<<

arXiv:2107.09495v1 [hep-th] 20 Jul 2021 eodLrnza Lorentzian Beyond nvriyo Edinburgh of University otro Philosophy of Doctor osGrassie Ross 2021

Declaration

I declare that this thesis was composed by myself and that the work contained therein is my own, except where explicitly stated otherwise in the text. No part of this thesis has been submitted here or elsewhere for any other degree or qualification. Chapter 3 of this thesis is based on a published paper [1] authored by Stefan Prohazka, my supervisor José Figueroa-O’Farrill, and myself. Additionally, Chapter 4 is based on a published paper [2] authored by my supervisor José Figueroa-O’Farrill and myself. Chapter 2 contains aspects from both of these papers.

(Ross Grassie)

iii In Memory of Winifred Whittle

iv Acknowledgements

I would like to start by thanking my supervisor José Figueroa-O’Farrill. His enthusiasm for and physics has certainly had a profound impact on my approach to research, teaching me not to be so serious all the time and remember that there is a lot of fun to be had in exploring new material.

I would also like to thank my family and friends for their constant support through what has been a very difficult few years. In particular, I would like to thank my partner Katie Meyer for always believing in me and picking me up when I felt as though the world was falling apart. I don’t think my PhD would have been half as fun as it was without my office mates, Fred Tomlinson, Graeme Auld and Ollie Burke. Whether it was going to the pub perhaps a bit too early on Friday or chatting complete nonsense for a few hours in the office, they made the last couple of years infinitely more enjoyable. I think a special mention should also be reserved for Sam Stern and Manya Sahni. If it were not for them, I would never have made it through my undergraduate of masters degrees to even start my PhD!

Finally, I would like to thank my gran, Winifred Whittle. I do not think anything I can write here can do justice to how much of a positive influence she has had on my life. It is primarily due to the work ethic she passed on to me that this thesis exists.

v vi Lay Summary

The goal for any physicist is to better understand the world around them. To complete this task, we need to construct models that estimate, characterise, or otherwise describe the par- ticular phenomenon we are interested in. This may be how an object moves through space, what happens to it over time, what happens when it is exposed to extreme temperatures or pressures, how it interacts with other things; whatever it may be, we want to describe its place in the world. The hope is that by using insightful and robust models, we will accurately pre- dict interesting aspects of future occurrences of the phenomenon. Equipped with the ability to produce these predictions, we may say we have crucial insight into how the world works. However, our understanding of these phenomena will be heavily influenced by the model we choose. Therefore, we need to make sure we are constructing models which are fit for purpose.

Constructing a model for a particular phenomenon can be thought of as consisting of two parts. The first part captures the initial conditions of the system. Using the movement of a planet as an example, we would like to know its starting position and how fast it is moving. The second part contains what we may call the “laws of physics” or the “laws of nature”. These are the ideas we believe hold true irrespective of the initial conditions; for example, Newton’s gravitational laws. As the initial conditions will necessarily be different each time we conduct an experiment, the second part of our model gives us the desired predictive power to better understand the world around us. However, building this part of our models is challenging; therefore, we would like to recognise and exploit any inherent structure and coherence in these natural laws. Since mathematics is the chosen language of physicists, we want a mathematical principle that will capture this information for us. The current principle of choice, which guides us in producing our models of the world, is symmetry.

In this thesis, we will classify symmetries and construct models of space and time that em- ploy these symmetries. Our hope is that future researchers will use these classifications to accurately describe and better understand the world we live in.

vii viii Abstract

This thesis presents a framework in which to explore kinematical symmetries beyond the stan- dard Lorentzian case. This framework consists of an algebraic classification, a geometric classi- fication, and a derivation of the geometric properties required to define physical theories on the classified spacetime . The work completed in substantiating this framework for kine- matical, super-kinematical, and super-Bargmann symmetries constitutes the body of this thesis.

To this end, the classification of kinematical Lie algebras in spatial dimension D = 3, as pre- sented in [3,4], is reviewed; as is the classification of spatially-isotropic homogeneous spacetimes of [5]. The derivation of geometric properties such as the non-compactness of boosts, soldering forms and vielbeins, and the space of invariant affine connections is then presented.

We move on to classify the N = 1 kinematical Lie superalgebras in three spatial dimensions, finding 43 isomorphism classes of Lie superalgebras. Once these algebras are determined, we classify the corresponding simply-connected homogeneous (4|4)-dimensional superspaces and show how the resulting 27 homogeneous superspaces may be related to one another via geo- metric limits.

Finally, we turn our attention to generalised Bargmann superalgebras. In the present work, these will be the N = 1 and N = 2 super-extensions of the Bargmann and Newton-Hooke alge- bras, as well as the centrally-extended static kinematical , of which the former three all arise as deformations. Focussing solely on three spatial dimensions, we find 9 isomorphism classes in the N = 1 case, and we identify 22 branches of superalgebras in the N = 2 case.

ix x Contents

Acknowledgements v

Lay Summary vii

Abstract ix

Contents xii

1 Introduction 1 1.1 ABriefHistoryofKinematicalSymmetry ...... 1 1.2 BeyondLorentzianSymmetries ...... 2 1.2.1 ClassesofKinematicalSpacetime ...... 3 1.2.2 Supersymmetry ...... 4 1.3 OutlineofThesis ...... 5

2 Mathematical Preliminaries 7 2.1 Algebra...... 7 2.1.1 LieAlgebrasandLieSuperalgebras...... 7 2.1.2 KinematicalLieAlgebras ...... 8 2.1.3 AristotelianLieAlgebras...... 9 2.1.4 GeneralisedBargmannAlgebras ...... 9 2.1.5 KinematicalLieSuperalgebras ...... 10 2.1.6 GeneralisedBargmannSuperalgebras ...... 11 2.2 ...... 12 2.2.1 LieGroupsandtheirLieAlgebras ...... 13 2.2.2 ExponentialCoordinatesonaLieGroup ...... 14 2.2.3 CosetSpaces...... 15 2.2.4 HomogeneousSpaces ...... 16 2.2.5 KinematicalSpacetimes ...... 18 2.2.6 Lie Supergroups and Homogeneous Superspaces ...... 19 2.3 GeometricProperties ...... 21 2.3.1 The Action and the Fundamental Vector Fields ...... 21 2.3.2 InvariantConnections ...... 23 2.3.3 The Soldering Form and the Canonical Connection ...... 24 2.3.4 InvariantTensors ...... 25

3 Kinematical Spacetimes 27 3.1 Classification of Kinematical Lie Algebras ...... 27 3.1.1 Classification of Aristotelian Lie Algebras ...... 29 3.2 Classification of Kinematical Spacetimes ...... 29 3.2.1 Admissibility...... 30 3.2.2 Effectivity ...... 31 3.2.3 GeometricRealisability...... 31 3.2.4 Classification...... 31 3.3 LimitsBetweenSpacetimes...... 33 3.3.1 ContractionLimits ...... 33

xi 3.3.2 Remaining Galilean Spacetimes ...... 37 3.3.3 LimitoftheCarrollianLightCone ...... 37 3.3.4 AristotelianLimits ...... 37 3.3.5 ANon-ContractingLimit ...... 38 3.3.6 Summary...... 38 3.4 Geometric Properties of Kinematical Spacetimes ...... 39 3.4.1 TheActionoftheRotations...... 40 3.4.2 TheActionoftheBoosts...... 41 3.4.3 Invariant Connections, Curvature, and Torsion for Reductive Spacetimes 43 3.4.4 Pseudo-Riemannian Spacetimes and their Limits ...... 47 3.4.5 TorsionalGalileanSpacetimes ...... 58 3.4.6 AristotelianSpacetimes ...... 67 3.4.7 Carrollian Light Cone (LC)...... 70 3.5 Conclusion ...... 74

4 Kinematical Superspaces 75 4.1 Classification of Kinematical Superalgebras ...... 75 4.1.1 SetupfortheClassification ...... 75 4.1.2 Classification of Kinematical Lie Superalgebras ...... 82 4.1.3 Summary...... 91 4.1.4 Classification of Aristotelian Lie Superalgebras ...... 93 4.1.5 CentralExtensions ...... 95 4.1.6 Automorphisms of Kinematical Lie Superalgebras ...... 96 4.2 Classification of Kinematical Superspaces ...... 107 4.2.1 AdmissibleSuperLiePairs ...... 109 4.2.2 EffectiveSuperLiePairs...... 112 4.2.3 Aristotelian Homogeneous Superspaces...... 116 4.2.4 Summary...... 116 4.2.5 Low-RankInvariants ...... 118 4.3 LimitsBetweenSuperspaces ...... 119 4.3.1 ContractionsoftheAdSSuperalgebra ...... 120 4.3.2 RemainingGalileanSuperspaces ...... 121 4.3.3 AristotelianLimits ...... 124 4.3.4 ANon-ContractingLimit ...... 125 4.3.5 Summary...... 125 4.4 Conclusion ...... 127

5 Generalised Bargmann Superspaces 129 5.1 Classification of N = 1 Generalised BargmannSuperalgebras ...... 131 5.1.1 Setup for the N = 1Calculation...... 131 5.1.2 Classification...... 135 5.1.3 Summary...... 136 5.2 Classification of N = 2 Generalised BargmannSuperalgebras ...... 138 5.2.1 Setup for the N = 2Calculation...... 138 5.2.2 EstablishingBranches ...... 145 5.2.3 Classification...... 150 5.2.4 Summary...... 182 5.3 Conclusion ...... 183

6 Conclusion 185

Bibliography 192

xii Chapter 1

Introduction

This thesis is concerned with the classification of (super-)kinematical symmetry algebras and their corresponding spacetime (super)geometries, as well as the determination of various ge- ometric properties of these (super)geometries. As such, this chapter will give a brief history of kinematical symmetry, introducing the key concepts underlying the research presented here and discussing why Lorentz symmetry became the primary example of kinematical symme- try. Additionally, there will be a discussion on why we may be interested in extending beyond Lorentzian symmetry which will include the introduction of the five kinematical spacetime classes and supersymmetry. We conclude by outlining the rest of the thesis.

1.1 A Brief History of Kinematical Symmetry

People have been using symmetry to try and describe the world around them for thousands of years. The idea of symmetry in nature can be found in the Timaeus by Plato and Euclid’s Ele- ments [6]. However, symmetry, in the sense in which it is used in modern physics, is a relatively new concept. In 1905, Einstein published his seminal work on special relativity, introducing the profound paradigm shift that placed symmetry at the core of how we think about the world. Prior to this work, physicists such as Newton and Maxwell had built their models of the world by first seeking to write down natural laws. Therefore, the invariance of these laws under some symmetry was recognised, but it was not seen as particularly important. However, with the advent of special relativity, the roles of symmetry and natural laws were reversed; in particular, we now seek to derive laws of nature from symmetry considerations, rather than derive the sym- metry of natural laws [7]. In the special relativity paper, Einstein showed that one could derive the transformation properties of an electromagnetic field using Lorentz invariance alone, rather than deriving them from Maxwell’s equations [8]. Thus he took the known Lorentz symmetry of Maxwell’s equations and demonstrated how we could view it as something more fundamental; we may view it as a symmetry of the spacetime in which the electromagnetic field is propagating.

From this historical perspective, if we want to understand symmetries of spacetime, also called kinematical symmetries, we need to look at the classical laws of nature, such as Newton’s laws of motion and Maxwell’s electromagnetic equations. By determining what type of symmetries they are invariant under, we may better understand what kinematical symmetries we can have. The key observation here is that many of the known laws of nature take the form of differential equations [9]; therefore, to understand the allowed kinematical symmetries, we must under- stand the allowed symmetries of differential equations. Sophus Lie undertook this programme of study in the early 1870s [10]. This work, which Lie saw as a direct generalisation of Galois’s earlier work on applying to algebraic equations, led to his theory of continuous groups, now called Lie groups [11]. Thus we find that the Lie groups provide a natural setting for describing kinematical symmetries.

One of Lie’s early collaborators on this project was a mathematician called . Inter- ested in applying group theory, not to differential equations, but differential geometry, Klein’s

1 research diverged from Lie’s, and he set out his Erlanger Programme in 1872. This programme aimed to describe a using its transformation group, defining any geometric objects on the manifold by the subgroup which left the object invariant. Although the Erlanger Pro- gramme was influential, forming a uniform framework to characterise classical geometries, it was only with Élie Cartan’s work on generalised spaces that the programme’s connection to spacetime was made manifest [12]. In particular, when Einstein’s general theory of relativ- ity emerged in 1915, it was shown that spacetime might be viewed as a Lorentzian manifold. Importantly, this means that, according to Einstein, the objects in spacetime, such as elec- tromagnetic fields, must transform under the . Klein noticed that the Erlanger Programme might be related to this geometric picture of space and time [10]; however, it was Cartan, building upon the Erlanger programme, making more explicit use of Lie groups due to their relation to kinematics, that sufficiently generalised Einstein’s theory. Notably, this led him to a reformulation of Newtonian gravity in the geometric language of general relativity [13,14].1

Due to the overwhelming success of general relativity in describing phenomena outside the reach of classical Newtonian gravity (see [17]), Lorentzian symmetry, and its classical, Galilean limit, were the primary kinematical symmetries considered for much of the 20th century. How- ever, in the 1960s, people began to ask whether there may be other kinematical group choices. The first paper to attempt to classify all the possible kinematical symmetries was [18]. How- ever, this classification imposed time-reversal and parity symmetries, which are not strictly necessary from a purely algebraic perspective. On removing these conditions, the classification was completed by Bacry and Nuyts in [19]. These papers show that we can split kinematical symmetries into five classes: Lorentzian, Euclidean, Galilean, Carrollian, and Aristotelian. As mentioned above, the Lorentzian and Galilean cases are the most prevalent examples of kine- matical symmetries and Euclidean symmetries, owing to their close connection with Lorentzian symmetries, frequently appear in the literature. Therefore, the novel classes were the Carrol- lian and Aristotelian symmetries; although it was believed that they might be “without much physical application”. As we will argue in the next section, each class of kinematical symmetry is interesting in its own right, and we will outline a few examples of how they are used in the literature.

1.2 Beyond Lorentzian Symmetries

Following Einstein’s lead in placing kinematical symmetries at the forefront of our physical theories, and assuming we have defined the correct notion of kinematical symmetry, as described in [3, 4, 20–22], we can reasonably ask,

(i) how would physical objects, such as particles and electromagnetic fields, act in spacetimes described by these symmetries?, (ii) do we find natural systems that are described by these symmetries?, and (iii) is there any way we could extend the kinematical symmetries to other physically interesting symmetries?

Systematically answering the first question is the purpose of Chapter 3 in this thesis; therefore, we will defer this conversation until then. However, we can answer the second question more succinctly here; in particular, this will be the topic of Section 1.2.1. Furthermore, the basis for the investigations of Chapters 4 and 5 into extending the kinematical symmetries will be reviewed in Section 1.2.2.

1At this point, it may be interesting to make the following comment. The term kinematics was coined by Ampère in the late 1820s with the explicit intent of merging the study of mechanics with geometry [15]. Thus, the geometric picture of spacetime began to fall under the label of kinematics. Klein’s introduction of group theory into geometry then provided the setting for Einstein’s symmetry-first approach to describing the movement of objects in a spacetime geometry. For a discussion on the historical connection between the study of kinematics and Einstein’s theory of gravity, see [16].

2 1.2.1 Classes of Kinematical Spacetime There is a growing body of literature on applications of every type of kinematical symmetry. We will now briefly review some of the research utilising these different symmetries and their associated spacetime models.

Lorentzian Most of the progress in 20th-century physics was made with the assumption of underlying Lorentz symmetry. It is built into gravity theories and quantum theories, and, therefore, lies at the foundation of some of the century’s most famous ideas. There are far too many research fields influenced by Lorentz symmetry to recount here. Therefore, we will only briefly give an account of how Lorentz symmetry appears in and impacts our understanding of the two pillars of modern physics, general relativity and quantum field theory. This review aims to highlight the parts of these theories that may be altered by imposing one of the other types of kinematical symmetry.

As mentioned above, Lorentz symmetry emerged as a fundamental symmetry of spacetime through Einstein’s enquiries into electromagnetism [23]. In particular, it arises from a desire to have a fixed speed of light, and impose that the laws of physics look the same in all inertial ref- erence frames. With special relativity being built into general relativity as the local description of spacetime, the general theory of relativity thus requires the spacetime geometry to admit a Lorentzian structure.2 This condition on the geometry has significant consequences, which were not present in the preceding Newtonian picture of gravity. One of the critical implications of Lorentzian symmetry is that there exists a restricted subspace in spacetime with which we can be in causal contact. This means that interactions, such as gravitational forces, are no longer instantaneous in Einstein’s picture.

Quantum field theory is the typical language used to describe the interactions between sub- atomic particles; therefore, it is widely used in particle physics, atomic physics, condensed matter physics, and astrophysics [24, 25]. In this theory, fields, whose excitations define parti- cles, propagate in a fixed, Lorentzian spacetime geometry. This propagation is described by the transformation group, or relativity group, of the underlying spacetime, which, in a flat space- time as described by Einstein, means that the matter must live in a module of the Poincaré group. After quantisation, the representation of the Lorentz subgroup that acts on a particular matter field will be labelled by a (half-)integer known as the spin of the field. The half-integer spin fields are called fermions, and the integer spin fields are called bosons [25]. Bosons and fermions are then our basic building blocks for any quantum field theory we wish to write down; thus, Lorentz symmetry is built into the foundation of our ideas on quantum theory. As we will discuss in Section 1.2.2, wishing to replace the Lorentz group with the Galilean or Carrollian groups has a profound impact on our modelling of these basic constituents of matter.

Euclidean In modern physics, Euclidean symmetries are frequently used as a computational tool [25]. Many calculations in gravitational and quantum theories are hard in the Lorentzian case, but

2By Lorentzian structure, we mean the following. Let M be a (D+1)-dimensional real smooth manifold. The frame bundle of M is then a principle GL(D + 1, R) bundle over M. Let ι : SO(D, 1) GL(D + 1, R) be the Lie → group monomorphism which embeds the Lorentz group inside GL(D + 1, R). With this data, we may construct a principle Lorentz bundle over M. We call this reduction of the frame bundle a Lorentz structure. Using this structure, we may define a Lorentzian metric g and call (M, g) a Lorentzian manifold. This process holds for all kinematical groups, leading to Euclidean, Galilean, Carrollian, and Aristotelian structures in the other instances.

Notice, this requirement of a Lorentz structure arises from the equivalence principle: the fact that lo- cally, we should recover special relativity. However, there is an additional principle in general relativity, which leads to another form of symmetry. The general principle of relativity states that the laws of physics must look the same in any coordinate system. Translated into our geometric language, this statement says that the action we write down for our gravitational theory should be invariant under diffeomorphisms. When describing general relativity as a gauge theory, it is the group of diffeomorphisms that is presented as the gauge group, not the Lorentz group.

3 become tractable when the time coordinate is Wick rotated, such that we arrive at a Euclidean description [26]. Thus, Euclidean symmetries have been indispensable in driving research due to their close connection to the computationally less friendly Lorentz symmetries.

Galilean Although Galilean symmetry was historically “superseded” by Lorentzian symmetry, there are still numerous physical systems that admit a cleaner description when described using non- relativistic symmetries. In particular, classical Newtonian mechanics does not require the full machinery of Lorentzian symmetry; the non-relativistic limit provides a far nicer framework for these systems. However, there are more complex systems that make use of Galilean symmetries. In condensed matter theory, the (fractional) quantum Hall effect admits a description as an ef- fective field theory invariant under Galilean symmetries [27–30]. Non-relativistic spacetimes have also been useful in extending the holographic duality beyond the AdS/CFT correspon- dence [31–35]. Recently, non-relativistic spacetimes have also been incorporated into string theory [36, 37], and supersymmetric quantum field theories [38, 39], and have been interpreted using double field theory [40–43]. They have also been systematically studied in relation to JT gravity [44, 45] in the hope they may elucidate some outstanding problems in the search for a quantum theory of gravity. Furthermore, numerous papers have gauged known non-relativistic algebras, or extensions thereof, to arrive at novel gravity theories [33, 37, 46–65].

Carrollian These ultra-relativistic spacetimes have received increased interest over recent years due to their connection with null hypersurfaces in Lorentzian spacetimes. Namely, the restriction of the manifold’s Lorentzian structure in a D + 1-dimensional spacetime to a D-dimensional null hypersurface induces a Carrollian structure on the surface. Therefore, interesting null hypersurfaces such as black hole horizons and past and future null infinity carry a Carrollian structure by construction [66, 67]. This connection has lead to exciting results regarding the physics of black hole horizons [68] and has advanced research into flat space holography [69–72].

Aristotelian The spacetime models in this class are quite distinct from the four above. In particular, these spacetimes do not allow for a change of reference frame: we are always considering the world from one fixed coordinate system. In the physics nomenclature, this fixed reference frame corresponds to the absence of boosts. By having a fixed system, Aristotelian geometries are useful for describing phenomena with a preferred reference frame or phenomena where we would like to avoid thinking about which type of boosts we will allow. This property of being “boost agnostic” has recently been used to consider hydrodynamic systems [73, 74].

1.2.2 Supersymmetry This classification of kinematical spacetimes provides an already fertile ground for exploration into symmetries beyond the usual Lorentzian case; however, we can still think of pushing this classification a little further. This section will introduce supersymmetry, giving some historical context to how it first appeared in physical theories and briefly reviewing some of the outstand- ing problems it could help to solve. We then describe how the kinematical symmetries may be generalised to super-kinematical symmetries.

Inspired by Einstein’s symmetry-first approach in building general relativity, the particle physi- cists of the 1920s built their models of particle interactions by introducing a new type of symmetry [75]. Equipped with the concept of gauge symmetry, developments in the 1960s and 1970s showed that three of the four fundamental forces, the strong, electromagnetic and weak interactions, may all be described using this language. Thus, these forces admit a unified treat- ment in the Standard Model. While this model has been incredibly successful, many physicists would like to see gravitational forces included in such a unified description. Unfortunately,

4 gravitational interactions, as presented by general relativity, do not have an analogous descrip- tion in terms of gauge symmetry [11].3 Attempting to circumvent the obstructions to such a unified theory, we may look to exploit new symmetries with properties which allow them to bypass any no-go theorems, such as the Coleman-Mandula theorem [76].4 One such proposal is supersymmetry.

In the Lorentzian setting, supersymmetry is a spacetime symmetry that allows us to inter- change fermionic and bosonic degrees of freedom [78]. It achieves this task by introducing a new type of symmetry generator with half-integer spin. Some of the consequences of this new type of symmetry and striking and profound. As mentioned above, the introduction of super- symmetry allows us to construct theories which may unify gravity with the other fundamental forces. It also provides an elegant solution to the hierarchy problem and suggests a wide range of new particles which may be used to describe dark matter.5 Furthermore, supersymmetric theories are often easier to analyse than their traditional counterpart; thus, supersymmetry also provides researchers with a useful test-bed for exploring new physics.6

The first paper looking to incorporate supersymmetry into the classification of kinematical spacetimes presented above was [81]. However, this paper applies the same “by no means convincing” assumptions of parity and time-reversal symmetries as [18]. Other papers which extend the kinematical symmetries of [18] include [82] and [83]. Later papers such as [84–86] tackled this problem through the process of contractions; however, to the best of this author’s knowledge, the first full classification of super-kinematical symmetries and superspaces was presented in [2].

It is perhaps interesting to note at this stage that supersymmetry, beyond the Lorentzian and Euclidean cases, is not a well-explored or well-defined concept. We notice this fact instantly by considering the symmetry generators it has in addition to the classical case. These are defined to have half-integer spin; however, if we do not have Lorentzian or Euclidean symmetry, the concept of spin is not well-defined. Therefore, what it means to have Galilean or Carrollian supersymmetry is a priori unclear. In Chapter 2, we will give one possible interpretation for these generators; namely, we will use the spin associated with the subalgebra of spatial rota- tion. However, it should be noted that other researchers assume the new symmetry generators to have vanishing spin [87, 88].

1.3 Outline of Thesis

Having reviewed some critical chapters in the history of symmetry and geometry and briefly summarised why they are so fundamental to our ideas of spacetime, we can now turn our at- tention towards using this knowledge to explore kinematical symmetries beyond the standard Lorentzian case. In particular, we will build our understanding of (super-)kinematical symme- tries in the following framework. First, we will consider algebraic classifications, such as those found in [19]. These investigations will furnish us with the basic building blocks from which we may construct physical theories. Next, we will classify spacetime (super)geometries mod- elled on these algebras. We may view this geometric classification as a physical realisation of Klein’s Erlanger Programme, in the spirit of Cartan. Finally, we will investigate the geometric properties required to define physical theories on each spacetime. The three main chapters in the body of this thesis will present the work completed towards fulfilling this framework in the kinematical, super-kinematical, and super-Bargmann instances, respectively. In more detail,

3Perhaps I should be more explicit here. General relativity does not admit a description as a renormalisable Yang-Mills-type gauge theory, as the strong, electromagnetic and weak interactions do. 4Note, the Coleman-Mandula theorem, first presented in [77], states that the internal symmetries describing the particle interactions and the spacetime describing particle movement cannot be combined in a non-trivial way for a Lorentzian-relativistic quantum field theory. 5Note, the hierarchy problem is the name given to the fact the experimental value for the mass of the Higgs boson is smaller than predicted. 6There are countless papers, books, and reviews on supersymmetry and its uses; for some of the better known instances see [79, 80].

5 this thesis will run as follows.

Chapter 2 This chapter contains a discussion on the basic mathematical objects required in this thesis. A graduate-level knowledge of differential geometry and algebra is assumed. The chapter is divided into three sections, algebra, geometry, and geometric properties, to align with the framework we wish to substantiate. In Section 2.1, we first introduce the notion of a Lie (super)algebra before defining each of the classes of Lie (super)algebra we will use in later chapters. Section 2.2 then describes how we may integrate these Lie algebras to form Lie groups and some associated geometries. In addition, this section defines our characterisation of a supermanifold and its relation to the corresponding Lie superalgebra and underlying classical geometry. Finally, Section 2.3 provides a detailed explanation of how we will define the geometric properties of kinematical spacetimes.

Chapter 3 This chapter gives a full substantiation of the framework set out above for the case of kinematical symmetries. In particular, we review the algebraic classification for kinematical Lie algebras from [3, 4] in Section 3.1, before showing how these algebras were integrated into kinematical spacetimes in Section 3.2. In Section 3.3, there is a discussion on how the kinematical spacetimes are connected via geometric limits. Section 3.4 then defines the geometric properties of these spacetimes.

Chapter 4 In this chapter, we turn our attention to the super-kinematical case. These symmetries have not been studied as thoroughly as the classical, kinematical symmetries; therefore, we can only present an algebraic and geometric classification. These may be found in Sections 4.1 and 4.2, respectively. Section 4.3 then describes how the classified superspaces may be connected via limits analogously to the kinematical spacetimes of Chapter 3. The study of the geometric properties of the resulting superspaces is left to future work.

Chapter 5 The last symmetries we consider are the super-Bargmann symmetries. This case is the least studied of the three, so we can only present an algebraic classification in this instance. The direct generalisation of the super-kinematical classification in Section 4.1 is found in Section 5.1. Section 5.2 then extends this classification by adding additional supersymmetric generators.

Chapter 6 This final chapter offers some concluding remarks, noting the possible extensions of the frame- work set out here, areas of further study within the framework, and highlighting some exciting ways the classified symmetries, both algebras and spacetimes, may be utilised.

6 Chapter 2

Mathematical Preliminaries

This chapter will introduce all of the mathematical objects used to construct, describe, and explore our (super-)kinematical spacetime models. As such, this chapter is divided into three sections. In Section 2.1, we begin by defining Lie algebras, gradually introducing more levels of complexity to arrive at concrete definitions for kinematical Lie algebras, generalised Bargmann algebras, kinematical Lie superalgebras, and generalised Bargmann superalgebras. In Sec- tion 2.2, we demonstrate how to integrate these Lie algebras to arrive at spacetime models which hold the relevant symmetries. Finally, in Section 2.3, we describe how to obtain geomet- ric properties for kinematical spacetimes, including fundamental vector fields, soldering forms, vielbeins, and invariant connections. Note, Einstein summation will be assumed throughout.

2.1 Algebra

In this section, we will first introduce the concept of a Lie (super)algebra. This brief discus- sion will define many ideas that will be alluded to throughout the later chapters of this thesis, including the notion of a Lie (super)algebra, Lie (super)algebra homomorphism, Lie subalge- bra and ideal. Once these basics have been reviewed, we move on to discuss kinematical Lie (super)algebras and generalised Bargmann (super)algebras. First, we discuss kinematical Lie algebras, as these are the primary objects all the other types of Lie algebra will generalise. We then discuss generalised Bargmann algebras, which may be viewed as one-dimensional abelian extensions of the underlying kinematical Lie algebra. Kinematical Lie superalgebras are then presented as the supersymmetric extensions of the kinematical algebras. These are the Lie superalgebras s=s¯ s¯, which have a kinematical Lie algebra k as s¯. Finally, we combine 0 ⊕ 1 0 the Bargmann and supersymmetric extensions into the definition of a generalised Bargmann superalgebra. Crucially, in this section, we will present the quaternionic formalism that will play such a dominant role in Chapters 4 and 5.

2.1.1 Lie Algebras and Lie Superalgebras This section will briefly summarise some of the key definitions in Lie (super)algebra theory used throughout this thesis. The material presented here is very well established and may be found in numerous places (see [89–91] for just a few examples). Therefore, we will state definitions, leaving any discussion or elaboration to the sources mentioned above.

Definition 1. An n-dimensional real Lie algebra g consists of an n-dimensional real vector space V equipped with an anti-symmetric, R-bilinear bracket

[−, −] : V V V × → (2.1.1.1) (a, b) [a, b], 7→ which satisfies the Jacobi identity:

[a, [b, c]] = [[a, b], c]+[b, [a, c]] a, b, c V. (2.1.1.2) ∀ ∈

7 Using the typical overloading of notation, we will write both the Lie algebra g=(V, [−, −]) and its underlying vector space as g from now on. Also, g and h should always be taken as Lie algebras, and we will assume we are working over R.

Definition 2. A Lie algebra homomorphism is an R-linear map f : g h which preserves the → Lie bracket: f([a, b])=[f(a), f(b)] a, b, g. (2.1.1.3) ∀ ∈ With this definition of a Lie algebra homomorphism, we note that a Lie algebra isomorphism i :g h is an injective, ker i = 0, surjective, im i = h, Lie algebra homomorphism. → Definition 3. A Lie subalgebra h is an R-linear subspace h g, such that [h, h] h. ⊂ ⊂ Definition 4. An ideal h g is an R-linear subspace such that [g, h] h. ⊂ ⊂ Let h g be an ideal and let g g/h be the canonical projection. The vector space g/h has a ⊂ → unique Lie algebra structure which makes the canonical projection a Lie algebra homomorphism.

Having established some basic definitions for Lie algebras, we now turn to their supersym- metric generalisation.

Definition 5. An (m|n)-dimensional real Lie superalgebra s consists of an (m|n)-dimensional Z2-graded real vector space V = V¯ V¯, equipped with an R-bilinear bracket, which preserves 0 ⊕ 1 the Z2 grading:

[−, −] : Vi Vj Vi+j × → (2.1.1.4) (a, b) [a, b]. 7→ This bracket is anti-symmetric in the super sense,

[λ, µ] = −(−1)ij[µ, λ], (2.1.1.5) where λ s and µ s , and it obeys the super-Jacobi identity: ∈ i ∈ j [λ, [µ, ν]] = [[λ, µ], ν]+(−1)ij[µ, [λ, ν]], (2.1.1.6) where λ s and µ s . ∈ i ∈ j As in the Lie algebra case, we will use the standard overloading of notation, calling both the Lie superalgebra s=(V, [−, −]) and its underlying vector space s. For our purposes, we will usually state the dimension of the Lie superalgebra as d = m + n. The definitions of Lie super- algebra homomorphism, Lie subalgebra, and ideal (in the super sense) follow mutatis mutandis from definitions 2, 3, and 4; therefore, we will omit them here.

The Z2 grading of the Lie superalgebra has some profound consequences, which we will ex- ploit in Chapters 4 and 5. In particular, it states that s0¯ must be a Lie subalgebra of s, and s1 must be an s0¯ module under the adjoint action.

2.1.2 Kinematical Lie Algebras Having established some basic definitions, we now define the primary object that our later investigations will centre on.

1 Definition 6. A kinematical Lie algebra (KLA) k in D spatial dimensions is a 2 (D + 1)(D + 2)- dimensional real Lie algebra containing a rotational subalgebra r isomorphic to so(D) such that, under the adjoint action of r, it decomposes as

k=r 2V R, (2.1.2.1) ⊕ ⊕ where V is a D-dimensional so(D) vector module and R is a one-dimensional so(D) scalar module.

8 We will denote the real basis for these Lie algebras as {Jij,Bi,Pi, H}, where Jij are the generators for the subalgebra r, Bi and Pi span our two copies of V, and H spans the so(D) scalar module. Implicit in this characterisation of kinematical Lie algebras is the assumption of space isotropy, which implies that all the generators transform as expected under the spatial rotations:

[Jij,Jkl]= δjkJil − δikJjl − δjlJik + δilJjk,

[Jij,Bk]= δjkBi − δikBj, (2.1.2.2) [Jij,Pk]= δjkPi − δikPj,

[Jij, H]= 0.

Note that in D spatial dimensions, we always have the so(D) invariant tensors δij and ǫi1i2...iD with which we can define our structure constants. Therefore, when D = 3, we may write the above expressions more concisely. Namely, we may define an so(3) vector module Ji through Jjk =−ǫijkJi, and write

[Ji,Jj]= ǫijkJk, [Ji,Bj]= ǫijkBk, [Ji,Pj]= ǫijkPk, [Ji, H]= 0. (2.1.2.3)

Throughout this thesis, we will frequently use the following abbreviated notation.

[Jij,Bk]= δjkBi − δikBj is equivalent to [J, B]= B,

[H, Bi]= Pi is equivalent to [H, B]= P, (2.1.2.4)

[Bi,Pj]= δijH + Jij is equivalent to [B, P]= H + J, et cetera.

2.1.3 Aristotelian Lie Algebras An Aristotelian Lie algebra (ALA) a in D spatial dimensions may be thought of as a kinematical Lie algebra k with only one copy of the so(D) vector module V. More explicitly, we have the following definition. 1 Definition 7. An Aristotelian Lie algebra (ALA) a in D spatial dimensions is a ( 2 D(D − 1)+ D + 1)-dimensional real Lie algebra containing a rotational subalgebra r isomorphic to so(D) such that, under the adjoint action of r, it decomposes as

a=r V R, (2.1.3.1) ⊕ ⊕ where V is a D-dimensional so(D) vector module and R is a one-dimensional so(D) scalar module.

We will denote the real basis for these Lie algebras as {Jij,Pi, H}, where Jij are the generators for the subalgebra r,Pi span our copy of V, and H spans the so(D) scalar module. Implicit in this characterisation of Aristotelian Lie algebras is the assumption of space isotropy, which implies that all the generators transform as expected under the spatial rotations:

[Jij,Jkl]= δjkJil − δikJjl − δjlJik + δilJjk,

[Jij,Pk]= δjkPi − δikPj, (2.1.3.2)

[Jij, H]= 0.

Since Aristotelian Lie algebras are so similar to kinematical Lie algebras, we will frequently use the term kinematical Lie algebra when referring to both.

2.1.4 Generalised Bargmann Algebras A generalised Bargmann algebra (GBA) ˆk in D spatial dimensions may be thought of as a real one-dimensional abelian extension of a kinematical Lie algebra k. Therefore, we may think of the generalised Bargmann algebras as sitting in short exact sequences

0 R ˆk k 0. (2.1.4.1) → → → →

9 More explicitly, we have the following definition. ˆ 1 Definition 8. A generalised Bargmann algebra (GBA) k in D spatial dimensions is a ( 2 (D + 1)(D + 2)+ 1)-dimensional real Lie algebra containing a rotational subalgebra r isomorphic to so(D) such that, under the adjoint action of r, it decomposes as

ˆk=k R =r 2V 2R, (2.1.4.2) ⊕ ⊕ ⊕ where V is a D-dimensional so(D) vector module and R is a one-dimensional so(D) scalar module. Additionally, we require the GBA to have the following bracket. Denote the basis for the vector modules as Xi and Yi, where 1 6 i 6 D, and let one of the scalar modules have a basis element Z. The required bracket is then [Xi, Yj]= δijZ. (2.1.4.3) As in Section 2.1.2, we will denote the real basis for the kinematical Lie algebra k as {Jij,Bi,Pi, H}, where Jij are the generators for the subalgebra r, Bi and Pi span our two copies of V, and H spans the so(D) scalar module. We will choose to denote the generator of the one-dimensional extension as Z. Given this definition, we inherit the kinematical brackets of (2.1.2.2); however, for completeness, we will restate them here alongside the new brackets brought about by the introduction of Z.

[Jij,Jkl]= δjkJil − δikJjl − δjlJik + δilJjk,

[Jij,Bk]= δjkBi − δikBj, [Jij,Z]= 0, (2.1.4.4) [Jij,Pk]= δjkPi − δikPj, [Bi,Pj]= δijZ.

[Jij, H]= 0,

The brackets of these algebras may also be summarised using the abbreviated notation intro- duced in Section 2.1.2:

[Bi,Pj]= δijZ is equivalent to [B, P]= Z. (2.1.4.5)

2.1.5 Kinematical Lie Superalgebras An N-extended kinematical Lie superalgebra (KLSA) s in three spatial dimensions is a real Lie superalgebra s=s¯ s¯, such that s¯ =k is a kinematical Lie algebra for which D = 3, and s¯ 0 ⊕ 1 0 1 consists of N copies of S, the real four-dimensional spinor module of the rotational subalgebra r =∼ so(3). More explicitly, we have the following definition. Definition 9. An N-extended kinematical Lie superalgebra (KLSA) s in three spatial dimen- sions is a (10 + 4N)-dimensional real Lie superalgebra containing a rotational subalgebra r isomorphic to so(3) such that, under the adjoint action of r, it decomposes as

s=s¯ s¯ =r 2V R NS, (2.1.5.1) 0 ⊕ 1 ⊕ ⊕ ⊕ where V is the three-dimensional so(3) vector module, R is a one-dimensional so(3) scalar module, and S is the four-dimensional so(3) spinor module.

As in Section 2.1.2, we will denote the real basis for the kinematical Lie algebra s0¯ = k as {Ji,Bi,Pi, H}, where Ji are the generators for the subalgebra r, Bi and Pi span our two copies 1 A of V, and H spans the so(D) scalar module. We will choose to denote generators of s1¯ as {Qa }, where 1 6 A 6 N and 1 6 a 6 4. Given this definition, we inherit the kinematical brackets of (2.1.2.3); however, for completeness, we will restate them here alongside the new brackets A brought about by the introduction of {Qa }.

[Ji,Jj]= ǫijkJk,

[Ji,Bj]= ǫijkBk, A 1 A B b [Ji,Qa ]=− 2 δB Qb Γi a , (2.1.5.2) [Ji,Pj]= ǫijkPk,

[Ji, H]= 0,

1 Since we are in the special case of D = 3, we will adopt Ji, defined by Jjk = −ǫijkJi, from the outset.

10 where we can define Γi using the Pauli matrices σi as

iσ 0 0 σ 0 σ Γ = 2 , Γ = 3 , and Γ = 1 . (2.1.5.3) 1 0 iσ 2 −σ 0 3 −σ 0  2  3   1  For now, the only other requirement we will state is that we want to focus on the instances where [Q, Q] = 0. The brackets of these algebras may also be summarised using the abbreviated 6 notation introduced in Section 2.1.2:

A A B b [Ji,Qa ]= δB Qb Γi a is equivalent to [J, Q]= Q. (2.1.5.4)

Quaternionic Notation In Chapter 4, when discussing kinematical Lie superalgebras, we will make extensive use of the following quaternionic formalism. Notice that r =∼ so(3) =∼ sp(1) =∼ Im H, where H denotes the quaternions. We let i, j, and k be the quaternion units such that ij = k and ji = −k. Under the isomorphism r =∼ sp(1), V may be described as a copy of Im H and S may be described as a copy of H where, in both instances, r acts via left quaternion multiplication. Using this isomorphism, we may rewrite the brackets in (2.1.5.2) by invoking the following injective R-linear maps:2

J : Im(H) s¯ such that J(ω)= ω J where ω = ω1i + ω2j + ω3k Im(H), → 0 i i ∈ B : Im(H) s0¯ such that B(β)= βiBi where β = β1i + β2j + β3k Im(H), → ∈ (2.1.5.5) P : Im(H) s¯ such that P(π)= π P where π = π1i + π2j + π3k Im(H), → 0 i i ∈ N N Q : H s¯ such that Q(θ)= θ Q where θ H . → 1 α α ∈ With these maps defined, the brackets of (2.1.5.2) become3

1 [J, J]= J = [J(ω), J(ω′)] = J([ω, ω′]), ⇒ 2 [J, B]= B = [J(ω), B(β)] = 1 B([ω, β]), ⇒ 2 [J, Q]= Q = [J(ω), Q(θ)] = 1 Q(ωθ), [J, P]= P = [J(ω), P(π)] = 1 P([ω, π]), ⇒ 2 ⇒ 2 [J, H]= 0 = [J(ω), H]= 0, ⇒ (2.1.5.6) where ω, β, π Im(H), θ HN, [ω, β] := ωβ−βω, and ωβ is given by quaternion multiplication. ∈ ∈ 2.1.6 Generalised Bargmann Superalgebras An N-extended generalised Bargmann superalgebra (GBSA)s ˆ in three spatial dimensions is a real Lie superalgebra s=s¯ s¯, such that s¯ = ˆk is a generalised Bargmann algebra for 0 ⊕ 1 0 which D = 3, and s1¯ consists of N copies of S, the real four-dimensional spinor module of the rotational subalgebra r =∼ so(3). More explicitly, we have the following definition.

Definition 10. An N-extended generalised Bargmann superalgebra (GBSA)s ˆ in three spatial dimensions is a (11 + 4N)-dimensional real Lie algebra containing a rotational subalgebra r isomorphic to so(3) such that, under the adjoint action of r, it decomposes as

s=s¯ s¯ =r 2V 2R NS, (2.1.6.1) 0 ⊕ 1 ⊕ ⊕ ⊕ where V is the three-dimensional so(3) vector module, R is a one-dimensional so(3) scalar module, and S is the four-dimensional so(3) spinor module.

2 It may be important to note at this stage that θ = θ4 + θ1i + θ2j + θ3k is just a quaternion with real components θi, there are no Grassmann variables. We have used the N = 1 case as an example here, and, as written in the text, we’ve introduced a new index α = (A, a) to capture the basis for the individual quaternions a and the number of quaternionic directions A. We will never use these indices explicitly, so perhaps these comments are not important, but introducing new indices, even for one line, without explaining them seems impolite. 3Solely to keep the notation consistent, when referring to the so(3) scalar module basis element H in this formalism, we will use H. Therefore, if we consider J(ω) = ωiJi to be the map between J and J, the map between H and H is H = H.

11 As these algebras combine the kinematical Lie algebras’ two previous generalisations, we will inherit all the notation we have already established.. In particular, we will denote the real basis for the underlying kinematical Lie algebra k as {Ji,Bi,Pi, H}, where Ji are the generators 4 for the subalgebra r,Bi and Pi span our two copies of V, and H spans the so(D) scalar module. We will choose to denote the generator of the one-dimensional, Bargmann extension as Z, and A we label the generators of s1¯ as {Qa }. The brackets characterising our generalised Bargmann superalgebra ˆs are then

[Ji,Jj]= ǫijkJk, A A B b [Ji,Qa ]= δB Qb Γi a , [Ji,Bj]= ǫijkBk, [Ji,Z]= 0, (2.1.6.2) [Ji,Pj]= ǫijkPk, [Bi,Pj]= δijZ, [Ji, H]= 0, where, in addition, we want to impose the condition [Q, Q] = 0. We will also use the abbreviated 6 notation set up in Sections 2.1.2, 2.1.4, and 2.1.5.

D = 3 Quaternionic Formalism In Chapter 5, when we discuss generalised Bargmann superalgebras in more detail, we will make extensive use of the quaternionic formalism first considered in Section 2.1.5. In particular, we would like to extend this formalism to the case of generalised Bargmann algebras. This is a straightforward procedure. Since Z is an so(3) scalar, the brackets of the generalised Bargmann superalgebras become5

1 [J, J]= J = [J(ω), J(ω′)] = J([ω, ω′]), ⇒ 2 [J, Q]= Q = [J(ω), Q(θ)] = 1 Q(ωθ), [J, B]= B = [J(ω), B(β)] = 1 B([ω, β]), ⇒ 2 ⇒ 2 [J,Z]= 0 = [J(ω), Z]= 0, [J, P]= P = [J(ω), P(π)] = 1 P([ω, π]), ⇒ ⇒ 2 [B, P]= Z = [B(β), P(π)] = Re(βπ¯ )Z. [J, H]= 0 = [J(ω), H]= 0, ⇒ ⇒ (2.1.6.3)

2.2 Geometry

This section will provide the necessary definitions and discussions to understand how the dif- ferent types of Lie algebra, introduced in Section 2.1, can be integrated to describe spacetime geometries. In particular, we will arrive at a homogeneous space description for the kinematical spacetimes and a homogeneous supermanifold description for the kinematical superspaces. The fact we arrive at such a description is an artefact of our choice of approach. Klein’s Erlanger Programme uses a homogeneous space description of geometry; therefore, we arrive at homo- geneous spacetimes by applying this programme to the kinematical story.

A useful result, which will be utilised when defining and exploring the kinematical spacetimes, is an association between homogeneous spaces with respect to a G and coset spaces G/H, where H G is a Lie subgroup. To arrive at the desired definition of a kinematical ⊂ spacetime and define its association to a corresponding kinematical Lie algebra, we will take the following path. First, we will cover some standard material concerning Lie groups, coset spaces, and homogeneous spaces and how they relate to Lie algebras. Although we may find this content in numerous places (see [89–92] for just a few examples), we include it here for com- pleteness. Additionally, this material allows us to establish the notation we will use throughout the rest of this thesis. Anticipating the requirement to specify geometric objects from algebraic objects, we will also define exponential coordinates, which will prove useful for this purpose. Finally, we introduce supermanifolds and Lie supergroups, and show how we may construct superisations of our kinematical spacetimes.

4 As in Section 2.1.5, because we are already in the special case of D = 3, we will use the generator Ji, defined by Jjk = −ǫijkJi, from the outset. 5As in the kinematical Lie algebra case, to keep the notation consistent in this formalism the so(3) scalars H and Z will be denoted H and Z, respectively.

12 2.2.1 Lie Groups and their Lie Algebras This section will briefly summarise some key definitions in Lie (super)group theory used through- out this thesis. The material presented here is very well established and may be found in nu- merous places (see [89–91] for just a few examples). Therefore, we will state definitions, leaving any discussion or elaboration to the sources mentioned above.

Definition 11. An n-dimensional real Lie group G is a group endowed with the structure of a real n-dimensional C (smooth) manifold, such that both the inversion map

∞ s : G G → (2.2.1.1) g g−1, 7→ and multiplication map

m : G G G × → (2.2.1.2) (g, h) gh, 7→ are smooth.

From now on, G and H should always be taken as Lie groups, and we will assume we are working over R.

Definition 12. A Lie group homomorphism f : G H is a smooth map that is also a group → homomorphism.

We can now demonstrate the relationship between a Lie group G and its associated Lie algebra. In particular, we will see that the left-invariant vector fields of G form a Lie algebra, which we will call g. Before proceeding to define the left-invariant vector fields of G, we first require the following definition.

Definition 13. Let ϕ : G H be a Lie group homomorphism and X X (G). We will denote → ∈ the tangent vector produced by X acting on p G as X . We define ∈ p dϕ : TG TH → (2.2.1.3) (p, X ) dϕ(X ). p 7→ p The vector field X is then ϕ-related to Y X (H) if6 ∈ dϕ(X )= Y p G. (2.2.1.4) p ϕ(p) ∀ ∈ For our purposes, it is worth noting that when acting on a smooth function f C (H), the 7 ∈ above expression becomes ∞

dϕ(X )(f)= X (f ϕ)= Y (f). (2.2.1.5) p p ◦ ϕ(p) Using the multiplication map on G, we can define a left translation by g G as ∈

Lg : G G → (2.2.1.6) h m(g, h)= gh. 7→ Definition 14. Let X X (G). Then X is called a left-invariant vector field if it is L -related ∈ g to itself for all g G: ∈ dL X = X L . (2.2.1.7) g ◦ ◦ g 6This expression may also be written as dϕ X = Y ϕ if we do not want to make explicit reference to p G. 7 ◦ ◦ ∈ As above, it is also useful to note that this expression may be written as dϕ X(f) = Y(f) ϕ if we do not ◦ ◦ want to make explicit use of p G. ∈

13 We will denote the space of left-invariant vector fields on G as X (G)L. With this definition, we notice that each left-invariant vector field X is uniquely defined by its value at the identity e G. Therefore, we have an isomorphism T G =∼ X (G)L. An important consequence of this ∈ e isomorphism is that X (G)L takes the form of a real vector space with the same dimension as the Lie group, G.

For any smooth manifold M, there exists a defined on the space of vector fields. This object may be defined as follows.

Definition 15. Let M be a C manifold and X (M) be its set of vector fields. The commutator is an anti-symmetric R-bilinear∞ map, defined

[−, −] : X (M) X (M) X (M) × → (2.2.1.8) (X, Y) [X, Y], 7→ which satisfies the Jacobi identity

[X, [Y, Z]] = [[X, Y], Z]+[Y, [X, Z]] X, Y, Z X (M), (2.2.1.9) ∀ ∈ and acts on smooth functions of the manifold as

[X, Y](f)= X(Y(f))− Y(X(f)) f C (M). (2.2.1.10) ∀ ∈ ∞ Notice, this definition is almost identical to that of a Lie bracket, given in (2.1.1.1). In- deed, we will now show that the commutator of two left-invariant vector fields on G is itself a left-invariant vector field. Thus, the commutator restricts from X (G) to X (G)L. Combining this result with the fact X (G)L forms a dim(G)-dimensional real vector space, we notice that (X (G)L, [−, −]) forms a Lie algebra associated to G, which we call g.

Proving that the commutator restricts to X (G)L is perhaps best seen by first considering the more general setting of ϕ-related vector fields. Explicitly, let X and X′ be ϕ-related to Y and Y , and ϕ : G H be a Lie group homomorphism. If we can show that [X, X ] is ϕ-related ′ → ′ to [Y, Y ′], then we arrive at the desired result by setting ϕ = Lg, Y = X, and Y ′ = X′. Therefore, we want to show dϕ([X, X′]p)(f)=[Y, Y ′]ϕ(p)(f), (2.2.1.11) where p G and f C (H). Taking the left-hand side (L.H.S), we have ∈ ∈ ∞ dϕ([X, X′] )(f)=[X, X′] (f ϕ) p p ◦ = X (X′(f ϕ)) − X′ (X(f ϕ)) p ◦ p ◦ = X (dϕ X′(f)) − X′ (dϕ X(f)) p ◦ p ◦ = Xp(Y ′(f) ϕ)− X′ (Y(f) ϕ) (2.2.1.12) ◦ p ◦ = dϕ(Xp)(Y ′(f)) − dϕ(Xp′ )(Y(f))

= Yϕ(p)(Y ′(f))− Yϕ′ (p)(Y(f))

=[Y, Y ′]ϕ(p)(f).

