin-torsion theories ultimately arrived at all involve
a curved-space picture of gravitational interactions, even if they started out as a gauge
theory in at space. Furthermore, the separation into local translations and rotations is
considerably cleaner in the theory developed here, as all transformations considered are
155
which contracts to give
R = 0: (7:152)
Our
eld equations are therefore
D^ h(a) = h(r^ a) (7:153)
R(a) = 0:
As was discussed in the previous section, if we expand in a basis then the equations for
the coordinates are the same as those of general relativity. It follows that any solution
to (7.153) will generate a metric which solves the Einstein equations. But the converse
does not hold | the additional physical constraints at work in our theory rule out certain
solutions that are admitted by general relativity. This is illustrated by a comparison of
the Schwarzschild metric used in general relativity with the class of radially-symmetric
static solutions admitted in the present theory. Throughout the following sections we use
units with G = 1.
159
It can now be con
rmed that @a R(a ^ b) = 0. Indeed, one can simultaneously check
both the
eld equations and the symmetry properties of R(B ), since R(a) = 0 and
@a ^R(a ^ b) = 0 combine into the single equation
@aR(a ^ b) = 0: (7:186)
This equation greatly facilitates the study of the Petrov classi
cation of vacuum solutions
to the Einstein equations, as is demonstrated in Chapter 3 of Hestenes & Sobczyk 24].
There the authors refer to @aR(a^b) as the contraction and @a^R(a^b) as the protraction.
The combined quantity @aR(a ^ b) is called simply the traction. These names have much
to recommend them, and are adopted wherever necessary.
Verifying that (7.185) satis
es (7.186) is a simple matter, depending solely on the
result that, for an arbitrary bivector B ,
@a(a ^ b + 3Ba ^ bB ;1) = @a(a ^ b + 3BabB ;1 ; 3Ba bB ;1)
= @a(a ^ b ; 3a b)
= @a(ab ; 4a b)
= 0: (7.187)
The compact form of the Riemann tensor (7.185), and the ease with which the
eld
equations are veri
ed, should serve to demonstrate the power of the STA approach to
relativistic physics.
Radially-Symmetric Gauge Transformations
From a given solution in the set (7.181) we can generate further solutions via radially-
symmetric gauge transformations. We consider Lorentz rotations
rst. All rotations
leave the metric terms g = g g unchanged, since these are de
ned by invariant inner
products, so g12 ; g22, f12 ; f22, f1g2 ; f2g1 and det h are all invariant. Since the
elds
are a function of x^et only, the only Lorentz rotations that preserve symmetry are those
that leave x ^ et unchanged. It is easily seen that these leave the Riemann tensor (7.185)
unchanged as well. There are two such transformations to consider! a rotation in the
e ^ e plane and a boost along the radial axis. The rotors that determine these are as
follows:
Rotation: R = exp((r)ieret=2)! (7.188)
Radial Boost: R = exp((r)eret=2): (7.189)
Both rotations leave +t untransformed, but introduce an +r and transform the + and
+ terms.
If we take the solution in the form (7.184) and apply a radial boost determined by the
rotor
R = exp M 2r r t
e e (7:190)
we arrive at the following, highly compact solution
h(a) = a + Mr a e;e;
+(a) = M r2 (e; ^ a + 2e; aeret) (7.191)
160
where
e; = et ; er : (7:192)
Both the forms (7.191) and (7.184) give a metric which, in GR, is known as the (advanced-
time) Eddington-Finkelstein form of the Schwarzschild solution,
ds2 = (1 ; 2M=r)dt2 ; (4M=r)dr dt ; (1 + 2M=r)dr2 ; r2(d 2 + sin2 d2): (7:193)
There are also two types of transformation of position dependence to consider. The
161
and
det f = u0(u=r)2: (7:207)
The new function hy = hx f ;1 has an additional dilation in the e e plane, and the
y
162
and
+T (a) = ; M2 et (e; ^ (etaet) + 2e; (etaet)er et) et
r
M
= r2 (2a e+eret ; e+ ^ a) (7.216)
where e+ = et + er . This new solution reproduces the metric of the retarded-time
Eddington-Finkelstein form of the Schwarzschild solution. Time reversal has therefore
switched us from a solution where particles can cross the horizon on an inward journey,
but cannot escape, to a solution where particles can leave, but cannot enter. Covariant
quantities, such as the
eld strength (7.169), are, of course, unchanged by time reversal.
From the gauge-theory viewpoint, it is natural that the solutions of the
eld equations
should fall into sets which are related by discrete transformations that are not simply con-
nected to the identity. The solutions are simply reproducing the structure of the Poincare
group on which the theory is constructed.
Behaviour near the Horizon
For the remainder of this section we restrict the discussion to solutions for which det h = 1.
For these the line element takes the form
ds2 = (1 ; 2M=r)dt2 ; (f1g2 ; f2g1)2dr dt ; (f12 ; f22)dr2
;r2(d 2 + sin2 d2): (7.217)
The horizon is at r = 2M , and at this distance we must have
g1 = g2: (7:218)
But, since det h = f1g1 ; f2g2 = 1, we must also have
f1g2 ; f2g1 = 1 at r = 2M (7:219)
so an o-diagonal term must be present in the metric at the horizon. The assumption that
this term can be transformed away everywhere does not hold in our theory. This resolves
the problem of the Schwarzschild discontinuity discussed at the start of this section. The
Schwarzschild metric does not give a solution that is well-de
ned everywhere, so lies
outside the set of metrics that are derivable from (7.181). Outside the horizon, however,
it is always possible to transform to a solution that reproduces the Schwarzschild line
element, and the same is true inside. But the transformations required to do this do
not mesh at the boundary, and their derivatives introduce -functions there. Because
the Schwarzschild line element is valid on either side of the horizon, it reproduces the
correct Riemann tensor (7.185) on either side. Careful analysis shows, however, that
the discontinuities in the + and +
elds required to reproduce the Schwarzschild line
element lead to -functions at the horizon in R(a ^ b).
The fact that the f1g2 ; f2g1 term must take a value of 1 at the horizon is interesting,
since this term changes sign under time-reversal (7.213). Once a horizon has formed, it
is therefore no longer possible to
nd an h such that the line element derived from it is
163
invariant under time reversal. This suggests that the f1g2 ; f2g1 term retains information
about the process by which the horizon formed | recording the fact that at some earlier
time matter was falling in radially. Matter infall certainly picks out a time direction,
and knowledge of this is maintained after the horizon has formed. This irreversibility is
apparent from the study of test particle geodesics 9]. These can cross the horizon to
the inside in a
nite external coordinate time, but can never get back out again, as one
expects of a black hole.
The above conclusions dier strongly from those of GR, in which the ultimate form
of the Schwarzschild solution is the Kruskal metric. This form is arrived at by a series
of coordinate transformations, and is motivated by the concept of \maximal extension"
| that all geodesics should either exist for all values of their ane parameter, or should
terminate at a singularity. None of the solutions presented here have this property. The
solution (7.191), for example, has a pole in the proper-time integral for outgoing radial
geodesics. This suggests that particles following these geodesics would spend an in
nite
coordinate time hovering just inside the horizon. In fact, in a more physical situation
this will not be the case | the eects of other particles will tend to sweep all matter
back to the centre. The solutions presented here are extreme simpli
cations, and there
is no compelling physical reason why we should look for \maximal" solutions. This is
important, as the Kruskal metric is time-reverse symmetric and so must fail to give a
globally valid solution in our theory. There are a number of ways to see why this happens.
For example, the Kruskal metric de
nes a spacetime with a dierent global topology to
at spacetime. We can reach a similar conclusion by studying how the Kruskal metric
is derived from the Schwarzschild metric. We assume, for the time being, that we are
outside the horizon so that a solution giving the Schwarzschild line element is
g1 = -1=2 g2 = 0 (7:220)
f1 = - ;1=2
f2 = 0
where
- = 1 ; 2M=r: (7:221)
The
rst step is to re-interpret the coordinate transformations used in general relativity
as active local translations. For example, the advanced Eddington-Finkelstein metric is
reached by de
ning
ty ; ry = t ; (r + 2M ln(r ; 2M )) (7.222)
ry = r (7.223)
or
xy = x ; 2M ln(r ; 2M )et (7:224)
which is now recognisable as a translation of the type of equation (7.194). The result of
this translation is the introduction of an f2y function
f2y = ; 2M ;1=2
r - (7:225)
which now ensures that f1yg2y ; f2yg1y = 1 at the horizon. The translation (7.224), which
is only de
ned outside the horizon, has produced a form of solution which at least has a
164
chance of being extended across the horizon. In fact, an additional boost is still required
to remove some remaining discontinuities. A suitable boost is de
ned by
R = exp(er et=2) (7:226)
where
sinh = 12 (-;1=2 ; -1=2) (7:227)
and so is also only de
ned outside the horizon. The result of this pair of transformations
is the solution (7.191), which now extends smoothly down to the origin.