Thus, we find the desired result. Setting ϕ = Lg, Y = X, and Y ′ = X′, the commutator restricts to the Lie bracket on the dim(G)-dimensional real vector space of left-invariant vector fields, giving the Lie algebra g.

2.2.2 Exponential Coordinates on a Lie Group The above story demonstrates how we may recover a Lie algebra from a Lie group; however, our interests will lie in determining Lie groups’ geometric properties based on their Lie algebras. Therefore, we need to identify a method of utilising our Lie algebra knowledge to explore an associated Lie group. As there may be several Lie groups with the same Lie algebra, we want to clarify from the outset which Lie groups we will be discussing. It is a well-known result that

14 their exists a unique (up to isomorphism) connected, simply-connected Lie group G such that Lie(G)=g, and that all other connected Lie groups G′ with Lie(G′)=g are discrete quotients G′ = G/Z, where Z G is a discrete, central subgroup [91]. Therefore, to make the mapping ⊂ between g and G unique, we will always consider the simply-connected Lie groups.

Anticipating the need to identify geometric objects on G from algebraic objects on Lie(G)=g, we begin by showing how we will construct a local chart (U, σ−1) around any point o G. This ∈ task may be achieved using the exponential map exp :g G; however, to understand how this → map is being used, it helps to first see the following.

Consider a curve γ : R G such that γ(0)= e, where e G is the identity element. In partic- → ∈ ular, we will define this curve such that it is a Lie group homomorphism: γ(s + t) = γ(s)γ(t). Taking the derivative of this curve, we acquire dγ : TR TG. Notice that, at the identity, the → derivative gives us a map R g. Letting the sole basis element of R map to one of the basis → elements of g, let us call it X, it can show that γ is the unique integral curve for the chosen left-invariant vector field [90]. Since the curve is a Lie group homomorphism, and is unique toa particular Lie algebra element X, the image of γ is called the one-parameter subgroup generated by X.

The above story allowed us to take a single Lie algebra element and uniquely determine a Lie group structure associated with it. Using this knowledge, we can determine our map g G. → Definition 16. Let X g and γ be its unique integral curve. The exponential map is a smooth ∈ X map defined

exp :g G → (2.2.2.1) X exp(X)= γ (1). 7→ X With this definition, the one-parameter subgroup generated by X can be understood as having the group multiplication

exp((s + t)X)= exp(sX) exp(tX) s, t, R. (2.2.2.2) ∀ ∈ Now, choosing a point o G, we can establish a local chart using the exponential map. Let ∈ exp :g G, such that o → exp (X)= exp(X) o X g. (2.2.2.3) o ∀ ∈ This map defines a local diffeomorphism from a neighbourhood V of (0,0,...,0) g and a ∈ neighbourhood U of o G. Choosing a basis {e } for g, where 1 6 i 6 n, we can define ∈ i σ : Rn G, where σ(c) = exp (c e ). We now have a local coordinate chart (U, σ−1), such → o i i that o U G maps to (0,0,...,0) V g. We can move this chart around G by using the ∈ ⊂ ∈ ⊂ group multiplication to change the origin, giving us an atlas for G. Later in this section, it will be shown how we may utilise this method of producing coordinates on a Lie group to give us spacetime coordinates.

2.2.3 Coset Spaces Having established the definition of a Lie group G, shown its connection to an associated Lie algebra g, and taken some first steps towards defining geometric objects on G from algebraic objects on g, we now turn to the important topic of defining coset spaces. However, before defining a coset space, we need the concept of a Lie subgroup. Definition 17. A Lie subgroup H G is a submanifold such that H is also a subgroup. ⊂ For our purposes, we will take submanifold to mean that there exists a closed embedding φ : H G. That is → 1. dφ : TH TG is globally injective, → 2. φ is a homeomorphism, and

15 3. φ(H) G is closed. ⊂ Wanting to connect our ideas of Lie subgroups with Lie subalgebras, we remark that there is a one-to-one correspondence between Lie subalgebras h g and connected immersed subgroups ⊂ H G: connected subgroups H G which satisfy condition 1 above, but not 2 or 3 [91]. ⊂ ⊂ Therefore, although knowledge of Lie subalgebras is useful in finding the possible Lie subgroups, additional work is required if we are to determine which subalgebras integrate to subgroups.

Definition 18. Let H G be a Lie subgroup. The coset space G/H is a smooth manifold ⊂ equipped with the quotient topology and the group action inherited from the canonical projec- tion ̟ : G G/H. → For the coset space G/H to be a smooth manifold, we require the H-action on G to be free and proper. From a purely group-theoretic perspective, we know that H is a Lie subgroup of G; therefore, the H-action on G will be free as the action of G on itself is free. Additionally, we note that a closed embedding is equivalent to a proper injective immersion. Therefore, using our definition of a submanifold, we know that φ : H G must be a proper map. This map → being proper means that G/H is a Hausdorff space. Combining this result with the fact the H-action is free, we find that G/H is indeed a smooth manifold.

It is interesting to consider how the exponential coordinate construction, defined earlier for G, may be adapted to the coset space G/H. Let g and h be the Lie algebras of G and H, respectively, and consider exponential coordinates around the identity element e G, such that ∈ expo = exp. From the above discussion, we know that h must be a Lie subalgebra, so we can think of writing g = m h, where m is some vector space complement to h. We can now ⊕ use the fact that there exists a local diffeomorphism exp, which is a slight modification of the exponential map g exp : Vm Vh U × → (2.2.3.1) (X, Y) exp(X) exp(Y), g 7→ from a neighbourhood Vm of (0,0,...,0) m and neighbourhood Vh of (0,0,...,0) h to ∈ ∈ a neighbourhood U of e G [90]. Letting ̟(e) = eH G/H be the identity coset, notice ∈ ∈ ̟ exp(X, Y)= exp|m(X)= exp(X). Choosing a basis {e } for m, where 1 6 i 6 n, we can define ◦ i σ : Rn G /H, where σ(c) = exp(c e ). We now have a local coordinate chart (U, σ−1), such → i i thatgeH G/H mapsg to (0,0,...,0) m. ∈ ∈ 2.2.4 Homogeneous Spaces Above this point, everything is solely about Lie groups and pertains directly to them. Now we shift gear to come in direct contact with Klein’s Erlanger Programme, and discuss the necessary language for our spacetime models.

Definition 19. A Lie group G is called the Lie transformation group of a smooth manifold M if there exists a smooth G-action G M M × → (2.2.4.1) (g, m) g m, 7→ · such that (g1g2) m = g1 ( m) for all g1, g2 G and m M. In particular, if this G-action · · · ∈ ∈ is transitive, that is, for all m, n M, there exists a g G such that g m = n, then M is called ∈ ∈ · a homogeneous space with respect to G. Furthermore, the G-action is said to be effective, if the kernel of the action N = {g G | g m = m, m M} is the identity element, N = {e}, or locally ∈ · ∀ ∈ effective if N G is a discrete subgroup. ⊂ Choosing a distinguished point p in a homogeneous space M, we may define the stabiliser subgroup which fixes the point as

StabG(p)= {g G | g p = p}. (2.2.4.2) ∈ ·

16 This group is sometimes called the isotropy group at p. An important theorem states that not only is H = StabG(p) a Lie subgroup of G, but there exists a local diffeomorphism at p such that M =∼ G/H [90]. Given this mapping, we may seek to describe the homogeneous space M using the Lie algebras g and h. We can achieve such a description using the following definition.

Definition 20. A Klein pair, or Lie pair, (g, h) consists of a Lie algebra g and a Lie subalgebra h, such that the pair is (geometrically) realisable. The pair is geometrically realisable if there exists a Lie group G′ with Lie algebra g′ and Lie subgroup H′ G′ with Lie algebra h′ such that ⊂ G /H describes a homogeneous space, and there exists a Lie algebra isomorphism ϕ : g g ′ ′ ′ → such that ϕ(h′) = h. The homogeneous space G′/H′ is called a (geometric) realisation of (g, h).

Given a Lie pair (g, h), we say that it is effective if h does not contain any non-zero ideals of g. Putting all these definitions together, we have the following important result. There is a one-to-one correspondence between effective Lie pairs (g, h) and homogeneous spaces M =∼ G/H, when we take G to be the unique connected, simply-connected Lie group associated with g act- ing effectively on M. The conditions of geometric realisability and effectiveness are required for this mapping’s existence and uniqueness, respectively.

At this stage, we may introduce some useful definitions which will be alluded to through- out the rest of the thesis. Assuming we have a Lie pair (k, h) with a vector space decomposition k = m h, we call the pair reductive if [h, h] h and [h, m] m. A reductive pair may, in addition, ⊕ ⊂ ⊂ be symmetric if [m, m] h. Alternatively, if [m, m] m, then the corresponding homogeneous ⊂ ⊂ space is called a principal homogeneous space. The intersection of these two instances then defines an affine pair. Explicitly, an affine Lie pair has [m, m] = 0.

The above mapping between Lie pairs and homogeneous spaces tells us that if we want to classify certain types of homogeneous space, we can do so purely at the Lie algebraic level. What we require is a classification of Lie algebras g followed by consistency checks, which make sure we have a suitable Lie subalgebra h, with which we can form an effective Lie pair (g, h). It is this procedure that will be utilised in Section 3.2 to find the possible kinematical spacetimes.

Before proceeding to apply this framework to the kinematical case explicitly, we show how the exponential coordinates for a coset space may be thought of through the lens of homo- geneous spaces. Choose a point o M and let H be the isotropy group at o, such that the ∈ homogeneous space may be described locally as M = G/H. Let g and h be the Lie algebras of G and H, respectively; and write g = m h, where m is some vector space complement to h. ⊕ Using the restricted exponential map exp|m : m M identified in Section 2.2.3, we may write → exp : m M, defined such that o → g exp (X)= exp|m(X) o = exp(X) o X m. (2.2.4.3) o ∀ ∈ This map defines a local diffeomorphism from a neighbourhood V of (0,0,...,0) m and a g ∈ neighbourhood U of o M = G/H. Choosing a basis {e } for m, where 1 6 i 6 n, we can define ∈ i σ : Rn M, where σ(c) = exp (c e ). We now have a local coordinate chart (U, σ−1), such → o i i that o U M maps to (0,0,...,0) V M. We can move this chart around M by using the ∈ ⊂ ∈ ⊂ group action to change the origin, giving us an atlas for M.

There are some natural questions one can ask about the local diffeomorphism exp : m M o → or, equivalently, the local diffeomorphism σ : Rn M. One can ask how much of M is covered → by the image of expo. We say that M is exponential if M = expo(m) and weakly exponential if M = expo(m), where the bar denotes topological closure. Similarly, we can ask about the domain of validity of exponential coordinates: namely, the subspace of Rn where σ remains injective. In particular, if σ is everywhere injective, does it follow that σ is also surjective? We know very little about these questions for general homogeneous spaces, even in the reductive case. However, there are some general theorems for the case of M a .

Theorem 2.2.1 (Voglaire [93]). Let M = G/H be a connected symmetric space with symmetric decomposition g = m h and define exp : m M. Then the following are equivalent: ⊕ o →

17 1. exp : m M is injective o → 2. exp : m M is a global diffeomorphism o → 3. M is simply-connected and for no X m, does ad : g g have purely imaginary eigen- ∈ X → values. Since our homogeneous spaces are by assumption simply-connected, the last criterion in the theorem is infinitesimal and, therefore, easily checked from the Lie algebra. This result makes it a relatively simple task to determine for which of the symmetric spaces the last criterion holds.

Concerning the (weak) exponentiality of symmetric spaces, we will make use of the follow- ing result. Theorem 2.2.2 (Rozanov [94]). Let M = G/H be a symmetric space with G connected. Then 1. If G is solvable, then M is weakly exponential. 2. M is weakly exponential if and only if M = G/H is weakly exponential, where G = G/ Rad(G) and similarly for H, where the radical Rad(G) is the maximal connected solvable normal subgroup of G. c b b b b The Lie algebra of Rad(G) is the radical of the Lie algebra g, which is the maximal solvable ideal, and can be calculated efficiently via the identification rad g=[g, g]⊥, namely, the radical is the perpendicular subspace (relative to the , which may be degenerate) of the first derived ideal.

These two theorems will be used when demonstrating that the action of the boosts are non- compact for our symmetric kinematical spacetimes. Note, this is a very desirable property: if the boosts were compact, they would be more suitably interpreted as additional rotations. We first find those spacetimes which satisfy the third criterion of theorem 2.2.1, determining the instances for which the exponential coordinates define a global chart. It will be shown that, in these cases, showing the non-compactness of the boosts only requires solving a linear ODE. We then find the symmetric kinematical spacetimes which satisfy criterion 2 of theorem 2.2.2. The weak exponentiality of these spacetimes is then exploited to determine the non-compactness of their boosts. The remaining spacetimes require a variety of arguments to demonstrate the non- compactness of their boosts; however, the majority of cases are covered by these two theorems.

2.2.5 Kinematical Spacetimes Now that we have seen how we may describe homogeneous spaces M =∼ G/H in terms of the Lie algebras Lie(G)=g and Lie(H) = h, we may apply this story to the kinematical case. In particular, we wish to use our knowledge of the possible kinematical Lie algebras to define homogeneous spacetime geometries which hold these symmetries. The first step in this procedure is to define, explicitly, what we mean by spacetime geometry. Definition 21. A (homogeneous) kinematical spacetime M is a homogeneous space with respect to a kinematical group K, such that • M is a connected, smooth manifold, • K acts transitively and locally effectively on M with a stabiliser subgroup H, and

• H K is a Lie subgroup whose Lie algebra h contains a rotational subalgebra r =∼ so(D) ⊂ and decomposes as h=r V under the adjoint action of r, where V is a D-dimensional ⊕ so(D) vector module. Notice that not all Lie subgroups H K may be used to describe a kinematical spacetime. ⊂ To distinguish the Lie subalgebras h k and the Lie subgroups H K which may be used ⊂ ⊂ to describe a kinematical spacetime, we will call these subalgebras and subgroups admissible. More explicitly, a Lie subalgebra h k, and its corresponding subgroup H K, will be called ⊂ ⊂ admissible if

18 1. H K is a Lie subgroup, and ⊂ 2. h decomposes under the adjoint action of r as h=r V. ⊕ The first condition ensures that (k, h) defines a Lie pair, and the second condition ensures that h is of the correct form. Thus, we have the following definition.

Definition 22. A kinematical Lie pair (k, h) is a Lie pair consisting of a kinematical Lie algebra k and an admissible Lie subalgbera h. From our previous discussions, we know that the connected, simply-connected kinematical spacetimes M˜ = K˜ /H will be in one-to-one correspondence with kinematical Lie pairs (k, h).

With this prescription for kinematical spacetimes, we may employ the exponential coordinates defined for homogeneous spaces to provide a uniform foundation from which to investigate the differences in spacetime geometry. Explicitly, let H K be the isotropy group at the point ⊂ o M. The group H will always be taken as the Lie subgroup generated by the spatial ro- ∈ tations Jij and the boosts Bi; therefore, m = spanR {H, Pi} for a kinematical Lie algebra. Our coordinates are then defined by the map σ : RD+1 M, such that σ(t, x)= exp (tH + x P). → p · This map gives us a local chart (U, σ−1) centred on o M. ∈ Aristotelian Spacetimes Although the above story holds for the majority of the spacetime classes, Aristotelian spacetimes requires a slightly different treatment. In particular, owing to the absence of the generator B, we need to amend what we mean by an admissible subalgebra in this instance.

Definition 23. A (homogeneous) Aristotelian spacetime M is a homogeneous space with respect to a Aristotelian group A, such that • M is a connected, smooth manifold,

• A acts transitively and locally effectively on M with a stabiliser subgroup R, and

• R A is a Lie subgroup whose Lie algebra is the rotational subalgebra r =∼ so(D). ⊂ Note, only the Lie subalgebras and Lie subgroups which satisfy the above conditions will be deemed admissible, in the Aristotelian sense. With this definition of an Aristotelian spacetime M = A/R, we may specify the form of a Lie pair (a, r) which will be associated to M.

Definition 24. An Aristotelian Lie pair (a, r) is a Lie pair consisting of an Aristotelian Lie algebra a and an admissible Lie subalgbera r. Although these Lie pairs are generally distinct from the kinematical cases, there are a subset of Aristotelian Lie pairs which can arise from kinematical Lie pairs through the following procedure. Let (k, h) be a Lie pair containing a kinematical Lie algebra k and a Lie subalgebra h. If this pair is not effective, h must contain an ideal of k; in particular, since h=r V, the only ⊕ possible ideal is b= spanR {V}. To make this Lie pair effective, we may take the quotient with respect to b to arrive at the pair (k/b, h/b). This effective pair then describes an Aristotelian Lie pair.

2.2.6 Lie Supergroups and Homogeneous Superspaces A Lie supergroup is defined with respect to a supermanifold in an analogous manner to how a Lie group is defined with respect to a classical manifold. With this characterisation, the relationship between Lie superalgebras and Lie supergroups is analogous to the classical set- ting. Due to the increased complexity of the objects involved, this correspondence is highly non-trivial, and a full treatment of this correspondence is not necessary our current purposes. Therefore, we note that a good introduction to this topic is found in [95] and leave this version of the story to the interested reader. To proceed, we still need the notion of a supermanifold and Lie supergroup together with an idea of how we may tie these objects to an associated Lie

19 superalgebra. The method presented here will have a stronger focus on the underlying classical manifold than the one demonstrated in [95]. Such a method is preferable for the current story since the underlying manifold describes our spacetime, which is the primary object of interest.

The rest of this section is written as follows. First, we will introduce our definition of a su- permanifold before introducing Harish-Chandra pairs, which are equivalent to Lie supergroups. We then define the superisation of a homogeneous space, which will be the geometric object describing the supersymmetric generalisations of the kinematical spacetimes of Chapter 3. Fi- nally, we introduce the idea of a super Lie pair, which will be the key object in the classification of kinematical superspaces in Chapter 4.

Before defining our notion of a Lie supergroup, we first need to introduce supermanifolds. We will take our definition of a supermanfiold from [96], such that we arrive at the following. Definition 25. An (m|n)-dimensional real supermanifold is a pair (M, O), where the body M is a smooth m-dimensional real manifold, and the structure sheaf O is a sheaf of supercommutative superalgebras, extending the sheaf of smooth functions C by the subalgebra of nilpotent elements N; that is, we have an exact sequence of sheaves of∞ supercommutative superalgebras:

0 N O C 0, (2.2.6.1) → → → → ∞ where, for every point p M, there is a neighbourhood p U M such that ∈ ∈ ⊂ O(U) =∼ C (U) Λ[θ1, θ2,..., θn]. (2.2.6.2) ⊗ ∞ In the physics literature, superspace is typically referred to through superfields, which are functions on superspace understood in terms of their expansion as a power series in Grassmann coordinates. These Grassmann coordinates are precisely the nilpotent basis elements {θi}. Thus, in the physics nomenclature, the structure sheaf defined above is simply the space of superfields.

A Lie supergroup may be defined as a group object in the category of supermanifolds; however, for our purposes, we will use the following characterisation of Lie supergroups. The category of Lie supergroups is equivalent to the category of Harish-Chandra pairs (G, s), where G is a Lie ∼ group and s is a Lie superalgebra such that s0¯ = Lie(G)=g and the action of g on s1¯ lifts to an action of G on s1¯ by automorphisms [96,97]. The structure sheaf of the Lie supergroup cor- responding to (G, s) is then the sheaf of smooth functions G Λ s¯, which may be interpreted → • 1 as the the smooth sections of a trivial vector bundle G Λ•s¯ over G [97]. × 1 We can now consider the case where the Lie group K in our Harish-Chandra pair (K, s) has an associated homogeneous space M = K/H described by the Lie pair (k, h). Recall that for this mapping between homogeneous space and Lie pair to be unique, K must be connected and simply-connected, with H K closed. We know that s=s¯ s¯ where s¯ = k and s¯ ⊂ 0 ⊕ 1 0 1 must be an s0¯-module; therefore, since K is simply-connected, s1¯ is also a K-module and, by restriction, a H-module. This knowledge allows us to construct the homogeneous vector bundle E = K H s¯. Notice, we may now define a supermanifold (M, O), where the body is M = K/H × 1 and the structure sheaf O is the smooth sections of Λ•E. We will call this supermanifold the superisation of the homogeneous space M defined by the Lie superalgebra s [98].

It is perhaps interesting to note that the superisations presented above all have the form of a split supermanifold; that is, the structure sheaf O is isomorphic to the sheaf of sections of the exterior algebra bundle of a homogeneous vector bundle E M. Letting U M be an open → ⊂ subset, we have C (U)= Γ(U, ΛpE) and N(U)= Γ(U, ΛpE). (2.2.6.3) >0 >1 ∞ M M This result may not be surprising given a theorem by Batchelor, stating that any smooth su- permanifold admits a splitting; although the splitting may not be canonical [99].

It is also interesting to note that any homogeneous supermanifold must be of this form. In-

20 deed, it was shown in [98] that the homogeneous superisation of K/H has the H-equivariant smooth functions K Λ•s¯ as structure sheaf, which are precisely the smooth section of the → 1 homogeneous vector bundle Λ•E over M = K/H, where E = K H s¯. × 1 Since all homogeneous superisations of K/H are of this form, we may think of associating a unique pair (s, h) to each homogeneous superisation. Restricting ourselves to think solely of kinematical spacetimes K/H, we arrive at the following definition.

Definition 26. A super Lie pair (s, h) consists of a kinematical Lie superalgebra s and an admissible Lie subalgebra h, such that the Lie pair (s0¯, h) is geometrically realisable; that is, h s¯ contains the rotational subalgebra r, decomposes as h=r V under the adjoint action ⊂ 0 ⊕ of r, where V s¯ is a vector r module, and h integrates to a Lie subgroup H K. ⊂ 0 ⊂ As in the non-supersymmetric case, a super Lie pair (s, h) is called effective if h does not contain any non-zero ideals of s. We observe that the condition of being geometrically realisable is not associated with supersymmetry, whereas the condition for being effective does take s1¯ into account. We can, therefore, have effective super Lie pairs (s, h) where the underlying Lie pair (s0¯, h) is not effective. In these cases, the copy of V in h acts trivially on the body M of the supermanifold, but acts non-trivially on the odd coordinates. Using physics nomenclature, V generates R-symmetries in these instances.

As in the classical case, there is a one-to-one correspondence between effective super Lie pairs and homogeneous superisations of homogeneous . To the best of our knowledge, this result is part of the mathematical folklore and we are not aware of any reference where this result is proved or even stated as such.

Just as the one-to-one correspondence between effective kinematical Lie pairs and kinematical spacetimes lifts to the supersymmetric case, the correspondence between non-effective kinemat- ical Lie pairs and Aristotelian spacetimes also lifts to the supersymmetric case. Explicitly, we define an Aristotelian super Lie pair (sa, r) as consisting of an Aristotelian Lie superalgebra sa, where sa0¯ = a is an Aristotelian Lie algebra, and a Lie subalgebra r, which is admissible in the Aristotelian sense. We may then form an Aristotelian super Lie pair (sa, r) from a non-effective super Lie pair (s, h) by taking the quotient with respect to the ideal b= spanR {V}, where V is the vector module in the Lie subalgebra h=r V. ⊕

2.3 Geometric Properties

In this final section, we will introduce the geometric properties of homogeneous spaces necessary for beginning to explore the physics of each kinematical spacetime. In particular, we will intro- duce fundamental vector fields, soldering forms, vielbeins, invariant connections and canonical connections. We will see that the fundamental vector fields tell us how the rotations, boosts and spacetime translations act on our spacetime manifold; the soldering forms and vielbeins will allow us to translate between the Lie algebra and geometry; and, the various connections will tell us how to move from one point in spacetime to another. Note, this section deals exclusively with the geometric properties in the non-supersymmetric case. Although a supersymmetric generalisation is possible, it is lies beyond the scope of this thesis.

2.3.1 The Group Action and the Fundamental Vector Fields The action of the group K on M is induced by left multiplication on the group. Indeed, we have a commuting square

L Kg K τ ̟ = ̟ L , (2.3.1.1) ̟ ̟ g ◦ ◦ g τ MMg

21 where L is the diffeomorphism of K given by left multiplication by g K and τ is the g ∈ g diffeomorphism of M given by acting with g. In terms of exponential coordinates, we have g (t, x)=(t′, x′) where ·

g exp(tH + x P)= exp(t′H + x′ P)h, (2.3.1.2) · · for some h H, which typically depends on g, t, and x.8 ∈ If g = exp(X) with X h and if A = tH + x P m, the following identity will be useful: ∈ · ∈

exp(X) exp(A)= exp (exp(adX)A) exp(X). (2.3.1.3)

If M is reductive, so that [h, m] m, then ad A m and, since m is a finite-dimensional vector ⊂ X ∈ space and hence topologically complete, exp(ad )A m as well. In this case, we may act on X ∈ the origin o M, which is stabilised by H, to rewrite equation (2.3.1.3) as ∈

exp(X) expo(A)= expo (exp(adX)A) , (2.3.1.4) or, in terms of σ,

exp(X)σ(t, x)= σ(exp(ad )(tH + x P)) = σ(t′, x′). (2.3.1.5) X · This latter way of writing the equation shows the action of exp(X) on the exponential coordinates (t, x), namely

(t, x) (t′, x′) where t′H + x′ P := exp(ad )(tH + x P). (2.3.1.6) 7→ · X · As we will show in Section 3.4.1, the rotations act in the usual way: they leave t invariant and rotate x, so we will normally concentrate on the action of the boosts and translations. This requires calculating, for example,

i exp(v Pi)σ(t, x)= σ(t′, x′)h. (2.3.1.7)

In some cases, this calculation is not practical and instead we may take v to be very small and work out t′ and x′ to first order in v. This approximation then gives the vector field ξPi generating the infinitesimal action of P . To be more concrete, let X k and consider i ∈

exp(sX)σ(t, x)= σ(t′, x′)h (2.3.1.8)

2 for s small. Since for s = 0, t′ = t, x′ = x, and h = e, we may write (up to O(s ))

exp(sX)σ(t, x)= σ(t + sτ, x + sy) exp(Y(s)), (2.3.1.9) for some Y(s) h with Y(0)= 0, and where τ and y do not depend on s. Equivalently, ∈ exp(sX)σ(t, x) exp(−Y(s)) = σ(t + sτ, x + sy), (2.3.1.10) again up to terms in O(s2). We now differentiate this equation with respect to s at s = 0. Since the equation holds up to O(s2), the differentiated equation is exact.

To calculate the derivative, we recall the expression for the differential of the exponential map (see, e.g., [100]) d exp(X(s)) = exp(X(0))D(ad )X′(0) , (2.3.1.11) ds X(0) s=0 where D is the Maclaurin series corresponding to the analytic function

1 − e−z D(z)= = 1 − 1 z + O(z2). (2.3.1.12) z 2

8Note, as stated, the exponential coordinates here are defined with respect to the Lie group, exp : k K. →

22 (We have abused notation slightly and written equations as if we were working in a matrix group. This is only for clarity of exposition: the results are general.)

Let A = tH + x P. Differentiating equation (2.3.1.10), we find ·

X exp(A)− exp(A)Y ′(0)= exp(A)D(ad )(τH + y P), (2.3.1.13) A · and multiplying through by exp(−A) and using that D(z) is invertible as a power series with inverse the Maclaurin series corresponding to the analytic function F(z)= z/(1 − e−z), we find

G(ad )X − F(ad )Y ′(0)= τH + y P, (2.3.1.14) A A · where we have introduced G(z) = e−zF(z) = z/(ez − 1). It is a useful observation that the analytic functions F and G satisfy the following relations: z z F(z)= K(z2)+ and G(z)= K(z2)− , (2.3.1.15) 2 2

1 2 for some analytic function K(ζ)= 1+ 12 ζ+O(ζ ). To see this, simply notice that F(z)−G(z)= z and that the analytic function F(z)+ G(z) is invariant under z −z. 7→ Equation (2.3.1.14) can now be solved for τ and y on a case-by-case basis. To do this, we need to compute G(adA) and F(adA) on Lie algebra elements. Often a pattern emerges which allows us to write down the result. If this fails, one can bring adA into Jordan normal form and then apply the usual techniques from operator calculus. A good check of our calculations is that the linear map k X (M), sending X to the vector field → ∂ ∂ ξ = τ + yi , (2.3.1.16) X ∂t ∂xi should be a Lie algebra anti-homomorphism: namely,

[ξX, ξY]=−ξ[X,Y]. (2.3.1.17)

We have an anti-homomorphism since the action of k on M is induced from the vector fields which generate left translations on K and these are right-invariant, hence obeying the opposite Lie algebra.

2.3.2 Invariant Connections

Let (k, h) be a Lie pair associated to a reductive homogeneous space. We assume that (k, h) is effective so that h does not contain any non-zero ideals of k. We let k = h m denote a reductive ⊕ split, where [h, m] m. This split makes m into an h-module relative to the linear isotropy ⊂ representation λ : h gl(m), where → λ Y =[X, Y] X h and Y m. (2.3.2.1) X ∀ ∈ ∈ As shown in [101], one can uniquely characterise the invariant affine connections on (k, h) by their Nomizu map α : m m m, an h-equivariant bilinear map; that is, such that for all X h × → ∈ and Y, Z m, ∈ [X, α(Y, Z)] = α([X, Y], Z)+ α(Y, [X, Z]). (2.3.2.2) The torsion and curvature of an invariant affine connection with Nomizu map α are given, respectively, by the following expressions for all X, Y, Z m, ∈

Θ(X, Y)= α(X, Y)− α(Y, X)−[X, Y]m, (2.3.2.3) Ω(X, Y)Z = α(X, α(Y, Z))− α(Y, α(X, Z))− α([X, Y]m, Z)−[[X, Y]h, Z],

23 where [X, Y]=[X, Y]h +[X, Y]m is the decomposition of [X, Y] k = h m. In particular, for the ∈ ⊕ canonical invariant connection with zero Nomizu map, we have

Θ(X, Y)=−[X, Y]m and Ω(X, Y)Z =−λ[X,Y]h Z. (2.3.2.4)

For kinematical spacetimes, we can determine the possible Nomizu maps in a rather uniform way. Rotational invariance determines the form of the Nomizu map up to a few parameters and then we need only study the action of the boosts. We will see that the action of the boosts is common to all spacetimes within a given class: Lorentzian, Riemannian, Galilean, and Carrollian; although the curvature and torsion of the invariant connections of course do depend on the spacetime in question.

2.3.3 The Soldering Form and the Canonical Connection On the Lie group K there is a left-invariant k-valued one-form ϑ: the (left-invariant) Maurer– Cartan one-form [89, 90]. This connection is the canonical invariant connection on K and it obeys the structure equation 1 dϑ =− 2 [ϑ, ϑ], (2.3.3.1) where the notation hides the wedge product in the right-hand side (R.H.S). Using exponential coordinates, we can pull back ϑ to a neighbourhood of the origin on M. The following formula, which follows from equation (2.3.1.11), shows how to calculate it:

σ∗ϑ = D(ad )(dtH + dx P), (2.3.3.2) A · where, as before, A = tH + x P and D is the Maclaurin series corresponding to the analytic · function in (2.3.1.12).

The pull-back σ∗ϑ is a one-form defined near the origin on M with values in the Lie alge- bra k. Since ϑ is the canonical invariant connection on K, we can ask whether we can construct a canonical invariant connection on M using this pull-back. Note, such a one-form must be H-invariant and should take values in k/h. Let m be a vector space complement to h in k so that as a vector space k = h m. This split allows us to write ⊕

σ∗ϑ = θ + ω , (2.3.3.3) where θ is m-valued and and ω is h-valued. Notice that our requirement of H invariance tells us that not only are m and k/h isomorphic as vector spaces, but they are also isomorphic as h modules. Notice, this is precisely the reductive condition. Additionally, since the Maurer- Cartan one-form defines the canonical invariant connection, with vanishing Nomizu map, on K, the one-form we recover from the pullback, corresponds to the canonical invariant connection on M. Therefore, if the Lie pair (k, h) is reductive then ω is the one-form corresponding to the canonical invariant connection on M. The soldering form is then given by θ.

The torsion and curvature of ω are easy to calculate using the fact that ϑ obeys the Maurer– Cartan structure equation (2.3.3.1).9 Indeed, the torsion two-form Θ is given by

1 Θ = dθ +[ω, θ]=− 2 [θ, θ]m (2.3.3.4) and the curvature two-form Ω by

1 1 Ω = dω + 2 [ω, ω]=− 2 [θ, θ]h, (2.3.3.5) which agree with the expressions in equation (2.3.2.4).

In the non-reductive case, ω does not define a connection, but we may still project the lo-

9Let us emphasise that in this work, curvature always refers to the curvature of an invariant affine connection and hence should not be confused with the curvature of the associated Cartan connection, which is always flat for the homogeneous spaces.

24 cally defined k-valued one-form σ∗ϑ to k/h. The resulting local one-form θ with values in k/h is a soldering form which defines an isomorphism T M k/h for every o M near the origin. o → ∈ Wherever θ is invertible, the exponential coordinates define an immersion, which may however fail to be an embedding or indeed even injective. In practice, it is not easy to determine injec- tivity, but it is easy to determine where θ is invertible by calculating the top exterior power of θ and checking that it is non-zero. Provided that θ is invertible, the inverse isomorphism is the vielbein E, where E(o):k/h T M for every o M near the origin. The vielbein allows → o ∈ us to transport tensors on k/h to tensor fields on M and, as we now recall, it takes H-invariant tensors on k/h to K-invariant tensor fields on M.

2.3.4 Invariant Tensors It is well-known that K-invariant tensor fields on M = K/H are in one-to-one correspondence with H-invariant tensors on k/h, and, if H is connected, with h-invariant tensors on k/h. We may assume that H is indeed connected, passing to the universal cover of M, if necessary. In r s practice, given an (r, s)-tensor T on k/h—that is, an element of (k/h)⊗ ((k/h)∗)⊗ —we can ⊗ turn it into an (r, s)-tensor field T on M by contracting with soldering forms and vielbeins r s as appropriate to arrive, for every o M, at T (o) (T M)⊗ (T ∗M)⊗ . Moreover, if T is ∈ ∈ o ⊗ o H-invariant, T is K-invariant.

Our choice of basis for k is such that J and B span h and therefore P := P mod h and H := H mod h span k/h. In the reductive case, k = h m and m =∼ k/h as h-modules. We will let η and a ⊕ π denote the canonical dual basis for (k/h)∗.

Invariant non-degenerate metrics are in one-to-one correspondence with h-invariant non-degenerate symmetric bilinear forms on k/h and characterise, depending on their signature, Lorentzian or Riemannian spacetimes. On the other hand, invariant Galilean structures10 consist of a pair 2 (τ, h), where τ (k/h)∗ and h S (k/h) are h-invariant, h has co-rank 1 and h(τ, −) = 0, if we ∈ ∈ think of h as a symmetric bilinear form on (k/h)∗. On M, τ gives rise to an invariant clock one-form and h to an invariant spatial metric on one-forms. Carrollian structures are dual to Galilean structures and consist of a pair (κ, b), where κ k/h defines an invariant vector field 2 ∈ and b S (k/h)∗ is an invariant symmetric bilinear form of co-rank 1 and such that b(κ, −) = 0. ∈ Homogeneous Aristotelian spacetimes admit an invariant Galilean structure and an invariant Carrollian structure simultaneously.

Invariance under h implies, in particular, invariance under the rotational subalgebra, which is non-trivial for D > 2. Assuming that D > 2, it is easy to write down the possible rotationally invariant tensors, and, therefore, we need only check invariance under B. The action of B is induced by duality from the action on k/h which is given by

λBa (H)= [Ba, H] and λBa (Pb)= [Ba,Pb], (2.3.4.1) with the brackets being those of k. In practice, we can determine this from the explicit expression of the Lie brackets by computing the brackets in k and simply dropping any B or J from the right-hand side. The only possible invariants in k/h are proportional to H, which is invariant provided that [B, H]= 0 mod h. Dually, the only possible invariants in (k/h)∗ are proportional to η, which is invariant provided that there is no X k such that H appears in [B, X]. Omitting ∈ the tensor product symbol, the only rotational invariants in S2(k/h) are linear combinations of 2 2 ab 2 2 2 a b H and P := δ PaPb, whereas in S (k/h)∗ are η and π = δabπ π .

10We will not distinguish notationally the H-invariant tensor from the K-invariant tensor field.

25 26 Chapter 3

Kinematical Spacetimes

In this chapter, we consider the first of our three types of symmetry, kinematical symmetry. These symmetries are the best-studied of the three; therefore, we can present the algebra classi- fication, spacetime classification and explore the spacetimes’ geometric properties. Each of the following two chapters will show the progress made towards recovering a similar description in the super-kinematical and super-Bargmann cases, respectively; however, it is the kinematical case with the fullest picture to-date. Therefore, we may view this chapter as showing the direc- tion in which we hope to take the other two types of symmetry. The kinematical classifications have been derived for dimensions D > 1; however, the classifications of the supersymmetric algebras and spacetimes are limited to D = 3. To keep dimension consistent throughout the thesis, we will focus solely on the D = 3 kinematical spacetimes.

We begin in Section 3.1 by reviewing the classification of kinematical Lie algebras, as pre- sented in [3, 4]. Section 3.2 shows how Figueroa-O’Farrill and Prohazka generated a spacetime classification from the preceding Lie algebra classification in their paper [5]. Section 3.3 then demonstrates how the classified kinematical spacetimes are connected via geometric limits. Fi- nally, in Section 3.4, we go through each of the spacetime geometries identified in Section 3.2 and determine some of their geometric properties.

3.1 Classification of Kinematical Lie Algebras

This section reviews the classification of kinematical Lie algebras, as presented in the papers of Figueroa-O’Farrill [3,4]. These papers aimed to find a methodology for deriving this classifica- tion that would allow us to extend beyond the D = 3 case, which had been completed previously by Bacry and Nuyts in [21]. The methodology that allowed for this generalisation was to con- sider taking deformations of the static kinematical Lie algebra a; that is, the kinematical Lie algebra which has only the kinematical brackets,

[Jij,Jkl]= δjkJil − δikJjl − δjlJik + δilJjk,

[Jij,Bk]= δjkBi − δikBj, (3.1.0.1) [Jij,Pk]= δjkPi − δikPj,

[Jij, H]= 0 : all other brackets vanish. We will not go into the details of this classification as they are beyond the scope of our discussion; we will only briefly note that each non-vanishing deformation gives rise to a new non-vanishing bracket.1

Using this deformation theoretic method, Figueroa-O’Farrill arrived at the classification of kinematical Lie algebras in spatial dimension D > 3, presented in Table 3.1, and the classifica- tion of kinematical Lie algebras unique to spatial dimension D = 3, presented in Table 3.3.

1The Lie algebra one reaches when we can no longer add any more brackets is sometimes called a rigid Lie algebra in the literature [102–104].

27 To make sense of these tables, a few comments are required. First, these tables are util-

Table 3.1: Kinematical Lie Algebras for D > 3

Label Non-zero Lie brackets in addition to [J, J]= J, [J, B]= B, [J, P]= P Comments K1 a K2 [H, B]= P g K3 [H, B]= γB [H, P]= P γ (−1,1) ∈ K4 [H, B]= B [H, P]= P K5 [H, B]=−B [H, P]= P n− K6 [H, B]= B + P [H, P]= P K7χ [H, B]= χB + P [H, P]= χP − B χ> 0 K8 [H, B]= P [H, P]=−B n+ K9 [B, P]= H c K10 [H, B]=−εP [B, B]= εJ [B, P]= H ε = 1 p ± e K11 [H, B]= B [H, P]=−P [B, P]= H + J so(D + 1,1) K12 [H, B]=−εP [H, P]= εB [B, B]= εJ [B, P]= H [P, P]= εJ ε = 1 so(D,2) ± so(D+2)

ising the abbreviated notation first presented in Section 2.1.2. In the final column of each table, we state the names of the known Lie algebras found in the classification. See Table 3.2 for a key. Finally, the kinematical Lie algebras of Table 3.3 all contain brackets not possible when D = 3; in particular, they have either [B, B]= P or [B, B]= P. These brackets are exclusive to 6 D = 3 since, in this dimension, we have the so(3)-invariant vector product εi1i2i3 . Explicitly, this lets us write

[B, B]= B which is equivalent to [Bi,Bj]= εijkBk. (3.1.0.2)

Table 3.2: Notation Summary

Notation Name Notation Name p Poincaré e Euclidean g Galilean c Carroll

n− (Elliptic) Newton-Hooke a Static

n+ (Hyperbolic) Newton-Hooke so Special Orthogonal

Table 3.3: Kinematical Lie Algebras Unique to D = 3

Label Non-zero Lie brackets in addition to [J, J]= J, [J, B]= B, [J, P]= P Comments K13 [B, B]= B [P, P]= ε(B − J) ε = 1 ε ± K14 [B, B]= B K15 [B, B]= P K16 [H, P]= P [B, B]= B K17 [H, B]=−P [B, B]= P K18 [H, B]= B [H, P]= 2P [B, B]= P

28 3.1.1 Classification of Aristotelian Lie Algebras The discussion in this section follows appendix A in [5]. Before moving on to the classification of the spacetime models associated with the kinematical Lie algebras in Tables 3.1 and 3.3, we will pause here for a short discussion on the classification of Aristotelian Lie algebras. This classification is separated from the above story due to the Aristotelian algebras having a different underlying vector space from the other four kinematical symmetry classes; namely, they contain only one copy of the so(D) vector module V. Owing to the relative simplicity of the Aristotelian Lie algebras, their classification only requires the Jacobi identity. In particular, given the fixed Aristotelian brackets

[Jij,Jkl]= δjkJil − δikJjl − δjlJik + δilJjk,

[Jij,Pk]= δjkPi − δikPj, (3.1.1.1)

[Jij, H]= 0, we may write down the most general form for the possible remaining brackets [H, P] and [P, P], to obtain a two-dimensional R vector space, V . We then impose the Jacobi identity to cut out an algebraic variety J V . The possible brackets in D = 3 are ⊂

[H, Pi]= αPi and [Pi,Pj]= βJij + γǫijkPk, (3.1.1.2) where α, β, γ R. Note, we may consistently set γ to zero under a suitable choice of basis; ∈ thus, V is a two-dimensional vector space. Imposing the Jacobi identities, we arrive at the classification shown in Table 3.4.

Table 3.4: Aristotelian Lie Algebras for D = 3

Label Non-zero Lie brackets in addition to [J, J]= J, [J, P]= P Comments A1 a A2 [H, P]= P A3 [P, P]= εJ ε = 1 ±

This method of forming an R vector space V representing the remaining possible brackets and utilising the Jacobi identity to cut out an algebraic variety J V will be generalised and ⊂ employed in the supersymmetric classifications of Chapters 4 and 5.

3.2 Classification of Kinematical Spacetimes

Now that we have seen the classification of the kinematical Lie algebras, we may review how Figueroa-O’Farrill and Prohazka took these algebras and produced a classification of spacetime geometries in [5]. Unlike the Lie algebra classification, we will present a (nearly) complete discussion on the method utilised in acquiring this classification since its direct generalisation will be employed in the supersymmetric case in Chapter 4.

As stated in Section 2.2.5, not all kinematical Lie algebras k necessarily produce a kinematical spacetime. Recall, we are choosing to model our spacetime geometries as homogeneous spaces M with respect to the kinematical Lie group K with Lie(K)=k. Our homogeneous space description then relies on the choice of a point o M with isotropy group H, such that, locally, ∈ we may use the diffeomorphism M = K/H. We then have a unique algebraic description of the connected, simply-connected manifold M˜ = G˜/H in terms of the effective kinematical Lie pair (k, h), where h =∼ Lie(H).

29 The criteria which must be met for the pair (k, h) to form a kinematical spacetime are as follows.

1. (admissibility) The Lie subalgebra h k must contain the rotational Lie subalgebra r =∼ ⊂ so(D), such that, under the adjoint action of r, it decomposes as h=r V, where V is an ⊕ so(D) vector module. 2. (effectivity) The Lie subalgebra h k must not contain any non-zero ideals of k. ⊂ 3. (geometric realisability) The Lie subalgebra h k must integrate to a connected Lie ⊂ subgroup H K˜ . ⊂ Therefore, the task in this section will be to take each of the kinematical Lie algebras k presented in Section 3.1 and check for suitable Lie subalgebras h k. We will now go through each of the ⊂ above criteria, discussing how we can check for Lie algebras which satisfy them.

A Note of Aristotelian Spacetimes A quick inspection of the above criteria highlights that each of the Aristotelian Lie algebras in Table 3.4 will give rise to a unique connected, simply-connected Aristotelian spacetime. First, since all Aristotelian Lie algebras necessarily contain a unique copy of the rotational subalgebra r =∼ so(D), there is only one choice of admissible subalgebra in each instance. Second, the rotational subalgebra is semi-simple, therefore, cannot contain any non-zero ideals of a, making the subalgebra trivially effective. Finally, r =∼ so(D) integrates to a compact subgroup, and compact subgroups are always closed; therefore, (a, r) is always geometrically realisable. Thus, the classification of Aristotelian Lie algebras in Table 3.4 is also a classification of the connected, simply-connected Aristotelian spacetimes up to isomorphism. With this in mind, the rest of the section will concentrate solely on the other classes of kinematical spacetime.

3.2.1 Admissibility Determining the number of admissible Lie subalgebras h k corresponding to a particular ⊂ kinematical Lie algebra k amounts to determining the number of Aut(k)-orbits in the space of so(D) vector modules, spanned by B and P. To see why this is the case, we need to be more specific about which automorphisms we are considering and we need to revisit the definition of an admissible Lie subalgebra. For our current purposes, we will focus exclusively on the automorphisms which fix the rotational subalgebra r; that is, the automorphisms that send the generator J back to itself. By fixing the rotational subalgebra r, the only choice we have in determining our admissible subalgebra h=r V is in our choice of V. Therefore, the number ⊕ of admissible Lie subalgebras h k corresponding to a particular kinematical Lie algebra k will ⊂ be the number of ways we can choose V. We will now show that the number of choices we have for V is exactly the number of Aut(k)-orbits on the space of so(D) vector modules in k.

Since k contains two copies of the so(D) vector module V, we have a two-dimensional real vector space of so(D) vector modules with a basis (B, P). The group Aut(k) then acts on this space as a b (B, P) (B, P) =(aB + cP, bB + dP), (3.2.1.1) 7→ c d   such that Aut(k) GL(2, R). With this prescription, it transpires that there are three possible ⊂ forms for the above matrix, see [5]. These are

a b a 0 a 0 case 1 = , case2 = , and case3 = . (3.2.1.2) c d c d 0 d       We will now see that each of these cases gives rise to a different number of Aut(k)-orbits, and, therefore, a different number of admissible Lie subalgebras h k. To show this result, we ⊂ introduce an arbitrary vector module V = αB + βP =(α, β), where α, β R. In case 1, we have ∈ αB + βP (aα + bβ)B +(cα + dβ)P. (3.2.1.3) 7→

30 Notice that for a suitable choice of a, b, c, and d, we can bring the transformed vector into any form we choose. Let us send it to (1,0). Thus, any kinematical Lie algebra k with an of this form will only have a single admissible Lie subalgebra associated with it. Explicitly, it will have a Lie pair with admissible Lie subalgebra h=r V, where ⊕ V = spanR {B}. In contrast, case 2 gives the transformation

αB + βP aαB +(cα + dβ)P. (3.2.1.4) 7→ Notice that if α = 0, we can always choose to bring the vector into the form (1,0), as in case 6 1; however, if α = 0, we can only produce a vector module of the form (0,1). Thus, any kinematical Lie algebra k with an automorphism group of this form will have two admissible Lie pairs associated with it. Explicitly, it will have a Lie pair with admissible Lie subalgebra h=r V, where V = span {B}, and a Lie pair with admissible Lie subalgebra h=r V, where ⊕ R ⊕ V = spanR {P}. Finally, in case 3, we have

αB + βP aαB + dβP. (3.2.1.5) 7→ Notice that if α = 0, β = 0, we can, again, choose the vector (1,0). Alternatively, if α = 0, β = 0, 6 6 we can choose the vector (0,1). This covers the same two instances as case 2. However, here, we have a third choice. Letting αβ = 0, we can bring our vector into the form (1,1), which, 6 in this case, is a representative of a different Aut(k)-orbit than (1,0) and (0,1). Explicitly, in case 3, we will have a Lie pair with admissible Lie subalgebra h=r V, where V = span {B}, ⊕ R a Lie pair with admissible Lie subalgebra h=r V, where V = span {P}, and a Lie pair with ⊕ R admissible Lie subalgebra h=r V, where V = span {B + P}. ⊕ R Anticipating the final geometric interpretation of the admissible subalgebra’s basis elements, we will want to write h as the span of the spatial rotations J and the boosts B. Therefore, after determining the possible admissible Lie subalgebras, the basis may be relabelled such that h always has the desired form. This procedure for finding the admissible subalgebras will be carried out explicitly in the super-kinematical case in Section 4.2.1.

3.2.2 Effectivity Having determined the Lie pairs (k, h) which contain admissible Lie subalgebras h k, finding ⊂ the effective Lie pairs is relatively straightforward. Since h = spanR {J, B}, the only possible non-zero ideal of k in h is b = spanR {B}. Therefore, since [J, B] = B, we only need to check whether [b, X] b for X {H, B, P}. If this property holds, the pair will not be effective; if the ⊂ ∈ property does not hold, the pair will be effective. For those Lie pairs that are not effective, we can obtain an effective Lie pair by taking the quotient with respect to the ideal, (k/b, h/b).2 Determining the effective Lie pairs is then a matter of inspecting the Lie brackets for each Lie pair containing an admissible Lie subalgebra.

3.2.3 Geometric Realisability Unfortunately, unlike the admissibility and effectivity criteria above, there is no “one-size-fits- all” method of determining a Lie pair’s geometric realisability. For this reason, we will not labour over the details and instead refer the reader to section 4.2 in [5] for the full discussion on this criterion.

3.2.4 Classification Having gone through each of the criteria, the remaining Lie pairs (k, h) may be called kine- matical Lie pairs, and each corresponds to a unique connected, simply-connected homogeneous space, which is taken as the geometry for the associated kinematical spacetime. The spacetimes which arise in this manner are listed in Table 3.5, taken from [5]. Notice, this table has been

2Notice, however, that this Lie pair will no longer be admissible in the kinematical sense. But it may still form an Aristotelian Lie pair.