In a similar manner, it is possible to reach the retarted-time Eddington-Finkelstein
metric by starting with the translation de
ned by
ty + ry = t + (r + 2M ln(r ; 2M )) (7.228)
ry = r: (7.229)
The Kruskal metric, on the other hand, is reached by combining the advance and retarded
coordinates and writing
ty ; ry = t ; (r + 2M ln(r ; 2M )) (7.230)
ty + ry = t + (r + 2M ln(r ; 2M )) (7.231)
which de
nes the translation
xy = x etet + (r + 2M ln(r ; 2M ))er : (7:232)
This translation is now of the type of equation (7.202), and results in a completely dif-
ferent form of solution. The transformed solution is still only valid for r > 2M , and the
transformation (7.232) has not introduced the required f1g2 ; f2g1 term. No additional
boost or rotation manufactures a form which can then be extended to the origin. The
problem can still be seen when the Kruskal metric is written in the form
ds2 = 32M
3
;r=2M
(dw2 ; dz2) ; r2(d 2 + sin2 d2)
r e (7:233)
where
z 2 ; w 2 = 2M 1 (r ; 2M )e;r=2M (7.234)
w = tanh t
z 4M (7.235)
which is clearly only de
ned for r > 2M . The loss of the region with r < 2M does not
present a problem in GR, since the r-coordinate has no special signi
cance. But it is a
problem if r is viewed as the distance from the source of the
elds, as it is in the present
theory, since then the
elds must be de
ned for all r. Even in the absence of torsion, the
at-space gauge-theory approach to gravity produces physical consequences that clearly
dier from general relativity, despite the formal mathematical similarities between the
two theories.
165
7.2.2 Kerr-Type Solutions
We now briey discuss how the Kerr class of solutions
t into the present theory. The
detailed comparisons of the previous section will not be reproduced, and we will simply
illustrate a few novel features. Our starting point is the Kerr line element in Boyer-
Lindquist form 87]
2.5
1.5
0.5
0
-3
-2
-1
0
y
1
-2 -3
2 0 -1
2 1
3 3 x
Figure 7.1: Incoming light paths for the Kerr solution I | view from above. The paths
terminate over a central disk in the z = 0 plane.
where n is a spatial vector. The explicit forms of n and can be found in Schier et al. 89]
and in Chapter 6 of \The mathematical theory of black holes" by S. Chandrasekhar 91].
These forms will not be repeated here. From the
eld equations it turns out that n
satis
es the equation 89]
n rn = 0: (7:248)
The integral curves of n are therefore straight lines, and these represent the possible
paths for incoming light rays. These paths are illustrated in
gures (7.1) and (7.2). The
paths terminate over a central disk, where the matter must be present. The fact that the
solution (7.246) must represent a disk of matter was originally pointed out by Kerr in a
footnote to the paper of Newman and Janis 88]. This is the paper that
rst gave the
derivation of the Kerr metric via a complex coordinate transformation. Kerr's observation
is ignored in most modern texts (see 91] or the popular account 92]) where it is claimed
that the solution (7.246) represents not a disk but a ring of matter | the ring singularity,
where the Riemann tensor is in
nite.
The transformations taking us from the solution (7.246) to solutions with a point
singularity involve the translation
f (x) = x0 x ; Lr x (i3) (7:249)
which implies that
(r0)2 = r2 + L2 cos2 : (7:250)
Only the points for which r satis
es
r jL cos j (7:251)
168
3
z 0
-1
-2
-3
-1
-0.5
0
y
0.5 -2 -3
1 0 -1
1 3 2 x
Figure 7.2: Incoming null geodesics for the Kerr solution II | view from side on.
are mapped onto points in the transformed solution, and this has the eect of cutting out
the central disk and mapping it down to a point. Curiously, the translation achieves this
whilst keeping the total mass
xed (i.e. the mass parameter M is unchanged). The two
types of solution (7.242) and (7.246) represent very dierent matter con
gurations, and
it is not clear that they can really be thought of as equivalent in anything but an abstract
mathematical sense.
eld) stress-energy tensors are no longer conserved. To recover a conservation law, one
must either replace directional derivatives by covariant derivatives, or realise that it is
only the total stress-energy tensor that is conserved. The same is true for gravity. Once
gravitational eects are turned on, the only quantity that one expects to be conserved is
the sum of the individual matter and gravitational stress-energy tensors. But the
eld
equations ensure that this sum is always zero, so conservation of total energy-momentum
ceases to be an issue.
If, however, a global time-like symmetry is present, one can still sensibly separate
the total (zero) energy into gravitational and matter terms. Each term is then separately
conserved with respect to this global time. For the case of the star, the total 4-momentum
is the sum of the individual uxes of 4-momentum in the et direction. We therefore de
ne
the conserved momentum P by Z
P = d3x T (et) (7:275)
and the total angular momentum J by
Z
J = d3x x ^T (et): (7:276)
Concentrating on P
rst, we
nd that
P = Mrotet (7:277)
where
ZR Z h i
Mrot = 2 dr d r2 sin (r) cosh2 !(r ) + p(r) sinh2 !(r ) : (7:278)
0 0
The eective mass Mrot reduces to M when the rotation vanishes, and rises with the
magnitude of !, showing that the internal energy of the star is rising. The total 4-
momentum is entirely in the et direction, as it should be. Performing the J integral next,
we obtain
ZR Z
J = ;i3 2 dr d r3 sin2 ((r) + p(r)) sinh !(r ) cosh !(r ) (7:279)
0 0
173
so the angular momentum is contained in the spatial plane de
ned by the ^et direc-
tion. Performing an active radial boost has generated a
eld con
guration with suitable
momentum and angular momentum properties for a rotating star.
Unfortunately, this model cannot be physical, since it does not tie down the shape of
the star | an active transformation can always be used to alter the shape to any desired
con
guration. The missing ingredient is that the particles making up the star must satisfy
their own geodesic equation for motion in the
elds due to the rest of the star. The simple
rotation (7.270) does not achieve this.
Attention is drawn to these points for the following reason. The boost (7.270) produces
a Riemann tensor at the surface of the star of
R(B ) = ; M2rrot3 B + 3er (cosh ! et + sinh ! ^)Ber(cosh ! et + sinh ! ^) (7:280)
which is that for a rotated Schwarzschild-type solution, with a suitably modi
ed mass.
This form is very dierent to the Riemann tensor for the Kerr solution (7.240), which
contains a complicated duality rotation. Whilst a physical model will undoubtedly require
additional modi
cations to the Riemann tensor (7.280), it is not at all clear that these
modi
cations will force the Riemann tensor to be of Kerr type. Indeed, the dierences
between the respective Riemann tensors would appear to make this quite unlikely. The
suggestion that a rotating star does not couple onto a Kerr-type solution is strengthened
by the fact that, in the 30 or so years since the discovery of the Kerr solution 90], no-one
has yet found a solution for a rotating star that matches onto the Kerr geometry at its
boundary.
7.4 Conclusions
The gauge theory of gravity developed from the Dirac equation has a number of inter-
esting and surprising features. The requirement that the gravitational action should be
consistent with the Dirac equation leads to a unique choice for the action integral (up to
the possible inclusion of a cosmological constant). The result is a set of equations which
are
rst-order in the derivatives of the
elds. This is in contrast to general relativity,
which is a theory based on a set of second-order partial dierential equations for the met-
ric tensor. Despite the formal similarities between the theories, the study of point-source
solutions reveals clear dierences. In particular, the
rst-order theory does not admit
solutions which are invariant under time-reversal.
The fact that the gauge group consists of active Poincare transformations of spacetime
elds means that gauge transformations relate physically distinct situations. It follows
that observations can determine the nature of the h and +
elds. This contrasts with
Yang-Mills theories based on internal gauge groups, where one expects that all observables
should be gauge-invariant. In this context, an important open problem is to ascertain
how the details of radial collapse determine the precise nature of the h and +
elds around
a black hole.