31 Table 3.5: Simply-Connected Spatially-Isotropic Homogeneous Spacetimes

Label Nonzero Lie brackets in addition to [J, J]= J, [J, B]= B, [J, P]= P Comments M4 [H, B]=−P [B, B]=−J [B, P]= H Minkowski dS4 [H, B]=−P [H, P]=−B [B, B]=−J [B, P]= H [P, P]= J de Sitter AdS4 [H, B]=−P [H, P]= B [B, B]=−J [B, P]= H [P, P]=−J Anti-de Sitter E4 [H, B]= P [B, B]= J [B, P]= H Euclidean S4 [H, B]= P [H, P]=−B [B, B]= J [B, P]= H [P, P]= J Sphere H4 [H, B]= P [H, P]= B [B, B]= J [B, P]= H [P, P]=−J Hyperbolic Space G [H, B]=−P Galilean spacetime dSG [H, B]=−P [H, P]=−B Galilean de Sitter (dSG = dSGγ=−1) dSG [H, B]=−P [H, P]= γB +(1 + γ)P Torsional Galilean de Sitter (γ (−1,1]) γ ∈ AdSG [H, B]=−P [H, P]= B Galilean Anti-de Sitter (AdSG = AdSGχ=0) 2 AdSGχ [H, B]=−P [H, P]=(1 + χ )B + 2χP Torsional Galilean Anti-de Sitter (χ> 0) C [B, P]= H Carrollian Spacetime dSC [H, P]=−B [B, P]= H [P, P]= J Carrollian de Sitter AdSC [H, P]= B [B, P]= H [P, P]=−J Carrollian Anti-de Sitter LC [H, B]= B [H, P]=−P [B, P]= H − J Carrollian Light Cone A Aristotelian Static TA [H, P]= P Torsional Aristotelian Static R S3 [P, P]= J Einstein Static Universe × R H3 [P, P]=−J Hyperbolic Einstein Static Universe × divided into the five kinematical symmetry classes; namely, from top to bottom, Lorentzian, Riemannian, Galilean, Carrollian, and Aristotelian. Now that we have all these spacetime mod- els collected in one place, we can highlight the distinguishing algebraic features of each class. In particular, we note that it is possible to determine the class of spacetime by inspecting only the [H, B] and [B, P] brackets.

Starting from the bottom, we can see that all Aristotelian algebras are without either of these brackets. This fact is not surprising given that Aristotelian algebras do not have the B generators; still, when analysing kinematical algebras, the absence of [H, B] and [B, P] tells us we have an Aristotelian spacetime. Next, we have the Carrollian spacetimes, which are distinguished by the brackets [H, B] = 0 and [B, P] = H. Thus, we may think of the pres- ence of the bracket [B, P] = H as telling us that the kinematical spacetime has an associated Carrollian structure. Conversely, for the Galilean spacetimes, this bracket vanishes; we have [H, B]=−P and [B, P] = 0. In an analogous manner, we may say that the presence of the bracket [H, B]=−P tells us that the kinematical spacetime has an associated Galilean struc- ture. Putting the Galilean and Carrollian brackets together, such that we have [H, B]=−P and [B, P] = H, gives us a Lorentzian spacetime. The fact a Lorentzian spacetime combines Galilean and Carrollian kinematics has a nice interpretation in terms of limits. In brief, when we let the speed of light c go to zero, we recover Carrollian kinematics, and when we take the limit c , we recover Galilean kinematics. This story will be reviewed in the next section. → Finally, Riemannian spacetimes are characterised by the brackets [H, B] = P and [B, P] = H. Notice that∞ the only change from the Lorentzian case is the change in sign for [H, B]. This sign change in the Lie brackets changes the signature of the h-invariant non-degenerate symmetric bilinear form in k, such that it integrates to a Riemannian as opposed to a Lorentzian metric. This discussion on the characteristic Lie brackets of the various kinematical symmetry classes is summarised in Table 3.6.

Table 3.6: The Characteristic Lie Brackets of the Kinematical Symmetry Classes

Class [H, B] [B, P] Lorentzian −P H Riemannian P H Galilean −P 0 Carrollian 0 H Aristotelian - -

32 3.3 Limits Between Spacetimes

Now that we have seen the kinematical spacetimes we can construct from the possible kine- matical Lie algebras, we can investigate how the different models are related to one another. In this section, we will see that the spacetimes in Table 3.5 are related via geometric limits. It will also be shown that most of these limits may be viewed as geometric interpretations of contractions of the underlying kinematical Lie algebras.

To set up this discussion on limits, we will first introduce the idea of a Lie algebra contrac- tion. Equipped with this method of relating the kinematical Lie algebras, we begin with the (semi-)simple Lie algebras so(D + 1,1), so(D,2), and so(D + 2), for the de Sitter, anti-de Sitter, and round sphere, respectively. We then show how each of the other kinematical Lie algebras arises as a contraction from one of these starting points. We end this section by explaining an additional limit which does not arise from a contraction.

Contractions To define a Lie algebra contraction, we first choose to interpret the Jacobi identity as cutting out 2 an algebraic varietyJ in the real vector space Λ V∗ V of all possible anti-symmetric R-bilinear ⊗ maps. Notice, each point in this variety will represent a different Lie algebra structure on the underlying vector space V.3 However, many of these Lie algebra structures may be equivalent. In particular, basis changes in the underlying vector space, given by GL(V), will give rise to isomorphic Lie algebra structures. Therefore, we need to investigate the GL(V)-orbits in J . We will see that Lie algebra contractions arise quite naturally from these investigations.

Recall, an n-dimensional real Lie algebra consists of an n-dimensional real vector space V equipped with an anti-symmetric, R-bilinear bracket ϕ : Λ2V V, which satisfies the Jacobi → identity. As stated above, we may think of the Jacobi identity as cutting out an algebraic vari- 2 ety J Λ V∗ V. Notice, the basis changes of the underlying vector space GL(V) will have ⊂ ⊗ 2 an induced action on Λ V∗ V; in particular, since the action is tensorial, it will preserve J . ⊗ Each GL(V)-orbits in J then represents a unique (up to isomorphism) Lie algebra structure k on V. Now, these orbits may, or may not, be closed with respect to the induced topology on J . If the closure of the orbit contains Lie algebra structures which are not isomorphic to the original structure k, then these non-isomorphic structures are called “degenerations”. Lie algebra contractions are then a specific form of Lie algebra degeneration; in particular, they are limits of curves in the GL(V)-orbit. Let, g : (0,1] GL(V), mapping t g , be a continuous → 7→ t curve with g1 = 1V . We can then define a curve of isomorphic Lie algebras (V, φt), where

φ (X, Y) := g−1 φ (X, Y)= g−1 (φ(g X, g Y)) . (3.3.0.1) t t · t t t  If the limit φ0 = limt 0 φ exists, it defines a Lie algebra g0 =(V, φ0) which is then a contraction → of g=(V, φ1).

3.3.1 Contraction Limits Now that we know how to formulate a Lie algebra contraction, we may see how various con- tractions can implement geometric limits between our spacetimes. Recall, we wished to start our exploration of the various contraction limits with the Lie algebras so(D + 1,1), so(D,2), D+2 and so(D + 2). To avoid any repetition, consider R , and define a basis eM = (ei, e0˜, e1˜), where 1 6 i 6 D. On this space, we define the inner product such that η(ei, ej) = δij, and, κ η(e0˜, e0˜) =: σ and η(e1˜, e1˜) =: are the only other non-vanishing entries. Using this basis, we can view the typical kinematical Lie algebra generators as being embedded into the higher- dimensional generators {JMN}, for 1 6 M, N, 6 D + 2, as

Jij := Jij, Bi := J0˜i,Pi := Ji1˜, H := J0˜1˜. (3.3.1.1)

3This interpretation of the Jacobi identities has already been alluded to in Section 3.1.1. Additionally, it will be generalised to the supersymmetric case in Chapters 4 and 5, playing a crucial role in our classification of the kinematical Lie superalgebras and generalised Bargmann superalgebras.

33 The Lie brackets of our (D + 2)-dimensional Lie algebra then produce the following kinematical Lie algebra k,

[Jij,Jkl]= δjkJil − δikJjl − δjlJik + δilJjk, [H, Pi]=−κBi

[Jij,Bk]= δjkBi − δikBj, [Bi,Pj]= δijH (3.3.1.2) [Jij,Pk]= δjkPi − δikPj, [Bi,Bj]=−σJij

[H, Bi]= σPi [Pi,Pj]=−κJij.

Given these brackets, we can now investigate how our choice of σ and κ affect our possible spacetime models. This picture is perhaps clearest when looking at

[Bi,Bj]=−σJij and [Pi,Pj]=−κJij. (3.3.1.3)

The first bracket above tells us that, for σ = 0, the J and B generators close to form a Lie 6 subalgebra h k, which will be isomorphic to either so(D,1) or so(D + 1) depending on the ⊂ sign of σ. In the geometric picture, h is our admissible Lie subalgebra in the kinematical Lie pair (k, h). From the discussion in Section 2.3.4, we know that the h-invariant tensors on k/h integrate to K-invariant tensor fields on M =∼ K/H; therefore, the change in the signature of this algebra will directly impact the signature of our metric on the spacetime model. In par- ticular, if σ =−1, we will induce a Lorentzian spacetime model, and, if σ =+1, we will induce a Riemannian spacetime model. This connection between σ and the invariant structure on the spacetime will be made more precise in Section 3.4.4.

The second bracket in (3.3.1.3) gives us an analogous algebraic perspective of κ; in partic- ular, for κ = 0, the J and P generators close to form a Lie subalgebra h k, which is isomorphic 6 ⊂ to either so(D,1) or so(D+1). However, since P is not a generator of the admissible Lie subalge- bra for a kinematical Lie pair, we arrive at a different geometric perspective. The fundamental vector fields associated with P will move us around the spacetime manifold M, whereas, since J h, the fundamental vector fields associated with J will implement transformations at a ∈ point; see Section 3.4. Thus, this bracket is capturing the curvature of M. In particular, if κ = −1, we induce a hyperbolic model, such as anti-de Sitter spacetime, and, if κ = +1, we induce an elliptic model, such as de Sitter spacetime.

Now that we have a kinematical Lie algebra k which can capture all three starting points, so(D + 1,1), so(D,2), and so(D + 2), we can introduce the three-parameter family of linear transformations gκ,c,τ, which will allow us to take the desired Lie algebra contractions. These transformations act on the generators by

τ κ g , , J = J, g , , B = B, g , , P = P, g , , H = τκH. (3.3.1.4) κ c τ · κ c τ · c κ c τ · c κ c τ · Putting these definitions into the brackets of k, we have the transformed brackets

2 [Jij,Jkl]= δjkJil − δikJjl − δjlJik + δilJjk, [H, Pi]=−κ κBi 1 [Jij,Bk]= δjkBi − δikBj, [Bi,Pj]= c2 δijH τ 2 (3.3.1.5) [Jij,Pk]= δjkPi − δikPj, [Bi,Bj]=−( c ) σJij 2 κ 2κ [H, Bi]= τ σPi [Pi,Pj]=−( c ) Jij. We now want to take the limits κ 0, c , and τ 0 in turn, corresponding to the flat, → → → non-relativistic, and ultra-relativistic limits, respectively. We will see that by taking multiple limits, we arrive at a “web” of relations, tying∞ all of the kinematical spacetimes together. Since the kinematical brackets are unaffected by these limits, we will omit them from the following discussion, but it should be clear that they still hold for all contractions.

34 Taking the first limit, κ 0, we reduce the set of non-vanishing brackets to → 2 [H, Bi]= τ σPi 1 [Bi,Pj]= c2 δijH (3.3.1.6) τ 2 [Bi,Bj]=−( c ) σJij. This contraction explains multiple geometric limits depending on our choices for κ and σ. These limits are summarised in Table 3.7. From here, we can now take either the non-relativistic or

Table 3.7: Flat Limits

σ κ Geometric Limit −1 1 dS M → −1 −1 AdS M → 1 1 S E → 1 −1 H E → ultra-relativistic limits. First taking c , we are left with only → 2 [H,∞ Bi]= τ σPi. (3.3.1.7) Letting τ be finite and non-vanishing, we thus arrive at the kinematical Lie algebra for the Galilean spacetime G. Notice that the sign of this bracket will depend on our starting point, whether it was a Lorentzian or Riemannian spacetime. Irrespective of the sign, both are equally fair descriptions of the spacetime G; however, as we shall see, this sign does impact the non- relativistic limits of curved spacetimes.

Returning to the kinematical Lie algebra given by the brackets in (3.3.1.6), we can take the ultra-relativistic limit, τ 0, giving → 1 [Bi,Pj]= c2 δijH. (3.3.1.8)

Letting c be finite and non-vanishing, we thus arrive at the kinematical Lie algebra for the Carrollian spacetime C.

Notice that taking both the non-relativistic and ultra-relativistic limits causes all the brackets in (3.3.1.6) to vanish. This produces a non-effective Lie pair, which may be used to describe an Aristotelian Lie pair; in particular, it describes the Aristotelian Lie pair associated with the static spacetime, A.

Now taking the non-relativistic limit of the brackets in (3.3.1.5), we find

2 2 [H, Bi]= τ σPi and [H, Pi]=−κ κBi. (3.3.1.9)

This contraction explains multiple geometric limits depending on our choices for κ and σ. These limits are summarised in Table 3.8. Notice that the non-relativistic spacetimes we contract to

Table 3.8: Non-Relativistic Limits

σ κ Geometric Limit −1 1 dS dSG → −1 −1 AdS AdSG → 1 1 S AdSG → 1 −1 H dSG → switch depending on whether we begin with a Lorentzian or Riemannian spacetime. This result

35 is an artefact of the sign change in [H, B] bracket discussed earlier. Since contraction limits commute, we already know that taking the flat limit will take us to G, thus we find the limits dSG G, and AdSG G. The ultra-relativistic limit then takes us back to A, as described → → previously.

Finally, we consider the ultra-relativistic limit of the brackets in (3.3.1.5) and arrive at

2 [H, Pi]=−κ κBi 1 [Bi,Pj]= c2 δijH (3.3.1.10) κ 2κ [Pi,Pj]=−( c ) Jij. This contraction explains multiple geometric limits depending on our choices for κ and σ. These limits are summarised in Table 3.9. Since the limits commute, we know that taking a subse-

Table 3.9: Ultra-Relativistic Limits

σ κ Geometric Limit −1 1 dS dSC → −1 −1 AdS AdSC → 1 1 S dSC → 1 −1 H AdSC → quent flat limit will give us the spacetime C; thus, we find the limits dSC C and AdSC C. → → Taking the non-relativistic limit from C then takes us to A, as discussed previously.

Table 3.10 summarises the spacetimes discussed in this section so far. They can be characterised as those homogeneous kinematical spacetimes which are symmetric, in the sense described in Section 2.2.4. The table divides into four sections corresponding, from top to bottom, to Lorentzian, Riemannian, Galilean and Carrollian symmetric spacetimes.4

Table 3.10: Symmetric Spacetimes

σ κ c−1 Spacetime −1 0 1 Minkowski (M) −1 1 1 de Sitter (dS) −1 −1 1 anti-de Sitter (AdS) 1 0 1 euclidean (E) 1 1 1 sphere (S) 1 −1 1 hyperbolic (H) 1 0 0 Galilean (G) ∓ 1 1 0 Galilean de Sitter (dSG) ∓ ± 1 1 0 Galilean anti-de Sitter (AdSG) ∓ ∓ 0 0 1 Carrollian (C) 0 1 1 Carrollian de Sitter (dSC) 0 −1 1 Carrollian anti-de Sitter (AdSC)

This discussion accounts for the majority of the limits between the classified spacetimes in Table 3.5; however, it does not include the torsional Galilean spacetimes dSGγ and AdSGχ, the

4Note, the Arisotelian spacetimes will not fit into this picture as they do not have the generator B. Ad- ditionally, we have taken both the Lorentzian and Riemannian contractions into account when describing the parameters for the Galilean spacetimes.

36 Carrollian light cone LC, or the Aristotelian spacetimes TA, R S3, and R H3.5 We will now × × consider each of the remaining classes of limit in turn.

3.3.2 Remaining Galilean Spacetimes In this short section, we will demonstrate how to recover the Galilean spacetime G as a geo- metric limit from the torsional Galilean spacetimes dSGγ and AdSGχ. In particular, we will see that this geometric limit is induced by a contraction of the underlying Lie algebras.

Consider the following linear transformations

g J = J, g B = B, g P = tP, g H = tH. (3.3.2.1) t · t · t · t ·

Under these maps, the kinematical Lie algebras for dSGγ, and AdSGχ are transformed to

[H, B]=−P [H, B]=−P and (3.3.2.2) [H, P]= t2γB + t(1 + γ)P [H, P]= t2(1 + χ2)B + 2tχP, respectively. Notice that taking t 0 produces the kinematical Lie algebra for the spacetime → G in both instances. Thus, we arrive at the geometric limits dSG G and AdSG G. γ → χ → 3.3.3 Limit of the Carrollian Light Cone In this brief section, we will demonstrate how to recover the Carrollian spacetime C and tor- sional static spacetime TA as a geometric limit from the Carrollian light cone LC. In particular, we will see that these geometric limits are induced by a contraction of the underlying Lie alge- bras.

For the first limit, LC C, consider the following linear transformations → g J = J, g B = B, g P = tP, g H = tH. (3.3.3.1) t · t · t · t · Under these maps, the kinematical Lie algebra for LC is transformed to

[H, B]= tB [H, P]= tP [B, P]= H + tJ. (3.3.3.2)

Notice that taking t 0 produces the kinematical Lie algebra for the spacetime C. Thus, we ar- → rive at the geometric limit LC C from a contraction of the underlying kinematical Lie algebra. → The final limit for the Carrollian light cone is LC TA. This will involve taking a quotient → after the appropriate contraction. In particular, transform the basis as

g J = J, g B = B, g P = tP, g H = H. (3.3.3.3) t · t · t · t · The transformed brackets for LC are written

[H, B]= B, [H, P]= P, [B, P]= tH + tJ. (3.3.3.4)

Taking t 0 leaves only the first two brackets. Notice, that [H, B]= B means that the span of → the generators B form an ideal in the Lie subalgebra h; thus, we have a non-effective Lie pair. Quotienting by this ideal, we arrive at an Aristotelian Lie pair isomorphic to the one associated with TA.

3.3.4 Aristotelian Limits The only contraction limits remaining are those associated with the Aristotelian spacetimes A, TA, R S3, and R H3. Here we will show that the Aristotelian Lie algebras associated with × × 5It should be noted that here the use of the term “torsional” is not related to its use in the physics literature, where it states that the clock one-form of a Newton-Cartan structure satisfies dτ = 0. 6

37 the latter three spacetimes all have the Lie algebra corresponding to A as a contraction.

First, let us demonstrate the limit TA A. Consider the linear transformations → g J = J, g P = P, g H = tH. (3.3.4.1) t · t · t · The transformed bracket for TA is written

[H, P]= tP. (3.3.4.2)

Taking the limit t 0, this final bracket vanishes, leaving the Aristotelian Lie algebra for A. → The last limits to describe are R S3 A and R H3 A. Notice that the Aristotelian × → × → analogue of the brackets in (3.3.1.2) are given by

[Jij,Jkl]= δjkJil − δikJjl − δjlJik + δilJjk,

[Jij,Pk]= δjkPi − δikPj, (3.3.4.3)

[Pi,Pj]=−κJij, where κ = +1 describes R S3 and κ = −1 describes R H3. Now, transform these brackets × × using the linear transformations gt, which act on the basis as

g J = J, g P = tP, g H = H. (3.3.4.4) t · t · t · The transformed brackets are then

[Jij,Jkl]= δjkJil − δikJjl − δjlJik + δilJjk,

[Jij,Pk]= δjkPi − δikPj, (3.3.4.5) 2 [Pi,Pj]=−t κJij.

Letting t 0, we have the contraction giving rise to the geometric limits R S3 A and → × → R H3 A. × →

3.3.5 A Non-Contracting Limit The discussion in the previous sections covers the majority of the geometric limits between the spacetimes of Table 3.5; however, there is one final limit which does not arise as a contraction. This limit is found by considering the limit χ for the torsional Galilean algebra AdSG → χ 2 [H, B]=−P [H, P]=(∞1 + χ )B + 2χP. (3.3.5.1) To take this limit, we must transform the basis as follows

H˜ = χ−1H, B˜ = B, P˜ = χ−1P. (3.3.5.2)

The new brackets take the form

[H,˜ B˜]=−P˜, [H,˜ P˜]=(1 + χ−2)B˜ + 2P˜. (3.3.5.3)

Letting χ , we arrive at the Lie algebra corresponding to the spacetime dSG1. → ∞ 3.3.6 Summary The picture resulting from the above discussion is given in Figure 3.1. There are several types of limits displayed in Figure 3.1:

• flat limits in which the curvature of the canonical connection goes to zero: AdS M, → dS M, AdSC C, dSC C, AdSG G and dSG G; → → → → →

38 • non-relativistic limits in which the speed of light goes to infinity: M G, AdS AdSG → → and dS dSG; → In this limit there is still the notion of relativity, it just differs from the standard Lorentzian one. Therefore, it might be more appropriate to call it the “Galilean limit”. • ultra-relativistic limits in which the speed of light goes to zero: M C, AdS AdSC and → → dS dSC. → • limits to non-effective Lie pairs which, after quotienting by the ideal generated by the boosts, result in an Aristotelian spacetime: the dotted arrows LC TA, C A and → → G A; → • LC C, which is a contraction of so(D + 1,1); → • dSG G and AdSG G, which are contractions of the corresponding kinematical Lie γ → γ → algebras;

• limits between Aristotelian spacetimes TA A, R SD A and R HD A; and → × → × →

• a limit limχ AdSGχ = dSG1, which is not due to a contraction of the kinematical Lie → algebras. ∞ Note, we can compose these limits like arrows in a commutative diagram.

AdSC AdS

Lorentzian LC C M Galilean Carrollian AdSG = AdSG0 Aristotelian dS

dSC G AdSGχ>0

A

TA

dSG dSG1 = AdSG dSG γ∈[−1,1] ∞

R × SD R × HD

Figure 3.1: Homogeneous Spacetimes in Dimension D > 3 and Their Limits.

3.4 Geometric Properties of Kinematical Spacetimes

The two previous sections reviewed not only the classification of the kinematical spacetimes, but also described a method of relating the spacetimes through geometric limits. From our perspective, the crucial aspect of this discussion was that both the classification and the limits arose from underlying algebraic procedures. We will extend this connection between the alge- bra and geometric aspects of these spacetimes in this section; in particular, we will now derive invariant connections, invariant structures, fundamental vector fields, soldering forms and viel- beins using techniques which make extensive use of the underlying kinematical Lie algebra.

This section is organised as follows. In Section 3.4.1, we begin by showing that the rotations J act as expected on every kinematical spacetime. This uniform treatment of fundamental vector fields is only possible for the rotations since, by definition, the rotations act the same in every kinematical spacetime; namely, their action is given by the kinematical brackets in (2.1.2.2). In Section 3.4.2, we turn our attention to the action of the boosts B. Here, we will argue that the boosts have non-compact orbits in all Lorentzian, Galilean and Carrollian spacetimes,

39 although, for some instances, complete proofs are postponed until later in the chapter. In Sec- tion 3.4.3, we derive and discuss the space of invariant affine connections for each kinematical spacetime. It will be shown that this space is heavily dependent on the class of the kinematical spacetime being considered, therefore, this discussion is split to describe each class separately. In Section 3.4.4, we derive the geometric properties of the symmetric spacetimes highlighted in Table 3.10. As shown in Section 3.3, these spacetimes may be considered together, with each spacetime being recovered by taking relevant limits. This method of taking limits will be utilised in deriving the invariant structures, fundamental vector fields, soldering forms and vielbeins for these spacetimes. Unfortunately, this method is not available when consider the torsional Galilean spacetimes, Aristotelian spacetimes, or the Carroll light cone; therefore, the geometric properties of these spacetimes are derived on a case-by-case basis in Sections 3.4.5, 3.4.6, and 3.4.7, respectively. Note, we will use the exponential coordinates described in Sec- tion 2.2.2 throughout.

3.4.1 The Action of the Rotations

In this short section, we will derive the fundamental vector field ξJij , corresponding to the action of the rotations on our kinematical spacetimes. We will see rotations act in the way we may naively expect on the exponential coordinates: namely, t is a scalar and xi is a vector.

The infinitesimal action of the rotational generators Jij on the exponential coordinates can be deduced from [Jij, H]= 0 and [Jij,Pk]= δjkPi − δikPj. (3.4.1.1)

To be concrete, consider J12, which rotates P1 and P2 into each other:

[J12,P1]=−P2 and [J12,P2]= P1, (3.4.1.2)

2 but leaves H and P3, ,P inert. We see that ad P =−P for i = 1,2, so that exponenti- ··· D J12 i i ating,

1 2 3 D exp(θ adJ )(tH + x P)= tH + x (cos θP1 − sin θP2)+ x (cos θP2 + sin θP1)+ x P3 + x P 12 · ··· D 1 2 2 1 3 D = tH +(x cos θ + x sin θ)P1 +(x cos θ − x sin θ)P2 + x P3 + + x P . ··· D (3.4.1.3)

1 2 1 2 Restricting attention to the (x , x ) plane, we see that the orbit of (x0, x0) under the one- parameter subgroup exp(θJ12) of rotations is

x1(θ) cos θ sin θ x1 = 0 . (3.4.1.4) x2(θ) − sin θ cos θ · x2      0 Differentiating (x1(θ), x2(θ)) with respect to θ yields

dx1 dx2 = x2 and =−x1, (3.4.1.5) dθ dθ so that ∂ ∂ ξ = x2 − x1 . (3.4.1.6) J12 ∂x1 ∂x2 In the general case, and in the same way, we find

∂ ∂ ξ = xj − xi , (3.4.1.7) Jij ∂xi ∂xj which can be checked to obey the opposite Lie algebra

[ξJij , ξJkl ]=−δjkξJil + δjlξJik + δikξJjl − δilξJjk =−ξ[Jij,Jkl]. (3.4.1.8)

40 3.4.2 The Action of the Boosts For a homogeneous space M = K/H of a kinematical Lie group K to admit a physical interpre- tation as a genuine spacetime, one would seem to require that the boosts act with non-compact orbits [20]. Otherwise, it would be more suitable to interpret them as (additional) rotations. In other words, if (k, h) is the Lie pair describing the homogeneous spacetime, with h the subalgebra spanned by the rotations and the boosts, then a desirable geometrical property of M is that for all X = wiB h the orbit of the one-parameter subgroup B H generated by X should be i ∈ X ⊂ homeomorphic to the real line. Of course, this requirement is strictly speaking never satisfied: the “origin” of M is fixed by H and, in particular, by any one-parameter subgroup of H, so its orbit under any BX consists of just one point. Therefore the correct requirement is that the generic orbits be non-compact. It is interesting to note that we impose no such requirements on the space and time translations.6

With the exception of the Carrollian light cone LC, which will have to be studied separately, the action of the boosts are uniform in each class of spacetimes: Lorentzian, Riemannian, Galilean and Carrollian. (There are no boosts in Aristotelian spacetimes.) We can read the action of the boosts (infinitesimally) from the Lie brackets: • Lorentzian: [B, H]= P, [B, P]= H and [B, B]= J; (3.4.2.1)

• Riemannian: [B, H]=−P, [B, P]= H and [B, B]=−J; (3.4.2.2)

• Galilean: [B, H]= P; (3.4.2.3)

• (Reductive) Carrollian: [B, P]= H; (3.4.2.4)

• and Carrollian Light Cone (LC):

[B, H]=−B and [B, P]= H + J. (3.4.2.5)

Below we will calculate the action of the boosts for all spacetimes except for the Carrollian light cone which will be studied separately.

In order to simplify the calculation, it is convenient to introduce the parameters σ and c from Section 3.3.1, and write the infinitesimal action of the boosts as 1 [B , H]=−σP and [B ,P ]= δ H. (3.4.2.6) i i i j c2 ij Then (σ, c−1) = (−1,1) for Lorentzian, (σ, c−1)=(1,1) for Riemannian, (σ, c−1) = (−1,0) for Galilean and (σ, c−1)=(0,1) for (reductive) Carrollian spacetimes.

The action of the boosts on the exponential coordinates is given by equation (2.3.1.6), which in this case becomes tH + x P exp(adw B)(tH + x P). (3.4.2.7) · 7→ · · From (3.4.2.6), we see that 1 adw B H =−σw P adw B P = wH · · · c2 2 1 and (3.4.2.8) 2 2 1 adw B H =− 2 σw H, · c adw B P =− 2 σw(w P), · c · 6We could wonder whether compact orbits for space and time translations might impose that the boost orbits should be compact; however, given that our choice of coordinates is not adapted to the group action, we can say very little about how the orbits of the different group elements interact. This point may be interesting to investigate in future studies.

41 3 1 2 w w so that in all cases adw B =− c2 σw ad B. This allows us to exponentiate ad B easily: · · ·

sinh z cosh z − 1 2 w w exp(ad B)= 1 + ad B + 2 adw B, (3.4.2.9) · z · z ·

2 1 2 where z =− c2 σw , and hence sinh z exp(adw B)tH = t cosh zH − σt w P, · z · (3.4.2.10) 1 sinh z cosh z − 1 exp(adw B)x P = x P + x wH + (x w)w P. · · · c2 z · w2 · ·

Therefore, the orbit of (t0, x0) under exp(sw B) is given by · 1 sinh(sz) t(s)= t0 cosh(sz)+ 2 x0 w, c z · (3.4.2.11) sinh(sz) cosh(sz) x(s)= x⊥ − σt0 w + (x0 w)w, 0 z w2 ·

x0 w where we have introduced x0⊥ := x0 − w·2 w to be the component of x0 perpendicular to w. It follows from this expression that x⊥(s)= x0⊥, so that the orbit lies in a plane spanned by w and the time direction.

Differentiating these expressions with respect to s, we arrive at the fundamental vector field ξB . Indeed, differentiating (t(s), x(s)) with respect to s at s = 0, we obtain the value of ξw B i · at the point (t0, x0). Letting (t0, x0) vary we obtain that

1 ∂ ∂ ξ = xi − σt . (3.4.2.12) Bi c2 ∂t ∂xi In particular, notice that one of the virtues of the exponential coordinates, is that the funda- mental vector fields of the stabiliser h – that is, of the rotations and the boosts – are linear and, in particular, they are complete. This will be useful in determining whether or not the generic orbits of one-parameter subgroup of boosts are compact.

Let exp(sw B), s R, be a one-parameter subgroup consisting of boosts. Given any o M, · ∈ ∈ its orbit under this subgroup is the image of the map c : R M, where c(s) := exp(sw B)o. → · As we just saw, in the reductive examples (all but LC) the fundamental vector field ξw B is linear in the exponential coordinates, and hence it is complete. Therefore, its integral curves· are one-dimensional connected submanifolds of M and hence either homeomorphic to the real line (if non compact) or to the circle (if compact). The compact case occurs if and only if the map c is periodic.

If the exponential coordinates define a global coordinate chart (which means, in particular, that the homogeneous space is diffeomorphic to RD+1), then it is only a matter of solving a linear ODE to determine whether or not c is periodic. In any case, we can determine whether or not this is the case in the exponential coordinate chart centred at the origin. For the special case of symmetric spaces, which are the spaces obtained via limits from the Riemannian and Lorentzian maximally symmetric spaces, we may use Theorem 2.2.1, which gives an infinites- imal criterion for when the exponential coordinates define a global chart. In particular, by inspecting Table 3.5 and studying the eigenvalues of adH and adPi on k, we may easily deter- mine the spacetimes which satisfy criterion (3) of Theorem 2.2.1, and hence the spacetimes for which the exponential coordinates defines a diffeomorphism to RD+1. These spacetimes are M, E, H, G, dSG, C, and AdSC. Using exponential coordinates, we will see that the orbits of boosts in E and H are compact, whereas the generic orbits of boosts in the other cases are non-compact.

The remaining symmetric spacetimes dS, AdS, S, AdSG, and dSC do not satisfy the infinitesimal criterion (3) in Theorem 2.2.1, and hence the exponential coordinates are not a global chart. It may nevertheless still be the case that the image of expo covers the homogeneous spacetime (or

42 a dense subset). It turns out that S is exponential and AdSG is weakly exponential. The result for S is classical, since the sphere is a compact Riemannian symmetric space, and the case of AdSG follows from Theorem 2.2.2. For AdSG, we have the radicals rad k= spanR {B, P, H} and ∼ ∼ rad h= spanR {B}. Therefore, k/ rad k = so(D) = h/ rad h. Therefore, with K := K/ Rad(K) and similarly for H, K/H is trivially weakly exponential and hence, by Theorem 2.2.2, so is K/H. We will see that boosts act with compact orbits in S, but with non-compactb orbits in AdSG. b b b Among the symmetric spaces in Table 3.5, this leaves dS, AdS, and dSC. We treat those cases using the same technique, which will also work for the non-symmetric LC. Let M be a simply-connected homogeneous spacetime and q : M M a covering map which is equivariant → under the action of (the universal covering group of) K. By equivariance, q(exp(sw B)o) = · exp(sw B)q(o), so the orbit of o M under the boost is sent by q to the orbit of q(o) M. · ∈ ∈ Since q is continuous it sends compact sets to compact sets, so if the orbit of q(o) M is ∈ not compact then neither is the orbit of o M. For M one of dS, AdS, dSC, or LC, there is ∈ some covering q : M M such that we can equivariantly embed M as a hypersurface in some → pseudo-Riemannian space where K acts linearly. It is a simple matter to work out the nature of the orbits of the boosts in the ambient pseudo-Riemannian space (and hence on M), with the caveat that what is a boost in M need not be a boost in the ambient space. Having shown that the boost orbit is non-compact on M, we deduce that the orbit is non-compact on M. In this way, we will show that the generic boost orbits are non-compact for dS, AdS, dSC, and LC.

Finally, this still leaves the torsional Galilean spacetimes dSGγ and AdSGχ, which require a different argument to be explained when we discuss these spacetimes in Section 3.4.5.

3.4.3 Invariant Connections, Curvature, and Torsion for Reductive Spacetimes

In this section, we determine the invariant affine connections for the reductive spacetimes in Tables 3.5. This is equivalent to determining the space of Nomizu maps which, can be done uniformly, a class at a time. We also calculate the curvature and torsion of the invariant con- nections.

For reductive homogeneous spaces there always exists, besides the canonical connection with vanishing Nomizu map, another interesting connection. It is given by the torsion-free connec- 7 1 tion defined by α(X, Y) = 2 [X, Y]m. The canonical and the natural torsion-free connections coincide for symmetric spaces.

For any spacetime the Nomizu maps need to be rotationally invariant, which, when D = 3, gives us α(H, H)= µH α(H, Pi)= νPi (3.4.3.1) α(Pi,Pj)= ζδijH + ζ′ǫijkPk α(Pi, H)= ξPi for some real parameters µ, ν, ζ, ζ′, ξ. Now we simply impose invariance under Bi.

Nomizu Maps for Lorentzian Spacetimes

The Lorentzian spacetimes in Table 3.5 all share the same action of the boosts:

λBi H = Pi and λBi Pi = δijH. (3.4.3.2)

We will impose invariance explicitly in this case to illustrate the calculation and only state the results in all other cases.

7It is the unique Nomizu map with α(X, X) = 0 for all X m and vanishing torsion and called “canonical ∈ affine connection of the first kind” in [105].

43 We calculate (remember (2.3.2.2))

(λBk α)(Pi,Pj)= ζδijPk + ζ′ǫijkH − νδikPj − ξδjkPi, (3.4.3.3) whose vanishing requires ζ = ζ′ = ν = ξ = 0, as can be seen by considering i = j = k, i = k = j, 6 6 and j = k = i in turn. Finally, 6 (λBk α)(H, H)= µPk, (3.4.3.4) whose vanishing imposes µ = 0 and hence the only invariant Nomizu map is the zero map.

Nomizu Maps for Riemannian Spacetimes The situation here is very similar to the Lorentzian case. Now the boosts act as

λBi H =−Pi and λBi Pj = δijH. (3.4.3.5)

The results are as in the Lorentzian case: the only invariant connection is the canonical con- nection.

Nomizu Maps for Galilean Spacetimes On a Galilean spacetime, the boosts act as

λBi H = Pi, (3.4.3.6) and the Pi are invariant. This results in the following invariant Nomizu maps:

α(H, H)=(ν + ξ)H α(H, Pi)= νPi (3.4.3.7) α(Pi,Pj)= 0 α(Pi, H)= ξPi

We will now analyse the curvature and torsion for these Nomizu maps for each Galilean space- time.

Galilean Spacetime (G) The torsion and curvature of the resulting connection have the following non-zero components:

2 Θ(H, Pi)=(ν − ξ)Pi and Ω(H, Pi)H =−ξ Pi. (3.4.3.8)

There is a unique torsion-free, flat invariant connection corresponding to the canonical connec- tion with ν = ξ = 0.

Galilean de Sitter Spacetime (dSG) The torsion and curvature, given by equation (2.3.2.3), have the following non-vanishing components:

2 Θ(H, Pi)=(ν − ξ)Pi and Ω(H, Pi)H =(1 − ξ )Pi. (3.4.3.9)

Therefore, there are two torsion-free, flat invariant connections corresponding to ν = ξ = 1. ± The Nomizu maps for these two connections are

α(H, H)= 2H α(H, H)=−2H

α(H, Pi)= Pi and α(H, Pi)=−Pi (3.4.3.10)

α(Pi, H)= Pi α(Pi, H)=−Pi.

Galilean anti-de Sitter Spacetime (AdSG) The torsion and curvature have have the fol- lowing non-zero components:

2 Θ(H, Pi)=(ν − ξ)Pi and Ω(H, Pi)H = −(1 + ξ )Pi (3.4.3.11)

There are torsion-free connections, but none are flat.

44 Torsional Galilean de Sitter Spacetime (dSGγ=1) The torsion has the following non-zero components Θ(H, Pi)=(ν − ξ − 2)Pi, (3.4.3.12) whereas the only non-zero component of the curvature is

2 Ω(H, Pi)H = −(1 + ξ) Pi. (3.4.3.13)

Therefore, there exists a unique invariant connection with zero torsion and curvature corre- sponding to ν = 1 and ξ =−1:

α(H, Pi)= Pi and α(Pi, H)=−Pi. (3.4.3.14)

Torsional Galilean de Sitter Spacetime (dSGγ=1) In this instance, the torsion is given 6 by Θ(H, Pi)=(ν − ξ −(1 + γ))Pi (3.4.3.15) and the curvature by Ω(H, Pi)H = −(ξ + 1)(ξ + γ)Pi. (3.4.3.16) Therefore, there are precisely two torsion-free, flat invariant connections, with Nomizu maps

α(H, H)=(γ − 1)H α(H, H)=(1 − γ)H

α(H, Pi)= γPi and α(H, Pi)= Pi (3.4.3.17)

α(Pi, H)=−Pi α(Pi, H)=−γPi.

Torsional Galilean anti-de Sitter Spacetime (AdSGχ) The torsion and curvature of the connection corresponding to this Nomizu map are given by the following non-zero components:

2 Θ(H, Pi)=(ν − ξ − 2χ)Pi and Ω(H, Pi)H = −(1 +(ξ + χ) )Pi. (3.4.3.18)

Therefore, we see that there are no flat invariant connections; although there is a one-parameter family of torsion-free invariant connections.

Nomizu Maps for Carrollian Spacetimes

On a Carrollian spacetime, the boosts act as

λBi Pj = δijH, (3.4.3.19) and H is invariant. This results in the following invariant Nomizu maps:

α(H, H)= 0 α(H, Pi)= 0 (3.4.3.20) α(Pi,Pj)= ζδijH α(Pi, H)= 0.

Carrollian Spacetimes (C) In this case, the corresponding invariant connections are flat and torsion-free for all values of ζ.

(Anti-)de Sitter Carrollian Spacetimes (dSC and AdSC) We will treat these two space- times together by introducing κ = 1. Carrollian de Sitter spacetime (dSC) corresponds to ± κ = 1 and Carrollian anti-de Sitter spacetime (AdSC) to κ =−1.

The torsion vanishes and the curvature has the following non-zero components:

Ω(H, Pi)Pj = κδijH and Ω(Pi,Pj)Pk = κ(δjkPi − δikPj), (3.4.3.21) which is never flat. Both of these results are independent of the Nomizu map.

45 Carrollian Light Cone (LC) As show in [5], this homogeneous spacetime does not admit any invariant connections for D > 3.

Nomizu Maps for Aristotelian Spacetimes

In this section, we study the space of invariant affine connections for the Aristotelian spacetimes of Table 3.5. They are all reductive, so there is a canonical invariant connection, and any other invariant connection is determined uniquely by its Nomizu map. The Nomizu maps α : m m m are only subject to equivariance under rotations and are given by (3.4.3.1). × → They depend only on the dimension D and not on the precise Aristotelian spacetime; although, of course, the precise expression for the torsion and curvature tensors does depend on the spacetime. We will calculate the torsion and curvature for each spacetime below.

Static spacetime (A) The torsion and curvature of the most general invariant connection have the following non-zero components:

Θ(H, Pi)=(ν − ξ)Pi,

Θ(Pi,Pj)= 2ζ′ǫijkPk,

Ω(H, Pi)H = ξ(ν − µ)Pi, (3.4.3.22) Ω(H, Pi)Pj = ζ(µ − ν)δijH,

Ω(Pi,Pj)H = 2ξζ′ǫijkPk, and 2 Ω(Pi,Pj)Pk =(ζξ − ζ′ )(δjkPi − δikPj)+ 2ζζ′ǫijkH.

The torsion-free condition implies that ζ′ = 0. With this value of ζ′, the above components reduce to

Θ(H, Pi)=(ν − ξ)Pi,

Ω(H, Pi)H = ξ(ν − µ)Pi, (3.4.3.23) Ω(H, Pi)Pj = ζ(µ − ν)δijH, and

Ω(Pi,Pj)Pk = ζξ(δjkPi − δikPj).

There are then three classes of torsion-free, flat invariant connections in addition to the canonical connection:

1. ζ = 0 and µ = ν = ξ = 0, 6 2. ν = ξ = ζ = 0 and µ = 0, and 6 3. µ = ν = ξ = 0 and ζ = 0. 6 Since the remaining Aristotelian spacetimes all have the same Nomizu maps as this static case, all of them will have the above torsion and curvature components as a base, with a few additional terms included due to the additional non-vanishing brackets of the specific spacetime.

Torsional static spacetime (TA) In this instance, we get the following non-vanishing torsion and curvature components:

Θ(H, Pi)=(ν − ξ − 1)Pi,

Θ(Pi,Pj)= 2ζ′ǫijkPk,

Ω(H, Pi)H = ξ(ν − µ − 1)Pi, (3.4.3.24) Ω(H, Pi)Pj = ζ(µ − ν − 1)δijH − ζ′ǫijkPk,

Ω(Pi,Pj)H = 2ξζ′ǫijkPk, and 2 Ω(Pi,Pj)Pk =(ζξ − ζ′ )(δjkPi − δikPj)+ 2ζζ′ǫijkH.

46 Imposing the torsion-free condition makes ζ′ vanish such that we arrive at

Θ(H, Pi)=(ν − ξ − 1)Pi,

Ω(H, Pi)H = ξ(ν − µ − 1)Pi, (3.4.3.25) Ω(H, Pi)Pj = ζ(µ − ν − 1)δijH, and

Ω(Pi,Pj)Pk = ζξ(δjkPi − δikPj).

As in the static case, we again find three classes of torsion-free, flat invariant connection:

1. ξ = ζ = 0, and ν = 1,

2. µ = ξ = ν − 1, and ζ = 0, and,

3. ξ = 0, ν = 1, and µ = 2.

Aristotelian spacetime (A23ε) The non-vanishing torsion and curvature components are

Θ(H, Pi)=(ν − ξ)Pi,

Θ(Pi,Pj)= 2ζ′ǫijkPk,

Ω(H, Pi)H = ξ(ν − µ)Pi, (3.4.3.26) Ω(H, Pi)Pj = ζ(µ − ν)δijH,

Ω(Pi,Pj)H = 2ξζ′ǫijkPk, and 2 Ω(Pi,Pj)Pk =(ζξ + ε − ζ′ )(δjkPi − δikPj)+ 2ζζ′ǫijkH.

As in the static and torsional static cases, imposing the torsion-free condition sets ζ′ = 0. This means the above components become

Θ(H, Pi)=(ν − ξ)Pi,

Ω(H, Pi)H = ξ(ν − µ)Pi, (3.4.3.27) Ω(H, Pi)Pj = ζ(µ − ν)δijH, and

Ω(Pi,Pj)Pk =(ζξ + ε)(δjkPi − δikPj).

Imposing flatness, we find that this requires ε to vanish; therefore, since ε = 1, we find no ± torsion-free, flat invariant connections.

3.4.4 Pseudo-Riemannian Spacetimes and their Limits In this section, we wish to use the geometric limits that arose from contractions, discussed in Section 3.3, to give us a unified treatment of the geometric properties of the spacetimes admitting such a description. Recall, we had the following brackets in addition to the standard kinematical brackets of (2.1.2.2)

2 2κ 1 τ 2 κ 2κ [H, B]= τ σP [H, P]=−κ B [B, P]= c2 H [B, B]=−( c ) σJ [P, P]=−( c ) J. (3.4.4.1) We will choose to absorb τ and κ into our definition of σ and κ, such that we have 1 σ κ [H, B]= σP, [H, P]=−κB, [B, P]= H, [B, B]=− J, and [P, P]=− J. c2 c2 c2 (3.4.4.2) As before, the parameter σ corresponds to the signature: σ = 1 for Riemannian and σ = −1 for Lorentzian. The parameter κ corresponds to the curvature, so κ = 1,0, −1 for positive, zero and negative curvature, respectively.8 The limit c corresponds to the non-relativistic → limit. In the computations below we will work with unspecified values of σ, κ, c and only at the end will we set them to appropriate values to recover∞ the results for particular spacetimes.

8This definition is tentative due to the possibility of having the curved Galilean spacetimes defined with either κ = 1 or κ = −1.

47 Some of the expressions will have (removable) singularities whenever σ or κ vanish, so will have to think of those cases as limits: the ultra-relativistic limit σ 0 and the flat limit κ 0. → → Invariant Structures We will determine the form of the invariant tensors of small rank. If k = h m is a reductive split ⊕ then, as explained in Section 2.3.4, invariant tensor fields on a simply-connected homogeneous space M = K/H are in bijective correspondence with H-invariant tensors on m, and since H is connected, these are in bijective correspondence with h-invariant tensors on m.

The action of h on m is the linear isotropy representation, which is the restriction to h of the adjoint action: Bi H =−σPi Jij H = 0 · · and 1 (3.4.4.3) Jij Pk = δjkPi − δikPj B P = δ H. · i · j c2 ij

With respect to the canonical dual basis η, πi for m∗, the dual linear isotropy representation is the restriction of the coadjoint action: 1 Jij η = 0 B η =− π · i · c2 i k k k and (3.4.4.4) Jij π =−δi πj + δj πi B πj = σδjη. · i · i It follows that H is invariant in the σ 0 limit, whereas η is invariant in the c limit. → → Concerning the rotationally invariant tensors of second rank, let us observe that ∞

2 2 1 α1H + β1P is invariant σα1 = β1 (3.4.4.5) ⇐⇒ c2 and 2 2 1 α2η + β2π is invariant α2 = σβ2. (3.4.4.6) ⇐⇒ c2 It is interesting to note that the sign κ of the curvature has played no role thus far.

We shall now specialise to the different classes of spacetimes and determine whether and how the structures are induced in the limit.

Lorentzian and Riemannian Case It is clear that for the (pseudo-)Riemannian case, where 1 σ = 0 = c2 , only the metric and its co-metric are invariant. Keeping in mind that we wish the 6 6 1 limit in which the parameters σ and c tend to zero to exist, we set α1 = c2 and β1 = σ and similarly for the co-metric, which leads to the invariants 1 1 H2 + σP2 and ση2 + π2 . (3.4.4.7) c2 c2 For negative (positive) σ this is the invariant Lorentzian (Riemannian) structure. The metric and the co-metric are not per se the inverse of each other, although using definite values for the limiting parameters they can be made to be.

Non- and Ultra-Relativistic Limits Let us now investigate the limits. Taking the non- relativistic limit (c ) of the metrics leads to the invariants → 2 2 ∞ σP and ση , (3.4.4.8) which can be interpreted as the invariants that properly arise from the Lorentzian structure. However, as (3.4.4.4) shows also η itself is an invariant in this limit. This does not follow from the contractions, but can be anticipated from the metrics. We could now take the ultra- relativistic limit (σ 0) of (3.4.4.8) leading to no invariant tensor. Of course, this spacetime → 2 has the invariants H, P , η, π2, but none of these arise from the limit of the original Lorentzian

48 and Riemannian metrics. For the ultra-relativistic limit, we may apply the same logic.

2 1 π2 Concluding, we have the Galilean structure η, σP and the Carrollian structure H, c2 , where we have left the contraction parameters for the invariants that arise from a limit.

Action of the Boosts

The actions of the boosts for all the Lorentzian, Riemannian, Galilean, and reductive Carrollian spacetimes were determined in Section 3.4.2, where we arrived at equation (3.4.2.11) for the orbit of (t0, x0) under the one-parameter family of boosts generated by w B, which we rewrite · here as follows: 1 sinh(sz) t(s)= t0 cosh(sz)+ 2 x0 w c z · (3.4.4.9) sinh(sz) (x0 w) x(s)= x⊥ − σt w + cosh(sz) · w, 0 0 z w2

x0 w 2 1 2 where x⊥ := x0 − ·2 w and z := − 2 σw . Notice that the orbits of (0, x0) with x0 w = 0 0 w c · are point-like. To understand the nature of the other (generic) orbits, we choose values for the parameters. Notice that in our coset parametrisation the boosts do not depend on κ, but only on σ and c. Therefore, we shall be able to treat each class of spacetime uniformly.

−1 2 w2 w Lorentzian boosts Here we take σ =−1 and keep c non-zero. Then z = c2 , so z = c , and the orbits of the boosts are

w 1 sinh(s ) t(s)= t cosh(s w )+ c x w 0 c c2 w 0 c · w (3.4.4.10) sinh (s c ) w (x0 w) x(s)= x⊥ + t0 w + cosh (s ) · w. 0 w c w2 c

Let x = x⊥ + yw, where x⊥ w = 0. Then x⊥(s)= x⊥ for all s and the orbit takes place in the · 0 (t, y) plane. Letting |w| = 1 and c = 1, we find

t(s)= t0 cosh(s)+ sinh(s)y0 and y(s)= t0 sinh(s)+ cosh(s)y0, (3.4.4.11) which is either a point (if t0 = y0 = 0), a straight line (if t0 = y0 = 0), or a hyperbola ± 6 (otherwise). The nature of the orbits in the exponential coordinates is clear, but only in the case of Minkowski spacetime do the exponential coordinates provide a global chart and hence only in that case can we deduce from this calculation that the generic orbits are not compact. For (anti-)de Sitter spacetime, we must argue in a different way.

Let dS denote the quotient of dS which embeds as a quadric hypersurface in Minkowski space- time. The covering map dS dS relates the orbits of the boosts on dS and in the quotient → dS and since continuous maps send compact sets to compact sets, it is enough to show the non-compactness of the orbits in dS. The embedding dS RD+1,1 is given by the quadric ⊂ x2 + + x2 + x2 − x2 = R2, (3.4.4.12) 1 ··· D D+1 D+2 which is acted on transitively by SO(D+1,1). The stabiliser Lie algebra of the point (0, , 0, R,0) ··· is spanned by the so(D + 1,1) generators Jij and Ji,D+2, so that Bi = Ji,D+2, which is a boost in RD+1,1. We have just shown that boosts in Minkowski spacetime have non-compact orbits; therefore, this is the case in dS and hence also in dS.