A strong point in favour of the approach developed here is the great formal clarity that
geometric algebra brings to the study of the equations. This is illustrated most clearly in
the compact formulae for the Riemann tensor for the Schwarzschild and Kerr solutions
174
and for radially-symmetric stars. No rival method (tensor calculus, dierential forms,
Newman-Penrose formalism) can oer such concise expressions.
For 80 years, general relativity has provided a successful framework for the study of
gravitational interactions. Any departure from it must be well-motivated by sound phys-
ical and mathematical reasons. The mathematical arguments in favour of the present
approach include the simple manner in which transformations are handled, the alge-
braic compactness of many formulae and the fact that torsion is perhaps better viewed
as a spacetime
eld than as a geometric eect. Elsewhere, a number of authors have
questioned whether the view that gravitational interactions are the result of spacetime
geometry is correct (see 94], for example). The physical motivation behind the present
theory is provided by the identi
cation of the h and +
elds as the dynamical variables.
The physical structure of general relativity is very much that of a classical
eld theory.
Every particle contributes to the curvature of spacetime, and every particle moves on
the resultant curved manifold. The picture is analogous to that of electromagnetism, in
which all charged particles contribute to an electromagnetic
eld (a kind of global ledger).
Yet an apparently crucial step in the development of Q.E.D. was Feynman's realisation
(together with Wheeler 95, 96]) that the electromagnetic
eld can be eliminated from
classical electrodynamics altogether. A similar process may be required before a quantum
multiparticle theory of gravity can be constructed. In the words of Einstein 97]
: : : the energy tensor can be regarded only as a provisional means of represent-
ing matter. In reality, matter consists of electrically charged particles : : :
The status of the h and +
elds can be regarded as equally provisional. They may simply
represent the aggregate behaviour of a large number of particles, and as such would not
be of fundamental signi
cance. In this case it would be wrong to attach too strong a
physical interpretation to these
elds (i.e. that they are the result of spacetime curvature
and torsion).
An idea of how the h
eld could arise from direct interparticle forces is provided by the
two-particle Dirac action constructed in Section 6.3. There the action integral involved
the dierential operator r1=m1 + r2=m2, so that the vector derivatives in each particle
space are weighted by the mass of the particle. This begins to suggest a mechanism by
which, at the one-particle level, the operator h(r) encodes an inertial drag due to the
other particle in the universe. This is plausible, as the h
eld was originally motivated
by considering the eect of translating a
eld. The theory presented here does appear
to contain the correct ingredients for a generalisation to a multiparticle quantum theory,
though only time will tell if this possibility can be realised.
175
Bibliography
1] A.N. Lasenby, C.J.L. Doran, and S.F. Gull. Grassmann calculus, pseudoclassical
mechanics and geometric algebra. J. Math. Phys., 34(8):3683, 1993.
2] C.J.L. Doran, D. Hestenes, F. Sommen, and N. van Acker. Lie groups as spin groups.
J. Math. Phys., 34(8):3642, 1993.
3] C.J.L. Doran, A.N. Lasenby, and S.F. Gull. Grassmann mechanics, multivector
derivatives and geometric algebra. In Z. Oziewicz, A. Borowiec, and B. Jancewicz,
editors, Spinors, Twistors and Cli ord Algebras, page 215. Kluwer, 1993.
4] A.N. Lasenby, C.J.L. Doran, and S.F. Gull. 2-spinors, twistors and supersymmetry
in the spacetime algebra. In Z. Oziewicz, A. Borowiec, and B. Jancewicz, editors,
Spinors, Twistors and Cli ord Algebras, page 233. Kluwer, 1993.
5] S.F. Gull, A.N. Lasenby, and C.J.L. Doran. Imaginary numbers are not real | the
geometric algebra of spacetime. Found. Phys., 23(9):1175, 1993.
6] C.J.L. Doran, A.N. Lasenby, and S.F. Gull. States and operators in the spacetime
algebra. Found. Phys., 23(9):1239, 1993.
7] A.N. Lasenby, C.J.L. Doran, and S.F. Gull. A multivector derivative approach to
Lagrangian
eld theory. Found. Phys., 23(10):1295, 1993.
8] S.F. Gull, A.N. Lasenby, and C.J.L. Doran. Electron paths, tunnelling and diraction
in the spacetime algebra. Found. Phys., 23(10):1329, 1993.