Similarly, let AdS denote the quotient of AdS which embeds in RD,2 as the quadric

x2 + + x2 − x2 − x2 =−R2. (3.4.4.13) 1 ··· D D+1 D+2

49 The Lie algebra so(D,2) acts transitively on this quadric and the stabiliser Lie algebra at the point (0, ,0,0, R) is spanned by the so(D,2) generators J and J , +1, so that B = J , +1 ··· ij i D i i D which is a “boost” in RD,2. The calculation of the orbit, in this case, is formally identical to the one for Minkowski spacetime (in fact, it takes place in the Lorentzian plane with coordinates (xi, xD+2)) and we see that they are non-compact, so the same holds in AdS and thus also in AdS.

−1 2 w2 w Euclidean “boosts” Here we take σ = 1 and keep c non-zero. Then z =− c2 , so z = i c , and the orbits of the boosts are

w 1 sin(s ) t(s)= t cos(s w )+ c x w 0 c c2 w 0 c · w (3.4.4.14) sin (s c ) w (x0 w) x(s)= x⊥ − t0 w + cos (s ) · w. 0 w c w2 c

As before, letting x = x⊥ + yw, and choosing |w| = 1 and c = 1, we find that the orbit is such that x⊥ is constant and (t, y) evolve as

t(s)= t0 cos(s)+ sin(s)y0 and y(s)=−t0 sin(s)+ cos(s)y0, (3.4.4.15) which is either a point (if t0 = y0 = 0) or a circle (otherwise) and in any case compact. This suffices for E and H since the exponential coordinates give a global chart. For S it is clear that the boosts act with compact orbits because the kinematical Lie group SO(D + 2) is itself compact, therefore, so are the one-parameter subgroups.

Galilean boosts Here we take the limit c and, for definiteness, σ =−1. The orbits of → the boosts are then the limit c of equation (3.4.4.10): → ∞ t(s)= t0 ∞ (3.4.4.16) x(s)= x0 + st0w.

Here the orbits of (0, x0) are point-like. The generic orbit (t0 = 0) is not periodic and hence not 6 compact. This suffices for G and dSG, since the exponential coordinates define a global chart. For AdSG we need to argue differently and this is done later in this section.

Carrollian boosts Here we keep c−1 non-zero, but take the limit σ 0 in equation (3.4.2.11): → 1 t(s)= t0 + s x0 w c2 · (3.4.4.17) x(s)= x0.

Here the orbits (t0, x0) with x0 w = 0 are point-like, but the other orbits are not periodic, · hence not compact. This settles it for AdSC, since the exponential coordinates give a global chart. For the other Carrollian spacetimes we can argue in a different way.

As shown in [106], a Carrollian spacetime admits an embedding as a null hypersurface in a Lorentzian spacetime. For the homogeneous examples in this thesis, this was done in [5] follow- ing the embeddings of the Carrollian spacetimes C and LC as null hypersurfaces of Minkowski spacetime described already in [106].

As explained in Section 3.4.2, for dSC it is enough to work with the discrete quotient dSC, which embeds as a null hypersurface in the hyperboloid model dS of de Sitter spacetime, which itself is a quadric hypersurface in Minkowski spacetime. In [5], it was shown that the boosts in dSC can be interpreted as null rotations in the (higher-dimensional) pseudo-orthogonal Lie group and the orbits of null rotations are never compact. This is done in detail in Section 3.4.7 for LC.

50 Fundamental Vector Fields

The fundamental vector fields for rotations and boosts are linear in exponential coordinates and given by equations (3.4.1.7) and (3.4.2.12), respectively. To determine the fundamental vector fields for the translations we must work harder.

Now let A = tH + x P. Then we have that · κ κ ad2 H = κσtx P − x2H adA H = x B A · c2 · κ κ 1 2 j κ 2 ad B = σtP − x H adA Bi = σtx Jij − σt Bi − xix B A i i c2 i c2 c2 · κ and 1 κ κ (3.4.4.18) ad P = J xj − κtB ad2 P =−κ( x2 + σt2)P + x x P + tx H A i c2 ij i A i c2 i c2 i · c2 i κ 2 adA Jij = xiPj − xjPi ad J =−κt(x B − x B )+ xk(x J − x J ), A ij i j j i c2 i jk j ik so that in general we have 1 ad3 =−κ( x2 + σt2) ad . (3.4.4.19) A c2 A κ 1 2 2 Letting x denote the two complex square roots of − ( c2 x + σt ), with x− = −x+, we can ± 3 2 rewrite this equation as adA = x+ adA.

n Now, if f(z) is analytic in z and admits a power series expansion f(z)= n=0 cnz , then P∞ 1 2k+1 1 2k 2 f(adA)= f(0)+ c2k+1x+ adA + 2 c2kx+ adA . (3.4.4.20) x+ ∞ x ∞ =0 + =1 kX kX Observing that

2k+1 1 2k 1 c2k+1x+ = (f(x+)− f(x−)) and c2kx+ = (f(x+)+ f(x−)− 2f(0)), (3.4.4.21) ∞ 2 ∞ 2 =0 =1 kX kX we arrive finally at

1 1 2 f(adA)= f(0)+ (f(x+)− f(x−)) adA + 2 (f(x+)+ f(x−)− 2f(0)) adA . (3.4.4.22) 2x+ 2x+ Introducing the shorthand notation:

+ 1 − 1 f := 2 (f(x+)+ f(x−)) and f := (f(x+)− f(x−)), (3.4.4.23) 2x+ equation (3.4.4.22) becomes

− 1 + 2 f(adA)= f(0)+ f adA + 2 (f − f(0)) adA . (3.4.4.24) x+

It follows from the above equation and equation (3.4.4.18), that for f(z) analytic in z,

−κ 1 + κ κ 2 f(adA)H = f(0)H + f x B + 2 (f − f(0)) σtx P − 2 x H · x+ · c − 1 1 + κ 2 κ κ j f(adA)Bi = f(0)Bi + f (σtPi − 2 xiH)+ 2 (f − f(0)) − σt Bi − 2 xix B + 2 σtJijx c x+ c · c + − κ κ j 1 + κ f(adA)Pi = f Pi + f (− tBi + 2 Jijx )+ 2 (f − f(0)) 2 xa(tH + x P)  c x+ c · − 1 + κ 1 k f(adA)Jij = f(0)Jij + f (xiPj − xjPi)+ 2 (f − f(0)) −t(xiBj − xjBi)+ 2 x (xiJjk − xjJik) . x+ c (3.4.4.25)

∂ i ∂ Let us calculate ξH = τ ∂t + y ∂xi , where by equation (2.3.1.14)

τH + y P = G(ad )H − F(ad )β B, (3.4.4.26) · A A ·

51 for some β. From equation (3.4.4.25), we have

−κ 1 + κ κ 2 τH + y P = H + G x B + 2 (G − 1) σtx P − c2 x H · · x+ ·  − 1 1 + 2 κ κ i j − β B + F (σtβ P − 2 x βH)+ (F − 1) −κσt β B − 2 x βx B + 2 σtJ β x . · · c · x2 · c · · c ij  +  (3.4.4.27)

By so(D)-covariance, β has to be proportional to x, since that is the only other vector appearing in the B terms, which means that the Jij term above vanishes. This leaves terms in B, H, and P, which allow us to solve for β, τ, and y, respectively. The B terms cancel if and only if

G− β = κx, (3.4.4.28) F+ which we can re-insert into the equation to solve for τ and y. Doing so we find

x coth x − 1 κ x coth x − 1 τ = 1 − + + x2 and ya = + + κσtxa, (3.4.4.29) x2 c2 x2  +   +  so that ∂ x coth x − 1 ∂ 1 ∂ ξ = + + + κ σtxa − x2 . (3.4.4.30) H ∂t x2 ∂xa c2 ∂t  +    ∂ i ∂ To calculate ξv P = τ + y i , equation (2.3.1.14) says we must solve · ∂t ∂x τH + y P = G(ad )v P − F(ad ) β B + 1 λijJ , (3.4.4.31) · A · A · 2 ij ij  for λ , β, τ, and y from the components along Jij, B, H, and P, respectively. The details of the calculation are not particularly illuminating. Let us simply remark that we find

ij i j j i i j j i λ = h1(v x − v x )+ h2(β x − β x ) (3.4.4.32) for κ − κ − 1 (F+ − ) σt G 2 x2 1 c2 h = c h = + 1 1 + κ 2 and 2 1 + κ 2 , (3.4.4.33) 1 − 2 (F − 1) 2 x 1 − 2 (F − 1) 2 x x+ c x+ c and G− β =− κtv, (3.4.4.34) F+ so that κ ij tanh(x+/2) i j j i λ =− 2 (v x − v x ). (3.4.4.35) c x+ Re-inserting these expressions into the equation we solve for τ and y, resulting in

x+ coth x+ − 1 κ τ = 2 2 tx v (3.4.4.36) x+ c · and κ i i x+ coth x+ − 1 i y = x+ coth(x+)v + 2 2 x vx . (3.4.4.37) x+ c · Finally, we have that

x coth x − 1 κ ∂ ∂ ∂ ξ = + + x t + xj + x coth x . (3.4.4.38) Pi x2 c2 i ∂t ∂xj + + ∂xi +  

52 κ 1 2 2 Let us summarise all the fundamental vector fields and remember that x+ = − ( c2 x + σt ) q ∂ ∂ ξ = xj − xi Jij ∂xi ∂xj 1 ∂ ∂ ξ = xi − σt Bi c2 ∂t ∂xi ∂ x coth x − 1 ∂ 1 ∂ (3.4.4.39) ξ = + + + κ σtxi − x2 H ∂t x2 ∂xi c2 ∂t  +    x coth x − 1 κ ∂ ∂ ∂ ξ = + + x t + xj + x coth x . Pi x2 c2 i ∂t ∂xj + + ∂xi +   We can now calculate the Lie brackets of the vector fields which indeed shows the anti- homomorphism with respect to (3.4.4.2)

1 σ κ [ξ , ξ ]=−σξ , [ξ , ξ ]= κξ , [ξ , ξ ]=− ξ , [ξ , ξ ]= ξ , and [ξ , ξ ]= ξ . H B P H P B B P c2 H B B c2 J P P c2 J (3.4.4.40) Let us emphasise that taking the limit of the vector fields and then calculating their Lie bracket leads to the same result as just taking just the limit of the Lie brackets, i.e., these operations commute.

Soldering Form and Connection One-Form

The soldering form and the connection one-form are the two components of the pull-back of the left-invariant Maurer–Cartan form on K. We will calculate it first for all the (pseu- do-)Riemannian cases and then take the flat, non-relativistic and ultra-relativistic limit. As we will see, the exponential coordinates are well adapted for that purpose, and the limits can then be systematically studied. That the limits are well-defined follows from our construction since the quantities we calculate are a power series of the contraction parameters, ǫ = c−1, κ, τ in the ǫ 0 limit and not of their inverse. → For the non-flat (pseudo-)Riemannian geometries our exponential coordinates are, except for the hyperbolic case, neither globally valid nor are quantities like the curvature very compact. Since coordinate systems for these cases are well studied, we will focus in the following mainly on the remaining cases. It is useful to derive the soldering form, the invariant connection and the vielbein in full generality since we take the limit and use them to calculate the remaining quantities of interest.

We start by calculating the Maurer–Cartan form via equation (2.3.3.2) for which we again use equation (3.4.4.25). We find that

−κ 1 + κ κ 2 θ + ω = dtH + D dtx B + 2 (D − 1) σtdtx P − c2 x dtH · x+ ·  + − κ κ i j 1 + κ + D dx P + D (− tdx B + c2 dx x Jij)+ 2 (D − 1) c2 x dx(tH + x P), (3.4.4.41) · · x+ · · which, using that

− 1 − cosh x+ + sinh x+ 1 + sinh x+ − x+ D = 2 , D = and hence 2 (D − 1)= 3 , (3.4.4.42) x+ x+ x+ x+ gives the following expressions:

sinh x+ sinh x+ − x+ κ 1 2 θ = dtH + dx P + 3 σtdtx P + c2 (t x dxH − x dtH + x dxx P) x+ · x+ · · · ·  1 − cosh x+ κ 1 i j ω = 2 dtx B − tdx B − c2 x dx Jij . x+ · ·  (3.4.4.43)

53 We can also evaluate the vielbein E = EHη + EP π which leads us to · κ x2 ∂ ∂ E = −σt2 − x csch x + σ (−1 + x csch x ) txi (3.4.4.44) H x2 c2 + + ∂t + + ∂xi +    κxi ∂ ∂ ∂ E = (−1 + x csch x ) t + xj + x csch x . (3.4.4.45) Pi c2x2 + + ∂t ∂xj + + ∂xi +  

Flat Limit, Minkowski (M) and Euclidean Spacetime (E)

In the flat limit κ 0 the soldering form and connection one-form are given by → θ = dtH + dx P and ω = 0, (3.4.4.46) · respectively, where (t, x) are global coordinates. The vielbein is given by

∂ ∂ E = η + π (3.4.4.47) ∂t ∂x · and the fundamental vector fields, taking the limit of (3.4.4.39), by

1 ∂ ∂ ∂ ∂ ξ = xi − σt , ξ = , and ξ = . (3.4.4.48) Bi c2 ∂t ∂xi H ∂t Pi ∂xi Using the soldering form and the vielbein we can now write the metric and co-metric, given in equation (3.4.4.7), in coordinates

1 1 ∂ ∂ 1 1 g = σdt2 + dx dx g˜ = + σδij . (3.4.4.49) c2 · c2 ∂t ⊗ ∂t ∂xi ⊗ ∂xj Since the connection one-form vanishes the torsion and curvature evaluate to

Ω = 0 Θ = 0. (3.4.4.50)

We can now set σ and c to definite values to obtain the Minkowski spacetime (σ =−1, c = 1), Euclidean space (σ =−1, c = 1), Galilean spacetime (σ = 1, c−1 = 0), and Carrollian spacetime (σ = 0, c = 1). This is obvious enough for the first two cases so that we go straight to the Galilean spacetime.

Galilean Spacetime (G)

For Galilean spacetimes we have the fundamental vector fields

∂ ∂ ∂ ξ = t ξ = ξ = , (3.4.4.51) Bi ∂xi H ∂t Pi ∂xi and the invariant Galilean structure which is characterised by the clock one-form τ = dt and the spatial metric on one-forms h = δij ∂ ∂ . ∂xi ⊗ ∂xj

Carrollian Spacetime (C)

The fundamental vector fields for the Carrollian spacetime are

∂ ∂ ∂ ξ = xi ξ = ξ = , (3.4.4.52) Bi ∂t H ∂t Pi ∂xi

∂ i j and the invariant Carrollian structure is given by κ = ∂t and b = δijdx dx .

54 Non-Relativistic Limit

2 In the non-relativistic limit c we get x+ = √−κσt and the soldering form and connection → one-form are given by ∞ sinh x+ sinh x+ − x+ κ θ = dtH + dx P + 3 σtdtx P x+ · x · + . (3.4.4.53) 1 − cosh x+ κ ω = 2 (dtx B − tdx B) x+ · · We take the non-relativistic limit of the vielbein and obtain ∂ xi ∂ = + ( − ) EH 1 x+ csch x+ i ∂t t ∂x (3.4.4.54) ∂ E = x csch x . Pi + + ∂xi We can now calculate the invariant Galilean structure which is given by the clock one-form and the spatial co-metric (h = σP2):

2 2 ij ∂ ∂ τ = η(θ)= σdt h = x csch x+δ . (3.4.4.55) + ∂xi ⊗ ∂xj The fundamental vector fields are given by

∂ ξ =−σt Bi ∂xi ∂ x coth x − 1 ∂ ξ = + + + κσtxi (3.4.4.56) H ∂t x2 ∂xi  +  ∂ ξ = x coth x . Pi + + ∂xi

Galilean de Sitter Spacetime (dSG)

We start be setting σ =−1 and κ = 1 so that x+ = t and see that

t − sinh(t) sinh(t) θ = dt H + x P + dx P t2 · t ·   (3.4.4.57) 1 − cosh(t) ω = (dtx B − tdx B) . t2 · · The soldering form is invertible for all (t, x), since sinh(t)/t = 0 for all t R. From the above 6 ∈ soldering form, it is easily seen that the torsion two-form vanishes and the curvature two-form is given by 1 Ω = sinh(t)B (dt ∧ dxi). (3.4.4.58) t i The vielbein is given by

∂ xi ∂ ∂ E = + (1 − t csch t) and E = t csch t . (3.4.4.59) H ∂t t ∂xi Pi ∂xi We can thus find the invariant Galilean structure: the clock one-form is given by τ = η(θ)= dt and the spatial co-metric is given by

∂ ∂ h = t2 csch2 tδij . (3.4.4.60) ∂xi ⊗ ∂xj

55 Finally, the fundamental vector fields are

∂ ξ = t (3.4.4.61) Bi ∂xi ∂ 1 ∂ ξ = + − coth(t) xi (3.4.4.62) H ∂t t ∂xi   ∂ ξ = t coth(t) . (3.4.4.63) Pi dxi We can change coordinates to bring the Galilean structure into a nicer form; in particular, let

i sinh(t) i t′ = t and x′ = x . (3.4.4.64) t

In these new coordinates we find τ = dt′ and

ij ∂ ∂ h = δ i j . (3.4.4.65) ∂x′ ⊗ ∂x′

Galilean Anti-de Sitter Spacetime (AdSG)

For σ =−1 and κ = 1 the soldering form and connection one-form for the canonical invariant connection are t − sin t sin t θ = dt H + x P + dx P t2 · t ·   (3.4.4.66) 1 − cos t ω = (tx B − tdx B) . t2 · · Because of the zero of sin(t)/t at t = π, the soldering form is an isomorphism for all x and ± for t (−π, π), so that the exponential coordinates are invalid outside of that region. Let ∈ D t0 (−π, π) and x0 R . The orbit of the point (t0, x0) under the one-parameter subgroup of ∈ ∈ boosts generated by w B is ·

t(s)= t0 and x(s)= x0 + st0w. (3.4.4.67)

The orbits are point-like for t0 = 0 and straight lines for t0 = 0. These orbits remain inside 6 the domain of validity of the exponential coordinates. The generic orbits are, therefore, non- compact.

The torsion two-form again vanishes and the curvature form is 1 Ω = sin tB (dt ∧ dxi). (3.4.4.68) t i The vielbein is given by

∂ t xi ∂ t ∂ E = + 1 − and E = , (3.4.4.69) H ∂t sin t t ∂xi Pi sin t ∂xi   so that the invariant Galilean structure has a clock one-form τ = η(θ) = dt and a spatial co-metric t 2 ∂ ∂ h = δij . (3.4.4.70) sin t ∂xi ⊗ ∂xj   The fundamental vector fields for Galilean AdS are ∂ ξ = t Bi ∂xi ∂ ξ = t cot t (3.4.4.71) Pi ∂xi ∂ ∂ ξ = + 1 − cot t xi . H ∂t t ∂xi  56 As in the case of dSG, we can bring the Galilean structure into a nicer form using a change of coordinates: i sin(t) i t′ = t and x′ = x . (3.4.4.72) t

With this change of coordinates, we have τ = dt′ and

ij ∂ ∂ h = δ i j . (3.4.4.73) ∂x′ ⊗ ∂x′

Ultra-Relativistic Limit

In the ultra-relativistic limit σ 0 to the Carrollian (anti-)de Sitter spacetimes we get x+ = κ 2 → − c2 x and the soldering form and invariant connection are p sinh x+ sinh x+ x dx θ = (dtH + dx P) + 1 − · (tH + x P) x · x x2 · +  +  (3.4.4.74) cosh x+ − 1 2 1 i j ω = c dtx B − tdx B − 2 J x dx . x2 · · c ij  The vielbein in the ultra-relativistic limit has the following form

∂ E = x csch x H + + ∂t (3.4.4.75) xi ∂ ∂ ∂ E = (1 − x csch x ) t + xj + x csch x . Pi x2 + + ∂t ∂xj + + ∂xi  

The ultra-relativistic limit leads to Carrollian structure consisting of κ = EH and the spatial 1 π2 metric b = c2 given by

2 2 2 1 sinh x+ 1 sinh x+ (x dx) b = dx dx + 1 − · . (3.4.4.76) c2 x · c2 x x2  +   +  ! The fundamental vector fields are 1 ∂ ξ = xi Bi c2 ∂t ∂ ξH = x+ coth x+ (3.4.4.77) ∂t xi ∂ ∂ ∂ ξ = (1 − x coth x ) t + xj + x coth x . Pi x2 + + ∂t ∂xj + + ∂xi  

(Anti-)de Sitter Carrollian Spacetimes (dSC and AdSC)

We will treat these two spacetimes together, such that κ = 1 corresponds to Carrollian de Sit- ter (dSC) and κ =−1 to Carrollian anti-de Sitter (AdSC) spacetimes. Furthermore we set c = 1.

We find that the soldering form is given by

κ sin |x| sin |x| x dx θ( =1) = (dtH + dx P)+ 1 − · (tH + x P) |x| · |x| x2 ·   (3.4.4.78) κ sinh |x| sinh |x| x dx θ( =−1) = (dtH + dx P)+ 1 − · (tH + x P). |x| · |x| x2 ·   sin |x| κ sinh |x| These soldering forms are invertible whenever the functions |x| (for = 1) or |x| (for κ =−1) are invertible. The latter function is invertible for all x, whereas the former function is invertible in the open ball |x| < π.

57 The connection one-form is given by

(κ=1) cos |x| − 1 i j ω = 2 (dtx B − tdx B + dx x Jij) x · · (3.4.4.79) cosh |x| − 1 ω(κ=−1) = (dtx B − tdx B + dxixjJ ). x2 · · ij The canonical connection is torsion-free, since (A)dSC is symmetric, but it is not flat. The curvature is given by

2 κ sin |x| sin |x| sin |x| x B Ω( =1) = dt ∧ dx B − − 1 · dt ∧ dx x+ |x| · |x| |x| x x ·     · sin |x| 2 2 sin |x| sin |x| J dxi ∧ dxj + − 1 (xkxjJ − txjB )dxi ∧ dxj, |x| ij |x| |x| ik i     2 κ sinh |x| sinh |x| sinh |x| x B Ω( =−1) =− dt ∧ dx B + − 1 · dt ∧ dx x− |x| · |x| |x| x x ·     · sinh |x| 2 2 sinh |x| sinh |x| J dxi ∧ dxj − − 1 (xkxjJ − txjB )dxi ∧ dxj. |x| ij |x| |x| ik i     (3.4.4.80)

Using the soldering form, we find the vielbein E to have components

i κ ∂ κ x ∂ ∂ ∂ E( =1) = |x| csc |x| and E( =1) = (1 − |x| csc |x|) t + xj + |x| csc |x| , H ∂t Pi x2 ∂t ∂xj ∂xi   i κ ∂ κ x ∂ ∂ ∂ E( =−1) = |x| csch |x| and E( =−1) = (1 − |x| csch |x|) t + xj + |x| csch |x| . H ∂t Pi x2 ∂t ∂xj ∂xi   (3.4.4.81)

The invariant Carrollian structure is given by κ = EH and the spatial metric

2 2 2 (κ=1) sin |x| sin |x| (x dx) b = dx dx + 1 − · 2 |x| · |x| ! x     (3.4.4.82) 2 2 2 κ sinh |x| sinh |x| (x dx) b( =−1) = dx dx + 1 − · . |x| · |x| x2     ! Finally, the fundamental vector field of our curved ultra-relativistic algebras are

∂ ξ = xi Bi ∂t κ ∂ ξ( =1) = |x| cot |x| H ∂t κ ∂ ξ( =−1) = |x| coth |x| H ∂t (3.4.4.83) i κ x ∂ ∂ ∂ ξ( =1) = (1 − |x| cot |x|) t + xj + |x| cot |x| Pi x2 ∂t ∂xj ∂xi   i κ x ∂ ∂ ∂ ξ( =−1) = (1 − |x| coth |x|) t + xj + |x| coth |x| . Pi x2 ∂t ∂xj ∂xi  

3.4.5 Torsional Galilean Spacetimes Unlike the Galilean symmetric spacetimes discussed in section 3.4.4, some Galilean spacetimes do not arise as limits from the (pseudo-)Riemannian spacetimes: namely, the torsional Galilean de Sitter (dSGγ) and anti-de Sitter (AdSGχ) spacetimes, which are the subject of this section. Galilean spacetimes can be seen as null reductions of Lorentzian spacetimes one dimension higher and it would be interesting to exhibit these Galilean spacetimes as null reductions. We hope to return to this question in the future.

58 Torsional Galilean de Sitter Spacetime (dSGγ=1) 6

The additional brackets not involving J for dSGγ are [H, B]=−P and [H, P] = γB +(1 + γ)P, where γ (−1,1). ∈

Fundamental vector fields We start by determining the expressions for the fundamental vector fields ξBi , ξPi , and ξH relative to the exponential coordinates. The boosts are Galilean and hence act in the usual way, with fundamental vector field

∂ ξ = t . (3.4.5.1) Bi ∂xi

To determine the other fundamental vector fields we must work harder. The matrix adA in this basis is given by 0 γ ad = t , (3.4.5.2) A −1 1 + γ   which is diagonalisable (since γ = 1) with eigenvalues 1 and γ, so that ad = S∆S−1, with 6 A t 0 γ 1 ∆ = and S = . (3.4.5.3) 0 tγ 1 1     Therefore if f(z) is analytic, f(t) 0 f(ad )= S S−1, (3.4.5.4) A 0 f(γt)   so that f(γt)− γf(t) f(γt)− f(t) f(ad )B = B + P A 1 − γ 1 − γ (3.4.5.5) γ(f(γt)− f(t)) γf(γt)− f(t) f(ad )P = B + P. A γ − 1 γ − 1

On the other hand, ad H =−γx B −(1 + γ)x P, so if f(z)= 1 + zf(z), then A · · f(ad )H = H − γf(ad )x B −(1 + γ)f(ad )x P e A A · A · γ 1 (3.4.5.6) = H + γf(γt)− f(t) x B + γ2f(γt)− f(t) x P, 1e− γ e· 1 − γ ·     e e e e where f(t)=(f(t)− 1)/t. With these expressions we can now use equation (2.3.1.14) to solve for the fundamental vector fields. e Put X = v P and Y ′(0)= β B in equation (2.3.1.14) to obtain that τ = 0 and · · 1 y P = γ−1 [γ (G(γt)− γG(t)) v B + (γG(γt)− G(t)) v P] · · · (3.4.5.7) − 1 [(F(γt)− γF(t)) β B + (F(γt)− F(t)) β P] . 1−γ · · This requires G(γt)− G(t) β =−γ v, (3.4.5.8) F(γt)− γF(t) and hence, substituting back into the equation for y and simplifying, we obtain

(γ − 1)et y = t −1 + v, (3.4.5.9) eγt − et   so that (γ − 1)et ∂ ξ = t −1 + . (3.4.5.10) Pi eγt − et ∂xi  

59 Finally, let X = H and Y (0)= β B in equation (2.3.1.14) to obtain that τ = 1 and ′ · γ 1 2 y P = 1−γ (γh(γt)− h(t)) x B + 1−γ γ h(γt)− h(t) x P · · · (3.4.5.11) − 1 (F(γt)− γF(t)) β B − 1 (F(γt)− F(t)) β P, 1−γ · 1−γ · where h(t)=(G(t)− 1)/t. This requires

γh(γt)− h(t) β = γ x (3.4.5.12) F(γt)− γ, F(t) so that 1 (1 − γ)et y = 1 + + x. (3.4.5.13) t eγt − et   This means that ∂ 1 (1 − γ)et ∂ ξ = + 1 + + xi . (3.4.5.14) H ∂t t eγt − et ∂xi  

We can easily check that [ξH, ξBi ]= ξPi and [ξH, ξPi ]=−γξBi −(1 + γ)ξPi .

Soldering form and canonical connection This homogeneous spacetime is reductive, so we have not just a soldering form, but also a canonical invariant connection, which can be determined via equation (2.3.3.2):

θ + ω = D(ad )(dtH + dx P) A · = dt(H + γ (γD(γt)− D(t))x B + 1 (γ2D(γt)− D(t))x P (3.4.5.15) 1−γ · 1−γ · + γ (D(γt)− D(t))dx B + 1 (γD(γt)− D(t))dx P, γ−1 e e · γ−1 e e · where now D(z)=(D(z)− 1)/z. Substituting D(z)=(1 − e−z)/z, we find that the soldering form is given by e 1 e−t − e−γt θ = dt H + x P + (dtx − tdx) P, (3.4.5.16) t · t2(1 − γ) ·   from where it follows that θ is invertible for all (t, x). The canonical invariant connection is given by 1 γe−t − e−γt ω = + (dtx − tdx) B. (3.4.5.17) t2 t2(1 − γ) ·   The torsion and curvature of the canonical invariant connection are easily determined from equations (2.3.3.4) and (2.3.3.5), respectively:

1 + γ e−t − e−γt γ e−t − e−γt Θ = dt ∧ dx P and Ω = dt ∧ dx B. (3.4.5.18) 1 − γ t · 1 − γ t ·     This spacetime admits an invariant Galilean structure with clock form τ = η(θ)= dt and spatial ij metric on one-forms h = δ EP EP , where E is the vielbein obtained by inverting the soldering i ⊗ j form: ∂ 1 γ − 1 ∂ t(γ − 1) ∂ E = + − xi and E = . (3.4.5.19) H ∂t t e−t − e−tγ ∂xi Pi e−t − e−γt ∂xi   Therefore, the spatial co-metric of the Galilean structure is given by

t2(γ − 1)2 ∂ ∂ h = δij . (3.4.5.20) (e−t − e−γt)2 ∂xi ⊗ ∂xj

We can simplify this expression using the following coordinate transformation.

−t −γt i e − e i t′ = t and x′ = x . (3.4.5.21) t(γ − 1)

60 In these new coordinates, we have τ = dt′ and

ij ∂ ∂ h = δ i j . (3.4.5.22) ∂x′ ⊗ ∂x′

Torsional Galilean de Sitter Spacetime (dSGγ=1)

This is dSG1, which is the γ 1 limit of the previous example. Some of the expressions in the → previous section have removable singularities at γ = 1, so it seems that treating that case in a separate section leads to a more transparent exposition.

The additional brackets not involving J are now [H, B]=−P and [H, P]= 2P + B. We start by determining the expressions for the fundamental vector fields ξBi , ξPi , and ξH relative to the exponential coordinates (t, x), where σ(t, x)= exp(tH + x P). ·

Fundamental vector fields The bracket [H, B]=−P shows that B acts as a Galilean boost. We can, therefore, immediately write down

∂ ξ = t . (3.4.5.23) Bi ∂xi To find the other fundamental vector fields requires solving equation (2.3.1.14) with A = tH+x P · and Y ′(0)= β B (for this Lie algebra) for X = P and X = H. To apply equation (2.3.1.14) we · i must first determine how to act with f(adA) on the generators, where f(z) is analytic in z.

We start from ad H =−x B − 2x P A · · adA P = 2tP + tB (3.4.5.24)

adA B =−tP.

It follows from the last two expressions that

0 t ad B P = B P , (3.4.5.25) A −t 2t     where the matrix 0 1 M = (3.4.5.26) −1 2   is not diagonalisable, but may be brought to Jordan normal form M = SJS−1, where

1 0 1 −1 J = and S = S−1 = . (3.4.5.27) 1 1 0 −1     It follows that for f(z) analytic in z,

f(adA) B P = B P Sf(tJ)S. (3.4.5.28)

n   If f(z)= n=0 cnz , ∞ n 1 0 f(t) 0 P f(tJ)= cnt = . (3.4.5.29) ∞ n 1 tf′(t) f(t) n=0 X     Performing the matrix multiplication, we arrive at

f(adA)B =(f(t)− tf′(t))B − tf′(t))P (3.4.5.30) f(adA)P = tf′(t)B +(f(t)+ tf′(t))P.

Similarly, f(ad )H = f(0)H − 2x f(ad )P − x f(ad )B, (3.4.5.31) A · A · A e e 61 where f(z)=(f(z)− f(0))/z.

We aree now ready to apply equation (2.3.1.14). Let X = v P. Then equation (2.3.1.14) · becomes

τH + y P = G(adA)v P − F(adA)β B · · · (3.4.5.32) =(G(t)+ tG′(t))v P + tG′(t)v B −(F(t)− tF′(t))β B + tF′(t)β P, · · · · from where we find that τ = 0,

tG (t) F(t)G(t)+ t(F(t)G (t)− F (t)G(t)) β = ′ v and hence y = ′ ′ v =(1−t)v, (3.4.5.33) F(t)− tF′(t) F(t)− tF′(t) so that ∂ ξ =(1 − t) , (3.4.5.34) Pi ∂xi which is indeed the limit γ 1 of equation (3.4.5.10). → Now let X = H, so that equation (2.3.1.14) becomes

τH + y P = G(ad )H − β F(ad )B · A · A = H − 2x G(ad )P − x G(ad )B − β F(ad )B · A · A · A = H −(G(t)+ tG′(t))x B −(F(t)− tF′(t))β B −(2G(t)+ tG′(t))x P + tF′(t)β P, e · e · · · (3.4.5.35) e e e e from where τ = 1,

G(t)+ tG (t) t(F (t)G(t)− F(t)G (t))− 2F(t)G(t) β = ′ x and hence y = ′ ′ x = x. (3.4.5.36) tF′(t)− F(t) F(t)− tF′(t) e e e e e In summary, ∂ ∂ ξ = + xi , (3.4.5.37) H ∂t ∂xi which is indeed the γ 1 limit of equation (3.4.5.14). →

Soldering form and canonical connection To calculate the soldering form and the con- nection one-form for the canonical invariant connection, we apply equation (2.3.3.2):

σ∗ϑ = D(ad )(dtH + dx P) A · = dt H − 2x D(ad )P − x D(ad )B + dx D(ad )P · A · A · A   = dt H −(D(t)+ tD′(t))x B −(2D(t)+ tD′(t))x P +(D(t)+ tD′(t))dx P + tD′(t)dx B. e · e · · ·   (3.4.5.38) e e e e Performing the calculation,

1 − e−t θ = dt H + x P + e−tdx P t · ·   (3.4.5.39) 1 1 − e−t ω = − e−t (x Bdt − tdx B), t t · ·   which are equations (3.4.5.16) and (3.4.5.17) in the limit γ 1. Notice that θ is an isomor- → phism for all (t, x).

The torsion and curvature two-forms for the canonical invariant connection are given by

Θ =−2e−tdt ∧ dx P and Ω =−e−tdt ∧ dx B. (3.4.5.40) · ·

62 The vielbein E has components

∂ 1 − et ∂ ∂ E = + xi and E = et . (3.4.5.41) H ∂t t ∂xi Pa ∂xa The invariant Galilean structure has clock form τ = η(θ)= dt and inverse spatial co-metric

ij 2t ij ∂ ∂ h = δ EP EP = e δ . (3.4.5.42) i ⊗ j ∂xi ⊗ ∂xj This expression can be simplified using the coordinate transformation

i −t i t′ = t and x′ = e x . (3.4.5.43)

In these coordinates, we find τ = dt′ and ∂ ∂ = ij h δ i j . (3.4.5.44) ∂x′ ⊗ ∂x′

Torsional Galilean Anti-de Sitter Spacetime (AdSGχ)

In this instance, the additional non-vanishing brackets are [H, B]=−P and [H, P]=(1 + χ2)B + 2χP.

Fundamental vector fields Since B acts via Galilean boosts we can immediately write down

∂ ξ = t . (3.4.5.45) Bi ∂xi To calculate the other fundamental vector fields we employ equation (2.3.1.14). The adjoint action of A = tH + x P is given by · ad H = −(1 + χ2)x B − 2χx P A · · adA B =−tP (3.4.5.46) 2 adA P = t(1 + χ )B + 2tχP.

In matrix form, 0 (1 + χ2)t ad B P = B P . (3.4.5.47) A −t 2tχ     We notice that this matrix is diagonalisable:

0 (1 + χ2) χ + i χ − i χ − i 0 = S∆S−1, where S := and ∆ := . −1 2χ 1 1 0 χ + i       (3.4.5.48) So if f(z) is analytic in z,

−1 f(adA) B P = B P Sf(t∆)S , (3.4.5.49) or letting t := t(χ i),   ± ± i 1 f(adA)B = 2 (f(t+)− f(t−))(P + χB)+ 2 (f(t+)+ f(t−))B i 2 1 (3.4.5.50) f(adA)P =− 2 (f(t+)− f(t−))(χP +(1 + χ )B)+ 2 (f(t+)+ f(t−))P. Similarly, 1 f(adA)H = f(0)H + (f(adA)− f(0)) adA H adA (3.4.5.51) = f(0)H −(1 + χ2)x f(ad )B − 2χx f(ad )P, · A · A e e 63 where f(z) := (f(z)− f(0))/z. With these formulae we can now use equation (2.3.1.14) to find out the expressions for the fundamental vector fields ξH and ξ . Putting X = v P and bPi · Y ′(0)=eβ B in equation (2.3.1.14) we arrive at · −i(1 + χ2)(G(t )− G(t )) β = + − v (3.4.5.52) F(t+)+ F(t−)+ iχ(F(t+)− F(t−)) and hence ∂ ξ = t(cot t − χ) . (3.4.5.53) Pi ∂xi Similarly, putting X = H and Y (0)= β B in equation (2.3.1.14) we find ′ · iχ(G(t )− G(t ))−(G(t )+ G(t )) β = + − + − x (3.4.5.54) F(t+)+ F(t−)+ iχ(F(t+)− F(t−)) e e e e and hence ∂ ∂ ξ = + 1 + χ − cot t xi . (3.4.5.55) H ∂t t ∂xi 2 We check that [ξH, ξBi ] = ξPi and [ξH, ξPi ] = −(1 + χ )ξBi − 2χξPi , as expected. Another check is that taking χ 0, we recover the fundamental vector fields for Galilean anti-de Sitter → spacetime given by equation (3.4.4.71).

Soldering form and canonical connection Let us now use equation (2.3.3.2) to calculate the soldering form θ and the connection one-form ω for the canonical invariant connection:

θ + ω = D(adA)(dtH + dx P) · (3.4.5.56) = dt H −(1 + χ2)x D(ad )B − 2χx D(ad )P + dx D(ad )P, · A · A · A   where D(z)=(D(z)− 1)/z. Evaluatinge these expressions,e we find

(t − eχt sin t) 1 e θ = dt H + x P + e−χt sin tdx P (3.4.5.57) t2 · t ·   and 1 − e−χt(cos t + χ sin t) ω = (dtx B − tdx B). (3.4.5.58) t2 · · −χt Again, the zeros of e sin t at t = π invalidate the exponential coordinates for t (−π, π). t ± 6∈ The torsion and curvature of the canonical invariant connection are easily calculated to be 2χ Θ =− e−χt sin tdt ∧ dx P t · (3.4.5.59) (1 + χ2) Ω =− e−χt sin tdt ∧ dx B. t · As χ 0, the torsion vanishes and the curvature agrees with that of the Galilean anti-de Sitter → spacetime (S10) in equation (3.4.4.68).

The vielbein E has components

∂ 1 ∂ E = + − eχt csc t xi H ∂t t ∂xi   (3.4.5.60) ∂ E = teχt csc t , Pi ∂xi

64 whose χ 0 limit agrees with equation (3.4.4.69). The invariant Galilean structure has clock → form τ = η(θ)= dt and inverse spatial metric

∂ ∂ h = t2e2χt csc2 tδij , (3.4.5.61) ∂xi ⊗ ∂xj which again agrees with equation (3.4.4.70) in the limit χ 0. These expressions for the → Galilean structure can be simplified by changing coordinates. In particular, let

−χt e i t′ = t and x′i = sin(t)x . (3.4.5.62) t

In our new coordinates, the clock one-form is τ = dt′ and the co-metric becomes

ij ∂ ∂ h = δ i j . (3.4.5.63) ∂x′ ⊗ ∂x′

The Action of the Boosts

In this section, we show that the generic orbits of boosts are not compact in the torsional Galilean spacetimes discussed above. This requires a different argument to the ones we used for the symmetric spaces.

Let M be one of the torsional Galilean spacetimes discussed in this section; that is, dSGγ or AdSGχ, for the relevant ranges of their parameters. The following discussion applies verba- tim to the torsional Galilean (anti-)de Sitter.

Our default description of M is as a simply-connected kinematical homogeneous spacetime K/H, where K is a simply-connected kinematical Lie group and H is the connected subgroup generated by the boots and rotations. Our first observation is that we may dispense with the rotations and also describe M as S/B, where S is the simply-connected solvable Lie group gen- erated by the boosts and spatio-temporal translations and B is the connected abelian subgroup generated by the boosts. This restriction from K/H to S/B can be understood as follows. By definition, we know that we have a transitive K-action on M. Since S is a subgroup of K, we find that we also have a transitive S-action on M. The typical stabiliser subgroup for this new action is not H but B. This statement tells us that M =∼ S/B. By construction, the action of the boosts will be the same on both K/H and S/B, so although we started with the Klein pair (k, h), we may have equally started with (s, b) to get an equivalent geometric realisation of M, where the Lie algebra s of S is spanned by H, Bi,Pi and the Lie algebra b of B is spanned by Bi with non-zero brackets

[H, Bi]=−Pi and [H, Pi]= αBi + βPi , (3.4.5.64) for some real numbers α, β depending on the parameters γ, χ. We may identify s with the Lie subalgebra of gl(2D + 1, R) given by

0 tα1 y s= −t1 tβ1 x (t, x, y) R2D+1 , (3.4.5.65)   ∈  0 0 0     where 1 is the D D identity matrix and b with the Lie subalgebra × 0 0 y b= 0 0 0 y RD . (3.4.5.66)   ∈  0 0 0    

  65 The Lie algebras b s gl(2D + 1, R) are the Lie algebras of the subgroups B S GL(2D + ⊂ ⊂ ⊂ ⊂ 1, R) given by

a(t)1 b(t)1 y 1 0 y S = c(t)1 d(t)1 x (t, x, y) R2D+1 and B = 0 1 0 y RD ,   ∈    ∈  0 0 1 0 0 1         (3.4.5.67) for some functions a(t), b(t), c(t ), d(t), which are given explicitly by 

a(t) b(t) 1 γet − eγt γ (eγt − et) = (3.4.5.68) c(t) d(t) γ − 1 et − eγt γetγ − et     for dSG with γ (−1,1), γ ∈ a(t) b(t) et(1 − t) ett = (3.4.5.69) c(t) d(t) −ett et(1 + t)     for dSG1, and

a(t) b(t) etχ(cos t − χ sin t) etχ(1 + χ2) sin t = (3.4.5.70) c(t) d(t) −etχ sin t etχ(cos t + χ sin t)     for AdSGχ with χ> 0. The homogeneous space M = S/B, if not simply-connected, is neverthe- less a discrete quotient of the simply-connected M and, as argued at the end of Section 3.4.2, it is enough to show that the orbits of boosts in M are generically non-compact to deduce that the same holds for M.

Let us denote by g(t, x, y) S the generic group element ∈ a(t)1 b(t)1 y g(t, x, y)= c(t)1 d(t)1 x S, (3.4.5.71)   ∈ 0 0 1   so that the generic boost is given by

1 0 y g(0,0, y)= 0 1 0 B. (3.4.5.72)   ∈ 0 0 1   Parenthetically, let us remark that while it might be tempting to identify M with the subman- ifold of S consisting of matrices of the form g(t, x,0), this would not be correct. For this to hold true, it would have to be the case that given g(t, x, y), there is some g(0,0, w) such that g(t, x, y)g(0,0, w) = g(t′, x′,0) for some t′, x′. As we now show, this is only ever the case provided that a(t) = 0. Indeed, 6 g(t, x, y)g(0,0, w)= g(t, c(t)w + x, a(t)w + y), (3.4.5.73) and hence this is of the form g(t′, x′,0) if and only if we can solve a(t)w + y = 0 for w. Clearly dSG log γ AdSG this cannot be done if a(t)= 0, which may happen for γ (0,1) at t = γ−1 and for χ>0 at cos t = χ . ∈ ±√1+χ2 The action of the boosts on M is induced by left multiplication on S:

g(0,0, v)g(t, x, y)= g(t, x, y + v) (3.4.5.74) which simply becomes a translation y y + v in RD. This is non-compact in S, but we need 7→ to show that it is non-compact in M.

66 The right action of B is given by

g(t, x, y)g(0,0, w)= g(t, x + c(t)w, y + a(t)w), (3.4.5.75) which is again a translation (x, y) (x + c(t)w, y + a(t)w) in R2D. The quotient R2D/B is 7→ the quotient vector space R2D/B, where B R2D is the image of the linear map RD R2D ⊂ → sending w (c(t)w, a(t)w). Notice that (a(t), c(t)) =(0,0) for all t, since the matrices in S are → 6 invertible, hence B =∼ RD and hence the quotient vector space R2D/B =∼ RD. By the Heine–Borel theorem, it suffices to show that the orbit is unbounded to conclude that it is not compact. Let [(x, y)] R2D/B denote the equivalence class modulo B of (x, y) R2D. The distance d ∈ ∈ between [(x, y)] and the boosted [(x, y + v)] is the minimum of the distance between (x, y) and any point on the coset [(x, y + v)]; that is,

d = min (x + c(t)w, y + v + a(t)w)−(x, y) = min (c(t)w, v + a(t)w) . (3.4.5.76) w k k w k k Completing the square, we find

a 2 c2 (cw, v + aw) 2 =(a2 + c2) w + v + v 2, (3.4.5.77) k k a2 + c2 a2 + c2 k k

a whose minimum occurs when w =− a2+c 2 v, resulting in

|c(t)| d = v . (3.4.5.78) a(t)2 + c(t)2 k k

As we rescale v sv, this is unboundedp provided that c(t) = 0. From equations (3.4.5.68), 7→ 6 (3.4.5.69) and (3.4.5.70), we see that for dSGγ (−1,1], c(t)= 0 if and only if t = 0, whereas for ∈ AdSG 0, c(t) = 0 if and only if t = nπ for n Z, and hence, in summary, the generic orbits χ> ∈ are non compact.

Let us remark that for AdSG 0, if t = nπ for n = 0 then the exponential coordinate sys- χ> 6 tem breaks down, so that we should restrict to t (−π, π). Indeed, using the explicit matrix ∈ representation, one can determine when the exponential coordinates on M stop being injective; that is, when there are (t, x) and (t′, x′) such that exp(tH + x P)= exp(t′H + x′ P)B for some · · B B. In dSGγ (−1,1] this only happens when t = t′ and x = x′, but in AdSGχ>0 it happens ∈ ∈ whenever t = t′ = nπ (n = 0) and, if so, for all x, x′. 6

3.4.6 Aristotelian Spacetimes In this section, we study the Aristotelian spacetimes of Table 3.5. In particular, we derive their fundamental vector fields, vielbeins, soldering forms, and canonical connections.

Static Spacetime (A)

This is an affine space and the exponential coordinates (t, x) are affine, so that

∂ ∂ ξ = and ξ = . (3.4.6.1) H ∂t Pi ∂xi Similarly, the soldering form is θ = dtH + dx P, the canonical invariant connection vanishes, · and so does the torsion. The vielbein is

EH = ξH and EPi = ξPi . (3.4.6.2)

We now have both Galilean (τ, h) and Carrollian (κ, b) structures to define on this spacetime. First, we find that the clock one-form is given by τ = η(θ)= dt and the spatial co-metric of the Galilean structure is written ∂ ∂ h = δij . (3.4.6.3) ∂xi ⊗ ∂xj

67 ∂ The Carrollian structure then consists of κ = EH = ∂t and

b = dx dx. (3.4.6.4) ·

Torsional Static Spacetime (TA)

In this case, the additional non-vanishing brackets are [H, P]= P.

Fundamental vector fields Letting A = tH + x P, we find ad H =−x P and ad P = tP. · A · A Therefore, for any analytic function f, we conclude that 1 f(ad )P = f(t)P and f(ad )H = f(0)H − (f(t)− f(0))x P. (3.4.6.5) A A t · Applying this to equation (2.3.1.14), we find

∂ 1 1 ∂ ξ = + − − 1 xi H ∂t t et − 1 ∂xi   (3.4.6.6) t ∂ ξ = , Pi 1 − e−t ∂xi

which one can check obey [ξH, ξPi ]=−ξPi , as expected.

Soldering form and canonical connection Applying the same formula to equation (2.3.3.2), we find that the canonical invariant connection one-form vanishes in this basis and that the soldering form is given by

1 1 − e−t 1 − e−t θ = dt H + 1 − x P + dx P, (3.4.6.7) t t · t ·     so that the corresponding vielbein is

∂ 1 1 ∂ t ∂ E = + − xi and E = . (3.4.6.8) H ∂t t 1 − e−t ∂xi Pi 1 − e−t ∂xi   1−e−t It is clear from the fact that the function t is never zero that θ is invertible for all (t, x).

Although the canonical connection is flat, its torsion 2-form does not vanish:

e−t − 1 Θ = dt ∧ dx P. (3.4.6.9) t · Choosing to change coordinates such that

−t i 1 − e i t′ = t and x′ = x (3.4.6.10) t and letting 1 1 f(t′)= −t′ − , (3.4.6.11) 1 − e t′ we find that the Galilean structure has clock one-form τ = η(θ)= dt′ and spatial co-metric

ij ∂ ∂ h = δ i j . (3.4.6.12) ∂x′ ⊗ ∂x′ Additionally, the Carrollian structure consists of

∂ i ∂ 2 2 2 κ = − f(t′)x′ i and b = f(t′) x′ dt′ + 2f(t′)x′ dx′dt′ + dx′ dx′. (3.4.6.13) ∂t′ ∂x′ · ·

68 Aristotelian Spacetime A23ε

In this instance, the additional non-vanishing brackets are [Pi,Pj]=−εJij.

Fundamental vector fields Let A = tH + x P. Then ad H = 0 and ad P = εxjJ . · A A i ij Continuing, we find

ad2 P = εxix P − εx2P and ad3 P = (−εx2) ad P . (3.4.6.14) A i · i A i A i Therefore, an induction argument shows that

adn P = (−εx2) adn−2 P n > 3. (3.4.6.15) A i A i ∀

If f(z) is analytic in z, then f(adA)H = f(0)H and

i 1 1 x x P ε j f(adA)Pi = 2 (f(x+)+ f(x−)) Pi − 2 (f(x+)+ f(x−)− 2f(0)) 2· − (f(x+)+ f(x−)) x Jij, x 2x+ (3.4.6.16) where |x| ε =−1 x = −εx2 = ± (3.4.6.17) ± ±  i|x| ε = 1. p ± i j Similarly, adA Jij = x Pj − x Pi, so that

1 i j ε k i j f(adA)Jij = f(0)Jij + 2 f(x+)+ f(x−) (x Pj − x Pi)− f(x+)− f(x−) x (x Jkj − x Jki), 2x+     (3.4.6.18) e e e e where f(z)=(f(z)− f(0))/z.

Insertinge these formulae in equation (2.3.1.14) with X = H and Y ′(0)= 0, we see that

∂ ξ = . (3.4.6.19) H ∂t If instead X = v P and Y (0)= 1 λijJ , we see first of all that τ = 0 and that demanding that · ′ 2 ij the Jij terms cancel, −ε (G(x )− G(x )) λij = + − (xivj − xjvi), (3.4.6.20) x+ (F(x+)+ F(x−)) and reinserting into equation (2.3.1.14), we find that

(G(x )− G(x ))(F(x )− F(x )) yi = 1 G(x )+ G(x )− + − + − vi 2 + − F(x )+ F(x )  + −  (G(x+)− G(x−))(F(x+)− F(x−)) v x − 1 G(x )+ G(x )− 2 − · xi. (3.4.6.21) 2 + − F(x )+ F(x ) x2  + − 

From this we read off the expression for ξPi :

F(x )G(x )+ F(x )G(x ) ∂ F(x )G(x )+ F(x )G(x ) xixj ∂ ξ = + − − + + 1 − + − − + , (3.4.6.22) Pi F(x )+ F(x ) ∂xi F(x )+ F(x ) x2 ∂xj + −  + −  which simplifies to

∂ xixj ∂ ξ(ε=1) = |x| cot |x| +(1 − |x| cot |x|) Pi i 2 j ∂x x ∂x (3.4.6.23) ∂ xixj ∂ ξ(ε=−1) = |x| coth |x| +(1 − |x| coth |x|) . Pi ∂xi x2 ∂xj

69 Soldering form and canonical connection The soldering form and connection one-form for the canonical connection are obtained from equation (2.3.3.2), which says that

i θ + ω = dtH + dx D(adA)Pi 1 = dtH + (D(x+)+ D(x−))dx P 2 · (3.4.6.24) 1 x dx ε i j − 2 (D(x+)D(x−)− 2) · 2 x P − (D(x+)− D(x−))x dx Jij, x · 2x+ such that sin |x| sin |x| x dx θ(ε=1) = dtH + dx P + 1 − · x P |x| · |x| x2 ·   (3.4.6.25) sinh |x| sinh |x| x dx θ(ε=−1) = dtH + dx P + 1 − · x P |x| · |x| x2 ·   and

(ε=1) 1 − cos |x| i j ω = 2 x dx Jij x (3.4.6.26) 1 − cosh |x| ω(ε=−1) = xidxjJ . x2 ij It follows that if ε =−1 the soldering form is invertible for all (t, x), whereas if ε = 1 then it is invertible for all t but inside the open ball |x| < π. With these caveats in mind, we find that the vielbeins are

i κ ∂ κ x ∂ ∂ E( =1) = and E( =1) = (1 − |x| csc |x|) xj + |x| csc |x| , H ∂t Pi x2 ∂xj ∂xi i (3.4.6.27) κ ∂ κ x ∂ ∂ E( =−1) = and E( =−1) = (1 − |x| csch |x|) xj + |x| csch |x| . H ∂t Pi x2 ∂xj ∂xi

The torsion of the canonical connection vanishes, since [θ, θ]m = 0. The curvature is given by

sin2 |x| sin |x| sin |x| xjxk Ω(ε=1) = 1 dxi ∧ dxjJ + 1 − dxi ∧ dxkJ 2 x2 ij |x| |x| x2 ij   (3.4.6.28) sinh2 |x| sinh |x| sinh |x| xjxk Ω(ε=−1) =− 1 dxi ∧ dxjJ − 1 − dxi ∧ dxkJ . 2 x2 ij |x| |x| x2 ij   The Galilean structure in this instance has clock one-form τ = dt and spatial co-metric

∂ ∂ ∂ ∂ (κ=1) | | | | 2 ij | | | | 2 i j h =( x csc x ) δ i j +(1 −( x csc x ) )x x i j ∂x ⊗ ∂x ∂x ⊗ ∂x (3.4.6.29) ∂ ∂ ∂ ∂ h(κ=−1) =(|x| csch |x|)2δij +(1 −(|x| csch |x|)2)xixj . ∂xi ⊗ ∂xj ∂xi ⊗ ∂xj

∂ The Carrollian structure is then κ = ∂t and

2 2 2 (κ=1) sin |x| sin |x| (x dx) b = dx dx + 1 − · 2 |x| · |x| ! |x|     (3.4.6.30) 2 2 2 κ sinh |x| sinh |x| (x dx) b( =−1) = dx dx + 1 − · |x| · |x| |x|2     !

3.4.7 Carrollian Light Cone (LC)

The Carrollian light cone LC is a hypersurface in Minkowski spacetime, identifiable with the future light cone. It does not arise as a limit and has additional brackets [H, B]= B, [H, P]=−P and [B, P]= H + J, which shows that it is a non-reductive homogeneous spacetime.

70 Action of the Boosts

Although it might be tempting to use that the boosts in Minkowski spacetime act with generic non-compact orbits to deduce the same about the boosts in LC, one has to be careful because what we call boosts in LC might not be interpretable as boosts in the ambient Minkowski spacetime. Indeed, as we will now see, boosts in LC are actually null rotations in the ambient Minkowski spacetime.

We first exhibit the isomorphism between the LC Lie algebra and so(D + 1,1). In the LC Lie algebra, the boosts and translations obey the following brackets:

[H, B]= B, [H, P]=−P, and [B, P]= H + J. (3.4.7.1)

If we let JMN be the standard generators of so(D + 1,1) with M =(i, 0,˜ 1˜), 1 6 i 6 D, and with Lie brackets [JMN, JPQ]= ηNPJMQ − ηMPJNQ − ηNQJMP + ηMQJNP, (3.4.7.2) where ηij = δij, η0˜0˜ =−1, and η1˜1˜ = 1, then the correspondence is:

= J = 1 (J + J ) = 1 (J − J ) =−J Jij ij, Bi √2 0˜i i1˜ ,Pi √2 0˜i i1˜ , and H 0˜1˜. (3.4.7.3)

We see that, as advertised, the boosts Bi are indeed null rotations.

The boosts act linearly on the ambient coordinates XM in Minkowski spacetime, with fun- damental vector fields

1 0˜ ∂ i ∂ i ∂ 1˜ ∂ ζB = −X − X + X − X . (3.4.7.4) i √ ∂Xi ∂X0˜ ∂X1˜ ∂Xi 2   i 0˜ i i 1˜ Consider a linear combination B = w Bi and let T := X , X := w X , and Y := X , so that in 1 terms of these coordinates and dropping the factor of √2 ,

∂ ∂ ∂ ∂ ζ =−T − X + X − Y . (3.4.7.5) B ∂X ∂T ∂Y ∂X This allows us to examine the orbit of this vector field while focussing on the three-dimensional space with coordinates T, X, Y. The vector field is linear, so there is a matrix A such that

∂ ∂T 0 −1 0 ∂ ζB = T X Y A = A = −1 0 1 . (3.4.7.6)  ∂X  ⇒   ∂ 0 −1 0  ∂Y     The matrix A obeys A3 = 0, so its exponential is

1 2 1 2 1 + 2 s −s − 2 s exp(sA)= −s 1 s (3.4.7.7)  1 2 1 2 2 s −s 1 − 2 s   and hence the orbit of (T0, X0, Y0,... ) is given by

1 2 1 2 T(s)=(1 + 2 s )T0 − sX0 − 2 s Y0 X(s)=−sT0 + X0 + sY0 (3.4.7.8) 1 2 1 2 Y(s)= 2 s T0 − sX0 +(1 − 2 s )Y0, with all other coordinates inert, which is clearly non-compact in the Minkowski spacetime. But of course, this orbit lies on the future light cone (indeed, notice that −T(s)2 + X(s)2 + Y(s)2 = 2 2 2 −T0 +X0 +Y0 ), which is a submanifold, and hence the orbit is also non-compact on LC, provided with the subspace topology.

71 Fundamental Vector Fields

Let A = tH + x P and let us calculate the action of ad on the generators, this time with the · A indices written explicitly:

i j adA Bi = tBi − x H − x Jij

adA Pi =−tPi i (3.4.7.9) adA H = x Pi i j adA Jij = x Pj − x Pi.

In order to compute the fundamental vector fields using equation (2.3.1.14) and the soldering form using equation (2.3.3.2), we need to calculate the action of certain universal power series on adA on the generators. To this end, let us derive formulae for the action of f(adA), for f(z) an analytic function of z, on the generators. We will do this by first calculating powers of adA on generators. It is clear, first of all, that on P,

f(adA)P = f(−t)P. (3.4.7.10)

On H and J we just need to treat the constant term separately: 1 f(adA)H = f(0)H − (f(−t)− f(0)) x P t · (3.4.7.11) 1 f(ad )J = f(0)J − (f(−t)− f(0)) (xiP − xjP ). A ij ij t j i On B it is a little bit more complicated. Notice first of all that whereas

2 i j 2 adA Bi = t adA Bi − 2x x Pj + x Pi, (3.4.7.12)

3 2 adA Bi = t adA Bi. Therefore, by induction, for all n > 1,

tn−1 ad B n odd n = A i adA Bi n−1 n−2 2 i (3.4.7.13) t adA Bi + t (x Pi − 2x x P) n even, · and therefore

1 i j 1 1 2 i f(ad )B = f(t)B − (f(t)− f(0))(x H + x J )+ 2 ( (f(t)+ f(−t))− f(0))(x P − 2x x P). A i i t ij t 2 i · (3.4.7.14) Using these formulae, we can now apply equation (2.3.1.14) in order to determine the expression of the fundamental vector fields in terms of exponential coordinates.

Let us take X = v P in equation (2.3.1.14). We must take Y ′(0)= 0 here and find that · t y P = G(ad )v P = G(−t)v P = y = v, (3.4.7.15) · A · · ⇒ 1 − e−t resulting in t ∂ ξ = . (3.4.7.16) Pi 1 − e−t ∂xi

Taking X = H in equation (2.3.1.14), we again must take Y ′(0)= 0. Doing so, we arrive at

1 1 1 τH + y P = G(ad )H = H − (G(−t)− 1) x P = τ = 1 and y = − 1 − x, · A t · ⇒ t et − 1  (3.4.7.17) resulting in ∂ 1 1 ∂ ξ = + − 1 − xi . (3.4.7.18) H ∂t t et − 1 ∂xi  

One checks already that [ξH, ξPi ]= ξPi , as expected.

72 Finally, put X = v B in equation (2.3.1.14) and hence now Y (0) = β B + 1 λijJ . Sub- · ′ · 2 ij stituting this in equation (2.3.1.14) and requiring that the h-terms vanish, we find

G(t) 1 − e−t β = v = e−tv and λij = (vixj − vjxi). (3.4.7.19) F(t) t

Comparing the H terms, we see that

1 − e−t τ = x v, (3.4.7.20) t · whereas the P terms give

1 − e−t 1 − t − e−t y = x2v + x vx, (3.4.7.21) 2t t2 · resulting in 1 − e−t ∂ 1 − e−t ∂ 1 − t − e−t ∂ ξ = xi + x2 + xixj . (3.4.7.22) Bi t ∂t 2t ∂xi t2 ∂xj

One checks that, as expected, [ξH, ξBi ]=−ξBi and that [ξBi , ξPj ]=−δijξH − ξJij , where j ∂ i ∂ ξJij = x ∂xi − x ∂xj .

Soldering Form and Canonical Connection

The soldering form can be calculated from equation (2.3.3.2) and projecting the result to k/h:

D(−t)− 1 θ = D(ad )(dtH + dx P)= dt H − x P + D(−t)dx P A · t · ·   (3.4.7.23) 1 + t − et et − 1 = dtH + x Pdt + dx P. t2 · t ·

t It follows from the expression of θ that it is invertible for all (t, x), since e −1 = 0 for all t R. t 6 ∈ Its inverse, the vielbein E, has components

∂ 1 1 ∂ t ∂ E = + − xa and E = . (3.4.7.24) H ∂t t et − 1 ∂xa Pa et − 1 ∂xa   2 The invariant Carrollian structure is given by κ = EH and spatial metric b = π (θ, θ), given by

(1 + t − et)2 (et − 1)2 (et − 1)(1 + t − et) b = x2dt2 + dx dx + 2 x dxdt. (3.4.7.25) t4 t2 · t3 · This metric can be simplified using the following change of coordinates:

t i e − 1 i t′ = t x′ = x . (3.4.7.26) t In these coordinates, we find

2 b = dx′ dx′ − 2x′ dx′dt′ + x′ x′dt′ . (3.4.7.27) · · · One final change of coordinates,

i −t′ i ˆt = t′ andx ˆ = e x′ , (3.4.7.28) brings the metric into the form b = e2ˆtdxˆ dxˆ. (3.4.7.29) · With all of these changes to the coordinate system, the vielbein is also altered such that

∂ κ = . (3.4.7.30) ∂ˆt

73 3.5 Conclusion

This chapter discussed how we might arrive at geometric properties for various spacetime models from the classification of their underlying kinematical Lie algebras. In Section 3.1, we reviewed the classification of kinematical Lie algebras in spatial dimension D = 3 due to Figueroa-O’Farrill, before discussing the subsequent classification of (spatially-isotropic) simply- connected homogeneous spacetimes in Section 3.2. We gave a more detailed review of the latter classification as direct generalisations of the methods utilised here will be at the heart of the classification of kinematical superspaces in Chapter 4. Finally, in Section 3.4, we derived nu- merous geometric properties for each spacetime model by employing our knowledge of the underlying Lie algebra and its use in constructing the spacetime geometry. This procedure of algebraic classification, geometric classification, and geometric property derivation, gives us a rigorous framework in which to explore spacetime symmetries beyond the Lorentzian case. Indeed, Chapters 4 and 5 describe the progress towards substantiating this framework in the super-kinematical and super-Bargmann instances, respectively. Since both of these instances are generalisations of the kinematical case, we will see many of the foundational results pre- sented here, including those regarding geometric realisability and geometric limits, will prove invaluable in deriving similar results in these supersymmetric cases.

74 Chapter 4

Kinematical Superspaces

In the previous chapter, we saw how we might arrive at a systematic study of kinematical spacetimes and their geometric properties, starting from the classification of the spacetimes’ underlying Lie algebras. This chapter will highlight the progress made towards this goal in the super-kinematical case. In particular, we will give complete classifications of kinematical Lie superalgebras and their corresponding kinematical superspaces.

Recall, an N-extended kinematical Lie superalgebra (KLSA) s in three spatial dimensions is a real Lie superalgebra s=s¯ s¯, such that s¯ =k is a kinematical Lie algebra for which D = 3, 0 ⊕ 1 0 and s1¯ consists of N copies of S, the real four-dimensional spinor module of the rotational sub- algebra r =∼ so(3). Here, we will focus solely on the N = 1 case.

This chapter is organised as follows. In Section 4.1, we present the classification of N = 1 kinematical Lie superalgebras in D = 3. As part of this classification, we will demonstrate how to unpack our quaternionic formalism, using the N = 1 Poincaré superalgebra as our example. Additionally, Section 4.1 will contain discussions on the N = 1 Aristotelian Lie superalgebras in D = 3, as well as the central extensions and automorphisms of the classified kinematical and Aristotelian Lie superalgebras. In Section 4.2, we use the Lie superalgebras’ automorphisms to classify the possible super Lie pairs, and, thus, the possible kinematical superspaces. Finally, in Section 4.3, we demonstrate how the kinematical superspaces are related to one another via geometric limits.

4.1 Classification of Kinematical Superalgebras

In this section, we begin by setting up our classification problem. In particular, Section 4.1.1 starts by defining which kinematical Lie algebras we shall be extending and discussing how we will derive the supersymmetric brackets. Next, it gives some preliminary results, which will be useful for limiting repetition in our calculations, and describes the Lie superalgebras’ basis transformations, allowing us to identify isomorphic kinematical Lie superalgebras. With the setup established, we classify the kinematical Lie superalgebras in Section 4.1.2, summarising our findings and unpacking the quaternionic formalism in Section 4.1.3. Sections 4.1.4 and 4.1.5 classify the Aristotelian Lie superalgebras and the central extensions of the kinematical Lie superalgebras, respectively. Section 4.1.6 then determines the automorphisms of the kinematical Lie superalgebras, which we will use to classify the kinematical superspaces later.

4.1.1 Setup for the Classification For this chapter, we will combine the classifications in Tables 3.1 and 3.3, so we deal with all D = 3 kinematical Lie algebras at once. These algebras are summarised in Table 4.1. We will now outline how we aim to build N = 1 supersymmetric extensions of these kinematical Lie algebras.

75 Table 4.1: Kinematical Lie Algebras in D = 3

K# Non-zero Lie brackets (besides [J, −] brackets) Comment 1 static 2 [H, B]=−P Galilean

3γ [−1,1] [H, B]= γB [H, P]= P ∈ 4χ>0 [H, B]= χB + P [H, P]= χP − B 5 [H, B]= B + P [H, P]= P 6 [B, P]= H Carroll 7 [H, B]= P [B, P]= H [B, B]= J Euclidean 8 [H, B]=−P [B, P]= H [B, B]=−J Poincaré 9 [H, B]= B [H, P]=−P [B, P]= H − J so(4,1) 10 [H, B]= P [H, P]=−B [B, P]= H [B, B]= J [P, P]= J so(5) 11 [H, B]=−P [H, P]= B [B, P]= H [B, B]=−J [P, P]=−J so(3,2) 12 [B, B]= B [P, P]= B − J 13 [B, B]= B [P, P]= J − B 14 [B, B]= B 15 [B, B]= P 16 [H, P]= P [B, B]= B 17 [H, B]=−P [B, B]= P 18 [H, B]= B [H, P]= 2P [B, B]= P

Let s be a kinematical Lie superalgebra where s0¯ = k is a kinematical Lie algebra from Ta- ble 4.1. To determine s, we need to specify the additional Lie brackets: [H, Q], [B, Q], [P, Q] and [Q, Q], subject to the super-Jacobi identity. There are four components to the super-Jacobi identity in a Lie superalgebra s=s¯ s¯: 0 ⊕ 1

1. The (s0¯, s0¯, s0¯) super-Jacobi identity simply says that s0¯ is a Lie algebra, which in our case is one of the kinematical Lie algebras k in Table 4.1.

2. The (s0¯, s0¯, s1¯) super-Jacobi identity says that s1¯ is a representation of s0¯ and, by restric- tion, also a representation of any Lie subalgebra of s0¯: for example, r in our case. 2 3. The (s¯, s¯, s¯) super-Jacobi identity says that the component of the Lie bracket s¯ 0 1 1 1 → s¯ is s¯-equivariant. In particular, in our case, it is r-equivariant. 0 0 J

4. The (s1¯, s1¯, s1¯) component does not seem to have any representation-theoretic reformula- tion and needs to be checked explicitly.

Our strategy will be the following. We shall first determine the space of r-equivariant brackets [H, Q], [B, Q], [P, Q] and [Q, Q], which will turn out to be a 22-dimensional real vector space V . For each kinematical Lie algebra k=s0¯ in Table 4.1, we then determine the algebraic variety J V cut out by the super-Jacobi identity. We are eventually interested in supersymme- ⊂ try algebras and hence we will restrict attention to Lie superalgebras s for which [Q, Q] = 0, 6 which define a sub-variety S J .1 The isomorphism classes of kinematical Lie superalgebras ⊂ (with [Q, Q] = 0) are in one-to-one correspondence with the orbits of S under the subgroup 6 G GL(s¯) GL(s¯) which acts by automorphisms of k=s¯, since we have fixed k from the ⊂ 0 × 1 0 start. The group G contains not just the automorphisms of the kinematical Lie algebra k which act trivially on r, but also transformations which are induced by automorphisms of the quater- nion algebra. We shall return to an explicit description of such transformations below.2

1Note, we restrict ourselves to the cases where [Q, Q] = 0 as our interests lie in spacetime supersymmetry: 6 we would like supersymmetry transformations to generate geometric transformations of the spacetime. 2Notice, this strategy is a direct generalisation of the one outlined in Section 3.1.1, used to classify the Aristotelian Lie algebras.

76 Let us start by determining the r-equivariant brackets: [H, Q], [B, Q], [P, Q] and [Q, Q]. The bracket [H, Q] is an r-equivariant endomorphism of the spinor module Q. If we identify r with the imaginary quaternions and Q with the quaternions, the action of r on Q is via left quaternion multiplication. The endomorphisms of the representation S, which commute with the action of r, consist of left multiplication by reals and right multiplication by quaternions, but for real numbers, left and right multiplications agree, since the reals are central in the quaternion algebra. Hence the most general r-equivariant [H, Q] bracket takes the form3

[H, Q(s)] = Q(sh) for some h = h1i + h2j + h3k + h4 H. (4.1.1.1) ∈ The brackets [B, Q] and [P, Q] are r-equivariant homomorphisms V S S, where V and S ⊗ → are the vector and spinor modules of so(3). There is an r-equivariant map V S S given by ⊗ → the “Clifford action”, which in this language is left multiplication by Im H on H. Its kernel is 3 the 8-dimensional real representation W of r with spin 2 . Therefore, the space of r-equivariant homomorphisms V S S is isomorphic to the space of r-equivariant endomorphisms of S, ⊗ → which, as we saw before, is a copy of the quaternions. In summary, the [B, Q] and [P, Q] brackets take the form

[B(β), Q(s)] = Q(βsb) for some b = b1i + b2j + b3k + b4 H ∈ (4.1.1.2) [P(π), Q(s)] = Q(πsp) for some p = p1i + p2j + p3k + p4 H, ∈ for all β, π Im H and s H. ∈ ∈ Finally, we look at the [Q, Q] bracket, which is an r-equivariant linear map 2 S k= R 3V. 2 → ⊕ The symmetric square S is a 10-dimensional r-module which decomposes as R 3V. Indeed, J ⊕ on S, we have an r-invariant inner product given by J

s1, s2 = Re(s1s2) where s1, s2 H. (4.1.1.3) h i ∈

It is clearly invariant under left multiplication by unit quaternions: us1, us2 = s1, s2 for all 2 h i h i u Sp(1). We can use this inner product to identify S with the symmetric endomorphisms ∈ of S: linear maps λ : S S such that λ(s1), s2 = s1, λ(s2) . Letting Lq and Rq denote left → h i hJ i and right quaternion multiplication by q H, the space of symmetric endomorphisms of S is ∈ spanned by the identity endomorphism and LiRi, LiRj, LiRk, LjRi, LjRj, LjRk, LkRi, LkRj and LkRk. The nine non-identity symmetric endomorphisms transform under r according to three copies of V. Since r acts on S via left multiplication, it commutes with the Rq and hence the three copies of V are

span LiRi, LjRi, LkRi span LiRj, LjRj, LkRj span LiRk, LjRk, LkRk . (4.1.1.4) R ⊕ R ⊕ R   2  The space of r-equivariant linear maps S 3V R is thus isomorphic to the space of → ⊕ r-equivariant endomorphisms of R 3V = R (R3 V), which is given by ⊕ J ⊕ ⊗ 3 ∼ 3 Endr R (R V) = End(R) End(R ) 1 . (4.1.1.5) ⊕ ⊗ ⊕ ⊗ V The second component of this isomorphism simply states that the r-equivariant endomorphisms do not act on the so(3) vector indices of the vector modules and rotate the three vector modules into one another. In particular, since r acts via left quaternion multiplication, the r-equivariant maps act via right quaternion multiplication. In summary, the r-equivariant [Q, Q] bracket is given by polarisation from the following

2 [Q(s), Q(s)] = c0|s| H + Re(sJsc1)+ Re(sBsc2)+ Re(sPsc3), (4.1.1.6)

3See Section 2.1.5 for a discussion on the quaternionic formalism employed here. In this chapter, we will use s as opposed to θ to denote the quaternion parameterising our supercharges Q. There is nothing sinister in this change of notation; only s was required for different purposes in Chapter 5, and I latterly preferred using θ as opposed to s for parameterising the supercharges, hence its use in setting up the quaternionic formalism in Sections 2.1.5 and 2.1.6.

77 where c0 R, c1, c2, c3 Im H and where we have introduced the shorthands ∈ ∈

J = J1i + J2j + J3k, B = B1i + B2j + B3k, and P = P1i + P2j + P3k. (4.1.1.7)

Notice that if ω Im H, then J(ω)= Re(ω¯J), and similarly B(β)= Re(β¯B) and P(π)= Re(π¯P), ∈ for β, π Im H, so that we can rewrite the [Q, Q] bracket as ∈ 2 [Q(s), Q(s)] = c0|s| H − J(sc1s)− B(sc2s)− P(sc3s), (4.1.1.8) which polarises to give

Q Q H 1 J 1 B 1 P [ (s), (s′)] = c0 Re(ss′) − 2 (s′c1s + sc1s′)− 2 (s′c2s + sc2s′)− 2 (s′c3s + sc3s′). (4.1.1.9) In summary, we have that the r-equivariant brackets by which we extend the kinematical Lie algebra k live in a 22-dimensional real vector space of parameters h, b, p H, c1, c2, c3 Im H ∈ ∈ and c0 R. ∈

Preliminary Results In this brief section, we will go through each of the super-Jacobi identity components, deter- mining any possible universal conditions that may aid our classification. Note, since we are using a kinematical Lie algebra k as s0¯, we do not need to consider the (s0¯, s0¯, s0¯) component as this will automatically be satisfied.

(s0¯, s0¯, s1¯)

As mentioned above, this component of the super-Jacobi identity says that s1¯ is an s0¯-module, where s0¯ =k is the underlying kinematical Lie algebra. The super-Jacobi identity

[X, [Y, Q(s)]]−[Y, [X, Q(s)]] = [[X, Y], Q(s)] for all X, Y k (4.1.1.10) ∈ gives relations between the parameters h, b, p H appearing in the Lie brackets. ∈ Lemma 4.1.1. The following relations between h, b, p H are implied by the corresponding ∈ k-brackets: [H, B]= λB + µP = [b, h]= λb + µp ⇒ [H, P]= λB + µP = [p, h]= λb + µp ⇒ [B, B]= λB + µP + νJ = b2 = 1 λb + 1 µp + 1 ν (4.1.1.11) ⇒ 2 2 4 [P, P]= λB + µP + νJ = p2 = 1 λb + 1 µp + 1 ν ⇒ 2 2 4 [B, P]= λH = bp + pb = 0 and [b, p]= λh. ⇒ Proof. The [H, B, Q] super-Jacobi identity says for all β Im H and s H, ∈ ∈ [[H, B(β)], Q(s)]=[H, [B(β), Q(s)]]−[B(β), [H, Q(s)]], (4.1.1.12) which becomes λQ(βsb)+ µQ(βsp)= Q(βsbh)− Q(βshb). (4.1.1.13) Since Q is real linear and injective, it follows that

λβsb + µβsp = βs[b, h], (4.1.1.14) which, since it must hold for all β Im H and s H, becomes ∈ ∈ [b, h]= λb + µp, (4.1.1.15) as desired. Similarly, the [H, P, Q] super-Jacobi identity gives the second equation in the lemma. The third equation follows from the [B, B, Q] super-Jacobi identity, which says that for all

78 β, β Im H and s H, ′ ∈ ∈

[[B(β), B(β′)], Q(s)]=[B(β), [B(β′), Q(s)]]−[B(β′), [B(β), Q(s)]], (4.1.1.16) which becomes

1 Q 1 Q 1 Q Q 2 Q 2 2 λ ([β, β′]sb)+ 2 µ ([β, β′]sp)+ 4 ν ([β, β′]s)= (ββ′sb )− (β′βsb ). (4.1.1.17) Again by linearity and injectivity of Q, this is equivalent to

1 1 1 2 2 λ[β, β′]sb + 2 µ[β, β′]sp + 4 ν[β, β′]s =[β, β′]sb , (4.1.1.18) which, being true for all β, β Im H and s H, gives ′ ∈ ∈ 1 1 1 2 2 λb + 2 µp + 4 ν = b , (4.1.1.19) as desired. The fourth identity in the lemma follows similarly from the [P, P, Q] super-Jacobi identity. Finally, we consider the [B, P, Q] super-Jacobi identity, which says that for all β, π ∈ Im H and s H, ∈ [[B(β), P(π)], Q(s)]=[B(β), [P(π), Q(s)]]−[P(π), [B(β), Q(s)]], (4.1.1.20) which expands to − λ Re(βπ)Q(sh)= Q(βπspb)− Q(πβsbp) (4.1.1.21) or, equivalently, − λ Re(βπ)sh = βπspb − πβsbp, (4.1.1.22) for all β, π Im H and s H. For any two imaginary quaternions β, π, we have that ∈ ∈ 1 βπ = 2 [β, π]+ Re(βπ), (4.1.1.23) which allows us to rewrite equation (4.1.1.22) as

1 Re(βπ)s(λh − bp + pb)+ 2 [β, π]s(pb + bp)= 0. (4.1.1.24)

Taking β = π and s = 1 we see that λh =[b, p] and taking β and π to be orthogonal and s = 1, that pb + bp = 0, as desired.

(s0¯, s1¯, s1¯)

Due to the large number of parameters in the [s0¯, s1¯] and [s1¯, s1¯] brackets, the components [H, Q, Q], [B, Q, Q] and [P, Q, Q] of the super-Jacobi identity are best studied on a case-by-case basis.

(s1¯, s1¯, s1¯)

The last component of the super-Jacobi identity to consider is the (s1¯, s1¯, s1¯) case [Q, Q, Q], which gives the following universal condition.

Lemma 4.1.2. The [Q, Q, Q] component of the super-Jacobi identity implies

1 c0h = 2 c1 + c2b + c3p. (4.1.1.25)

Proof. The [Q, Q, Q] component of the super-Jacobi identity is totally symmetric and hence, by polarisation, it is uniquely determined by its value on the diagonal. In other words, it is equivalent to ! [[Q(s), Q(s)], Q(s)] = 0 for all s H. (4.1.1.26) ∈

79 Using equation (4.1.1.8), this becomes

2 ! [c0|s| H − J(sc1s)− B(sc2s)− P(sc3s), Q(s)] = 0, (4.1.1.27) which expands to

2 1 ! c0|s| Q(sh)− 2 Q(sc1ss)− Q(sc2ssb)− Q(sc3ssp) = 0. (4.1.1.28) Since Q is injective, this becomes

2 1 ! |s| s(c0h − 2 c1 − c2b − c3p) = 0.

This must hold for all s H, so in particular for s = 1, proving the lemma. ∈

Basis Transformations

As mentioned above, once we determine the sub-variety S cut out by the super-Jacobi iden- tity, we need to quotient by the action of the subgroup G GL(s¯) GL(s¯), which acts by ⊂ 0 × 1 automorphisms of s0¯ =k in order to arrive at the isomorphism classes of Lie superalgebras. In this section, we describe the subgroup G in more detail. There are two kinds of elements of G, those which act trivially on the rotational subalgebra r and those which do not. The latter consist of inner automorphisms of k, which are generated infinitesimally by the adjoint action of J, B and P. The ones generated by J are particularly easy to describe in the quaternionic formulation, and we shall do so now in more detail.

Let u Sp(1) be a unit norm quaternion. Conjugation by u defines a homomorphism Ad : ∈ Sp(1) Aut(H) whose kernel is the central subgroup of Sp(1) consisting of 1. It is a classical → ± result that these are all the automorphisms of H. Hence Aut(H) =∼ SO(3), acting trivially on the real quaternions and rotating the imaginary quaternions. The action of Aut(H) on s leaves H invariant and acts on the remaining generators by pre-composing the linear maps J, B, P and Q with Adu. More precisely, let H = H, J = J Adu, B = B Adu, P = P Adu and Q = Q Adu. ◦ ◦ ◦ ◦ Since the Lie brackets of k are given in terms of quaternion multiplication, this transformation is an automorphism of k, and wee have ae group homomorphisme Aute (H) Aut(k).e The action → on the remaining brackets (those involving Q) is as follows. The Lie brackets of s which involve Q are given by

[H, Q(s)] = Q(sh) 1 [J(ω), Q(s)] = 2 Q(ωs) [B(β), Q(s)] = Q(βsb) (4.1.1.29) [P(π), Q(s)] = Q(πsp) 2 [Q(s), Q(s)] = c0|s| H − J(sc1s)− B(sc2s)− P(sc3s), and hence under conjugation by u Sp(1), ∈ [H, Q(s)] = Q(sh) [J(ω), Q(s)] = 1 Q(ωs) e e e2 e [B(β), Q(s)] = Q(βsb) (4.1.1.30) e e e [P(π), Q(s)] = Q(πsp) e e e e 2 [Q(s), Q(s)] = c0|s| H − J(sc1s)− B(sc2s)− P(sc3s), e e e e where h = uhu¯ , b = ubu¯e , pe= upu¯ , andeci =e uc¯e iu forei =e 1,2,3.e e In other words, the scalar parameters c0, Re h, Re b and Re p remain inert, but the imaginary quaternion parameters Im h, Ime b, Im p, ce1,2,3 are simultaneouslye rotated.e

There are other automorphisms of k which do transform r: those are the inner automorphisms

80 generated by B and P. Their description depends on the precise form of k but they will not play a rôle in our discussion.

In addition to these, G also consists of automorphisms of k which leave r intact. If a linear map Φ : s s restricts to an automorphism of k, then it is in particular r-equivariant. The → most general r-equivariant linear map Φ :s s sends (J, H, B, P, Q) (J, H, B, P, Q), where → 7→ H = µH e e e e B(β)= aB(β)+ cP(β)+ eJ(β) e (4.1.1.31) B(β)= bB(β)+ dP(β)+ fJ(β) e Q(s)= Q(sq) e e 0 a b where µ GL(1, R)= R , q GL(1, H)= H and 0 c d GL(3, R). The automorphisms ∈ × ∈ ×   ∈ 1 e f (which fix r) of the kinematical Lie algebras K1-K11 in Table 4.1 were derived in [5, §§3.1]. The automorphisms of the remaining kinematical Lie algebras in the table are listed below (see Table 4.2). In particular, we find that, although the precise form of the automorphisms depends on k, a common feature is that the coefficients e, f are always zero, so we will set them to zero from now on without loss of generality.

a b Assuming that the pair (A = , µ) GL(2, R) R is an automorphism of k=s¯, c d ∈ × × 0   the brackets involving Q change as follows:

[H, Q(s)] = Q(sh) [B(β), Q(s)] = Q(βsb) e e e e (4.1.1.32) [P(π), Q(s)] = Q(πsp) e e e e 2 [Q(s), Q(s)] = c0|s| H − J(sc1s)− B(sc2s)− P(sc3s), e e e e where J(ω)= J(ω) ande e e e e e e e e e

e h = µqhq−1 c1 = qc1q b = q(ab + cp)q−1 1 e c2 = q(dc2 − bc3)q p = q(bb + dp)q−1 e ad − bc (4.1.1.33) e 2 1 |q| ec3 = q(ac3 − cc2)q. c0 = c0 ad − bc e µ e In summary, the groupe G by which we must quotient the sub-variety S , cut out by the super- Jacobi identity (and [Q, Q] = 0), acts as follows on the generators: 6

J J Adu 7→ ◦ B aB Adu +cP Adu 7→ ◦ ◦ P bB Adu +dP Adu (4.1.1.34) 7→ ◦ ◦ H µH 7→ Q Q Adu Rq 7→ ◦ ◦ a b where µ R and q H are non-zero, u Sp(1) and A := GL(2, R) with (A, µ) an ∈ ∈ ∈ c d ∈   automorphism of k.

Let Autr(k) denote the subgroup of GL(2, R) R× consisting of such (A, µ). These subgroups × are listed in [5, §3.1] for the kinematical Lie algebras K1-K11 in Table 4.1. We will collect them

81 in Table 4.2 for convenience and in addition also record them for the remaining kinematical Lie algebras K12-K18 in Table 4.1.

Table 4.2: Automorphisms of Kinematical Lie Algebras (Acting Trivially on r)

K# Typical (A, µ) GL(2, R) R ∈ × × a b 1 , µ c d    a 0 2 , d c d a    a 0 3γ (−1,1) ,1 ∈ 0 d    a 0 0 b 3 ,1 , , −1 −1 0 d c 0       a b 3 ,1 1 c d    a b 4 ,1 χ>0 −b a    a b a b 4 ,1 , , −1 0 −b a b −a       a 0 5 ,1 c a    a b 6 , ad − bc c d    1 0 −1 0 7,8 , d , , −d c d c d       a 0 0 b 9 ,1 , , −1 0 a−1 b−1 0       a b a b 10,11 ,1 , , −1 , a2 + b2 = 1 −b a b −a       1 0 1 0 12,13 , µ , , µ 0 1 0 −1       1 0 14 , µ 0 d    a 0 15 , µ c a2    1 0 16 ,1 0 d    a 0 17 , a c a2    a 0 18 ,1 0 a2   

4.1.2 Classification of Kinematical Lie Superalgebras We now proceed to analyse each kinematical Lie algebra k in Table 4.1 in turn and impose the super-Jacobi identity for the corresponding Lie superalgebras extending k. We recall that we are only interested in those Lie superalgebras where [Q, Q] = 0, so c0, c1, c2, c3 cannot all 6 simultaneously vanish.

82 Kinematical Lie Algebras Without Supersymmetric Extensions There are three kinematical Lie algebras which cannot be extended to a kinematical superalge- bra: so(4,1), so(5) and the Euclidean algebra (K7 in Table 4.1).

The Euclidean Algebra

2 1 From Lemma 4.1.1, we find that p = h = 0 and b = 4 , so, in particular, b R, and, from 1 ∈ Lemma 4.1.2, we find that c2b + 2 c1 = 0. The [H, Q, Q] component of the super-Jacobi identity shows that c2 = 0, so that also c1 = 0. The [P, Q, Q] component of the super-Jacobi identity is trivially satisfied, whereas the [B, Q, Q] component shows that c3 = 0, and also that c0 = 0. In summary, there is no kinematical superalgebra extending the Euclidean algebra for which [Q, Q] = 0; although there is a kinematical superalgebra where [B(β), Q(s)] = 1 Q(βs), where 6 ± 2 both choices of sign are related by an automorphism of k: e.g., time reversal (J, B, P, H) 7→ (J, −B, P, −H) or parity (J, B, P, H) (J, −B, −P, H). 7→ so(4,1) In this case, Lemma 4.1.1 gives that p = b = 0, but then the [B, P, Q] component of the super- Jacobi identity cannot be satisfied, showing that the so(3) representation on the spinor module S cannot be extended to a module of so(4,1). The result would be different for N = 2 extensions, since so(4,1) =∼ sp(1,1) does have an irreducible spinor module of quaternionic dimension 2. so(5) From Lemma 4.1.1, we find that p = [b, h] from [H, B] = P, and, in particular, p Im H. ∈ But then [P, P] = J says that p2 = 1 , so that in particular p R and non-zero, which is a 4 ∈ contradiction. Again, this shows that the spinor module S of so(3) does not extend to a module of so(5), and, again, the conclusion would be different for N = 2 extensions, since so(5) =∼ sp(2) does admit a quaternionic module of quaternionic dimension 2.

Lorentzian Kinematical Superalgebras The Poincaré Lie algebra (K8) and so(3,2) are the Lorentzian isometry Lie algebras of the Minkowski and anti-de Sitter spacetimes, respectively. It is of course well known that such spacetimes admit N = 1 superalgebras of maximal dimension. We treat them in this section for completeness.

The Poincaré Superalgebra

2 1 From Lemma 4.1.1, we see that p = h = 0 and that b = − 4 , so that in particular b Im H. 1 ∈ From Lemma 4.1.2, we see that 2 c1 + c2b = 0. The [P, Q, Q] component of the super-Jacobi identity is trivially satisfied, whereas the [H, Q, Q] component forces c1 = c2 = 0 and the [B, Q, Q] component says c3 = 2c0b. Demanding [Q, Q] = 0 requires c0 = 0. 6 6 1 Using the quaternion automorphism, we can rotate b so that b = 2 k and via the automor- phism of the Poincaré Lie algebra, which rescales H and P by the same amount, we can bring c0 = 1. In summary, we have a unique isomorphism class of kinematical Lie superalgebras extending the Poincaré Lie algebra which has the additional Lie brackets

1 2 [B(β), Q(s)] = 2 Q(βsk) and [Q(s), Q(s)] = |s| H − P(sks). (4.1.2.1)

The AdS Superalgebra Here, Lemma 4.1.1 and Lemma 4.1.2 give the following relations:

2 1 2 1 1 p =[h, b], b =[p, h], h =[b, p], b =− 4 , p =− 4 and c0h = 2 c1 + c2b + c3p, (4.1.2.2)

83 and in addition bp + pb = 0, which simply states that b p. These relations imply that ⊥ b, p, h Im H and that (2b,2p,2h) is an oriented orthonormal basis for Im H. The remaining ∈ super-Jacobi identities give

c2 =−2c0p, c3 = 2c0b = c1 =−2c0h, (4.1.2.3) ⇒ and some other relations which are identically satisfied. If c0 = 0 then [Q, Q] = 0, so we re- c1 c2 c3 quires c0 = 0. Hence ( , , ) defines a negatively oriented, orthonormal basis for Im H. The 6 c0 c0 c0 automorphism group of H acts transitively on the space of orthonormal oriented bases, so we can choose (2b,2p,2h)=(i, j, k) without loss of generality.

The resulting Lie superalgebra becomes

H Q 1 Q [ , (s)] = 2 (sk) [B(β), Q(s)] = 1 Q(βsi) 2 (4.1.2.4) P Q 1 Q [ (π), (s)] = 2 (πsj) 2 [Q(s), Q(s)] = c0 |s| H + J(sks)+ B(sj¯s)− P(sis) .  We may rescale Q to bring c0 to a sign, but we can then change the sign via the automorphism of k which sends (J, B, P, H) (J, P, B, −H) and the inner automorphism induced by the auto- 7→ morphism of H which sends (i, j, k) (j, i, −k). In summary, there is a unique kinematical Lie 7→ superalgebra with [Q, Q] = 0 extending k = so(3,2): namely, 6 1 [H, Q(s)] = 2 Q(sk) B Q 1 Q [ (β), (s)] = 2 (βsi) 1 (4.1.2.5) [P(π), Q(s)] = 2 Q(πsj) [Q(s), Q(s)] = |s|2H + J(sks)+ B(sjs)− P(sis).

To show that this Lie superalgebra is isomorphic to osp(1|4), we may argue as follows. We first prove that s0¯ leaves invariant a symplectic form on s1¯. The most general rotationally invariant bilinear form on s1¯ is given by

ω(Q(s1), Q(s2)) := Re(s1qs2) for some q H. (4.1.2.6) ∈ Indeed, if u Sp(1) then ∈ −1 −1 (u ω)(Q(s1), Q(s2)) = ω(u Q(s1), u Q(s2)) · · · = ω(Q(us1), Q(us2))

= Re(us1qs2u) (4.1.2.7)

= Re(s1qs2)

= ω(Q(s1), Q(s2)).

Demanding that ω be invariant under the other generators H, B, P, we find that q = µk for some µ R. Acting infinitesimally now, ∈

(H ω)(Q(s1), Q(s2))=−ω([H, Q(s1)], Q(s2)) − ω(Q(s1), [H, Q(s2)]) · 1 1 =− 2 ω(Q(s1k), Q(s2)) − 2 ω(Q(s1), Q(s2k)) 1 1 (4.1.2.8) =− 2 Re(s1kqs2)+ 2 Re(s1qks2) 1 = 2 Re(s1[q, k]s2), which must vanish for all s1, s2 S, so that [q, k]= 0 and hence q = λ1+µk for some λ, µ R. A ∈ ∈ similar calculation with B and P shows that q must anti-commute with i and j and thus q = µk. So the action of s¯ =∼ so(3,2) on s¯ defines a Lie algebra homomorphism so(3,2) sp(4, R), 0 1 → which is clearly non-trivial. Since so(3,2) is simple, it is injective and a dimension count shows 2 ∼ 2 that this is an isomorphism. But as representations of so(3,2), s1¯ = ∧ V, where V is the 2 ∼ ⊙ ∼ 5-dimensional vector representation of s0¯, and, since ∧ V = so(V) = s0¯, we have that there

84 2 is a one-dimensional space of s0¯-equivariant maps s1¯ s0¯. Since [Q, Q] = 0, the bracket 2 ⊙ → 6 s¯ s¯ is an isomorphism. Thus s is, by definition, isomorphic to osp(1|4). ⊙ 1 → 0

The Carroll Superalgebra For k the Carroll Lie algebra (K6 in Table 4.1), Lemma 4.1.1 implies that p = b = h = 0, and then Lemma 4.1.2 says that c1 = 0. The [B, Q, Q] super-Jacobi says that c3 = 0, and the [P, Q, Q] super-Jacobi says that c2 = 0. The only non-zero bracket involving Q is

2 [Q(s), Q(s)] = c0|s| H, (4.1.2.9) which is non-zero for c0 = 0. If so, we can set c0 = 1 via an automorphism of k which rescales 6 H and P, say, by c0. In summary, there is a unique Carroll superalgebra with brackets

[Q(s), Q(s)] = |s|2H, (4.1.2.10) in addition to those of the Carroll Lie algebra itself. This Lie superalgebra is a contraction of the Poincaré superalgebra. We will show this explicitly in Section 4.3.

The Galilean Superalgebras For k the Galilean Lie algebra (K2 in Table 4.1), Lemma 4.1.1 says that b = p = 0, and Lemma 4.1.2 says that c1 = 2c0h. The [B, Q, Q] super-Jacobi identity says that c1 = 0 and c0 = 0. The [P, Q, Q] super-Jacobi identity is now identically satisfied, whereas the [H, Q, Q] super-Jacobi identity gives

hc2 + c2h¯ = 0 and c2 + hc3 + c3h¯ = 0. (4.1.2.11)

Since c2 and c3 cannot both vanish, we see that this is only possible if h Im H; therefore, ∈ these equations become [h, c2]= 0 and c2 =[c3, h]. There are two cases to consider, depending on whether or not h vanishes. If h = 0, then c2 = 0 and c3 is arbitrary. If h = 0, then on the 6 one hand c2 is collinear with h, but also c2 =[c3, h], which means that c2 = 0 so that c3 = 0 is 6 collinear with h. In either case, c3 = 0 and h = ψc3, where ψ R can be zero. 6 ∈ This gives rise to the following additional brackets

[H, Q(s)] = ψQ(sc3) and [Q(s), Q(s)]=−P(sc3s). (4.1.2.12)

We may use the automorphisms of H to bring c3 = φk, for some non-zero φ R. We can set ∈ φ = 1 by an automorphism of k which rescales P and also B and H suitably. This still leaves the freedom to set ψ = 1 if ψ = 0. In summary, we have two Galilean superalgebras: 6 0 [H, Q(s)] = and [Q(s), Q(s)]=−P(sks). (4.1.2.13) Q(sk)

The first one (where [H, Q] = 0) is a contraction of the Poincaré superalgebra, whereas the second (where [H, Q] = 0) is not. 6

Lie Superalgebras Associated with the Static Kinematical Lie Algebra The static kinematical Lie algebra is K1 in Table 4.1. In this case, Lemma 4.1.1 says that b = p = 0 and Lemma 4.1.2 says that c1 = 2c0h. The [H, Q, Q] super-Jacobi identity says that h Im H and that [h, c ]= 0 for i = 1, 2, 3. Finally, either the [B, Q, Q] or [P, Q, Q] super-Jacobi ∈ i identities say that c1 = 0, so that hc0 = 0. This means that either h = 0 or else c0 = 0 (or both).

There are several branches:

1. If c0 = 0 and h = 0, c2 and c3 are collinear with h, but cannot both be zero. Using 6 automorphisms of the static kinematical Lie algebra, and the ability to rotate vectors, we

85 1 can bring h = 2 k, c2 = 0 and c3 = k, so that we have a unique Lie superalgebra in this case, with additional brackets

1 [H, Q(s)] = 2 Q(sk) and [Q(s), Q(s)]=−P(sks). (4.1.2.14)

2. If c0 = 0 and h = 0, c2 and c3 are unconstrained, but not both zero. We distinguish two cases, depending on whether or not they are linearly independent:

(a) If they are linearly dependent, so that they are collinear, then we can use automor- phisms to set c2, say, to zero and c3 = k. This results in the Lie superalgebra

[Q(s), Q(s)]=−P(sks). (4.1.2.15)

(b) If they are linearly independent, we can bring them to c2 = j and c3 = k, resulting in the Lie superalgebra

[Q(s), Q(s)]=−B(sjs)− P(sks). (4.1.2.16)

3. Finally, if c0 = 0, then h = 0 and, again, c2 and c3 are unconstrained, but can now be zero. 6 Moreover we can rescale H so that c0 = 1. We have three cases to consider, depending on whether they span a zero-, one- or two-dimensional real subspace of Im H:

(a) If c2 = c3 = 0, we have the Lie superalgebra

[Q(s), Q(s)] = |s|2H. (4.1.2.17)

(b) If c2 and c3 span a line, then we may use the automorphisms to set c2 = 0 and c3 = k, resulting in the Lie superalgebra

[Q(s), Q(s)] = |s|2H − P(sks). (4.1.2.18)

(c) Finally, if c2 and c3 are linearly independent, we may use the automorphisms to set c2 = j and c3 = k, resulting in the Lie superalgebra

[Q(s), Q(s)] = |s|2H − B(sjs)− P(sks). (4.1.2.19)

Lie Superalgebras Associated with Kinematical Lie Algebra K3γ

Here, Lemma 4.1.1 says that b = p = 0 and Lemma 4.1.2 says that c1 = 2c0h. The [B, Q, Q] super-Jacobi identity says that c1 = 0 and c0 = 0, whereas the [P, Q, Q] super-Jacobi identity offers no further conditions. Finally, the [H, Q, Q] super-Jacobi identity gives two conditions

γc2 = hc2 + c2h¯ and c3 = hc3 + c3h¯, (4.1.2.20) which are equivalent to

(γ − 2 Re h)c2 =[Im h, c2] and (1 − 2 Re h)c3 =[Im h, c3]. (4.1.2.21)

We see that we must distinguish two cases: γ = 1 and γ [−1,1). ∈ If γ = 1, then we have two cases, depending on whether Re h = 1 or Re h = 1 γ. In the 6 2 2 former case, c2 = 0 and Im h is collinear with c3 = 0, whereas, in the latter, c3 = 0 and Im h is 6 collinear with c2 = 0. 6 1 If γ = 1, then Re h = 2 and c2, Im h and c3 are all collinear, with at least one of c2 and c3 non-zero. When γ = 1, the automorphisms of k include the GL(2, R) acting on the two copies of the vector representation. Using this fact, we can always assume that c2 = 0 and c3 = 0. 6

86 In either case, all non-zero vectors are collinear and we can rotate them to lie along the k axis. In the case γ = 1, we have a one-parameter family of Lie superalgebras:

1 [H, Q(s)] = 2 Q(s(1 + λk)) and [Q(s), Q(s)]=−P(sks), (4.1.2.22) where we have used the freedom to rescale P in order to set c3 = k. This is also a Lie superal- gebra for γ = 1. 6 If γ = 1, we have an additional one-parameter family of Lie superalgebras: 6 1 [H, Q(s)] = 2 Q(s(γ + λk)) and [Q(s), Q(s)]=−B(sks). (4.1.2.23)

The parameter λ is essential; that is, Lie superalgebras with different values of λ are not isomorphic. One way to test this is the following. Let [−, −]λ denote the above Lie bracket. This satisfies the super-Jacobi identity for all λ R. Write it as [−, −] =(1−λ)[−, −]0+λ[−, −]1. ∈ λ The difference [−, −]1 − [−, −]0 is a cocycle of the Lie superalgebra with λ = 0. The parameter would be inessential if and only if it is a coboundary. One can check that this is not the case. This same argument shows that the parameters appearing in other Lie superalgebras are essential as well.