9] C.J.L. Doran, A.N. Lasenby, and S.F. Gull. Gravity as a gauge theory in the space-
time algebra. In F. Brackx and R. Delanghe., editors, Third International Confer-
ence on Cli ord Algebras and their Applications in Mathematical Physics., page 375.
Kluwer, 1993.
10] A.N. Lasenby, C.J.L. Doran, and S.F. Gull. Cosmological consequences of a at-
space theory of gravity. In F. Brackx and R. Delanghe., editors, Third International
Conference on Cli ord Algebras and their Applications in Mathematical Physics.,
page 387. Kluwer, 1993.
11] M.F. Atiyah, R. Bott, and A. Shapiro. Cliord modules. Topology, 3(Supp. 1):3,
1964.
12] I.R. Porteous. Topological Geometry. Van Nostrand Reinhold Company, 1969.
176
13] I.W. Benn and R.W. Tucker. An Introduction to Spinors and Geometry. Adam
Hilger, 1988.
14] C. Chevalley. The Algebraic Theory of Spinors. Columbia University Press, 1954.
15] H.B. Lawson and M.-L. Michelsohn. Spin Geometry. Princeton University Press,
1989.
16] F. Reese. Harvey. Spinors and Calibrations. Academic Press, San Diego, 1990.
17] D. Hestenes. Space-Time Algebra. Gordon and Breach, 1966.
18] D. Hestenes. Multivector calculus. J. Math. Anal. Appl., 24:313, 1968.
19] D. Hestenes. Vectors, spinors, and complex numbers in classical and quantum physics.
Am. J. Phys., 39:1013, 1971.
20] D. Hestenes. Proper dynamics of a rigid point particle. J. Math. Phys., 15(10):1778,
1974.
21] D. Hestenes. Observables, operators, and complex numbers in the Dirac theory. J.
Math. Phys., 16(3):556, 1975.
22] D. Hestenes and R. Gurtler. Consistency in the formulation of the Dirac, Pauli and
Schr&odinger theories. J. Math. Phys., 16(3):573, 1975.
23] D. Hestenes. Spin and uncertainty in the interpretation of quantum mechanics. Am.
J. Phys., 47(5):399, 1979.
24] D. Hestenes and G. Sobczyk. Cli ord Algebra to Geometric Calculus. D. Reidel
Publishing, 1984.
25] D. Hestenes. New Foundations for Classical Mechanics. D. Reidel Publishing, 1985.
26] D. Hestenes. A uni
ed language for mathematics and physics. In J.S.R. Chisholm
and A.K. Common, editors, Cli ord Algebras and their Applications in Mathematical
Physics, page 1. D. Reidel, 1986.
27] D. Hestenes. Cliord algebra and the interpretation of quantum mechanics. In J.S.R.
Chisholm and A.K. Common, editors, Cli ord Algebras and their Applications in
Mathematical Physics, page 321. D. Reidel, 1986.
28] D. Hestenes. Curvature calculations with spacetime algebra. Int. J. Theor. Phys.,
25(6):581, 1986.
29] D. Hestenes. On decoupling probability from kinematics in quantum mechanics.
In P. Foug.ere, editor, Maximum Entropy and Bayesian Methods, page 161. Kluwer,
1990.
30] D. Hestenes. The design of linear algebra and geometry. Acta Appl. Math., 23:65,
1991.
177
31] D. Hestenes and R. Ziegler. Projective geometry with Cliord algebra. Acta. Appli.
Math., 23:25, 1991.
32] D. Hestenes. Hamiltonain mechanics with geometric calculus. In Z. Oziewicz,
A. Borowiec, and B. Jancewicz, editors, Spinors, Twistors and Cli ord Algebras,
page 203. Kluwer, 1993.
33] D. Hestenes. Real Dirac theory. In Preparation, 1994.
34] D. Hestenes. Dierential forms in geometric calculus. In F. Brackx and R. Delanghe.,
editors, Third International Conference on Cli ord Algebras and their Applications
in Mathematical Physics. Kluwer, 1993.
35] F.A. Berezin. The Method of Second Quantization. Academic Press, 1966.
36] R. Penrose and W. Rindler. Spinors and space-time, Volume I: two-spinor calculus
and relativistic elds. Cambridge University Press, 1984.