Lie Superalgebras Associated with Kinematical Lie Algebra K4χ

Here, Lemma 4.1.1 says b = p = 0 and Lemma 4.1.2 says that c1 = 2c0h. Then either the [B, Q, Q] or [P, Q, Q] super-Jacobi identities force c1 = 0 and c0 = 0. The [H, Q, Q] super-Jacobi identity results in the following two equations:

χc2 − c3 = hc2 + c2h¯ and χc3 + c2 = hc3 + c3h¯, (4.1.2.24) or equivalently,

(χ − 2 Re h)c2 − c3 =[Im h, c2] and (χ − 2 Re h)c3 + c2 =[Im h, c3]. (4.1.2.25)

Taking the inner product of the first equation with c2 and of the second equation with c3 and adding, we find 2 2 (χ − 2 Re h)(|c2| + |c3| )= 0, (4.1.2.26) χ and since c2 and c3 cannot both be zero, we see that Re h = 2 , and hence that

[Im h, c2]=−c3 and [Im h, c3]= c2, (4.1.2.27) so that c3 c2. This shows that (Im h, c3, c2) is an oriented orthogonal (but not necessarily ⊥ orthonormal) basis. We can rotate them so that Im h = φj, c3 = ψk and c2 = 2φψi, but then 2 1 we see that φ = 4 . Using the automorphism of k which rescales B and P simultaneously by 1 the same amount, we can assume that c3 = k; then, if Im h = j, we find c2 = i. But the ± 2 ± two signs are related by the automorphism of H which sends (i, j, k) (−i, −j, k). In summary, 7→ we have a unique Lie superalgebra associated with this kinematical Lie algebra:

1 [H, Q(s)] = 2 Q(s(χ + j)) and [Q(s), Q(s)]=−B(sis)− P(sks). (4.1.2.28)

Lie Superalgebras Associated with Kinematical Lie Algebra K5

Here, Lemma 4.1.1 says that b = p = 0 and Lemma 4.1.2 says that c1 = 2c0h. The [B, Q, Q] super-Jacobi identity forces c0 = c1 = 0, which then satisfies the [P, Q, Q] super-Jacobi identity identically. The [H, Q, Q] super-Jacobi identity gives two further equations

c2 = hc2 + c2h¯ and c2 + c3 = hc3 + c3h¯. (4.1.2.29)

The first equation is equivalent to

(1 − 2 Re(h))c2 =[Im h, c2]. (4.1.2.30)

87 1 If c2 = 0, then Re h = and Im h is collinear with c2. But then the second equation says that 6 2 c2 = [Im h, c3], which is incompatible with c2 and Im h being collinear. Therefore, c2 = 0 and 1 the second equation then says that Re h = and that Im h is collinear with c3 = 0. In this 2 6 instance, we have the following additional brackets

1 [H, Q(s)] = 2 Q(s(1 + λc3)) and [Q(s), Q(s)]=−P(sc3s), (4.1.2.31) where λ R. We may rotate c3 to ψk, for some non-zero ψ R. We can then rescale P and B ∈ ∈ simultaneously by the same amount to set ψ = 1. In summary, we are left with the following one-parameter family of Lie superalgebras:

1 [H, Q(s)] = 2 Q(s(1 + λk)) and [Q(s), Q(s)]=−P(sks). (4.1.2.32)

As in the case of the Lie superalgebras associated with Lie algebra K3γ, the parameter λ is essential and Lie superalgebras with different values of λ are not isomorphic.

Lie Superalgebras Associated with Kinematical Lie Algebra K12 Lemma 4.1.1 says that b2 = 1 b, so that b R, [h, p]= 0 and p2 = 1 (b − 1 ), so that p Im H. 2 ∈ 2 2 ∈ (In particular, bp = 0.) Lemma 4.1.2 does not simplify at this stage. The [H, Q, Q] super-Jacobi identity says that c0 Re h = 0 and that hci + cih¯ = 0 for i = 1,2,3. The [B, Q, Q] super-Jacobi identity says that bc1 = 0, bc3 = 0 and c1 = (2b − 1)c2. Finally, the [P, Q, Q] super-Jacobi identity says that c0p = 0, among other conditions that will turn out not to play a rôle.

We have two branches depending on the value of b:

2 1 1. If b = 0, p = − 4 , so that c0 = 0. This means c1 + c2 = 0 and c3 = 2c1p and none of c1,2,3 can vanish. This means that Re h = 0 and that h and ci are collinear for all i = 1,2,3. Also, h and p are collinear, which is inconsistent, unless h = 0: indeed, if p and c are collinear with h = 0, then c3 = 2c1p cannot be satisfied, since the L.H.S. i 6 is imaginary but the R.H.S. is real and both are non-zero. Therefore, we conclude that −1 −1 h = 0. The condition c3 = 2c1p says that there exists ψ> 0 such that (ψ c1,2p, ψ c3) is an oriented orthonormal basis, which can be rotated to (i, j, k). In other words, we can 1 write c1 = ψi, p = 2 j and c3 = ψk, so that c2 =−ψi. We may rescale Q to bring ψ = 1 and we may rotate (i, j, k) (−i, j, −k) to arrive at the following Lie superalgebra: 7→ 1 [P(π), Q(s)] = 2 Q(sj) and [Q(s), Q(s)] = J(sis)− B(sis)+ P(sk¯s). (4.1.2.33)

1 2. If b = , then p = 0, c1 = c3 = 0, and c2 = 2c0h with c0 = 0. We have two sub-branches, 2 6 depending on whether or not h = 0.

(a) If h = 0, we have the following Lie superalgebra, after rescaling H to set c0 = 1:

1 2 [B(β), Q(s)] = 2 Q(βs) and [Q(s), Q(s)] = |s| H. (4.1.2.34)

(b) On the other hand, if h = 0, we may rotate it so that 2h = ψk for some ψ such that 6 ψc0 > 0. Then we may rescale H and Q in such that a way that we bring ψc0 = 1, thus arriving at the following Lie superalgebra:

1 1 2 [B(β), Q(s)] = 2 Q(βs), [H, Q(s)] = 2 Q(sk) and [Q(s), Q(s)] = |s| H − B(sks). (4.1.2.35)

Lie Superalgebras Associated with Kinematical Lie Algebra K13 Here, Lemma 4.1.1 says that b2 = 1 b, so that b R and p2 = − 1 (b − 1 ) R. Lemma 4.1.2 2 ∈ 2 2 ∈ does not simplify further at this stage. The [H, Q, Q] super-Jacobi identity says that c0 Re h = 0 and hci + cih¯ = 0 for i = 1,2,3. The [B, Q, Q] super-Jacobi identity says that bc1 = bc3 = 0, 1 1 whereas (b − 2 )c2 = 2 c1. Finally, the [P, Q, Q] super-Jacobi identity says that c1 = 2pc3, c3 =−2pc2 and c3 = 2pc1.

88 As usual we have two branches depending on the value of b:

2 1 1. If b = 0, then p = 4 . Due to the automorphism of k which changes the sign of P, 1 we may assume p = 2 without loss of generality. It follows that c1 = c0h and that c2 =−c1 =−c0h and that c3 = c1 = c0h. If c0 = 0, then ci = 0 for all i, so we must have c0 = 0. In that case, h Im H and h is collinear with all c for i = 1, 2, 3. We distinguish 6 ∈ i two cases, depending on whether or not h = 0:

−1 (a) If h = 0, we may rotate it so that h = ψk, where ψc0 > 0. We may rescale H ψ H 6 7→ (which is an automorphism of k) and rescale Q to bring ψc0 = 1. In summary, we arrive at the following Lie superalgebra:

1 2 [H, Q(s)] = Q(sk), [P(π), Q(s)] = 2 Q(πs) and [Q(s), Q(s)] = |s| H−J(sks)+B(sks)−P(sks). (4.1.2.36) (b) If h = 0, then we have the Lie superalgebra

1 2 [P(π), Q(s)] = 2 Q(πs) and [Q(s), Q(s)] = |s| H. (4.1.2.37)

1 2. If b = , then p = 0 and c1 = c3 = 0 with c2 = 2c0h with c0 = 0 and h Im H. Again, we 2 6 ∈ distinguish between vanishing and non-vanishing h:

(a) If h = 0, we may rotate it so that 2h = ψk with ψc0 > 0. We apply the k- 6 −1 automorphism H ψ H and rescale Q to bring ψc0 = 1, thus resulting in the 7→ Lie superalgebra

1 1 2 [H, Q(s)] = 2 Q(sk), [B(β), Q(s)] = 2 Q(βs) and [Q(s), Q(s)] = |s| H − B(sks). (4.1.2.38) (b) If h = 0, we arrive at the Lie superalgebra

1 2 [B(β), Q(s)] = 2 Q(βs) and [Q(s), Q(s)] = |s| H. (4.1.2.39)

Lie Superalgebras Associated with Kinematical Lie Algebra K14 Here, Lemma 4.1.1 says that p = 0 and 2b2 = b, so that b R. Lemma 4.1.2 says that 1 ∈ 2 c1 + c2b = c0h. The [P, Q, Q] super-Jacobi identity says that c1 = 0, so that c0h = c2b. The [B, Q, Q] super-Jacobi identity says that (2b − 1)c2 = 0 and bc3 = 0, whereas the [B, Q, Q] super-Jacobi identity says that hci + cih¯ = 0 for i = 2,3.

We have two branches, depending on the value of b:

1. If b = 0, then c2 = 0, and we have two sub-branches depending on whether or not c0 = 0:

(a) If c0 = 0, then c3 = 0, so that Re h = 0 and h is collinear with c3. We may rotate 6 c3 to lie along k, say, and then use automorphisms of k to set c3 = k. If h = 0, 6 we may also set it equal to k. In summary, we have two isomorphism classes of Lie superalgebras here:

0 [H, Q(s)] = and [Q(s), Q(s)]=−P(sks). (4.1.2.40) Q(sk)

(b) If c0 = 0, then h = 0 and c3 is free: if non-zero, we may rotate it to k and, rescaling P 6 with the automorphisms of k, we can bring it to k. Rescaling H we can bring c0 = 1. This gives two isomorphism classes of Lie superalgebras:

[Q(s), Q(s)] = |s|2H and [Q(s), Q(s)] = |s|2H − P(sks). (4.1.2.41)

1 2. If b = 2 , then c3 = 0 and c2 = 2c0h, and we have two cases, depending on whether or not h = 0.

89 (a) If h = 0, then c2 = 0, and then c0 = 0. Rescaling H, we can set c0 = 1 and we arrive 6 at the Lie superalgebra

1 2 [B(β), Q(s)] = 2 Q(βs) and [Q(s), Q(s)] = |s| H. (4.1.2.42)

(b) If h = 0, we can rotate and rescale Q such that c2 = 2c0h = k and then we can 6 rescale H so that c0 = 1. The resulting Lie superalgebra is now

1 1 2 [H, Q(s)] = 2 Q(sk), [B(β), Q(s)] = 2 Q(βs) and [Q(s), Q(s)] = |s| H − B(sks). (4.1.2.43)

Lie Superalgebras Associated with Kinematical Lie Algebra K15

Here, Lemma 4.1.1 says that b = p = 0, whereas Lemma 4.1.2 says that c1 = 2c0h. The [B, Q, Q] super-Jacobi identity says that c1 = c2 = 0, and hence the [P, Q, Q] component is identically satisfied. Finally, the [H, Q, Q] super-Jacobi identity says that hc3 + c3h¯ = 0, which expands to 2 Re(h)c3 +[Im h, c3]= 0. (4.1.2.44) We have two branches of solutions:

1. If c0 = 0, then c3 = 0, Re h = 0 and h is collinear with c3. We may rotate c3 to lie along k 6 and then rescale Q so that c3 = k. If h = 0, we may use automorphisms of k to set h = k 6 as well. In summary, we have two isomorphism classes of Lie superalgebras:

Q(sk) [H, Q(s)] = and [Q(s), Q(s)]=−P(sks). (4.1.2.45) 0

2. If c0 = 0, then h = 0 and c3 is unconstrained. If non-zero, we may rotate it to lie along 6 k, rescale Q so that c3 = k and then use automorphisms of k to set c0 = 1. In summary, we have two isomorphism classes of Lie superalgebras:

[Q(s), Q(s)] = |s|2H or [Q(s), Q(s)] = |s|2H − P(sks). (4.1.2.46)

Lie Superalgebras Associated with Kinematical Lie Algebra K16

1 Here, Lemma 4.1.1 says that p = 0 and b(b − 2 ) = 0, so that b R. Lemma 4.1.2 then says 1 ∈ that c0h = 2 c1 + c2b. Now the [P, Q, Q] super-Jacobi identity says that c0 = 0 and c1 = 0, so that c2b = 0. The [H, Q, Q] super-Jacobi identity says that hc2 + c2h¯ = 0 and hc3 + c3h¯ = c3. 1 Finally the [B, Q, Q] super-Jacobi identity says that bc3 = 0 and (b − 2 )c2 = 0.

1 Notice that if b = 2 then c3 = 0 and c2 = 0, contradicting [Q, Q] = 0, so we must have 16 b = 0. Now c2 = 0 and hence c3 = 0. It then follows that Re h = and Im h is collinear 6 2 with c3. We can rescale P (which is an automorphism of k) and rotate so that c3 = k, so that h = 1 (1 + λk) for λ R. The resulting one-parameter family of Lie superalgebras is then 2 ∈ 1 [H, Q(s)] = 2 Q(s(1 + λk)) and [Q(s), Q(s)]=−P(sks). (4.1.2.47)

As in the case of the Lie superalgebras associated with Lie algebras K3γ and K5, the parameter λ is essential and Lie superalgebras with different values of λ are not isomorphic.

Lie Superalgebras Associated with Kinematical Lie Algebra K17

Here, Lemma 4.1.1 simply sets b = p = 0 and Lemma 4.1.2 says c1 = 2c0h. The [P, Q, Q] super-Jacobi identity sets c1 = 0 and hence c0h = 0. The [B, Q, Q] super-Jacobi identity sets c0 = 0 and c2 = 0, whereas the [H, Q, Q] super-Jacobi identity says that h is collinear with ψ c3 = 0. We can rotate c3 to lie along k and rescale Q to effectively set it to k. Then h = k for 6 2 some ψ and rescaling H allows us to set ψ = 1. In summary, we have a unique Lie superalgebra

90 associated with this kinematical Lie algebra; namely,

1 [H, Q(s)] = 2 Q(sk) and [Q(s), Q(s)]=−P(sks). (4.1.2.48)

Lie Superalgebras Associated with Kinematical Lie Algebra K18

Here, Lemma 4.1.1 simply sets b = p = 0, and Lemma 4.1.2 says c1 = 2c0h. The [P, Q, Q] super-Jacobi identity sets c1 = 0 and c0 = 0, whereas the [B, Q, Q] super-Jacobi identity sets c2 = 0. Finally, the [H, Q, Q] super-Jacobi identity says that Re h = 1 and Im h = λc3 for some λ R. We can rotate c3 to lie along k and rescale Q to effectively set it to k. Then h = 1+λk. In ∈ summary, we have a one-parameter family of Lie superalgebras associated with this kinematical Lie algebra; namely,

[H, Q(s)] = Q(s(1 + λk)) and [Q(s), Q(s)]=−P(sks). (4.1.2.49)

As in the case of the Lie superalgebras associated with Lie algebras K3γ, K5 and K16, the parameter λ is essential and Lie superalgebras with different values of λ are not isomorphic.

4.1.3 Summary

Table 4.3 summarises the results. In that table, we list the isomorphism classes of kinematical Lie superalgebras (with [Q, Q] = 0). Recall that the Lie brackets involving Q are the [Q, Q] 6 bracket and also

[H, Q(s)] = Q(sh), [B(β), Q(s)] = Q(βsb), [P(π), Q(s)] = Q(πsp), (4.1.3.1) for some h, b, p H. In Table 4.3, we list any non-zero values of h, b, p and the [Q, Q] bracket. ∈ The first column is simply the label for the Lie superalgebra, the second column is the cor- responding kinematical Lie algebra, the next columns are h, b, p and [Q, Q]. The next four columns are the possible so(3)-equivariant Z-gradings (with J of degree 0) compatible with the Z2-grading; that is, such that the parity is the reduction modulo 2 of the degree. This requires, in particular, that q be an odd integer, which we can take to be −1 by convention, if so desired.

Despite all of the Lie superalgebras in Table 4.3 producing spacetime supersymmetry, there are some important qualitative differences between those Lie superalgebras that have the time translation generator H in the [Q, Q] bracket and those that do not. In particular, theories invariant under a supersymmetry algebra for which [Q, Q] = H are guaranteed to have a non- negative energy spectrum [88]. Therefore, for the construction of phenomenological models, these Lie superalgebras may be of more interest.

Unpacking the Quaternionic Notation

The quaternionic formalism we have employed in the classification of kinematical Lie superal- gebras, which has the virtue of uniformity and ease in computation, results in expressions that are perhaps unfamiliar and, therefore, might hinder comparison with other formulations. In this section, we will go through an example illustrating how to unpack the notation.

The non-zero, supersymmetric brackets of the Poincaré superalgebra S14 are given by

1 2 [B(β), Q(s)] = Q( 2 βsk) and [Q(s), Q(s)] = |s| H − P(sks), (4.1.3.2) where 3 4

B(β)= βiBi and Q(s)= saQa, (4.1.3.3) =1 a=1 Xi X and where β = β1i + β2j + β3k and s = s1i + s2j + s3k + s4. (4.1.3.4)

91 Table 4.3: Kinematical Lie Superalgebras (with [Q, Q] = 0) 6

S# k h b p [Q(s), Q(s)] wH wB wP wQ 1 1 K1 2 k −P(sks) 0 2m 2q q 2 K1 |s|2H − B(sjs)− P(sks) 2q 2q 2q q 3 K1 |s|2H − P(sks) 2q 2m 2q q 4 K1 |s|2H 2q 2m 2p q 5 K1 −B(sjs)− P(sks) 2n 2q 2q q 6 K1 −P(sks) 2n 2m 2q q 7 K2 k −P(sks) 0 2q 2q q 8 K2 −P(sks) 2n 2(q − n) 2q q 1 9γ [−1,1],λ R K3γ (1 + λk) −P(sks) 0 2m 2q q ∈ ∈ 2 K3 1 B 10γ [−1,1),λ R γ 2 (γ + λk) − (sks) 0 2q 2p q ∈ ∈ 1 11χ>0 K4χ 2 (χ + j) −B(sis)− P(sks) 0 2q 2q q 1 12λ R K5 (1 + λk) −P(sks) 0 2q 2q q ∈ 2 13 K6 |s|2H 2q 2m 2(q − m) q 1 2 14 K8 2 k |s| H − P(sks) 2q 0 2q q K11 1 1 1 2H J B P 15 2 k 2 i 2 j |s| + (sks)+ (sjs)− (sis) − − − − 1 16 K12 2 j J(sis)− B(sis)+ P(sks) − − − − 1 2 17 K12 2 |s| H 2q 0 0 q 1 1 2 18 K12 2 k 2 |s| H − B(sks) − − − − K13 1 2H J B P 19 k 2 |s| − (sks)+ (sks)− (sks) − − − − 1 2 20 K13 2 |s| H 2q 0 0 q 1 2 21 K13 2 |s| H 2q 0 0 q 1 1 2 22 K13 2 k 2 |s| H − B(sks) − − − − 23 K14 k −P(sks) 0 0 2q q 24 K14 −P(sks) 2n 0 2q q 25 K14 |s|2H 2q 0 2p q 26 K14 |s|2H − P(sks) 2q 0 2q q 1 2 27 K14 2 |s| H 2q 0 2p q K14 1 1 2H B 28 2 k 2 |s| − (sks) − − − − 29 K15 k −P(sks) − − − − 30 K15 −P(sks) − − − − 31 K15 |s|2H 2q 2m 4m q 32 K15 |s|2H − P(sks) − − − − 1 33λ R K16 (1 + λk) −P(sks) 0 0 2q q ∈ 2 1 34 K17 2 k −P(sks) − − − − 35λ R K18 1 + λk −P(sks) − − − − ∈ The first column is our identifier for s, whereas the second column is the kinematical Lie algebra k=s0¯ in Table 4.1. The next four columns specify the brackets of s not of the form [J, −]. Supercharges Q(s) are parametrised by s H, whereas J(ω), B(β) and P(π) are parametrised ∈ by ω, β, π Im H. The brackets are given by [H, Q(s)] = Q(sh), [B(β), Q(s)] = Q(βsb) and ∈ [P(π), Q(s)] = Q(πsp), for some h, b, p H. (This formalism is explained in Section 2.1.5.) The ∈ final four columns specify compatible gradings of s, with m, n, p, q Z and q odd. ∈

This allows us to simply unpack the brackets into the following expressions

4 3 1 b µ [Bi,Qa]= 2 Qbβi a and [Qa,Qb]= Pµγab. (4.1.3.5) =1 µ=0 bX X b Here, we have introduced P0 = H, and the matrices βi := [βi a] are given by

0 −1 0 iσ 1 0 β = , β = 2 and β = . (4.1.3.6) 1 −1 0 2 −iσ 0 3 0 −1    2   

92 µ µ Additionally, the symmetric matrices γ := [γab] are given by

1 0 0 1 0 −iσ −1 0 γ0 = , γ1 = , γ2 = 2 and γ3 = . (4.1.3.7) 0 1 1 0 iσ 0 0 1      2    As will be shown in Section 4.2.5, there is a two-parameter family of symplectic forms on the spinor module S which are invariant under the action of Bi and Ji. They are given by

ω(s1, s2) := Re(s1(αi + βj)s2), (4.1.3.8) for α, β R not both zero. We may normalise ω such that α2 + β2 = 1, resulting in a circle of ∈ symplectic structures. Relative to the standard real basis (i, j, k,1) for H, the matrix Ω of ω is −1 given by Ω = iσ2 (−ασ1 + βσ3), whose inverse is Ω =−Ω, due to the chosen normalisation. ⊗ µ µ a −1 ac µ Let us define endomorphisms γ of S such that (γ ) b =(Ω ) γcb. Explicitly, they are given by 0 2 γ = iσ2 (ασ1 − βσ3) γ =−1 (ασ3 + βσ1) 1 ⊗ 3 ⊗ (4.1.3.9) γ = σ3 (ασ1 − βσ3) γ = σ1 (ασ1 − βσ3). ⊗ ⊗ It then follows that these endomorphisms represent the Clifford algebra Cℓ(1,3):

γµγν + γνγµ =−2ηµν1. (4.1.3.10)

4.1.4 Classification of Aristotelian Lie Superalgebras Table 4.4 lists the Aristotelian Lie algebras (with three-dimensional space isotropy), classified in [5, App. A]. In this section, we classify the N = 1 supersymmetric extensions of the Aristotelian Lie algebras (with [Q, Q] = 0). 6 Table 4.4: Aristotelian Lie Algebras and their Spacetimes

A# Non-zero Lie brackets Spacetime 1 A 2 [H, P]= P TA 3 3+ [P, P]= J R S × 3 3− [P, P]=−J R H ×

Lie Superalgebras Associated with Aristotelian Lie Algebra A1 We start with the static Aristotelian Lie algebra A1, whose only non-zero brackets are [J, J]= J and [J, P]= P. Any supersymmetric extension s has possible brackets

2 [H, Q(s)] = Q(sh), [P(π), Q(s)] = Q(πsp) and [Q(s), Q(s)] = c0|s| H − J(sc1s)− P(sc3s), (4.1.4.1) for some h, p H, c0 R and c1, c3 Im H, using the same notation as in Section 4.1.2. We ∈ ∈ ∈ can re-use Lemmas 4.1.1 and 4.1.2, by setting b = 0 and c2 = 0 and ignoring B. Doing so, we find that p = 0 and that c1 = 2c0h. The [H, Q, Q] component of the super-Jacobi identity gives c0 Re h = 0 (which already follows from Lemma 4.1.2), c1h + hc1 = 0 and c3h + hc3 = 0. The [P, Q, Q] component of the super-Jacobi identity says that [sc1s, π] = 0 for all π Im H and ∈ s H, which says c1 = 0 and hence c0h = 0. This gives rise to two branches: ∈

1. If c0 = 0, then c3 = 0 and the condition c3h+hc3 = 0 is equivalent to [Im h, c3]=−2c3 Re h, 6 which says Re h = 0, and, therefore, that h and c3 are collinear. We can change basis so that c3 = k and h = k if non-zero. This leaves two possible Lie superalgebras depending on whether or not h = 0:

Q(sk) [H, Q(s)] = and [Q(s), Q(s)]=−P(sks). (4.1.4.2) 0

93 2. If c0 = 0, then h = 0 and c3 is free. We can set c0 = 1 and, if non-zero, we can also set 6 c3 = k. This gives two possible Lie superalgebras:

|s|2H [Q( ) Q( )] = s , s 2 (4.1.4.3) |s| H − P(sks).

Lie Superalgebras Associated with Aristotelian Lie Algebra A2 Let us now consider the Aristotelian Lie algebra A2, with additional bracket [H, P] = P. Lemma 4.1.1 says p = 0 and Lemma 4.1.2 says that c1 = 2c0h. The [H, Q, Q] component of the super-Jacobi identity implies that c0 Re h = 0 (which is redundant), c1h + hc1 = 0 and 2 c3h + hc3 = c3, whereas the [P, Q, Q] component results in [sc1s, π] = 2c0|s| π for all π Im H ∈ and s H. This can only be the case if c0 = 0; therefore, c1 = 0, which then forces c3 = 0. The ∈ 1 6 equation c3h + hc3 = c3 results in [Im h, c3]=(1 − 2 Re h)c3, which implies Re h = 2 and Im h collinear with c3. We can change basis so that c3 = k and we end up with a one-parameter family of Lie superalgebras with brackets

1 [H, Q(s)] = Q( 2 s(1 + λk)), [Q(s), Q(s)]=−P(sks) (4.1.4.4) for λ R, in addition to [H, P(π)] = P(π). ∈

Lie Superalgebras Associated with Aristotelian Lie Algebras A3 ± Finally, we consider the Aristotelian Lie algebras A3 with bracket [P, P] = J. Lemma 4.1.1 2 1 ± ±1 says that [h, p] = 0 and p = , whereas Lemma 4.1.2 says that c0h = c1 + c3p. The ± 4 2 [H, Q, Q] super-Jacobi says c0 Re h = 0, c1h + hc1 = 0 and c3h + hc3 = 0, whereas the [P, Q, Q] super-Jacobi gives the following relations:

1 1 c0 Re(sπsp)= 0, πspc3s − sc3psπ = [π, sc1s] and πspc1s − sc1psπ = [π, sc3s]. 2 ± 2 (4.1.4.5) We must distinguish two cases depending on the choice of signs.

1. Let’s take the + sign. Then p2 = 1 R. Without loss of generality, we can take p = 1 4 ∈ 2 by changing the sign of P if necessary. Then the [P, Q, Q] super-Jacobi equations say that c1 = c3 and hence c0h = c1. If c0 = 0, then c1 = c3 = 0, hence we take c0 = 0 and thus 6 Re h = 0. We can change basis so that c0 = 1 and hence h = c1 = c3. If non-zero, we can take them all equal to k. In summary, we have two possible Aristotelian Lie superalgebras 1 extending A3+, with brackets [P(π), P(π′)] = 2 J([π, π′]) and, in addition, either

1 2 [P(π), Q(s)] = Q( 2 πs), and [Q(s), Q(s)] = |s| H (4.1.4.6) or

1 2 [H, Q(s)] = Q(sk), [P(π), Q(s)] = Q( 2 πs) and [Q(s), Q(s)] = |s| H − J(sks)− P(sks). (4.1.4.7)

2. Let us now take the − sign. Here p2 = − 1 , so that p Im H (and p = 0) and Im h is 4 ∈ 6 collinear with p. The [H, Q, Q] super-Jacobi equations force h = 0 and the [P, Q, Q] super- 1 Jacobi equations force c0 = 0 and c3p =− 2 c1. This means that (c1,2p, c3) is an oriented orthonormal frame for Im H and hence we can rotate them so that (c1,2p, c3)=(−j, i, k), for later uniformity. This results in the Aristotelian Lie superalgebra extending A3− by 1 the following brackets in addition to [P(π), P(π′)]=− 2 J([π, π′]):

1 [P(π), Q(s)] = Q( 2 πsi) and [Q(s), Q(s)] = J(sjs)− P(sks). (4.1.4.8)

These results are summarised in Table 4.5 below, together with the possible compatible Z- gradings. This table also classifies the homogeneous Aristotelian superspaces for similar reasons to the non-supersymmetric case; see Section 3.2 for a definite explanation.

94 Table 4.5: Aristotelian Lie Superalgebras (with [Q, Q] = 0) 6

S# a h p [Q(s), Q(s)] wH wP wQ 36 A1 k −P(sks) 0 2q q 37 A1 −P(sks) 2n 2q q 38 A1 |s|2H 2q 2p q 39 A1 |s|2H − P(sks) 2q 2q q 1 40λ R A2 (1 + λk) −P(sks) 0 2q q ∈ 2 1 2 41 A3+ 2 |s| H 2q 0 q 1 2 42 A3+ k 2 |s| H − J(sks)− P(sks) − − − 1 43 A3− 2 i J(sjs)− P(sks) − − −

The first column is our identifier for s, whereas the second column is the Aristotelian Lie algebra a=s0¯ in Table 4.4. The next three columns specify the brackets of s not of the form [J, −]. Supercharges Q(s) are parametrised by s H, whereas J(ω) and P(π) are parametrised by ∈ ω, π Im H. The brackets are given by [H, Q(s)] = Q(sh) and [P(π), Q(s)] = Q(πsp), for some ∈ h, p H. (The formalism is explained in Section 2.1.5.) The final three columns are compatible ∈ gradings of s, with n, p, q Z and q odd. ∈

4.1.5 Central Extensions

In this section, we determine the possible central extensions of the kinematical and Aristotelian Lie superalgebras.

We start with the kinematical Lie superalgebras. Let s=s¯ s¯ be one of the Lie super- 0 ⊕ 1 algebras in Table 4.3. Bya central extension of s, we mean a short exact sequence of Lie superalgebras 0 z s s 0, (4.1.5.1) where z is central in s. We may choose a vector space splitting and view (as a vector space) b s=s z and the Lie bracket is given, for (X, z), (Y, z′) s z, by ⊕ ∈ ⊕ b b [(X, z), (Y, z′)]s = ([X, Y]s, ω(X, Y)) , (4.1.5.2) where ω : ∧2s z is a cocycle. (Here ∧ isb taken in the super sense, so that it is symmetric → on odd elements.) Central extensions of s are classified up to isomorphism by the Chevalley– Eilenberg cohomology group H2(s), which, by Hochschild–Serre, can be computed from the subcomplex relative to the rotational subalgebra r s¯. Indeed, we have the isomorphism [107] ⊂ 0 H2(s) =∼ H2(s, r). (4.1.5.3)

2 2 Let W = spanR {H, B, P, Q}. Then the cochains in C (s, r) are r-equivariant maps ∧ W R, or, 2 → equivalently, r-invariant vectors in ∧ W∗. This is a two-dimensional real vector space which, in quaternionic language, is given for x, y R by ∈

ω(B(β), P(π)) = x Re(βπ)=−ω(P(π), B(β)) and ω(Q(s1), Q(s2)) = y Re(s1s2). (4.1.5.4)

The cocycle conditions (i.e., the Jacobi identities of the central extension s) have several com- ponents. Letting V stand for either B or P, the cocycle conditions are given by b ω([H, V(α)], V(β))+ ω(V(α), [H, V(β)]) = 0, ω([V(α), V(β)], V(γ)) + cyclic = 0, (4.1.5.5) ω([H, Q(s)], Q(s)) = 0, 2ω([V(α), Q(s)], Q(s))+ ω([Q(s), Q(s)], V(α)) = 0.

95 The first two of the above equations only involve the even generators and hence depend only on the underlying kinematical Lie algebra, whereas the last two equations do depend on the precise superalgebra we are dealing with. In the case of Aristotelian Lie superalgebras, there is no B and hence V = P in the above equations and, of course, the cocycle can only modify the [Q, Q] bracket and hence the cocycle conditions are simply

ω([H, Q(s)], Q(s)) = 0 and ω([P(α), Q(s)], Q(s)) = 0. (4.1.5.6)

The calculations are routine, and we will not give any details, but simply collect the results in Table 4.6, where Z is the basis for the one-dimensional central ideal z = spanR {Z}, and where we list only the brackets which are liable to change under central extension.

Table 4.6: Central Extensions of Kinematical and Aristotelian Lie Superalgebras

S# [B(β), P(π)] [Q(s), Q(s)] 1 |s|2Z − P(sks) 4 − Re(βπ)Z |s|2H 5 |s|2Z − B(sjs)− P(sks) 6 |s|2Z − P(sks) 7 |s|2Z − P(sks) 8 |s|2Z − P(sks) 2 10γ=0,λ R |s| Z − B(sks) ∈ 2 11χ=0 |s| Z − B(sis)− P(sks) 13 − Re(βπ)(H + Z) |s|2H 23 |s|2Z − P(sks) 24 |s|2Z − P(sks) 29 |s|2Z − P(sks) 30 |s|2Z − P(sks) 34 |s|2Z − P(sks) 36 − |s|2Z − P(sks) 37 − |s|2Z − P(sks)

The first column is our identifier for s, whereas the other two columns are the possible central terms in the central extension s. Here β, π Im H and s H are (some of) the parameters ∈ ∈ defining the Lie brackets in the quaternionic formalism explained in Section 2.1.5. b

4.1.6 Automorphisms of Kinematical Lie Superalgebras In the next section, we will classify the homogeneous superspaces associated to the kinematical Lie superalgebras. As we will explain below, the first stage is to classify “super Lie pairs” up to isomorphism. To that end, it behoves us to determine the group of automorphisms of the Lie superalgebras in Table 4.3, to which we now turn.

Without loss of generality, we can restrict to automorphisms which are the identity when restricted to r: we call them r-fixing automorphisms. Following from our discussion in Sec- tion 4.1.1, these are parametrised by triples

a b A := , µ, q GL(2, R) R× H× (4.1.6.1) c d ∈ × ×     subject to the condition that the associated linear transformations leave the Lie brackets in s unchanged.

96 It is easy to read off from equation (4.1.1.33) what (A, µ, q) must satisfy for the r-equivariant linear transformation Φ :s s defined by them to be an automorphism of s, namely: → hq = µqh qc1q = c1 bq = q(ab + cp) qc2q = ac2 + bc3 (4.1.6.2) pq = q(bb + dp) 2 qc3q = cc2 + dc3. µc0 = |q| c0

It is then a straightforward – albeit lengthy – process to go through each Lie superalgebra in Table 4.3 and solve equations (4.1.6.2) for (A, µ, q). In particular, (A, µ) Autr(k) and they ∈ are given in Table 4.2. The results of this section are summarised in Tables 4.7 and 4.8, which list the r-fixing automorphisms for the Lie superalgebras S1- S15 and S16-S35, respectively, in Table 4.3.

The first six Lie superalgebras in Table 4.3 are supersymmetric extensions of the static kine- matical Lie algebra for which (A, µ) can be any element in GL(2, R) R×. ×

Automorphisms of Lie Superalgebra S1

1 Here h = 2 k, b = p = 0, c0 = 0, c1 = c2 = 0 and c3 = k. The invariance conditions (4.1.6.2) give µqk = kq, bk = 0 and dk = qkq. (4.1.6.3)

The second equation requires b = 0. The third equation says that the real linear map αq : H H → defined by αq(x)= qxq preserves the k-axis in Im H.

Lemma 4.1.3. Let qkq = dq for some d R. Then either d = |q|2 and q span {1, k} or ∈ ∈ R d =−|q|2 and q span {i, j}. ∈ R Proof. Taking the quaternion norm of both sides of the equation qkq = dq, and using that q = 0, we see that d = |q|2 and hence right multiplying by q, the equation becomes kq = qk. 6 ± ± If kq = qk, then q span {1, k} and d = |q|2, whereas if −kq = qk, then q span {i, j} and ∈ R ∈ R d =−|q|2.

Taking the quaternion norm of the first equation shows that µ = 1 and hence that d = µ|q|2. ± In summary, we have that the typical automorphism (A, µ, q) takes one of two possible forms:

a 0 A = , µ = 1 and q = q + q k c |q|2 4 3   (4.1.6.4) a 0 or A = , µ =−1 and q = q i + q j. c −|q|2 1 2  

Automorphisms of Lie Superalgebra S2

Here h = b = p = 0, c0 = 1, c1 = 0, c2 = j and c3 = k. The invariance conditions (4.1.6.2) give

µ = |q|2, aj + bk = qjq and cj + dk = qkq. (4.1.6.5)

The last two equations say that the real linear map αq : H H defined earlier preserves the → (j, k)-plane in Im H.

Lemma 4.1.4. The map αq : H H preserves the (j, k)-plane in Im H if and only if q → ∈ span {1, i} span {j, k}. R ∪ R 2 Proof. Since q = 0, we can write it as q = |q|u, for some unique u Sp(1) and αq = |q| αu. 6 ∈ The map αq preserves separately the real and imaginary subspaces of H, and αq preserves the (j, k)-plane if and only if αu does. But, for u Sp(1), αu acts on Im H by rotations and hence ∈ if αu preserves (j, k)-plane, it also preserves the perpendicular line, which, in this case, is the 2 i-axis. Additionally, since it must preserve length, αu(i)= i. It follows that αq(i)= |q| i, so ± ±

97 that αq too preserves the i-axis. By an argument similar to that of Lemma 4.1.3 it follows that q belongs either to the complex line in H generated by i or to its perpendicular complement.

From the lemma, we have two cases to consider: q = q4 + q1i or q = q2j + q3k. In each case, we can use the last two equations to solve for a, b, c, d in terms of the components of q. Summarising, we have that the typical automorphism (A, µ, q) takes one of two possible forms:

2 2 q4 − q1 2q1q4 2 2 A = 2 2 , µ = q1 + q4 and q = q4 + q1i −2q1q4 q4 − q1   (4.1.6.6) q2 − q2 2q q or A = 2 3 2 3 , µ = q2 + q2 and q = q j + q k. 2q q q2 − q2 2 3 2 3  2 3 3 2

Automorphisms of Lie Superalgebra S3

Here h = b = p = 0, c0 = 1, c1 = c2 = 0 and c3 = k. The invariance conditions (4.1.6.2) give

µ = |q|2, bk = 0 and dk = qkq. (4.1.6.7)

This is very similar to the case of the Lie superalgebra S1; in particular, Lemma 4.1.3 applies. The typical automorphism (A, µ, q) takes one of two possible forms:

a 0 A = , µ = |q|2 and q = q + q k c |q|2 4 3   (4.1.6.8) a 0 or A = , µ = |q|2 and q = q i + q j. c −|q|2 1 2  

Automorphisms of Lie Superalgebra S4

2 Here, h = b = p = 0, c0 = 1 and c1 = c2 = c3 = 0. The only condition is µ = |q| . Hence the typical automorphism (A, µ, q) takes the form

a b 2 A = , µ = |q| and q H×. (4.1.6.9) c d ∈  

Automorphisms of Lie Superalgebra S5

Here, h = b = p = 0, c0 = 0, c1 = 0, c2 = j and c3 = k. The invariance conditions (4.1.6.2) are here as for the Lie superalgebra S2, except that µ is unconstrained. In other words, the typical automorphism (A, µ, q) takes one of two possible forms:

2 2 q4 − q1 2q1q4 A = 2 2 , µ and q = q4 + q1i −2q1q4 q4 − q1   (4.1.6.10) q2 − q2 2q q or A = 2 3 2 3 , µ and q = q j + q k. 2q q q2 − q2 2 3  2 3 3 2

Automorphisms of Lie Superalgebra S6

Here, h = b = p = 0, c0 = 0, c1 = c2 = 0 and c3 = k. This is similar to Lie superalgebra S3, except that µ remains unconstrained. In summary, the typical automorphisms (A, µ, q) takes one of two possible forms:

a 0 A = , µ and q = q + q k c |q|2 4 3   (4.1.6.11) a 0 or A = , µ and q = q i + q j. c −|q|2 1 2  

98 The next two Lie superalgebras (S7 and S8) are supersymmetric extensions of the Galilean Lie algebra, where (A, µ) take the form

a b d A = and µ = . (4.1.6.12) c d a  

Automorphisms of Lie Superalgebra S7

Here, h = k, b = p = 0, c0 = 0, c1 = c2 = 0 and c3 = k. The invariance conditions (4.1.6.2) are

dqk = akq and dk = qkq. (4.1.6.13)

Multiplying the second equation on the right by q, using the first equation and the fact that q = 0, results in a = d2/|q|2, so that a > 0. Taking the quaternion norm of the first equation 6 shows that a = |d|, so that a = |q|2. The first equation now follows from the second, and that is solved by Lemma 4.1.3.

In summary the typical automorphism (A, µ, q) takes one of two possible forms:

|q|2 0 A = , µ = 1 and q = q + q k c |q|2 4 3   (4.1.6.14) |q|2 0 or A = , µ =−1 and q = q i + q j. c −|q|2 1 2  

Automorphisms of Lie Superalgebra S8

Here, h = b = p = 0, c0 = 0, c1 = c2 = 0 and c3 = k. The invariance conditions (4.1.6.2) reduce to just dk = qkq, which we solve by Lemma 4.1.3. In summary, the typical automorphism (A, µ, q) is the same here as in the previous Lie superalgebra, except that a is unconstrained (but non-zero). It can thus take one of two possible forms:

a 0 |q|2 A = , µ = and q = q + q k c |q|2 a 4 3   (4.1.6.15) a 0 |q|2 or A = , µ =− and q = q i + q j. c −|q|2 a 1 2   The next two classes of Lie superalgebras are associated with the one-parameter family of kinematical Lie algebras K3γ, whose typical automorphisms (A, µ) depend on the value of γ [−1,1]. In the interior of the interval, it takes the form ∈ a 0 A = and µ = 1. (4.1.6.16) 0 d   At the boundaries, this is enhanced: at γ =−1, one can also have automorphisms of the form

0 b A = and µ =−1, (4.1.6.17) c 0   whereas, at γ = 1, the typical automorphism takes the form

a b A = and µ = 1. (4.1.6.18) c d  

Automorphisms of Lie Superalgebra S9γ,λ 1 Here, h = 2 (1 + λk), b = p = 0, c0 = 0, c1 = c2 = 0 and c3 = k. The invariance condi- tions (4.1.6.2) reduce to b = 0 and, in addition,

µq(1 + λk)=(1 + λk)q and dk = qkq. (4.1.6.19)

99 Taking the norm of the first equation, we find that µ = 1. If µ = 1, then λ[k, q] = 0 so that ± either λ = 0, in which case q span {1, k}, or λ = 0, and q is not constrained by this equation. 6 ∈ R The second equation is dealt with by Lemma 4.1.3, which implies that d = |q|2, and since ± q = 0, d = 0. This precludes the case µ =−1 by inspecting the possible automorphisms (A, µ) 6 6 of k. In summary, for generic γ and λ, the typical automorphism (A, µ, q) takes the form

a 0 A = , µ = 1 and q = q + q k, (4.1.6.20) 0 |q|2 4 3   which is enhanced for γ = 1 (but λ still generic) to

a 0 A = , µ = 1 and q = q + q k. (4.1.6.21) c |q|2 4 3   If λ = 0, then the automorphisms are enhanced by the addition of (A, µ, q) of the form

a 0 A = , µ = 1 and q = q i + q j, (4.1.6.22) 0 −|q|2 1 2   for generic γ or, for γ = 1 only, also

a 0 A = , µ = 1 and q = q i + q j. (4.1.6.23) c −|q|2 1 2  

Automorphisms of Lie Superalgebra S10γ,λ

1 Here, h = 2 (γ + λk), b = p = 0, c0 = 0, c1 = c3 = 0 and c2 = k. The invariance condi- tions (4.1.6.2) imply that c = 0 and also

µq(γ + λk)=(γ + λk)q and ak = qkq. (4.1.6.24)

It is very similar to the previous Lie superalgebra, except here γ = 1. Lemma 4.1.3 now says 2 2 6 that either a = |q| and q = q4 + q3k or a =−|q| and q = q1i + q2j. In particular, since q = 0, 6 a = 0. From the expressions for the automorphisms (A, µ) of k, we see that µ = 1. This means 6 that the first equation says q commutes with γ + λk. If λ = 0, this condition is vacuous, but if 2 λ = 0, then it forces q = q4 + q3k and hence a = |q| . 6 In summary, for λ = 0, we have that (A, µ, q) takes the form 6 |q|2 0 A = , µ = 1 and q = q + q k, (4.1.6.25) 0 d 4 3   whereas, if λ = 0, it can also take the form

−|q|2 0 A = , µ = 1 and q = q i + q j. (4.1.6.26) 0 d 1 2  

The next Lie superalgebra is based on the kinematical Lie algebra K4χ, whose automorphisms (A, µ) take the form a b A = and µ = 1 (4.1.6.27) −b a   for generic χ, whereas, if χ = 0, then they can also be of the form

a b A = and µ =−1. (4.1.6.28) b −a  

100 Automorphisms of Lie Superalgebra S11χ

1 Here, h = 2 (χ + j), b = p = 0, c0 = 0, c1 = 0, c2 = i and c3 = k. The invariance condi- tions (4.1.6.2) reduce to

µq(χ + j)=(χ + j)q, qiq = ai + bk and qkq = ci + dk. (4.1.6.29)

The last two equations are solved via Lemma 4.1.4: either q = q4 + q2j or else q = q1i + q3k. This latter case can only happen when χ = 0. Substituting these possible expressions for q in the last two equations, we determine the entries of the matrix A.

In summary, (A, µ, q) takes the form

q2 − q2 −2q q A = 4 2 2 4 , µ = 1 and q = q + q j, (4.1.6.30) 2q q q2 − q2 4 2  2 4 4 2  and, (only) if χ = 0, it can also take the form

q2 − q2 2q q A = 1 3 1 3 , µ =−1 and q = q i + q k. (4.1.6.31) 2q q q2 − q2 1 3  1 3 3 1 The next Lie superalgebra is the supersymmetric extension of the kinematical Lie algebra K5, whose automorphisms (A, µ) are of the form

a 0 A = and µ = 1. (4.1.6.32) c a  

Automorphisms of Lie Superalgebra S12λ

1 Here, h = 2 (1 + λk), b = p = 0, c0 = 0, c1 = c2 = 0 and c3 = k. The invariance condi- tions (4.1.6.2) reduce to qh = hq and ak = qkq. (4.1.6.33) 2 The second equation is solved via Lemma 4.1.3, which says that either a = |q| and q = q4 +q3k 2 or a =−|q| and q = q1i + q2j. The first equation is identically satisfied if λ = 0, but otherwise 2 it forces q = q4 + q3k and hence a = |q| . In summary, for general λ, an automorphism (A, µ, q) takes the form |q|2 0 A = , µ = 1 and q = q + q k, (4.1.6.34) c |q|2 4 3   whereas if λ = 0, it can also take the form

−|q|2 0 A = , µ = 1 and q = q i + q j. (4.1.6.35) c −|q|2 1 2   The next Lie superalgebra is the supersymmetric extension of the Carroll algebra, whose auto- morphisms (A, µ) take the form

a b A = and µ = ad − bc. (4.1.6.36) c d  

Automorphisms of Lie Superalgebra S13

Here h = b = p = 0, c0 = 1 and c1 = c2 = c3 = 0. The invariance conditions (4.1.6.2) reduce to a single condition: ad − bc = |q|2. The automorphisms (A, µ, q) are of the form

a b 2 A = , µ = ad − bc = |q| and q H×. (4.1.6.37) c d ∈  

101 The next Lie superalgebra is the Poincaré superalgebra whose (r-fixing) automorphisms (A, µ) can take one of two possible forms:

1 0 A = and µ = d c d   (4.1.6.38) −1 0 or A = and µ =−d. c d  

Automorphisms of Lie Superalgebra S14

1 Here h = p = 0, b = 2 k, c0 = 1, c1 = c2 = 0 and c3 = k. The invariance conditions (4.1.6.2) translate into qk = kq, d = |q|2 and dk = qkq, (4.1.6.39) ± ± where the signs are correlated and the last equation follows from the first two.

2 Choosing the plus sign, qk = kq, so that q = q4 + q3k and d = |q| , whereas choosing the 2 minus sign, qk =−kq, so that q = q1i + q2j and d =−|q| .

In summary, automorphisms (A, µ, q) of the Poincaré superalgebra take the form

1 0 A = , µ = |q|2 and q = q + q k c |q|2 4 3   (4.1.6.40) −1 0 or A = , µ = |q|2 and q = q i + q j. c −|q|2 1 2   The next Lie superalgebra is the AdS superalgebra, whose (r-fixing) automorphisms (A, µ) are of the form a b A = and µ = 1, (4.1.6.41) Pb a ±  ±  where a2 + b2 = 1.

Automorphisms of Lie Superalgebra S15

1 1 1 Here h = 2 k, b = 2 i, p = 2 j, c0 = 1, c1 = k, c2 = j and c3 = i. The invariance condi- tions (4.1.6.2) include µ = |q|2, which forces µ = 1. Taking this into account, another of the invariance conditions (4.1.6.2) is qk = kq, which together with |q| = 1, forces q = eθk. The remaining invariance conditions are

aqi − bqj = iq, bqi + aqj = jq, ajq + biq = qj and aiq − bjq = qi. (4.1.6.42)

Given the expression for q, these are solved by a = cos2θ and b = sin2θ. In summary, the (r-fixing) automorphisms (A, µ, q) of the AdS superalgebra are of the form

cos2θ sin2θ A = , µ = 1 and q = eθk. (4.1.6.43) − sin2θ cos2θ   The next three Lie superalgebras in Table 4.3 are supersymmetric extensions of the kinematical Lie algebra K12 in Table 4.1, whose r-fixing automorphisms (A, µ) take the following form:

1 0 A = and µ R×. (4.1.6.44) 0 1 ∈  ± 

Automorphisms of Lie Superalgebra S16

1 Here h = b = 0, p = 2 j, c0 = 0, c1 =−i, c2 = i and c3 =−k. The invariance conditions (4.1.6.2) reduce to qj = jq, qk = kq and qiq = i. (4.1.6.45) ± ±

102 It follows from the last equation that |q| = 1 and hence that qi = iq. Depending on the (correlated) signs of the first two equations, we find that, for the plus sign, q commutes with i, j and k and hence q R, but since |q| = 1, we must have q = 1. For the minus sign, we ∈ ± find that q commutes with i but anticommutes with j and k, so that q = i, after taking into ± account that |q| = 1. In summary, the automorphisms (A, µ, q) of this Lie superalgebra take one of two possible forms:

1 0 A = , µ R× and q = 1 0 1 ∈ ±   (4.1.6.46) 1 0 or A = , µ R× and q = i. 0 −1 ∈ ±  

Automorphisms of Lie Superalgebra S17

1 Here h = p = 0, b = 2 , c0 = 1 and c1 = c2 = c3 = 0. There is only one invariance condition: namely, µ = |q|2, and hence the automorphisms (A, µ, q) take the form

1 0 2 A = , µ = |q| and q H×. (4.1.6.47) 0 1 ∈  ± 

Automorphisms of Lie Superalgebra S18

1 1 Here h = 2 k, b = 2 , p = 0, c0 = 1, c1 = c3 = 0 and c2 = k. The invariance conditions (4.1.6.2) reduce to µqk = kq, µ = |q|2 and k = qkq. (4.1.6.48) From the first equation we see that µ = 1, but from the second it must be positive, so µ = 1, ± which says implies that |q| = 1 and hence that q commutes with k. In summary, the typical automorphism (A, µ, q) takes the form

1 0 A = , µ = 1 and q = eθk. (4.1.6.49) 0 1  ±  The next four Lie superalgebras in Table 4.3 are supersymmetric extensions of the kinematical Lie algebra K13 in Table 4.1, whose typical r-fixing automorphisms (A, µ) take the form

1 0 A = and µ R×. (4.1.6.50) 0 1 ∈  ± 

Automorphisms of Lie Superalgebra S19

1 Here h = k, b = 0, p = 2 , c0 = 1, c1 = c3 = k and c2 =−k. The invariance conditions (4.1.6.2) are given by µqk = kq, µ = |q|2 and dq = q. (4.1.6.51) The last equation says that d = 1, whereas the first says that µ = 1, but from the second ± equation it is positive and thus µ = 1. This also means |q| = 1 and that qk = kq. In summary, the typical automorphism (A, µ, q) of s takes the form

1 0 A = , µ = 1 and q = eθk. (4.1.6.52) 0 1  

Automorphisms of Lie Superalgebra S20

1 Here h = b = 0, p = 2 , c0 = 1, c1 = c2 = c3 = 0. The invariance conditions (4.1.6.2) are given by dq = q and µ = |q|2. (4.1.6.53)

103 The first equation simply sets d = 1 and, in summary, the typical automorphism of s is takes the form 1 0 2 A = , µ = |q| and q H×. (4.1.6.54) 0 1 ∈  

Automorphisms of Lie Superalgebra S21

1 2 Here h = p = 0, b = 2 , c0 = 1 and c1 = c2 = c3 = 0. The only invariance condition is µ = |q| , so that the typical automorphism (A, µ, q) takes the form

1 0 2 A = , µ = |q| and q H×. (4.1.6.55) 0 1 ∈  ± 

Automorphisms of Lie Superalgebra S22

1 1 Here h = 2 k, b = 2 , p = 0, c0 = 1, c1 = c3 = 0 and c2 = k. The invariance conditions (4.1.6.2) reduce to µqk = kq and µ = |q|2. (4.1.6.56) The first equation says that µ = 1, but the second equation says it is positive, so that µ = 1 ± and |q| = 1. Furthermore, q commutes with k, so that q = eθk. In summary, the typical automorphism (A, µ, q) takes the form

1 0 A = , µ = 1 and q = eθk. (4.1.6.57) 0 1  ±  The next six Lie superalgebras in Table 4.3 are supersymmetric extensions of the kinematical Lie algebra K14 in Table 4.1, whose r-fixing automorphisms (A, µ) take the form

1 0 A = and µ R×. (4.1.6.58) 0 d ∈  

Automorphisms of Lie Superalgebra S23

Here h = k, b = p = 0, c0 = 0, c1 = c2 = 0 and c3 = k. The invariance conditions (4.1.6.2) reduce to µqk = kq and dk = qkq. (4.1.6.59) The first equation says that µ = 1, so that qk = kq. The second equation follows from 2 ± ± 2 Lemma 4.1.3: either d = |q| and hence q = q4 + q3k or d =−|q| and hence q = q1i + q2j. In summary, the typical automorphism (A, µ, q) takes one of two possible forms:

1 0 A = , µ = 1 and q = q + q k 0 |q|2 4 3   (4.1.6.60) 1 0 or A = , µ =−1 and q = q i + q j. 0 −|q|2 1 2  

Automorphisms of Lie Superalgebra S24

Here h = b = p = 0, c0 = 0, c1 = c2 = 0 and c3 = k. Hence the only invariance condition is 2 2 dk = qkq. Lemma 4.1.3 says that either d = |q| and hence q = q4 + q3k or else d =−|q| and hence q = q1i + q2j. In summary, the typical automorphism (A, µ, q) takes one of two possible forms: 1 0 A = , µ R× and q = q4 + q3k 0 |q|2 ∈   (4.1.6.61) 1 0 or A = , µ R× and q = q1i + q2j. 0 −|q|2 ∈  

104 Automorphisms of Lie Superalgebra S25

Here h = b = p = 0, c0 = 1 and c1 = c2 = c3 = 0, so that the only invariance condition is µ = |q|2. In summary, the typical automorphism (A, µ, q) takes the form

1 0 2 A = , µ = |q| and q H×. (4.1.6.62) 0 d ∈  

Automorphisms of Lie Superalgebra S26

Here h = b = p = 0, c0 = 1, c1 = c2 = 0 and c3 = k, so that there are two conditions in (4.1.6.2): µ = |q|2 and dk = qkq. (4.1.6.63) 2 The second equation can be solved via Lemma 4.1.3: either d = |q| and q = q4 + q3k or 2 d =−|q| and q = q1i + q2j. In summary, the automorphisms (A, µ, q) take one of two possible forms: 1 0 A = , µ = |q|2 and q = q + q k 0 |q|2 4 3   (4.1.6.64) 1 0 or A = , µ = |q|2 and q = q i + q j. 0 −|q|2 1 2  

Automorphisms of Lie Superalgebra S27

1 Here h = p = 0, b = 2 , c0 = 1 and c1 = c2 = c3 = 0, so that the only invariance condition is µ = |q|2. Therefore the typical automorphism (A, µ, q) takes the form

1 0 2 A = , µ = |q| and q H×. (4.1.6.65) 0 d ∈  

Automorphisms of Lie Superalgebra S28

1 Here h = b = 2 , p = 0, c0 = 1, c1 = c3 = 0 and c2 = k. The invariance conditions (4.1.6.2) reduce to the following: µ = |q|2, µq = q and k = qkq. (4.1.6.66) From the second equation we see that µ = 1, so that from the first |q| = 1 and hence kq = qk, so that q = eθk. In summary, the typical automorphism (A, µ, p) takes the form

1 0 A = , µ = 1 and q = eθk. (4.1.6.67) 0 d   The next four Lie superalgebras in Table 4.3 are supersymmetric extensions of the kinematical Lie algebra K15 in Table 4.1, whose r-fixing automorphisms (A, µ) take the form

a 0 A = and µ R×. (4.1.6.68) c a2 ∈  

Automorphisms of Lie Superalgebra S29

Here b = p = 0, h = k, c0 = 0, c1 = c2 = 0 and c3 = k. The invariance conditions (4.1.6.2) result in µqk = kq and a2k = qkq. (4.1.6.69) Taking the norm of the first equation, we see that µ = 1, and of the second equation, a2 = |q|2. ± This then says that q commutes with k, so that µ = 1 and q = q4 + q3k. In summary, the typical automorphism (A, µ, q) takes the form

|q| 0 A = ± , µ = 1 and q = q4 + q3k. (4.1.6.70) c |q|2  

105 Automorphisms of Lie Superalgebra S30

Here h = b = p = 0, c0 = 0, c1 = c2 = 0 and c3 = k. The only invariance condition is 2 2 2 a k = qkq. Taking the norm, a = |q| and hence kq = qk and thus q = q4 + q3k. Hence the typical automorphism (A, µ, q) takes the form

|q| 0 A = ± , µ R× and q = q4 + q3k. (4.1.6.71) c |q|2 ∈  

Automorphisms of Lie Superalgebra S31

Here h = b = p = 0, c0 = 1 and c1 = c2 = c3 = 0, so that the only invariance condition is µ = |q|2. In summary, the typical automorphism (A, µ, q) takes the form

a 0 2 A = , µ = |q| and q H×. (4.1.6.72) c a2 ∈  

Automorphisms of Lie Superalgebra S32

Here h = b = p = 0, c0 = 1, c1 = c2 = 0 and c3 = k, so that there are two invariance conditions:

µ = |q|2 and a2k = qkq. (4.1.6.73)

2 2 The second shows that a = |q| and hence q commutes with k, so that q = q4 + q3k. In summary, the typical automorphism (A, µ, q) takes the form

|q| 0 2 A = ± , µ = |q| and q = q4 + q3k. (4.1.6.74) c |q|2   The next Lie superalgebra in Table 4.3 is a one-parameter family of supersymmetric extensions of the kinematical Lie algebra K16 in Table 4.1, whose r-fixing automorphisms (A, µ) take the form 1 0 A = and µ = 1. (4.1.6.75) 0 d  

Automorphisms of Lie Superalgebra S33

1 Here h = 2 (1 + λk), b = p = 0, c0 = 0, c1 = c2 = 0 and c3 = k. There are two invariance conditions: q(1 + λk)=(1 + λk)q and dk = qkq. (4.1.6.76) For the second equation we use Lemma 4.1.3 and for the first equation we must distinguish 2 between λ = 0 and λ = 0. In the latter case, we have that q = q4 + q3k so that only the d = |q| 6 of the lemma survives. If λ = 0, both branches survive. In summary, for λ = 0, the typical 6 automorphism (A, µ, q) takes the form

1 0 A = , µ = 1 and q = q + q k, (4.1.6.77) 0 |q|2 4 3   whereas if λ = 0 we have additional automorphisms of the form

1 0 A = , µ = 1 and q = q i + q j. (4.1.6.78) 0 −|q|2 1 2   The next Lie superalgebra in Table 4.3 is the supersymmetric extension of the kinematical Lie algebra K17 in Table 4.1, whose r-fixing automorphisms (A, µ) take the form

a 0 A = and µ = a. (4.1.6.79) c a2  

106 Automorphisms of Lie Superalgebra S34

1 Here h = 2 k, b = p = 0, c0 = 0, c1 = c2 = 0 and c3 = k. The invariance conditions are

aqk = kq and a2k = qkq. (4.1.6.80)

Taking norms of the first equation gives a = 1 and hence qk = kq and of the second equation ± ± a2 = |q|2 and hence qk = kq. This shows that a = 1 and hence |q| = 1, so that q = eθk. In summary, the typical automorphism (A, µ, q) takes the form

1 0 A = , µ = 1 and q = eθk. (4.1.6.81) c 1   The last Lie superalgebra in Table 4.3 is a one-parameter family of supersymmetric extensions of the kinematical Lie algebra K18 in Table 4.1, whose r-fixing automorphisms (A, µ) take the form a 0 A = and µ = 1. (4.1.6.82) 0 a2  

Automorphisms of Lie Superalgebra S35

Here h = 1 + λk, b = p = 0, c0 = 0, c1 = c2 = 0 and c3 = k. The invariance conditions (4.1.6.2) reduce to q(1 + λk)=(1 + λk)q and a2k = qkq. (4.1.6.83) 2 2 Taking the norm of the second equation, a = |q| so that qk = kq and hence q = q4 + q3k. This also solves the first equation, independently of the value of λ. In summary, the typical automorphism (A, µ, q) takes the form

|q| 0 A = ± , µ = 1 and q = q4 + q3k. (4.1.6.84) 0 |q|2  

Summary

Tables 4.7 and 4.8 summarise the above discussion and lists the typical automorphisms of each of the Lie superalgebras in Table 4.3.

4.2 Classification of Kinematical Superspaces

Now that we have a complete classification of the possible N = 1 kinematical Lie superalgebras in three spatial dimensions, we can turn to the corresponding kinematical superspace classification. As discussed in Section 2.2.6, each homogeneous superisation of a kinematical spacetime K/H, corresponds to a unique effective super Lie pair (s, h), where s=s¯ s¯, with s¯ = k the 0 ⊕ 1 0 kinematical Lie algebra associated with K, and h s¯ = k is the Lie subalgebra associated ⊂ 0 with H. To establish the possible effective super Lie pairs, this section runs as follows. In Section 4.2.1, we will use the automorphisms derived in Section 4.1.6 to identify the admissible super Lie pairs for each KLSA. Then, in Section 4.2.2, we determine which of the admissible Lie pairs are effective. We then give a brief account of the possible Aristotelian homogeneous superspaces in Section 4.2.3 before summarising our findings in Section 4.2.4. In Section 4.2.5, we end by discussing the low-rank invariants of the kinematical superspaces. Notice the method used in determining the kinematical superspaces is a direct generalisation of the classification method used in the kinematical spacetime case: we find the admissible Lie subalgebras and then restrict ourselves to those Lie pairs which are effective. Geometric realisability does not need to be studied in this instance since we assume the underlying kinematical spacetime is geometrically realisable. This property then trivially extends to the superised geometry.

107 Table 4.7: Automorphisms of Kinematical Lie Superalgebras

S# Typical (A, µ, q) GL(2, R) R× H× ∈ × × a 0 a 0 1 , 1, q + q k , , −1, q i + q j c |q|2 4 3 c −|q|2 1 2       q2 − q2 2q q q2 − q2 2q q 2 4 1 1 4 , q2 + q2, q + q i , 2 3 2 3 , q2 + q2, q j + q k −2q q q2 − q2 1 4 4 1 2q q q2 − q2 2 3 2 3  1 4 4 1   2 3 3 2  a 0 a 0 3 , |q|2, q + q k , , |q|2, q i + q j c |q|2 4 3 c −|q|2 1 2       a b 4 , |q|2, q c d    q2 − q2 2q q q2 − q2 2q q 5 4 1 1 4 , µ, q + q i , 2 3 2 3 , µ, q j + q k −2q q q2 − q2 4 1 2q q q2 − q2 2 3  1 4 4 1   2 3 3 2  a 0 a 0 6 , µ, q + q k , , µ, q i + q j c |q|2 4 3 c −|q|2 1 2       |q|2 0 |q|2 0 7 , 1, q + q k , , −1, q i + q j c |q|2 4 3 c −|q|2 1 2       a 0 | |2 a 0 | |2 8 , q , q + q k , , − q , q i + q j c |q|2 a 4 3 c −|q|2 a 1 2       a 0 9γ=1,λ=0 2 , 1, q4 + q3k 6 6 0 |q|    a 0 9γ=1,λ=0 2 , 1, q4 + q3k 6 c |q|    a 0 a 0 9γ=1,λ=0 2 , 1, q4 + q3k , 2 , 1, q1i + q2j 6 0 |q| 0 −|q|       a 0 a 0 9 , 1, q + q k , , 1, q i + q j γ=1,λ=0 c |q|2 4 3 c −|q|2 1 2       |q|2 0 10γ,λ=0 , 1, q4 + q3k 6 0 d    |q|2 0 −|q|2 0 10 , 1, q + q k , , 1, q i + q j γ,λ=0 0 d 4 3 0 d 1 2       q2 − q2 −2q q 11 4 2 2 4 , 1, q + q j χ>0 2q q q2 − q2 4 2  2 4 4 2  q2 − q2 −2q q q2 − q2 2q q 11 4 2 2 4 , 1, q + q j , 1 3 1 3 , −1, q i + q k χ=0 2q q q2 − q2 4 2 2q q q2 − q2 1 3  2 4 4 2   1 3 3 1  |q|2 0 12λ=0 2 , 1, q4 + q3k 6 c |q|    |q|2 0 −|q|2 0 12 , 1, q + q k , , 1, q i + q j λ=0 c |q|2 4 3 c −|q|2 1 2       a b 13 , ad − bc = |q|2, q c d    1 0 −1 0 14 , |q|2, q + q k , , |q|2, q i + q j c |q|2 4 3 c −|q|2 1 2       cos2θ − sin2θ 15 , 1, eθk sin2θ cos2θ   

108 Table 4.8: Automorphisms of Kinematical Lie Superalgebras (continued)

S# Typical (A, µ, q) GL(2, R) R× H× ∈ × × 1 0 1 0 16 , µ, 1 , , µ, i 0 1 ± 0 −1 ±       1 0 17 , |q|2, q 0 1  ±   1 0 18 , 1, eθk 0 1  ±   1 0 19 , 1, eθk 0 1    1 0 20 , |q|2, q 0 1    1 0 21 , |q|2, q 0 1  ±   1 0 22 , 1, eθk 0 1  ±   1 0 1 0 23 , 1, q + q k , , −1, q i + q j 0 |q|2 4 3 0 −|q|2 1 2       1 0 1 0 24 , µ, q + q k , , µ, q i + q j 0 |q|2 4 3 0 −|q|2 1 2       1 0 25 , |q|2, q 0 d    1 0 1 0 26 , |q|2, q + q k , , |q|2, q i + q j 0 |q|2 4 3 0 −|q|2 1 2       1 0 27 , |q|2, q 0 d    1 0 28 , 1, eθk 0 d    |q| 0 29 ± , 1, q4 + q3k c |q|2    |q| 0 30 ± , µ, q4 + q3k c |q|2    a 0 31 , |q|2, q c a2    |q| 0 2 32 ± , |q| , q4 + q3k c |q|2    1 0 33λ=0 2 , 1, q4 + q3k 6 0 |q|    1 0 1 0 33 , 1, q + q k , , 1, q i + q j λ=0 0 |q|2 4 3 0 −|q|2 1 2       1 0 34 , 1, eθk c 1    |q| 0 35λ ± , 1, q4 + q3k 0 |q|2   

4.2.1 Admissible Super Lie Pairs We are now ready to classify the admissible super Lie pairs, up to isomorphism. We recall these are pairs (s, h), where s is one of the kinematical Lie superalgebras in Table 4.3 and h is a Lie

109 subalgebra h k=s¯ which is admissible in the sense of Section 2.2.5; that is, it contains the ⊂ 0 rotational subalgebra r and, as an r module, h=r V where V k is a copy of the vector mod- ⊕ ⊂ ule. Two super Lie pairs (s, h) and (s, h′) are isomorphic if there is an automorphism of s which maps h (isomorphically) to h′. As in Section 3.2, our strategy in classifying admissible super Lie pairs up to isomorphism will be to take each kinematical Lie superalgebra s in Table 4.3 in turn, determine the admissible subalgebras h and study the action of the automorphisms in Tables 4.7 and 4.8 on the space of admissible subalgebras in order to select one representative from each orbit. In particular, every admissible super Lie pair (s, h) defines a unique admissible Lie pair (k, h) which, if effective and geometrically realisable, is associated with a unique simply- connected kinematical homogeneous spacetime K/H. That being the case, we may think of the super Lie pair (s, h) as a homogeneous kinematical superspacetime which superises K/H.

Without loss of generality – since an admissible subalgebra h contains r – the vectorial com- plement V can be taken to be the span of αB + βP , i = 1,2,3, for some α, β R not both i i ∈ zero, since the spans of {J , αB + βP } and of {J , αB + βP + γJ } coincide for all γ R. i i i i i i i ∈ We will often use the shorthand V = αB + βP. The determination of the possible admissible subalgebras can be found in [5, §§3.1-2], but we cannot simply import the results of that pa- per wholesale because here we are only allowed to act with automorphisms of s and not just of k.

As in that paper, and as discussed in Section 3.2.1, we will eventually change basis in the Lie superalgebra s so that the admissible subalgebra h is spanned by J and B. Hence, in determining the possible super Lie pairs, we will keep track of the required change of basis, ensuring, where possible, that (s, h) is reductive; that is, such that H, Pi,Qa (defined by equa- tion (2.1.5.5)) span a subspace m s complementary to h and such that [h, m] m. This is ⊂ ⊂ equivalent to requiring that the span m¯ of H,P satisfies [h, m¯] m¯, since the Q span s¯ 0 i 0 ⊂ 0 a 1 and [h, s¯] s¯ by virtue of s being a Lie superalgebra. 1 ⊂ 1 It follows by inspection of [5, §§3.1-2] that the Lie superalgebras s whose automorphisms are listed in Table 4.7 are extensions of kinematical Lie algebras k for which any vectorial subspace V = αB + βP defines an admissible subalgebra h=r V k. It is then a simple matter to ⊕ ⊂ determine the orbits of the action of the automorphisms listed in Table 4.7 on the space of vectorial subspaces and hence to arrive at a list of possible inequivalent super Lie pairs (s, h) for such s.

It also follows by inspection of [5, §§3.1-2] that, of the remaining Lie superalgebras (i.e., those whose automorphisms are listed in Table 4.8), most are extensions of kinematical Lie algebras possessing a unique vectorial subspace V for which h=r V is an admissible subalgebra. The ⊕ exceptions are those Lie superalgebras S23–S28 and S33λ, which are extensions of the kinemati- cal Lie algebras K14 and K16, respectively, for which there are precisely two vectorial subspaces leading to admissible subalgebras.

Let us concentrate first on the Lie superalgebras S1–S15, whose automorphisms are listed in Table 4.7. As mentioned above, for V any vectorial subspace, h=r V is an admissible ⊕ subalgebra. We need to determine the orbits of the action of the automorphisms in Table 4.7. Since V = αB + βP, this is equivalent to studying the action of the matrix part A of the automorphism (A, µ, q) on non-zero vectors (α, β) R2. In fact, since (α, β) and (λα, λβ) for ∈ 0 = λ R denote the same vectorial subspace, we must study the action of the subgroup of 6 ∈ GL(2, R) defined by the matrices A in the automorphism group on the projective space RP1. The map (A, µ, q) A defines a group homomorphism from the automorphism group of a Lie 7→ superalgebra s to GL(2, R). We will let A denote the image of this homomorphism: it is a subgroup of GL(2, R) and it is the action of A on RP1 that we need to investigate. Of course, A depends on s, even though we choose not to overload the notation by making this dependence explicit.

It follows by inspection of Table 4.7, that for s any of the Lie superalgebras S2, S4, S5, S11χ>0, S13 and S15, the subgroup A GL(2, R) acts transitively on RP1 and hence for such Lie su- ⊂ peralgebras there is a unique admissible subalgebra spanned by J and B.

110 In contrast, if s is any of the Lie superalgebras S1, S3, S6, S7, S8, S9γ=1,λ R, S12λ R and ∈ ∈ S14, the subgroup A GL(2, R) acts with two orbits on RP1. For example, consider the Lie ⊂ superalgebra S1, for which any A A takes the form ∈ a 0 for some a, c, d R with a, d = 0, (4.2.1.1) c d ∈ 6   and act as α a 0 α aα = . (4.2.1.2) β 7→ c d β dβ + cα         If α = 0, we can choose c =−dβ/α to bring (α, β) to (aα,0), which is projectively equivalent to 6 (1,0). On the other hand, if α = 0, then we cannot change that via automorphisms and hence we have (0, β), which is projectively equivalent to (0,1). In summary, we have two inequivalent admissible subalgebras with vectorial subspaces V = B and V = P. The same result holds for the other Lie superalgebras in this list.

For the cases where V = P we change basis in the Lie superalgebra s so that the admissible sub- algebra h is spanned by J and B. This results in different brackets, which we now proceed to list.

Finally, if s is any of the Lie superalgebras S9γ=1,λ R and S10γ,λ R, the subgroup A GL(2, R) 6 ∈ ∈ ⊂

Table 4.9: Super Lie pairs (with V = P)

S# k brackets h p [Q(s), Q(s)] 1 1 2 k −B(sks) 3 |s|2H − B(sks) 6 −B(sks) 7 [H, P]=−B k −B(sks) 8 [H, P]=−B −B(sks) 1 9γ=1,λ R [H, B]= B [H, P]= P (1 + λk) −B(sks) ∈ 2 1 12λ R [H, B]= B [H, P]= B + P (1 + λk) −B(sks) ∈ 2 1 2 14 [H, P]= B [B, P]= H [P, P]=−J 2 k |s| H + B(sks)

acts with three orbits. Indeed, the matrices A A are now diagonal and of the form ∈ a 0 , (4.2.1.3) 0 d   where at least one of a, d can take any non-zero value. If (α, β) is such that α = 0 or β = 0, we cannot alter this via automorphisms and hence projectively we have either (1,0) or (0,1). If αβ = 0, then we can always bring it to (1,1) or (−1, −1) via an automorphism, but these are 6 projectively equivalent. In summary, we have three orbits, corresponding to V = B, V = P and V = B + P.

When V = P, the Lie brackets of S9γ=1,λ R in the new basis are given by 6 ∈ 1 [H, B]= B, [H, P]= γP, [H, Q(s)] = Q( 2 s(1 + λk)) and [Q(s), Q(s)]=−B(sks), (4.2.1.4) and those of S10γ,λ R by ∈ 1 [H, B]= B, [H, P]= γP, [H, Q(s)] = Q( 2 s(γ + λk)) and [Q(s), Q(s)]=−P(sks). (4.2.1.5)

111 On the other hand, when V = B + P, the Lie brackets of S9γ=1,λ R in the new basis are given by 6 ∈ [H, B]=−P [H, Q(s)] = Q( 1 s(1 + λk)) 2 (4.2.1.6) Q Q 1 B P [H, P]= γB +(1 + γ)P [ (s), (s)] = 1−γ (γ (sks)+ (sks)), and those of S10γ,λ R by ∈ [H, B]=−P [H, Q(s)] = Q( 1 s(γ + λk)) 2 (4.2.1.7) Q Q 1 B P [H, P]= γB +(1 + γ)P [ (s), (s)] = γ−1 ( (sks)+ (sks)).

Now we turn to the Lie superalgebras whose automorphisms are listed in Table 4.8. If s is one such Lie superalgebra, not every vectorial subspace leads to an admissible subalgebra. From the results in [5, §§3.1-2], we have that Lie superalgebras S16–S22 admit a unique admissible subalgebra with V = B, whereas for the Lie superalgebras S29–S32, S34 and S35λ R also admit ∈ a unique admissible subalgebra with V = P. Finally, the Lie superalgebras S23–S28 and S33λ R ∈ admit precisely two admissible subalgebras with V = B and V = P, which cannot be related by automorphisms.

Table 4.11 summarises the above results. For each Lie superalgebra s in Table 4.3 it lists

Table 4.10: More super Lie pairs (with V = P)

S# k brackets h p [Q(s), Q(s)] 23 [P, P]= P k −B(sks) 24 [P, P]= P −B(sks) 25 [P, P]= P |s|2H 26 [P, P]= P |s|2H − B(sks) 1 2H 27 [P, P]= P 2 |s| 1 1 2 28 [P, P]= P 2 k 2 |s| H − P(sks) 29 [P, P]= B k −B(sks) 30 [P, P]= B −B(sks) 31 [P, P]= B |s|2H 32 [P, P]= B |s|2H − B(sks) 1 33λ R [H, B]= B [P, P]= P (1 + λk) −B(sks) ∈ 2 1 34 [H, P]=−B [P, P]= B 2 k −B(sks) 35λ R [H, P]= P [H, B]= 2B [P, P]= B 1 + λk −B(sks) ∈ the admissible subalgebras h and hence the possible super Lie pairs (s, h). The notation for h is simply the generators of the vectorial subspace V h, where the span of αB + βP is ⊂ i i abbreviated as αB + βP. The blue entries correspond to effective super Lie pairs, whereas the green and greyed out correspond to non-effective super Lie pairs: the green ones giving rise to Aristotelian superspaces upon quotienting by the ideal b = spanR {B}, as described in Section 2.2.5. In Section 4.1.4, we classified Aristotelian Lie superspaces by classifying their corresponding Aristotelian Lie superalgebras (see Table 4.5) and in Section 4.2.3 we exhibit the precise correspondence between the Aristotelian non-effective super Lie pairs and the Aris- totelian superspaces (see Table 4.12).

4.2.2 Effective Super Lie Pairs Recall that a super Lie pair (s, h) is said to be effective if h does not contain an ideal of s. Since h k and contains the rotational subalgebra, which has non-vanishing brackets with Q, ⊂ the only possible ideal of s contained in h would be the vectorial subspace V h. It is then a ⊂ simple matter to inspect the super Lie pairs determined in the previous section and selecting those for which V is not an ideal of s. Those super Lie pairs have been highlighted in blue in

112 Table 4.11: Summary of super Lie pairs

s k V h s k V h s k V h ⊂ ⊂ ⊂ S1 K1 B P S12λ R K5 B P S24 K14 B P ∈ S2 K1 B S13 K6 B S25 K14 B P S3 K1 B P S14 K8 B P S26 K14 B P S4 K1 B S15 K11 B S27 K14 B P S5 K1 B S16 K12 B S28 K14 B P S6 K1 B P S17 K12 B S29 K15 P S7 K2 B P S18 K12 B S30 K15 P S8 K2 B P S19 K13 B S31 K15 P

S9γ [−1,1),λ R K3γ B P B + P S20 K13 B S32 K15 P ∈ ∈ S9γ=1,λ R K3γ=1 B P S21 K13 B S33λ R K16 B P ∈ ∈ S10γ [−1,1),λ R K3γ B P B + P S22 K13 B S34 K17 P ∈ ∈ S11χ>0 K4χ B S23 K14 B P S35λ R K18 P ∈

The blue pairs (e.g., B ) are effective; the green pairs (e.g., B ) though not effective, give rise to Aristotelian superspaces; whereas the greyed out pairs (e.g., B ) are not effective and will not be considered further.

Table 4.11. Additionally, we highlight in green the non-effective super Lie pairs that can give rise to Aristotelian superspaces. Though we could ignore these cases, leaving the identification of Aristotelian superspaces to Section 4.2.3, we identify them here for completeness.

We now take each effective super Lie pair in turn, change basis if needed so that V is spanned by B, and then list the resulting brackets in that basis. Every such super Lie pair (s, h) determines a Lie pair (k, h). If the Lie pair (k, h) is effective (and geometrically realisable), then (s, h) de- scribes a homogeneous superisation of one of the spatially-isotropic homogeneous spacetimes in Table 3.5. We remark that there are effective super Lie pairs (s, h) for which the underlying Lie pair (k, h) is not effective. In those cases, there are no boosts on the body of the superspacetime, but instead there are R-symmetries in the odd coordinates.

As usual, in writing the Lie brackets of s below, we do not include any bracket involving J, which are given in equation (2.1.5.2), and instead give any non-zero additional brackets.

Galilean Superspaces

Galilean spacetime is described by (k, h), where k has the additional bracket [H, B]=−P. There are two possible superisations (s, h), with brackets

Q(sk) [H, Q(s)] = and [Q(s), Q(s)]=−P(sks). (4.2.2.1) 0

These are associated with Lie superalgebras S7 and S8 in Table 4.3.

Galilean de Sitter Superspace

Galilean de Sitter spacetime is described by (k, h), where k has the additional brackets [H, B]= −P and [H, P]=−B. There are two one-parameter families of superisations (s, h), with brackets

[H, Q(s)] = Q( 1 s( 1 + λk)) and [Q(s), Q(s)]=− 1 (B(sks) P(sks)) (4.2.2.2) 2 ± 2 ∓ for λ R. They are associated with Lie superalgebras S9 =−1, and S10 =−1, , respectively. ∈ γ λ γ λ

Torsional Galilean de Sitter Superspaces

Torsional Galilean de Sitter spacetime is described by (k, h), where k has the additional brackets b[H, B]=−P and [H, P] = γB +(1 + γ)P, where γ (−1,1). There are two one-parameter ∈

113 families of superisations (s, h), with brackets

H Q Q 1 Q Q 1 B P [ , (s)] = ( 2 s(1 + λk)) and [ (s), (s)] = 1−γ (γ (sks)+ (sks)) (4.2.2.3) and H Q Q 1 Q Q 1 B P [ , (s)] = ( 2 s(γ + λk)) and [ (s), (s)] = γ−1 ( (sks)+ (sks)) (4.2.2.4) for λ R. The associated Lie superalgebras are S9 , and S10 , , respectively. ∈ γ λ γ λ For γ = 1, with additional brackets [H, B]=−P and [H, P]= B + 2P, there is a one-parameter family of superisations, with brackets

H Q Q 1 Q Q B P [ , (s)] = ( 2 s(1 + λk)) and [ (s), (s)] = (sks)+ (sks). (4.2.2.5)

The associated Lie superalgebras are S12λ.

Galilean Anti-de Sitter Superspace

Galilean anti-de Sitter spacetime is described by (k, h), where k has the additional brackets [H, B]=−P and [H, P]= B. It admits a superisation (s, h), with brackets

H Q Q 1 Q Q B P [ , (s)] = ( 2 sj) and [ (s), (s)]=− (sis)+ (sks), (4.2.2.6) which corresponds to the Lie superalgebra S11χ=0, after changing the sign of P.

Torsional Galilean Anti-de Sitter Superspace

Torsional Galilean anti-de Sitter spacetime is described by (k, h), where k has the additional brackets [H, B]= χB+P and [H, P]= χP−B, where χ> 0. There is a unique superisation (s, h), with brackets

1 [H, Q(s)] = Q( 2 s(χ + j)) and [Q(s), Q(s)]=−B(sis)− P(sks). (4.2.2.7) For uniformity, we change basis so that [H, B]=−P as for all Galilean spacetimes. Then the resulting super Lie pair (s, h) is determined by the brackets [H, B]=−P, [H, P]=(1+χ2)B+2χP and, in addition,

1 [H, Q(s)] = Q( 2 s(χ + j)) and [Q(s), Q(s)] = B(sk(χ + j)s)+ P(sks), (4.2.2.8) corresponding to the Lie superalgebra S11χ.

Carrollian Superspace

Carrollian spacetime is described by (k, h), where k has the additional brackets [B, P] = H. It admits a superisation (s, h), with brackets

[Q(s), Q(s)] = |s|2H, (4.2.2.9) which corresponds to the Lie superalgebra S13.

Minkowski Superspace

Minkowski superspace arises as a superisation of Minkowski spacetime, described by (k, h) with brackets [H, B]=−P, [B, P]= H and [B, B]=−J and, in addition,

1 2 [B(β), Q(s)] = Q( 2 βsk) and [Q(s), Q(s)] = |s| H − P(sks). (4.2.2.10) This is, of course, the Poincaré superalgebra S14.

114 Carrollian Anti-de Sitter Superspace Carrollian anti-de Sitter spacetime is described as (k, h), where the k brackets are given by [H, P]= B, [B, P]= H and [P, P]=−J. It admits a unique superisation (s, h) with brackets (we have rotated k to i)

1 2 [P(π), Q(s)] = Q( 2 πsi) and [Q(s), Q(s)] = |s| H + B(sis). (4.2.2.11) We remark that just as with Carrollian anti-de Sitter and Minkowski spacetimes, which are both homogeneous spacetimes of the Poincaré group, their superisations have isomorphic su- persymmetry algebras: namely, the Poincaré superalgebra S14.

Anti-de Sitter Superspace Anti-de Sitter spacetime is described kinematically as (k, h) with brackets

[H, B]=−P, [H, P]= B, [B, P]= H, [B, B]=−J and [P, P]=−J. (4.2.2.12)

It admits a unique superisation (s, h), with additional brackets (where we have rotated (i, j, k) 7→ (k, i, j) for uniformity)

1 1 1 [H, Q(s)] = Q( 2 sj), [B(β), Q(s)] = Q( 2 βsk), [P(π), Q(s)] = Q( 2 πsi) and [Q(s), Q(s)] = |s|2H + J(sjs)+ B(sis)− P(sks). (4.2.2.13)

The associated Lie superalgebra is S15, which is isomorphic to osp(1|4).

Super-Spacetimes Extending R S3 × These correspond to the effective super Lie pairs associated with the Lie superalgebras S21 and S22. The super Lie pairs (s, h) are effective, but the underlying Lie pair (k, h) is not. Indeed, the brackets of k are now [B, B] = B and [P, P] = J − B, from where we see that B spans an ideal of k; although not one of s, due to the brackets

1 2 [B(β), Q(s)] = Q( 2 βs) and [Q(s), Q(s)] = |s| H, (4.2.2.14) for s the Lie superalgebra S21 or

H Q Q 1 B Q Q 1 Q Q 2H B [ , (s)] = ( 2 sk), [ (β), (s)] = ( 2 βs) and [ (s), (s)] = |s| − (sks), (4.2.2.15) for s the Lie superalgebra S22. In both superspaces, B does not generate boosts but R- symmetries. The underlying spacetime in both cases is the Einstein static universe R S3.4 × Super-Spacetimes Extending R H3 × These correspond to the effective super Lie pairs associated with the Lie superalgebras S17 and S18. The super Lie pairs (s, h) are effective, but the underlying Lie pair (k, h) is not. Indeed, the brackets of k are [B, B]= B and [P, P]= B − J, so that B spans an ideal v k. The result- ⊂ ing Aristotelian spacetime (k/v, r) is the hyperbolic version of the Einstein static universe R H3. × For s the Lie superalgebra S17, the brackets are

1 2 [B(β), Q(s)] = Q( 2 βs) and [Q(s), Q(s)] = |s| H, (4.2.2.16)

4The naming of this manifold may be a slight misnomer. When referring to the Einstein static universe, we typically mean the Lorentzian manifold with topology R S3; however, here, we refer to an Aristotelian × manifold with the same topology. This discrepancy is an artefact of how we defined the classification problem: we classify only effective (super) Lie pairs, and the Lorentz action on the Einstein static universe is not effective. Therefore, the Lorentzian description of this manifold does not appear in our classification; instead, we find an Aristotelian description of this manifold since the rotational Lie subgroup does act effectively. In particular, there exists an SO(D)-equivariant diffeomorphism between K/H and A/R, where A is the Lie group generated by the rotations and spatio-time translations and R is the Lie subgroup generated by the rotations. A similar story holds for the R H3 case in the next section. ×

115 so that B does not span an ideal of s. In other words, B does not generate boosts in the underlying homogeneous spacetime, but rather R-symmetries.

A similar story holds for s the Lie superalgebra S18, with the additional brackets

1 1 2 [H, Q(s)] = Q( 2 sk), [B(β), Q(s)] = Q( 2 βs) and [Q(s), Q(s)] = |s| H − B(sks). (4.2.2.17) Again, the generator B is to be interpreted as an R-symmetry.

Super-Spacetimes Extending the Static Aristotelian Spacetime This corresponds to the Lie superalgebras S27 and S28. In either case the resulting super Lie pair (s, h) is effective, but the underlying Lie pair (k, h) is not since [B, B]= B spans an ideal of k. The homogeneous spacetime associated with the non-effective (k, h) is the Aristotelian static spacetime A.

As in the previous cases, the generators B do not act as boosts but rather as R-symmetries, as evinced by the brackets:

B Q Q 1 Q Q 2H [ (β), (s)] = ( 2 βs) and [ (s), (s)] = |s| (4.2.2.18) for s the Lie superalgebra S27, or

1 1 2 [H, Q(s)] = Q( 2 sk), [B(β), Q(s)] = Q( 2 βs) and [Q(s), Q(s)] = |s| H − B(sks) (4.2.2.19) for s the Lie superalgebra S28.

4.2.3 Aristotelian Homogeneous Superspaces The super Lie pairs (s, h) in green in Table 4.11 are such that the vectorial subspace V h is ∼ ⊂ an ideal v of s. Quotienting s by this ideal yields a Lie superalgebra sa = s/v with a= sa0¯ an Aristotelian Lie algebra (see Table 3.4 for a classification). The resulting Aristotelian super Lie pair (sa, r) is effective by construction and geometrically realisable. It is then a simple matter to identify the Aristotelian Lie superalgebra to which each of those non-effective super Lie pairs in Table 4.11 leads. We summarise this in Table 4.12, which exhibits the correspondence between Aristotelian super Lie pairs in Table 4.11 and Aristotelian Lie superalgebras in Table 4.5. We identify the super Lie pair (s, h) by the label for s as in Table 4.3 and the ideal v h. ⊂ Table 4.12: Correspondence Between Non-Effective Super Lie Pairs and Aristotelian Superal- gebras

s v sa s v sa s v sa S1 B S36 S25 B S38 S10γ [−1,0) (0,1),λ R P S40λ S2 B S39 ∈ ∪ ∈ S25 P S38 S10γ=0,λ=0 P S36 S3 B S39 6 S26 B S39 S10 P S37 S3 P S38 γ=0,λ=0 S26 P S38 S16 B S43 S4 B S38 S27 P S41 S19 B S42 S5 B S37 S28 P S42 S20 B S41 S6 B S37 S31 P S38 S23 B S36 S9γ [−1,1),λ R B S40λ S32 P S38 ∈ ∈ S24 B S37 S9γ=1,λ R B S40λ S33λ R B S40λ ∈ ∈

4.2.4 Summary Table 4.13 lists the homogeneous superspaces we have classified in this section. Each super- spacetime is a superisation of an underlying spatially-isotropic, homogeneous (kinematical or Aristotelian) spacetime, which we list in Table 3.5. Let us recall that Table 3.5 is divided

116 into five sections, corresponding to the different invariant structures which the homogeneous spacetimes admit, as discussed in Chapter 3. We have a similar division of Table 4.13: with the superisations of spacetimes admitting a Lorentzian, Galilean, Carrollian, Aristotelian (with R- symmetries) and Aristotelian (without R-symmetries) structures, respectively. All spacetimes admit superisations with the exception of the Riemannian spaces, de Sitter spacetime (dS4) and two of the Carrollian spacetimes: Carrollian de Sitter (dSC) and the Carrollian light-cone (LC).

Table 4.13: Simply-Connected Spatially-Isotropic Homogeneous Superspaces

SM# M s k (or a) h b p [Q(s), Q(s)] 4 1 2 1 M S14 K8 2 k |s| H − P(sks) 1 1 1 2 2 AdS4 S15 K11 2 j 2 k 2 i |s| H + J(sjs)+ B(sis)− P(sks) 3 G S7 K2 k −P(sks) 4 G S8 K2 −P(sks) 1 1 5λ R dSG S9−1,λ K3−1 (1 + λk) − (B(sks)− P(sks)) ∈ 2 2 1 1 6λ R dSG S10−1,λ K3−1 (−1 + λk) − (B(sks)+ P(sks)) ∈ 2 2 dSG S9 K3 1 1 B P 7γ (−1,1),λ R γ γ,λ γ (1 + λk) − (γ (sks)+ (sks)) ∈ ∈ 2 1 γ dSG S10 K3 1 1 B P 8γ (−1,1),λ R γ γ,λ γ (γ + λk) − ( (sks)+ (sks)) ∈ ∈ 2 γ 1 1 9λ R dSGγ=1 S12λ K31 (1 + λk) B(sks)+ P(sks) ∈ 2 1 10 AdSG S110 K40 2 j −B(sis)+ P(sks) AdSG S11 K4 1 B P 11χ>0 χ χ χ 2 (χ + j) (sk(χ + j)s)+ (sks) 12 C S13 K6 |s|2H 1 2 13 AdSC S14 K8 2 i |s| H + B(sis) 14 R H3 S17 K12 1 |s|2H × 2 15 R H3 S18 K12 1 k 1 |s|2H − B(sks) × 2 2 16 R S3 S21 K13 1 |s|2H × 2 17 R S3 S22 K13 1 k 1 |s|2H − B(sks) × 2 2 A S27 K14 1 2H 18 2 |s| 1 1 2 19 A S28 K14 2 k 2 |s| H − B(sks) 20 A S36 A1 k − −P(sks) 21 A S37 A1 − −P(sks) 22 A S38 A1 − |s|2H 23 A S39 A1 − |s|2H − P(sks) 1 24λ R TA S40λ A2 (1 + λk) − −P(sks) ∈ 2 3 1 2 25 R S S41 A3+ − 2 |s| H × 3 1 2 26 R S S42 A3+ k − 2 |s| H − J(sks)− P(sks) × 3 1 27 R H S43 A3− − i J(sjs)− P(sks) × 2 The first column is our identifier for the superspace, whereas the second column is the underlying homogeneous spacetime it superises. The next two columns are the isomorphism classes of kinematical Lie superalgebra and kinematical Lie algebra, respectively. The next columns specify the brackets of s not of the form [J, −] in a basis where h is spanned by J and B. As explained in Section 2.1.5, supercharges Q(s) are parametrised by s H, whereas J(ω), B(β) ∈ and P(π) are parametrised by ω, β, π Im H. The brackets are given by [H, Q(s)] = Q(sh), ∈ [B(β), Q(s)] = Q(βsb) and [P(π), Q(s)] = Q(πsp), for some h, b, p H. The table is divided ∈ into five sections from top to bottom: Lorentzian, Galilean, Carrollian, Aristotelian with R- symmetries and Aristotelian.

117 4.2.5 Low-Rank Invariants

In this section, we exhibit the low-rank invariants of the homogeneous superspaces in Ta- ble 4.13, all of which are reductive. Indeed, a homogeneous supermanifold with super Lie pair (s, h), where h k=s¯, is reductive if and only the underlying homogeneous manifold (k, h) ⊂ 0 is also reductive. This is because if k = h m is a reductive split, then so is s = h (m S), ⊕ ⊕ ⊕ with S = s¯: the bracket [h, m] m because (k, h) is reductive and the bracket [h, S] S be- 1 ⊂ ⊂ cause h s¯ and S = s¯. In[5], it is shown that all the homogeneous spacetimes in Table 3.5 ∈ 0 1 are reductive with the exception of the Carrollian light cone LC, which in any case does not admit any (4|4)-dimensional superisation. Hence all the superspaces in Table 4.13 are reductive.

Let (s, h) be the super Lie pair associated with one of the homogeneous superspaces in Ta- ble 4.13. We will write s = h m, where we have promoted m to a vector superspace m = m¯ m¯, ⊕ 0 ⊕ 1 with k = h m¯ a reductive split and m¯ =s¯ = S. ⊕ 0 1 1 Invariant tensors on the simply-connected superspace with super Lie pair (s, h) are in one-to- one correspondence with h-invariant tensors on m. Since h contains the rotational subalgebra r =∼ so(3), h-invariant tensors are in particular also rotationally invariant. It is not difficult to write down the rotationally invariant tensors of low order.

As an r module, m = R V S, where R is the trivial one-dimensional module, V is the ⊕ ⊕ vector three-dimensional module and S is the spinor four-dimensional module. Under the iso- morphism r = sp(1) = Im H, m = R Im H H, where the integrated action of a unit-norm ⊕ ⊕ quaternion u Sp(1) on (h, p, s) m is given by ∈ ∈ u (h, p, s)=(h, upu¯, us). (4.2.5.1) ·

Let H,Pi,Qa denote a basis for m, where Pi and Qa have been defined in equation (2.1.5.5). i a We let η, π , θ denote the canonically dual basis for m∗. There is a rotationally invariant line in m: namely, the span of H, which lives in m0¯. Dually, there is a rotationally invariant line in m∗, which is the span of η. These are all the rotationally invariant tensors of rank 1.

Let us now consider rank 2. As an Sp(1) module, m m has the following invariants. First of ⊗ 2 all, we have H2, which is the only invariant featuring H. Another invariant is P := P P , i i ⊗ i which corresponds to the Sp(1)-invariant inner product −, − : Im H Im H R given by h i × → P α, β = Re(αβ¯)=− Re(αβ). If q H is any quaternion, the real bilinear form h i ∈

ωq : H H R defined by ωq(s1, s2)= Re(s1qs2) (4.2.5.2) → → is Sp(1)-invariant: symmetric if q is real and symplectic if q is imaginary (and non-zero). This gives rise to four Sp(1)-invariants quadratic in Q: Q Q and the triplet I Q Q , a a ⊗ a a,b ab a ⊗ b J Q Q and K Q Q , where I, J, K are the matrices representing right- a,b ab a ⊗ b a,b ab a ⊗ b P P multiplication by the quaternions i, j, k; that is, P P 4 4 4

Q(si)= QaIabsb, Q(sj)= QaJabsb and Q(sk)= QaKabsb. (4.2.5.3) , =1 , =1 , =1 aXb aXb aXb 2 Similarly there are several rotational invariants in m∗ m∗: η and, in addition, the symmetric 2 2 ⊗ tensors π and θ , and the triplet of symplectic forms ωI, ωJ and ωK, defined as follows:

2 π (P(α′), P(α)) = Re(α′α¯)=− Re(α′α) 2 θ (Q(s′), Q(s)) = Re(s′s)

ωI(Q(s′), Q(s)) = Re(s′is) (4.2.5.4)

ωJ(Q(s′), Q(s)) = Re(s′js)

ωK(Q(s′), Q(s)) = Re(s′ks).

118 To investigate the invariant tensors on (s, h), we need to investigate the action of B on the a tensors. For the classical invariants (i.e., those not involving Qa or θ ), we may consult [5]: the Lorentzian metric (and the corresponding cometric) are invariant for the Lorentzian spacetimes, the clock one-form and spatial cometric for the Galilean spacetimes, the Carrollian vector and the spatial metric for the Carrollian spacetimes. The generators B act trivially on Aristotelian spacetimes, so the rotationally invariant tensors are the invariant tensors. For the invariants a involving Qa or θ , we need to examine how B acts on S.

As can be gleaned from Table 4.13, B acts trivially on Q in most cases. The exceptions are Minkowski and AdS superspaces and the Aristotelian superspaces where B acts via R- symmetries. Hence, in all other superspaces, the four rotational invariants in m¯ m¯ defined 1 ⊗ 1 above and θ2, ω , ω and ω in m m are h-invariant. This situation continues to hold for the I J K 1¯∗ ⊗ 1¯∗ Aristotelian superspaces with R-symmetry, namely SM14–SM19. Indeed, one can show that all the rotational invariants which are quadratic in Q or in the θa are also R-symmetry invariant. Indeed, the R-symmetry generator Bi acts on m1¯ in the same way as the infinitesimal rotation generator Ji.

Hence it is only for Minkowski and AdS superspaces that the h-invariants do not agree with the r-invariants. For both of these superspaces, h =∼ so(3,1), acting in the same way on the spinors:

B Q Q 1 [ (β), (s)] = ( 2 βsk). (4.2.5.5)

It is a simple calculation to see that the following are h-invariant: I Q Q , J Q a,b ab a⊗ b a,b ab a⊗ Q , ω and ω . b I J P P Since h is isomorphic to the Lorentz subalgebra, we recover the well-known fact that there are two independent Lorentz-invariant symplectic structures on the Majorana spinors. This does not contradict the fact that the Majorana spinor representation S of so(3,1) is irreducible as a real representation, since its complexification (the Dirac spinor representation) decomposes as a direct sum of the two Weyl spinor representations, each one having a Lorentz-invariant symplectic structure.

4.3 Limits Between Superspaces

In this section, we exhibit some limits between the superspaces in Table 4.13 and interpret them in terms of contractions of the underlying Lie superalgebras.

As we will show, a limit between two superspaces induces a limit of the underlying homo- geneous spacetimes. These were discussed in Section 3.3. Our discussion will closely follow that in Section 3.3. There, contractions of a Lie algebra g=(V, φ), where V is a finite-dimensional real vector space and φ : ∧2V V is a linear map satisfying the Jacobi identity, were defined → as limits of curves in the space of Lie brackets. If g : (0,1] GL(V), mapping t g , is a → 7→ t continuous curve with g1 = 1V , we can define a curve of isomorphic Lie algebras (V, φt), where

φ (X, Y) := g−1 φ (X, Y)= g−1 (φ(g X, g Y)) . (4.3.0.1) t t · t t t  If the limit φ0 = limt 0 φ exists, it defines a Lie algebra g0 =(V, φ0), which is then a contrac- → tion of g=(V, φ1).

In the current case, we will contract Lie superalgebras s=(V, φ), where V is now a real finite-dimensional super vector space and φ : ∧2V V is a linear map, where ∧2 is defined → in the super sense, satisfying the super-Jacobi identity. We will define contractions of s in a completely analogous manner.

119 4.3.1 Contractions of the AdS Superalgebra

We begin with the superalgebra for the AdS superspace SM2, whose generators J, B, P, H and Q satisfy the following brackets (in our abbreviated notation):

[H, B]=−P [J, J]= J [H, Q]= Q [H, P]= B [J, B]= B [B, Q]= Q [B, P]= H (4.3.1.1) [J, P]= P [P, Q]= Q [B, B]=−J [J, Q]= Q [Q, Q]= H + J + B − P. [P, P]=−J

Consider the following three-parameter family of linear transformations gκ,c,τ defined by

τ κ κτ g , , J = J, g , , B = B, g , , P = P, g , , H = τκH, g , , Q = Q. κ c τ · κ c τ · c κ c τ · c κ c τ · κ c τ · c (4.3.1.2) The action on the even generators is as in Section 3.3 and the action on Q is chosen to ensure that the bracket [Q, Q] has well-defined limits as κ 0, c or τ 0. → → → The brackets involving J remain unchanged for the above transformations∞ and the remaining brackets become

2 2 [B B]=− τ J τ [H, B]=−τ P , c2 [B, Q]= c Q [H, P]= κ2B κ2 [P, Q]= κ Q (4.3.1.3) [P, P]=− c2 J c 1 1 κτ [B, P]= c2 H [H, Q]= κτQ [Q, Q]= c H + c J + κB − τP.