37] R. Penrose and W. Rindler. Spinors and space-time, Volume II: spinor and twistor
methods in space-time geometry. Cambridge University Press, 1986.
38] A.O. Barut and N. Zanghi. Classical models of the Dirac electron. Phys. Rev. Lett.,
52(23):2009, 1984.
39] F.A. Berezin and M.S. Marinov. Particle spin dynamics as the Grassmann variant
of classical mechanics. Annals of Physics, 104:336, 1977.
40] M. Kline. Mathematical Thought from Ancient to Modern Times. Oxford University
Press, 1972.
41] H. Grassmann. Der ort der Hamilton'schen quaternionen in der ausdehnungslehre.
Math. Ann., 12:375, 1877.
42] W. K. Cliord. Applications of Grassmann's extensive algebra. Am. J. Math., 1:350,
1878.
43] M.F. Atiyah and I.M. Singer. The index of elliptic operators on compact manifolds.
Bull. A.M.S., 69:422, 1963.
44] T.G. Vold. An introduction to geometric algebra with an application to rigid body
mechanics. Am. J. Phys., 61(6):491, 1993.
45] T.G. Vold. An introduction to geometric calculus and its application to electrody-
namics. Am. J. Phys., 61(6):505, 1993.
46] R. Ablamovicz, P. Lounesto, and J. Maks. Second workshop on Cliord algebras and
their applications in mathematical physics. Found. Phys., 21(6):735, 1991.
47] E.T. Jaynes. Scattering of light by free electrons. In A. Weingartshofer and
D. Hestenes, editors, The Electron, page 1. Kluwer, 1991.
178
48] N. Salingaros. On the classi
cation of Cliord algebras and their relation to spinors
in n-dimensions. J. Math. Phys., 23(1):1, 1982.
49] I. Stewart. Hermann grassmann was right (News and Views). Nature, 321:17, 1986.
50] M. Barnabei, A. Brini, and G.-C. Rota. On the exterior calculus of invariant theory.
J. Algebra, 96:120, 1985.
51] F.A. Berezin. Introduction to Superanalysis. D. Reidel, 1987.
52] B. de Witt. Supermanifolds. Cambridge University Press, 1984.
53] R. Coquereaux, A. Jadczyk, and D. Kastler. Dierential and integral geometry of
Grassmann algebras. Reviews in Math. Phys., 3(1):63, 1991.
54] A. Connes and J. Lott. Particle models and non-commutative geometry. Nucl. Phys.
B (Proc. Suppl.), 18B:29, 1990.
55] G.C. Sherry. A generalised phase space for classical fermions arising out of Schon-
berg's geometric algebras. Found. Phys. Lett., 2(6):591, 1989.
56] G.C. Sherry. Algebraic model of a classical non-relativistic electron. Found. Phys.
Lett., 3(3):267, 1990.
57] J.F. Cornwell. Group Theory in Physics II. Academic Press, 1984.
58] A.O. Barut and A.J. Bracken. The remarkable algebra so (2n), its representations,
its Cliord algebra and potential applications. J. Phys. A, 23:641, 1990.
59] J.D. Bjorken and S.D. Drell. Relativistic Quantum Mechanics, vol 1. McGraw-Hill,
1964.
60] C. Itzykson and J-B. Zuber. Quantum Field Theory. McGraw-Hill, 1980.
61] S.F. Gull. Charged particles at potential steps. In A. Weingartshofer and D. Hestenes,
editors, The Electron, page 37. Kluwer, 1991.
62] E.M. Corson. Introduction to Tensors, Spinors, and Relativistic Wave-Equations.
Blackie & Son, Ltd. (Glasgow), 1953.
63] A.P. Galeao and P. Leal Ferreira. General method for reducing the two-body Dirac
equation. J. Math. Phys., 33(7):2618, 1992.
64] E.E. Salpeter and H.A. Bethe. A relativisitc equation for bound-state problems.
Phys. Rev., 84(6):1232, 1951.
65] G. Breit. The eect of retardation on the interaction of two electrons. Phys. Rev.,
34(4):553, 1929.
66] N. Kemmer. Zur theorie der neutron-proton wechselwirkung. Helv. Phys. Acta.,
10:48, 1937.
179
67] E. Fermi and C.N. Yang. Are mesons elementary particles? Phys. Rev., 76(12):1739,
1949.
68] Y. Koide. Exactly solvable model of relativistic wave equations and meson spectra.
Il Nuovo Cim., 70A(4):411, 1982.
69] W. Krolikowski. Dun-Kemmer-Petiau particle with internal structure. Acta Phys.
Pol. B, 18(2):111, 1987.
70] W. Krolikowski. Cliord algebras and the algebraic structure of fundamental
fermions. In Z. Oziewicz, A. Borowiec, and B. Jancewicz, editors, Spinors, Twistors
and Cli ord Algebras, page 183. Kluwer, 1993.
71] H. Goldstein. Classical Mechanics. Addison Wesley, 1950.
72] J.W. van Holten. On the electrodynamics of spinning particles. Nucl. Phys.,
B356(3):3, 1991.
73] P.G.O. Freund. Supersymmetry. Cambridge University Press, 1986.
74] R. Casalbuoni. The classical mechanics for Bose-Fermi systems. Il Nuovo Cimento.,
33A(3):389, 1976.
75] L. Brink, S. Deser, B. Zumino, P. di Vecchia, and P. Howe. Local supersymmetry
for spinning particles. Phys. lett. B, 64(4):435, 1976.
76] J.F. Cornwell. Group Theory in Physics III. Academic Press, 1989.
77] S. Coleman. Aspects of Symmetry. Cambridge University Press, 1985.
78] H. Tetrode. Der impuls-energiesatz in der Diracschen quantentheorie des elektrons.
Z. Physik, 49:858, 1928.
79] R. Utiyama. Invariant theoretical interpetation of interaction. Phys. Rev.,
101(5):1597, 1956.
80] T.W.B. Kibble. Lorentz invariance and the gravitational
eld. J. Math. Phys.,
2(3):212, 1961.
81] F.W. Hehl, P. von der Heyde, G.D. Kerlick, and J.M. Nestev. General relativity with
spin and torsion: Foundations and prospects. Rev. Mod. Phys., 48:393, 1976.
82] D. Ivanenko and G. Sardanashvily. The gauge treatment of gravity. Phys. Rep.,
94(1):1, 1983.
83] S.W. Hawking and G.F.R. Ellis. The Large Scale Structure of Space-Time. Cambridge
University Press, 1973.
84] R.T. Rauch. Equivalence of an R + R2 theory of gravity to Einstein-Cartan-Sciama-
Kibble theory in the presence of matter. Phys. Rev. D, 26(4):931, 1982.
180
85] R.D. Hecht, J. Lemke, and R.P. Wallner. Can Poincare gauge theory be saved? Phys.
Rev. D, 44(8):2442, 1991.
86] A.V. Khodunov and V.V. Zhytnikov. Gravitational equations in space-time with
torsion. J. Math. Phys., 33(10):3509, 1992.
87] R. d'Inverno. Introducing Einstein's Relativity. Oxford University Press, 1992.
88] E.T. Newman and A.I. Janis. Note on the Kerr spinning-particle metric. J. Math.
Phys., 6(4):915, 1965.
89] M.M. Schier, R.J. Adler, J. Mark, and C. Sheeld. Kerr geometry as complexi
ed
Schwarzschild geometry. J. Math. Phys., 14(1):52, 1973.
90] R.P. Kerr. Gravitational
eld of a spinning mass as an example of algebraically
special metrics. Phys. Rev. Lett., 11(5):237, 1963.
91] S. Chandrasekhar. The Mathematical Theory of Black Holes. Oxford University
Press, 1983.
92] W.J. Kaufmann. The Cosmic Frontiers of General Relativity. Penguin Books, 1979.
93] C.W. Misner, K.S. Thorne, and J.A. Wheeler. Gravitation. W.H. Freeman and
Company, 1973.
94] S. Weinberg. Gravitation and Cosmology. John Wiley and Sons, 1972.
95] J.A. Wheeler and R.P. Feynman. Interaction with the absorber as the mechanism of
radiation. Rev. Mod. Phys., 17:157, 1945.
96] J.A. Wheeler and R.P. Feynman. Classical electrodynamics in terms of direct inter-
particle action. Rev. Mod. Phys., 21(3):425, 1949.
97] A. Einstein. The Meaning of Relativity. Princeton University Press, New Jersey,
1945.
181