We now want to take the limits κ 0, c , and τ 0 in turn, corresponding to the flat, → → → non-relativistic, and ultra-relativistic limits, respectively. Notice that the limits of the brackets between the even generators will produce the∞ same Lie algebra contractions as in Section 3.3. Thus we cannot have a limit from one superspace to another unless there exists a limit between their underlying homogeneous spacetimes.

Taking the flat limit κ 0, we are left with → 2 1 τ2 τ 1 [H, B]=−τ P, [B, P]= c2 H, [B, B]=− c2 J, [B, Q]= c Q and [Q, Q]= c H − τP. (4.3.1.4) For τ = 0, this is the Poincaré superalgebra (S14). Thus, we obtain the limit SM2 SM1. c 6 → Subsequently taking the non-relativistic limit c , the brackets reduce to → 2 [H, B]=−τ P and ∞ [Q, Q]=−τP. (4.3.1.5) For τ = 0, this shows us that we have the limit SM1 SM4. 6 → Alternatively, we could have taken the ultra-relativistic limit τ 0, which, for c = 0, gives → 6 us the Carroll superalgebra (S13):

1 1 [B, P]= c2 H and [Q, Q]= c H. (4.3.1.6) Thus, we have SM1 SM12. → Returning to the AdS superalgebra (S15) and taking the non-relativistic limit c , we → find ∞ [H, B]=−τ2P, [H, P]= κ2B, [H, Q]= κτQ and [Q, Q]= κB − τP. (4.3.1.7)

For τκ = 0, this is S110 (under a suitable basis change). Therefore, we have SM2 SM10. 6 → Because these limits commute, we may now take the flat limit to arrive at SM4.

120 Finally, we may take the ultra-relativistic limit of AdS (S15). This limit leaves the brackets

2 1 κ2 κ 1 [H, P]= κ B, [B, P]= c2 H, [P, P]=− c2 J, [P, Q]= c Q and [Q, Q]= c H + κB, (4.3.1.8) for κ = 0. Thus, we arrive at SM13. Subsequently taking the flat limit, we find SM12, as c 6 expected.

We can also take limits from the superspaces discussed above to non-effective super Lie pairs, which will have associated Aristotelian superspaces. Since all of the above superspaces have either SM4 or SM12 as a limit, we will only show the limits to Aristotelian superspaces coming form these two cases. Beginning with SM4, we can use the transformation

g B = tB, g H = H, g P = P and g Q = Q (4.3.1.9) t · t · t · t · and the limit t 0 to obtain SM21. Using the same transformation and limit, we can also → start with SM12 and find SM22.

4.3.2 Remaining Galilean Superspaces We have shown that we obtain the other Lorentzian and two Carrollian superspaces as limits of the AdS superspace SM2: namely, Minkowski (SM1), Carroll (SM12) and Carrollian anti-de Sit- ter (SM13) superspaces. In addition, we also obtain two superisations of Galilean spacetimes: a superisation SM4 of the flat Galilean spacetime and the superisation SM10 of Galilean anti- de Sitter spacetime. But what about the superisations of other Galilean spacetimes?

Flat Galilean Superspaces From SM2, we obtained the Galilean superspace SM4. There is a second superisation SM3 of the flat Galilean homogeneous spacetime, from which we can also reach SM4. Indeed, using the transformations

g B = tB, g H = tH, g P = tP and g Q = √tQ, (4.3.2.1) t · t · t · t · on the Lie superalgebra for SM3, and taking the limit t 0, we find the Lie superalgebra for → SM4. Thus, we have SM3 SM4. → Beginning with SM3, we may also consider the transformation

g B = tB, g H = H, g P = tP and g Q = √tQ, (4.3.2.2) t · t · t · t · and the limit t 0. This procedure will give us a non-effective super Lie pair corresponding → to SM20.

Galilean de Sitter Superspaces

The superspaces SM5 and SM6 arise as the γ −1 limit of SM7 , and SM8 , , respectively. λ λ → γ λ γ λ This fact has already been noted in Section 4.2.2. Section 4.2.2 demonstrated that SM9λ is the γ 1 limit of SM7 , and SM8 , . → γ λ γ λ The superalgebras associated with these five superspaces take the general form

[H, B(β)]=−P(β) [H, Q(s)] = 1 Q(s(η + λk)) 2 (4.3.2.3) [H, P(π)] = γB(π)+(1 + γ)P(π) [Q(s), Q(s)] = ρB(sks)+ σP(sks) for some η, ρ, σ R, where γ [−1,1] and λ R are the parameters of the Lie superalgebras. ∈ ∈ ∈ Using the transformations

g B = B, g H = tH, g P = tP and g Q = √ωtQ, (4.3.2.4) t · t · t · t ·

121 where ω R, and taking the limit t 0, the above brackets become ∈ → [H, B(β)]=−P(β) and [Q(s), Q(s)] = ωσP(sks). (4.3.2.5)

Therefore, by choosing ω =−σ−1, we can always recover SM4.

There is a second superisation of the flat Galilean homogeneous spacetime, namely SM3. There does not seem to be any Lie superalgebra contraction that gives SM3, but as we will see below, there are non-contracting limits (involving taking λ ) which take the superspaces SM5 , → ± λ SM6λ, SM7γ,λ, SM8γ,λ and SM9λ to SM3. ∞

Galilean Anti-de Sitter Superspaces

The superspace SM10 is, by definition, the χ 0 limit of SM11 . These algebras take the form → χ [H, B(β)]=−P(β) [H, Q(s)] = 1 Q(s(χ + j)) 2 (4.3.2.6) [H, P(π)]=(1 + χ2)B(π)+ χP(π) [Q(s), Q(s)]=−B(sis)− P(sks), where χ > 0 is the parameter of the Lie superalgebra. Using the same transformations as in the Galilean de Sitter case, but with ω = 1, we find

[H, B(β)] = P(β) and [Q(s), Q(s)]=−P(sks). (4.3.2.7)

Thus, we find SM4 as a limit of both SM10 and SM11χ.

We cannot obtain SM3 as a limit of these superspaces as SM3 has collinear h and c3, whereas SM10 and SM11χ have orthogonal h and c3.

Non-Contracting Limits

In Section 3.3, it was shown that limχ AdSGχ = dSG1, but this limit is not induced by a Lie → algebra contraction since the Lie algebras are non-isomorphic for different values of χ. Does this limit extend to the superspaces? ∞

Beginning with SM11χ, change basis such that

−1 −1 −1/2 H′ = χ H, B′ = B, P′ = χ P and Q′ = χ Q, (4.3.2.8) under which the brackets become H Q 1 Q [H′, B′(β)]=−P′(β) [ ′, ′(s)] = 2χ ′(s(χ + j)) −2 −1 [H′, P′(π)] = 2P′(π)+(1 + χ )B′(π) [Q′(s), Q′(s)]=−χ B′(sis)+ B′(sks)+ P(sks). (4.3.2.9) Taking the limit χ , we find → 1 [H′, B′(β)]=−P′(β) [H′, Q′(s)] = Q′(s) ∞ 2 (4.3.2.10) [H′, P′(π)] = 2P′(π)+ B′(π) [Q′(s), Q′(s)]=−B′(sks)+ P(sks).

This Lie superalgebra is precisely that for SM90. Thus, we inherit this limit from the underlying homogeneous spacetimes.

The superspaces SM5λ, SM6λ, SM7γ,λ, SM8γ,λ and SM9λ all have an additional parameter λ, and we can ask what happens if we take the limit λ in these cases. This is again a → ± non-contracting limit, since the Lie superalgebras with different values of λ R are not isomor- ∈ phic. ∞

122 Using the general form of the brackets stated in (4.3.2.3), consider a change of basis

1 −1 −1 − B′ = B, H′ = 2λ H, P′ = 2λ P and Q′ = λ 2 Q. (4.3.2.11)

In our new basis, the brackets become

−1 [H′, B′(β)]=−P′(β) [H′, Q′(s)] = Q′(s(λ η + k)) H P −2 B −1 P Q Q −1 B σ P [ ′, ′(π)] = 4λ γ ′(π)+ 2λ (1 + γ) ′(π) [ ′(s), ′(s)] = λ ρ ′(sks)+ 2 ′(sks). (4.3.2.12) Taking either λ or λ − , we find → → H B P H Q Q Q Q σ P [ ′, ′(β)]=−∞ ′(β), [∞′, ′(s)] = ′(sk), [ ′(s), ′(s)] = 2 ′(sks). (4.3.2.13)

σ Rescaling both B′ and P′ by 2 , we recover the Lie superalgebra for SM3.

Figure 4.1 below illustrates the different superspaces and the limits between them. The families SM5λ, SM6λ, SM7γ,λ, SM8γ,λ and SM9λ fit together into a two-dimensional space, which also includes SM3 as their common limits λ . This space may be described as follows. If we → ± fix λ R, then ∈ lim SM7γ,λ = SM9λ whereas∞ lim SM7γ,λ = SM5λ. (4.3.2.14) γ 1 γ −1 → → Similarly, again fixing λ R, we have ∈

lim SM8γ,λ = SM9λ whereas lim SM8γ,λ = SM6λ. (4.3.2.15) γ 1 γ −1 → → This gives rise to the following two-dimensional parameter spaces: 3 3

5λ 7γ,λ 9λ 5λ 8γ,λ 9λ

3 3

We then flip the square on the right horizontally and glue the two squares along their common 9λ edge to obtain the following picture 3

5λ 7γ,λ 9λ 8−γ,λ 6λ

3

We now glue the top and bottom edges to arrive at the following cylinder:

90

7γ,λ 8−γ,λ 5λ 9λ 6λ

3

123 Finally, we collapse the “edge” labelled 3 to a point, arriving at the object in Figure 4.1.

4.3.3 Aristotelian Limits There are two kinds of superisations of Aristotelian spacetimes: the ones where B acts as R-symmetries and the ones where B acts trivially. We treat them in turn.

Aristotelian Superspaces with R-Symmetry

The homogeneous spacetimes R H3 and R S3 underlying the homogeneous superspaces SM14 × × - SM17 have A as their limit. Therefore, we could expect SM14 - SM17 to have either SM18 or SM19 as limits. The relevant contraction uses the transformation

g B = B, g H = H and g P = tP. (4.3.3.1) t · t · t · Taking the limit t 0, the [P, P] bracket vanishes leaving all other brackets unchanged. Thus, → we find SM14 SM18, SM16 SM18, SM15 SM19 and SM17 SM19. → → → → Taking into account the form of h, and the [Q, Q] bracket for each of these superspaces, we notice that each homogeneous spacetime has two superspaces associated with it. One for which

1 2 b = 2 and [Q(s), Q(s)] = |s| H, (4.3.3.2) and one for which

1 1 2 b = 2 , h = 2 k and [Q(s), Q(s)] = |s| H − B(sks). (4.3.3.3) Using transformations which act as

g H = tH, g Q = √tQ (4.3.3.4) t · t · and trivially on J, B, and P, we find the brackets of the latter superspaces described by

1 t Q Q 2H B b = 2 , h = 2 k, and [ (s), (s)] = |s| − t (sks). (4.3.3.5)

Therefore, taking the limit t 0, we find the former superspaces. Thus, we get the limits → SM15 SM14, SM17 SM16 and SM19 SM18. → → → All of the above superspaces have SM18 as a limit. Therefore, we will only consider the limits of this superspace to those Aristotelian superspaces without R-symmetry. Letting

g B = tB, g H = H, g P = P, g Q = Q, (4.3.3.6) t · t · t · t · and taking the limit t 0, we arrive at a non-effective super Lie pair corresponding to SM22. →

Aristotelian Superspaces without R-Symmetry

The Aristotelian homogeneous spacetimes R S3, R H3, and TA have A as their limit; therefore, × × we would expect their superisations to have have one or more of SM20-SM23 as limits. For TA to have A as its limit, we require the transformation

g B = B, g H = tH and g P = P. (4.3.3.7) t · t · t · Wanting to ensure [Q, Q] = 0, and that the limit t 0 is well-defined, we need g Q = √tQ. 6 → t · Taking this limit, we find SM24 SM21. λ → To get A from R S3, we need the transformation × g B = B, g H = H and g P = tP. (4.3.3.8) t · t · t ·

124 Using this transformation and taking the limit t 0, we find SM25 SM22. However, the limit → → is not well-defined for SM26 due to P in the expression for [Q, Q]. In this case, we additionally require g Q = √tQ. Then SM26 SM20. Another choice of transformation, t · → g B = B, g H = tH, g P = tP and g Q = √tQ, (4.3.3.9) t · t · t · t for SM26, gives SM23 in the limit t 0. Thus, we also have SM26 SM23. → → Finally, to get A from R H3, we use the transformation × g B = B, g H = H, g P = tP. (4.3.3.10) t · t · t · To ensure the limit t 0 is well-defined, we subsequently need g Q = √tQ. This transforma- → t · tion with the limit gives SM27 SM21. → There are only two underlying Aristotelian homogeneous spacetimes which have more than one superisation. These are A and R S3. In the latter case, we find the superisation SM25 as × the limit of SM26 using the transformation

g B = B, g H = tH, g P = P and g Q = √tQ, (4.3.3.11) t · t · t · t · and taking t 0. In the former case, the superisations SM22 and SM21 can be found as limits → of SM23 using the transformations

g B = B, g H = tH, g P = P and g Q = √tQ, (4.3.3.12) t · t · t · t · and g B = B, g H = H, g P = tP and g Q = √tQ, (4.3.3.13) t · t · t · t · respectively. We also have

g B = B, g H = tH, g P = P and g Q = Q, (4.3.3.14) t · t · t · t · giving the limit SM20 SM21. → 4.3.4 A Non-Contracting Limit

Use the following change of basis on the Lie superalgebra for SM24λ,

−1 B′ = B, H′ = 2λ H, P′ = P, Q′ = Q. (4.3.4.1)

The brackets then become

−1 −1 [H′, P(π)′]= 2λ P(π)′, [H′, Q′(s)] = Q′(s(λ + k)), [Q′(s), Q′(s)]=−P′(sks). (4.3.4.2)

Taking the limits λ , we find the superspace SM20. Therefore, the line of superspaces → ± SM24λ compactifies to a circle with SM20 as the point at infinity. ∞ 4.3.5 Summary The picture resulting from the above discussion is given in Figure 4.1. Except for SM3 SM4, → the limits from the families SM5λ, SM6λ, SM7γ,λ, SM8γ,λ, SM9λ and SM11χ to SM4 are not shown explicitly in order to improve readability. Neither is the limit between SM24λ and SM21 shown.

For comparison, we extract from Figure 3.1 the subgraph corresponding to spacetimes which admit superisations and show it in Figure 4.2. There are arrows between these two pictures: taking a superspace to its corresponding spacetime, but making this explicit seems beyond our artistic abilities.

125 13 2

Lorentzian

15 14 12 1 Galilean 10

18 19 Carrollian 11χ

27 4 Aristotelian 17 16 23 22 5λ Aristotelian+R 7γ,λ 25 9 21 λ 90 20 3 8−γ,λ 26

6 24λ λ

Figure 4.1: Homogeneous Superspaces and their Limits. (Numbers are hyperlinked to the corresponding superspaces in Table 4.13.)

AdSC AdS4

Lorentzian 4 C M Galilean Carrollian AdSG = AdSG0 Aristotelian

G AdSGχ>0

A

TA

dSG dSG1 = AdSG dSG γ∈[−1,1] ∞

R × S3 R × H3

Figure 4.2: Limits Between Superisable Spacetimes

Nevertheless, interpreting Figures 4.1 and 4.2 as posets, with arrows defining the partial order, the map taking a superspace to its underlying spacetime is surjective by construction (we con- sider only superisable spacetimes) and order preserving, as shown at the start of this section. As can be gleaned from Table 4.13, the fibres of this map are often quite involved, clearly showing the additional “internal” structure in the superspace which allows for more than one possible superisation of a spacetime.

We should mention that, despite appearances, superspaces SM3 and SM4 share the same under- lying spacetime: namely, the Galilean spacetime G. Notice that superspaces SM21 and SM22, which are “terminal” in the partial order, correspond to the static Aristotelian spacetime A. With the exception of limχ SM11χ = SM90, all other non-contracting limits between su- → perspaces induce limits between the underlying spacetimes, which arise from contractions of the kinematical Lie algebras:∞ the limits |λ| of SM5 and SM6 induce the contraction → λ λ dSG G, whereas the limits |λ| of SM7 , , SM8 , and SM9 induce the contractions → → γ λ γ λ λ dSGγ G, where γ = 1 for SM9λ. ∞ → ∞ 126 4.4 Conclusion

In this chapter, we discussed the classification of N = 1 kinematical Lie superalgebras in three spatial dimensions and, subsequently, the classification of the kinematical superspaces which arise from these algebras.

The Lie superalgebras were classified by solving the super-Jacobi identities in a quaternionic re- formulation, which made the computations no harder than multiplying quaternions and paying close attention to the action of automorphisms in order to ensure that there is no repetition in our list. Since we are interested in supersymmetry, we focussed on Lie superalgebras where the supercharges were not abelian: i.e., we demand that [Q, Q] = 0 and, subject to that condition, 6 we classified Lie superalgebras which extend either kinematical or Aristotelian Lie algebras. The results are contained in Tables 4.3 and 4.5, respectively.

There are two salient features of these classifications. Firstly, not every kinematical Lie algebra admits a supersymmetric extension: in some cases because of our requirement that [Q, Q] = 0, 6 but in other cases (e.g., so(5), so(4,1),...) because the four-dimensional spinor module of so(3) does not extend to a module of these Lie algebras.

Secondly, some kinematical Lie algebras admit more than one non-isomorphic supersymmetric extension. For example, the Galilean Lie algebra admits two supersymmetric extensions, but only one of them (S8) can be obtained as a contraction of osp(1|4). By far most of the Lie superalgebras in our classification cannot be so obtained and hence are not listed in previous classifications. Nevertheless, our “moduli space” of Lie superalgebras is connected, if not al- ways by contractions. For example, the other supersymmetric extension of the Galilean algebra (S7) can be obtained as a non-contracting limit of some of the multi-parametric families of Lie superalgebras in the limit as one of the parameters goes to , in effect compactifying one of ± the directions in the parameter space into a circle. ∞ We classified the corresponding superspaces via their super Lie pairs (s, h), where s is a kine- matical Lie superalgebra and h an admissible subalgebra. Every such pair “superises” a pair (k, h), where k=s0¯ is a kinematical Lie algebra. As discussed in Section 2.2.5, effective and geometrically realisable pairs (k, h) are in bijective correspondence with simply-connected ho- mogeneous spacetimes, and hence the super Lie pairs (s, h) are in bijective correspondence with superisations of such spacetimes. These are listed in Table 4.13.

There are several salient features of that table. Firstly, many spacetimes admit more than one inequivalent superisation. Whereas Minkowski and AdS spacetimes admit a unique (N=1) superisation, and so too do the (superisable) Carrollian spacetimes. Many of the Galilean spacetimes admit more than one and in some cases even a circle of superisations.

Secondly, there are effective super Lie pairs (s, h) for which the underlying pair (k, h) is not effective. This means that the “boosts” act trivially on the underlying spacetime, but non- trivially in the superspace: in other words, the “boosts” are actually R-symmetries. Since (k, h) is not effective, this means that it describes an Aristotelian spacetime and this gives rise to the class of Aristotelian superspaces with R-symmetry.

Thirdly, there are three superspaces in our list which also appear in [108]: namely, Minkowski (SM1) and AdS (SM2) superspaces, but also the Aristotelian superspace SM26, whose under- lying manifold appears in [108] as the Lorentzian Lie group R SU(2) with a bi-invariant metric. × Lastly, just like Minkowski (M4) and Carrollian AdS (AdSC) spacetimes are homogeneous under the Poincaré group, their (unique) superisations (SM1 and SM13, respectively) are homogeneous under the Poincaré supergroup, suggesting a sort of correspondence or duality.

127 128 Chapter 5

Generalised Bargmann Superspaces

In this chapter, we consider the last of our three types of symmetry, super-Bargmann symmetry. These symmetries are the least-studied of the three; therefore, we can only present the algebraic classification in this instance.

Before introducing any supersymmetry, recall that a generalised Bargmann algebra (GBA) ˆk in D spatial dimensions may be thought of as a real one-dimensional abelian extension of a kinematical Lie algebra k. In particular, this enhancement of a kinematical Lie algebra requires an additional so(3) scalar module in the underlying vector space. Let Z span this extra copy of R. The classification of these extensions was presented in [3] and followed a similar method to that used in the classification of kinematical Lie algebras, as discussed in Section 3.1. In particular, rather than classifying the Lie algebras we can recover as deformations of the static kinematical Lie algebra, they classified the Lie algebras we can recover as deformations of the static kinematical Lie algebra’s universal central extension. The centrally-extended static kinematical Lie algebra is spanned by J, B, P, H, and Z, with non-vanishing brackets

[J, J]= J [J, B]= B [J, P]= P [B, P]= Z. (5.0.0.1)

All other centrally-extended kinematical Lie algebras are then deformations of this algebra; therefore, these brackets are common to all such algebras. The results of this classification, for D = 3, are shown in Table 5.1, taken from [3].1 The three sections of this table, starting from the top, are the non-trivial central extensions, the trivial central extensions, and, finally, the non-central extensions of kinematical Lie algebras.

In this chapter, we will focus solely on the first of these sections, and, from now on, it shall be exclusively the algebras of this section that we are referring to when using the term gener- alised Bargmann algebras. It is useful for our later calculations to define a universal generalised Bargmann algebra. In addition to the standard kinematical brackets given in (2.1.2.3), this algebra has non-vanishing brackets

[B, P]= Z [H, B]= λB + µP [H, P]= ηB + εP, (5.0.0.2) where λ, µ, η, ε R.2 Setting these four parameters to certain values allows us to reduce to ∈ the four cases of interest. For example, ˆg is given by setting λ = η = ε = 0 and µ = −1. By beginning with the universal algebra, and only picking our parameters, and, thus, our algebra, when we can no longer make progress in the universal case, we reduce the amount of repetition in our calculations.

1As in Chapter 4, we will only consider the generalised Bargmann algebras in D = 3. 2Note, the universal generalised Bargmann algebra is not a Lie algebra for arbitrary λ, µ, η, ε. It is used here simply as a computational tool.

129 Table 5.1: Centrally-Extended Kinematical Lie Algebras in D = 3

KLA Non-zero Lie brackets in addition to [J, J]= J, [J, B]= B, [J, P]= P Comments 1 [B, P]= Z ˆa 2 [B, P]= Z [H, B]= B [H, P]=−P ˆn− 3 [B, P]= Z [H, B]= P [H, P]=−B ˆn+ 4 [B, P]= Z [H, B]=−P ˆg 5 [B, P]= H [H, B]= P [B, B]= J e RZ ⊕ 6 [B, P]= H [H, B]=−P [B, B]=−J p RZ ⊕ 7 [B, P]= H + J [H, B]=−B [H, P]= P so(4,1) RZ ⊕ 8 [B, P]= H [H, B]= P [H, P]=−B [P, P]= J [B, B]= J so(5) RZ ⊕ 9 [B, P]= H [H, B]=−P [H, P]= B [P, P]=−J [B, B]=−J so(3,2) RZ ⊕ 10 [B, P]= Z [H, B]= B [H, P]= P [H, Z]= 2Z 11 [B, P]= Z [H, B]= γB [H, P]= P [H, Z]=(γ + 1)Z γ (−1,1) ∈ 12 [B, P]= Z [H, B]= B + P [H, P]= P [H, Z]= 2Z 13 [B, P]= Z [H, B]= αB + P [H, P]=−B + αP [H, Z]= 2αZ α> 0 14 [B, P]= Z [Z, B]= P [H, P]= P [H, Z]= Z [B, B]= J co(4) ⋉ R4 15 [B, P]= Z [Z, B]=−P [H, P]= P [H, Z]= Z [B, B]=−J co(3,1) ⋉ R3,1

Table 5.2: Generalised Bargmann Algebras in D = 3

GBA Non-zero Lie brackets in addition to [J, J]= J, [J, B]= B, [J, P]= P Comments 1 [B, P]= Z ˆa 2 [B, P]= Z [H, B]= B [H, P]=−P ˆn− 3 [B, P]= Z [H, B]= P [H, P]=−B ˆn+ 4 [B, P]= Z [H, B]=−P ˆg

Our strategy for classifying generalised Bargmann superalgebras will be analogous to the strat- egy used in Section 4.1 to classify the kinematical Lie superalgebras. In particular, a super- extension s of one of our generalised Bargmann algebras ˆk will be a Lie superalgebra such that ˆ s0¯ = k. To determine the super-extensions of the generalised Bargmann algebras, we, therefore, begin by letting s0¯ be our universal generalised Bargmann algebra. We then need to specify the Lie brackets [H, Q], [Z, Q], [B, Q], [P, Q], and [Q, Q]. Each of the [s0¯, s1¯] components of the bracket must be an r-equivariant endomorphism of s1¯, while the [s1¯, s1¯] component must be an 2 r-equivariant map s¯ s¯. The space of possible brackets will be a real vector space V . 1 → 0 We then use the super-Jacobi identity to cut out an algebraic variety J V . Since we are J ⊂ exclusively interested in supersymmetric extensions, we restrict ourselves to those Lie superal- gebras for which [Q, Q] is non-vanishing, which define a sub-variety S J .3 This sub-variety ⊂ may be unique to each generalised Bargmann algebra; therefore, it is at this stage we start to set the parameters of the universal algebra, where applicable. The isomorphism classes of the remaining Lie superalgebras are then in one-to-one correspondence with the orbits of S under the subgroup G GL(s¯) GL(s¯). The group G contains the automorphisms of s¯ = ˆk and ⊂ 0 × 1 0 additional transformations which are induced by the endomorphism ring of s1¯. The form of this subgroup will be discussed in the N = 1 and N = 2 cases in sections 5.1.1 and 5.2.1, respectively.

In the N = 1 case, we will identify each orbit of S explicitly, giving a full classification of the generalised Bargmann superalgebras in this instance. However, in the N = 2 case, we will only identify the non-empty branches of S . Each branch will have a unique set of [s0¯, s1¯] and [s1¯, s1¯] brackets for the associated generalised Bargmann algebra. Thus, we can highlight the form of the possible super-extensions without spending too much time pinpointing exact coef- ficients.

The rest of this chapter is organised as follows. In Section 5.1, we classify the N = 1 generalised Bargmann superalgebras in D = 3. We begin by generalising the setup for the kinematical Lie

3As in Chapter 4, we restrict ourselves to the cases where [Q, Q] = 0 as our interests lie in spacetime 6 supersymmetry: we would like supersymmetry transformations to generate geometric transformations of the spacetime.

130 superalgebra classification presented in Section 4.1.1 before proceeding to the classification itself in Section 5.1.2 and summarising in Section 5.1.3. As part of our summary, we will demonstrate how to unpack our quaternionic formalism, using one of the N = 1 Bargmann superalgebras as our example. In Section 5.2, we move on to classify the N = 2 generalised Bargmann superalge- bras in D = 3. This case is considerably more involved than the N = 1 case; therefore, after an analogous discussion on the initial setup of the classification problem, we require an intermedi- ate step in which we define four possible branches of generalised Bargmann superalgebra in S . In Section 5.2.3, we go through the lengthy procedure of identifying the sub-branches which contain valid generalised Bargmann superalgebra structures, and we summarise our findings in Section 5.2.4.

5.1 Classification of N = 1 Generalised Bargmann Super- algebras

Our investigation into generalised Bargmann superalgebras begins with the simplest case, N = 1. Following on from Section 2.1.6, Section 5.1.1 will complete our set up for this case by specifying the precise form of the [s0¯, s1¯] and [s1¯, s1¯] brackets. We then give some preliminary results that will be useful in the classification of the N = 1 extensions, and define the group of basis transformations G GL(s¯) GL(s¯), which will allow us to pick out a single representative ⊂ 0 × 1 for each isomorphism class. In Section 5.1.2, the classification is given before we summarise the results in Section 5.1.3.

5.1.1 Setup for the N = 1 Calculation We note that, in addition to the standard kinematical Lie brackets, the brackets for the universal generalised Bargmann superalgebra are

[B(β), P(π)] = Re(βπ¯ )Z [H, B(β)] = λB(β)+ µP(β) (5.1.1.1) [H, P(π)] = ηB(π)+ εP(π), where β, π Im(H) and λ, µ, η, ε R. We now want to specify the possible [s¯, s¯] and [s¯, s¯] ∈ ∈ 0 1 1 1 brackets. From Section 4.1.1, we have

1 [J(ω), Q(θ)] = 2 Q(ωθ) [B(β), Q(θ)] = Q(βθb) (5.1.1.2) [P(π), Q(θ)] = Q(πθp) [H, Q(θ)] = Q(θh), where ω, π, β Im(H), θ, b, p, h H. Since Z is just another so(3) scalar module, and, therefore, ∈ ∈ the analysis of the bracket [Z, Q] will be identical to that of [H, Q], we know we can write

[Z, Q(θ)] = Q(θz), (5.1.1.3) where z H. Having added this additional generator, the possible [s1¯, s1¯] brackets are now ∈ 2 so(3)-equivariant elements of HomR( S, s¯) = HomR(3V R,3V 2R) = 9 HomR(V, V) 0 ⊕ ⊕ ⊕ 2 Hom (R, R). As may have been expected, given that we did not alter the vectorial sector of R J the algebra, the number of HomR(V, V) elements does not change. However, now that we have an additional so(3) scalar module, we have an additional scalar map, so

2 2 [Q(θ), Q(θ)] = n0|θ| H + n1|θ| Z − J(θn2θ¯)− B(θn3θ¯)− P(θn4θ¯), (5.1.1.4) where n0, n1 R and n2, n3, n4 Im(H). This expression polarises to ∈ ∈ ¯ ¯ ¯ ¯ ¯ ¯ ¯ ¯ [Q(θ), Q(θ′)] = n0 Re(θθ′)H+n1 Re(θθ′)Z−J(θ′n2θ+θn2θ′)−B(θ′n3θ+θn3θ′)−P(θ′n4θ+θn4θ′). (5.1.1.5)

131 Preliminary Results Following the example of Section 4.1.1, we will now consider the super-Jacobi identity and use its components to derive universal conditions that may aid our classification. We have three components of super-Jacobi identity to consider

1. (s0¯, s0¯, s1¯),

2. (s0¯, s1¯, s1¯), and

3. (s1¯, s1¯, s1¯).

We do not need to consider the (s0¯, s0¯, s0¯) case as we already know that these are satisfied by the generalised Bargmann algebras. Equally, we do not need to include J in our investiga- tions as the identities involving the rotational subalgebra r impose the so(3)-equivariance of the brackets, which we already have by construction. Now, let us consider each component of the identity in turn. In the following discussions, we will only write down explicitly those identities which are not trivially satisfied.

(s0¯, s0¯, s1¯)

By imposing these super-Jacobi identities, we ensure that s1¯ is an s0¯ module, not just an so(3) module. The (s0¯, s0¯, s1¯) identities can be summarised as follows.

Lemma 5.1.1. The following relations between h, z, b, p H are implied by the corresponding ∈ k-brackets: [H, Z]= λH + µZ = [z, h]= λh + µz ⇒ [H, B]= λB + µP = [b, h]= λb + µp ⇒ [H, P]= λB + µP = [p, h]= λb + µp ⇒ [Z, B]= λB + µP = [b, z]= λb + µp ⇒ (5.1.1.6) [Z, P]= λB + µP = [p, z]= λb + µp ⇒ [B, B]= λB + µP + νJ = b2 = 1 λb + 1 µp + 1 ν ⇒ 2 2 4 [P, P]= λB + µP + νJ = p2 = 1 λb + 1 µp + 1 ν ⇒ 2 2 4 [B, P]= λH + µZ = bp + pb = 0 and [b, p]= λh + µz. ⇒ Proof. All the results excluding Z are taken from Lemma 4.1.1, and the [Z, B] and [Z, P] results are the same mutatis mutandis as [H, B] and [H, P]. Therefore, the only new results are those for [H, Z] and [B, P]. The [H, Z, Q] identity is written

[H, [Z, Q(θ)] = [[H, Z], Q(θ)]+[Z, [H, Q(θ)]]. (5.1.1.7)

Substituting in the relevant brackets, we find

Q(θzh)= λQ(θh)+ µQ(θz)+ Q(θhz). (5.1.1.8)

Using the injectivity of Q, we arrive at

[z, h]= λh + µz. (5.1.1.9)

Finally, the [B, P, Q] identity is

[B(β), [P(π), Q(θ)]] = [[B(β), P(π)], Q(θ)]+[P(π), [B(β), Q(θ)]]. (5.1.1.10)

Substituting in the brackets from (5.1.1.2), we arrive at

Q(βπθpb)= Re(βπ¯ )(λQ(θh)+ µQ(θz)) + Q(πβθbp). (5.1.1.11)

132 Letting β = π = i, we find [b, p]= λh + µz. (5.1.1.12) Now, let β = i and π = j. In this case, the λ and µ terms vanish and we are left with

bp + pb = 0. (5.1.1.13)

(s0¯, s1¯, s1¯)

By imposing these super-Jacobi identities, we ensure that the [Q, Q] bracket is an s0¯-equivariant map s¯ s¯. The (s¯, s¯, s¯) identities can be difficult to study if we are trying to be 1 → 0 0 1 1 completely general; however, we know that all four algebras in Table 5.2 can be written as J specialisations of the universal generalised Bargmann algebra:

[B, P]= Z [H, B]= λB + µP [H, P]= ηB + εP, (5.1.1.14) where λ, µ, η, ε R. Therefore, we may use the brackets of this algebra to obtain the following ∈ result.

Lemma 5.1.2. The [H, Q, Q] identity produces the conditions

0 = n Re(h) where i {0,1} i ∈ 0 = hn2 + n2h¯ (5.1.1.15) λn3 + ηn4 = hn3 + n3h¯

µn3 + εn4 = hn4 + n4h¯.

The [Z, Q, Q] identity produces the conditions

0 = ni Re(h) where i {0,1} ∈ (5.1.1.16) 0 = hn + n h¯ where j {2,3,4}. j j ∈ The [B, Q, Q] identity produces the conditions

0 = n0 Re(θβθ¯ b)

0 = Re(βθ¯ (n4 + 2n1b¯)θ¯)

0 = θn2βθb + βθbn2θ¯ (5.1.1.17) 2 1 ¯ ¯ λn0|θ| β + 2 [β, θn2θ]= θn3βθb + βθbn3θ 2 µn0|θ| β = θn4βθb + βθbn4θ¯.

The [P, Q, Q] identity produces the conditions

0 = n0 Re(θπθ¯ p)

0 = Re(πθ¯ (n3 − 2n1p¯)θ¯)

0 = θn2πθp + πθpn2θ¯ (5.1.1.18) 2 ηn0|θ| π = θn3πθp + πθpn3θ¯ 2 1 ¯ ¯ εn0|θ| π + 2 [π, θn2θ]= θn4πθp + πθpn4θ.

Proof. The [H, Q, Q] super-Jacobi identity is written

[H, [Q(θ), Q(θ)]] = 2[[H, Q(θ)], Q(θ)]. (5.1.1.19)

133 Using (5.1.1.2) and (5.1.1.4), we find

−B(θ(λn3 + ηn4)θ¯)− P(θ(µn3 + εn4)θ¯)= 2n0 Re(θhθ)H + 2n1 Re(θhθ)Z

− J(θn2θh + θhn2θ¯)− B(θn3θh + θhn3θ¯)− P(θn4θh + θhn4θ¯). (5.1.1.20)

Comparing H, Z, J, B, and P coefficients, and using the injectivity and linearity of the maps J, B, and P, we find

0 = n Re(h) where i {0,1} i ∈ 0 = hn2 + n2h¯ (5.1.1.21) λn3 + ηn4 = hn3 + n3h¯

µn3 + εn4 = hn4 + n4h¯.

The calculations for the [Z, Q, Q] identity follows in an analogous manner. The key difference is this case is that the L.H.S. vanishes in all instances since Z commutes with all basis elements. The [B, Q, Q] identity is

[B(β), [Q(θ), Q(θ)]] = 2[[B(β), Q(θ)], Q(θ)]. (5.1.1.22)

Substituting in the relevant brackets, the L.H.S. becomes

2 1 ¯ ¯ ¯ L.H.S. =−λn0|θ| B(β)− 2 B([β, θn2θ]) − Re(βθn4θ)Z, (5.1.1.23) and the R.H.S. becomes

R.H.S. = 2n0 Re(θβθ¯ b)H + 2n1 Re(θβθ¯ b)Z (5.1.1.24) − J(θn2βθb + βθbn2θ¯)− B(θn3βθb + βθbn3θ¯)− P(θn4βθb + βθbn4θ¯).

Again, comparing coefficients and using the injectivity and linearity of our maps, we get

0 = n0 Re(θβθ¯ b)

0 = Re(βθ¯ (n4 + 2n1b¯)θ¯)

0 = θn2βθb + βθbn2θ¯ (5.1.1.25) 2 1 ¯ ¯ λn0|θ| β + 2 [β, θn2θ]= θn3βθb + βθbn3θ 2 µn0|θ| β = θn4βθb + βθbn4θ¯.

The [P, Q, Q] results follow in identical fashion by replacing b with p and β with π.

(s1¯, s1¯, s1¯)

The last super-Jacobi identity to consider is the (s1¯, s1¯, s1¯) case, [Q, Q, Q].

Lemma 5.1.3. The [Q, Q, Q] identity produces the condition

1 n0h + n1z = 2 n2 + n3b + n4p. (5.1.1.26)

Proof. The [Q, Q, Q] identity is 0 = [[Q(θ), Q(θ)], Q(θ)]. (5.1.1.27) Using (5.1.1.4), and the brackets in (5.1.1.2) and (5.1.1.3), we find

2 2 0 =[n0|θ| H + n1|θ| Z − J(θn2θ¯)− B(θn3θ¯)− P(θn4θ¯), Q(θ)] (5.1.1.28) 2 2 1 ¯ ¯ ¯ = n0|θ| Q(θh)+ n1|θ| Q(θz)− 2 Q(θn2θθ)− Q(θn3θθb)− Q(θn4θθp).

134 Since Q is injective, this gives us

1 n0h + n1z = 2 n2 + n3b + n4p. (5.1.1.29)

Basis Transformations

As well as modifying the super-Jacobi identities presented in Section 4.1.1, the new so(3) scalar also impacts the subgroup G GL(s¯) GL(s¯) of basis transformation for kinematical Lie ⊂ 0 × 1 superalgebras. All the automorphisms in G generated by so(3) remain the same for b, p, and h, but we may now add how z transforms. These automorphisms act by rotating the three imaginary quaternionic bases i, j, and k by an element of SO(3). In particular, we have a homomorphism Ad : Sp(1) Aut(H) defined such that for u Sp(1) and s H, Adu(s)= usu¯. → ∈ ∈ The map Adu then acts trivially on the real component of s and via SO(3) rotations on Im(H). Therefore, B˜ = B Adu, P˜ = P Adu, H˜ = H, Q˜ = Q Adu. Since Z is an so(3) scalar, Z˜ = Z. ◦ ◦ ◦ Substituting this with Q˜ = Q Adu into the [Z, Q] bracket, we find that ˜z = uzu¯ . Additionally, ◦ substituting these transformations into the [Q, Q] bracket, we see that n1 remains inert. The other type of transformations to consider are the so(3)-equivariant maps s s. Since we now → have two so(3) scalars, we can have

H˜ = aH + bZ a b where GL(2, R). (5.1.1.30) Z˜ H Z c d ∈ = c + d   The so(3) vector and spinor maps remain unchanged from those given in Section 4.1.1. In particular, Q˜(s)= Q(sq) for q H×. Substituting H˜, Z˜ and Q˜ into the brackets ∈ H˜ Q˜ Q˜ ˜ [ , (θ)] = (θh) 2 2 [Q˜(θ), Q˜(θ)] = n˜0|θ| H˜ + n˜1|θ| Z˜ − ˜J(θn˜2θ¯)− B˜(θn˜3θ¯)− P˜(θn˜4θ¯), [Z˜, Q˜(θ)] = Q˜(θ˜z) (5.1.1.31) we find |q|2 −1 n˜ = (dn − cn ) h˜ = q(ah + bz)q 0 ad − bc 0 1 (5.1.1.32) ˜z = q(ch + dz)q−1 |q|2 n˜ = (an − bn ). 1 ad − bc 1 0 These amendments mean that the transformation in G produce the following basis changes

J J Adu 7→ ◦ B eB Adu +fP Adu 7→ ◦ ◦ P hB Adu +iP Adu 7→ ◦ ◦ (5.1.1.33) H aH + bZ 7→ Z cH + dZ 7→ Q Q Adu Rq. 7→ ◦ ◦ These transformations may be summarised by (A = a b , C = e f , q, u) GL(R2) GL(R2) c d h i ∈ × × H H . × × ×  

5.1.2 Classification

The calculations for classifying the super-extensions of ˆn and ˆg all follow identically. It will, ± therefore, only be stated once below. However, the central extension of the static kinematical Lie algbera is a little different, so will be treated first.

135 ˆa

Using the preliminary results from Lemma 5.1.1 in Section 5.1.1, we find b = p = z = 0. Sub- stituting these quaternions into the [B, Q, Q] and [P, Q, Q] identities with the relevant brackets, we get n2 = 0, n3 = 0 and n4 = 0 . Then, wanting [Q, Q] = 0, the [H, Q, Q] conditions tells 6 us that Re(h)= 0. Finally, Lemma 5.1.3 reduces to n0h = 0. Therefore, we have two possible cases: one in which n0 = 0 and h Im(H) and another in which h = 0 and n0 is unconstrained. ∈ In the former case, the subgroup G GL(s¯) GL(s¯) can be used to set h = k and n1 = 1, ⊂ 0 × 1 such that the only non-vanishing brackets involving Q are

[H, Q(θ)] = Q(θk) and [Q(θ), Q(θ)] = |θ|2Z. (5.1.2.1)

Notice, however, that this case also allows for h = 0, leaving only

[Q(θ), Q(θ)] = |θ|2Z. (5.1.2.2)

In the latter case, we can use G to scale n0 and n1, so the non-vanishing brackets are

[Q(θ), Q(θ)] = |θ|2H + |θ|2Z. (5.1.2.3)

ˆn and ˆg ± Using the preliminary results of Lemmas 5.1.1 and 5.1.3, we instantly find b = p = z = 0, and, 1 subsequently, n0h = 2 n2. The super-Jacobi identity [B, Q, Q] then tells us that n0 = n2 = n4 = 0 and the identity [P, Q, Q] gives us n3 = 0. Thus, the (s1¯, s1¯, s1¯) condition is trivially satisfied. The only remaining condition is from [H, Q, Q], which tells us n1 Re(h) = 0. Since we want [Q, Q] = 0, we must have n1 = 0, therefore, h Im(H). Using G to set h = k and n1 = 1, we 6 6 ∈ have non-vanishing brackets

[H, Q(θ)] = Q(θk) and [Q(θ), Q(s)] = |θ|2Z. (5.1.2.4)

Similar to the ˆa case, the restriction h Im(H) does not remove the choice h = 0. Therefore, ∈ we may also have [Q(θ), Q(s)] = |θ|2Z (5.1.2.5) as the only non-vanishing bracket.

5.1.3 Summary

Table 5.3 lists all the N = 1 generalised Bargmann superalgebras we have classified. Each Lie superalgebra is an N = 1 super-extension of one of the generalised Bargmann algebras given in Table 5.2, taken from [3]. It is interesting to compare this list of N = 1 super-extensions of centrally-extended kinematical Lie algebras to the list of centrally-extended N = 1 kinematical Lie superalgebras given in Table 5.4. This table is a reduced and adapted version of one given in Section 4.1.5, where we have only kept those extensions built upon the static, Newton-Hooke, and Galilean algebras.

Notice that only one of the generalised Bargmann superalgebras is present in the classification of centrally-extended kinematical Lie superalgebras, S1. Although it does not match exactly, we can use the basis transformations in G GL(s¯) GL(s¯) to bring it into the same form as ⊂ 0 × 1 the second superalgebra in Table 5.4.

The fact that these tables only have one Lie superalgebra in common is, perhaps, unsurprising. The Lie superalgebras presented in Table 5.3 almost exclusively have [Q, Q]= Z. Thus, before introducing the new generator Z, these Lie superalgebras would have had [Q, Q] = 0. By con- struction, such Lie superalgebras were left out of the classification in Section 4.1.2. This may explain why there is so little crossover between these tables.

136 Table 5.3: N = 1 Generalised Bargmann Superalgebras (with [Q, Q] = 0) 6

S k h z b p [Q(θ), Q(θ)] 1 ˆa |θ|2Z 2 ˆa k |θ|2Z 3 ˆa |θ|2H + |θ|2Z 2 4 ˆn− |θ| Z 2 5 ˆn− k |θ| Z 2 6 ˆn+ |θ| Z 2 7 ˆn+ k |θ| Z 8 ˆg |θ|2Z 9 ˆg k |θ|2Z

The first column gives each generalised Bargmann superalgebra s a unique identifier, and the second column tells us the underlying generalised Bargmann algebra k. The next four columns tells us how the s0¯ generators H,Z, B, and P act on Q. Recall, the action of J is fixed, so we do not need to state this explicitly. The final column then specifies the [Q, Q] bracket.

Table 5.4: Central Extensions of N = 1 Kinematical Lie Superalgebras

k [B(β), P(π)] h z b p [Q(θ), Q(θ)] 1 2 ¯ a 2 k |θ| Z − P(θkθ) a Re(βπ¯ )Z |θ|2H a |θ|2Z − B(θjθ¯)− P(θkθ¯) a |θ|2Z − P(θkθ¯) g k |θ|2Z − P(θkθ¯) g |θ|2Z − P(θkθ¯) 1 2Z B ¯ P ¯ n+ 2 j |θ| − (θiθ)− (θkθ)

The first column identifies the kinematical Lie algebra k underlying the extensions. The second column indicates whether the central extension has been introduced in the [B, P] bracket. The next four columns show the [s0¯, s1¯] brackets for the KLSA. As we can see, the only non-vanishing case is [H, Q(θ)] = Q(θh), where θ H and h Im(H). The final column then tells us whether ∈ ∈ the central extension enters the [Q, Q] bracket.

Unpacking the Notation

Although the quaternionic formulation of these superalgebras is convenient for our purposes, it is perhaps unfamiliar to the reader. Therefore, in this section, we will convert one of the N = 1 super-extension for the Bargmann algebra (S9) into a more conventional format. The brackets for this algebra, excluding the s0¯ brackets, which are shown in Table 5.2, take the form

[H, Q(θ)] = Q(θk) and [Q(θ), Q(θ)] = |θ|2Z. (5.1.3.1)

4 Letting Q(θ)= a=1 θaQa, where θ = θ4 + θ1i + θ2j + θ3k, we can rewrite these brackets as

P 4 b [H, Qa]= QbΣ a and [Qa,Qb]= δabZ, (5.1.3.2) a=1 X where, for σ2 being the second Pauli matrix,

0 iσ Σ = 2 . (5.1.3.3) −iσ 0  2 

137 5.2 Classification of N = 2 Generalised Bargmann Super- algebras

Having established the introduction of the central extension Z into our classification problem in Section 5.1, we now look to introduce an additional so(3) spinor module. Section 5.2.1 will describe the setup up for the classification of the N = 2 generalised Bargmann superalgebras, before extending the preliminary results from Section 5.1.1 to this case. It is also in this section that the group of basis transformations G will be adapted for extended supersymmetry. The number of additional parameters in this case means that there are several branches of super- extension for each generalised Bargmann algebra. In Section 5.2.2, we use the preliminary results from the (s0¯, s0¯, s1¯) component of the super-Jacobi identity to identify four possible branches of generalised Bargmann superalgebra in the sub-variety S J . Each branch is ⊂ then explored in detail in Section 5.2.3 where we identify the non-empty sub-branches, which are summarised in Section 5.2.4.

5.2.1 Setup for the N = 2 Calculation

Recall that, in addition to the standard kinematical Lie brackets, the brackets for the universal generalised Bargmann superalgebra are

[B(β), P(π)] = Re(βπ¯ )Z [H, B(β)] = λB(β)+ µP(β) (5.2.1.1) [H, P(π)] = ηB(π)+ εP(π), where β, π Im(H) and λ, µ, η, ε R. Because we now have two spinor modules, the brackets ∈ ∈ involving s1¯ need to be adapted from those given in Section 5.1.1. We will continue to use the map Q for the odd dimensions; however, it now acts on θ, a vector in H2. We will choose H2 to be a left quaternionic vector space such that H acts linearly from the left and all 2 2 H × matrices act on the right. Therefore, writing θ out in its components, we have

θ = θ1 θ2 , (5.2.1.2)  where θ1, θ2 H. The [s¯, s¯] brackets are again the so(3)-equivariant endomorphisms of s¯. ∈ 0 1 1 Since we choose so(3) to act via left quaternionic multiplication, the commuting endomorphisms are all those that may act on the right. In the present case, these are elements of Mat2(H). Thus the brackets containing the so(3) scalars are

[H, Q(θ)] = Q(θH) (5.2.1.3) [Z, Q(θ)] = Q(θZ), where H, Z Mat2(H). In Section 5.1.1, [J, Q], [B, Q] and [P, Q] were described by Clifford ∈ multiplication V S S, which is given by left quaternionic multiplication by Im(H). The ⊗ → space of such maps was isomorphic to the space of r-equivariant maps of S, which is a copy of the quaternions. Now with s¯ = S S, we have four possible endomorphisms of this type. 1 ⊕ Labelling the two spinor modules S1 and S2, we may use the Clifford action to map S1 to S1, S1 to S2, S2 to S1, or S2 to S2. All of these maps may be summarised as follows

1 [J(ω), Q(θ)] = 2 Q(ωθ) [B(β), Q(θ)] = Q(βθB) (5.2.1.4) [P(π), Q(θ)] = Q(πθP).

Here, ω, β, π Im(H) and B, P Mat2(H). Finally, consider the [Q, Q] bracket. This will consist ∈ ∈ 2 of the so(3)-equivariant R-linear maps s¯ s¯. To write down these maps, we make use of 1 → 0 the so(3)-invariant inner product on s¯ 1J 2 T θ, θ ′ = Re(θθ†) where θ, θ ′ H θ† = θ¯ . (5.2.1.5) h i ∈

138 This bracket’s so(3)-invariance is clear on considering left multiplication by u sp(1) and noting ∼ 2 ∈ sp(1) = so(3). We can now use this bracket to identify s1¯ with the symmetric R-linear ∼ 2 ∼ 2 2 2 endomorphisms of s¯ = S = H , i.e. the maps µ : H H such that µ(θ), θ ′ = θ, µ(θ ′) . A 1 → J h i h i general R-linear map of H2 may be written

µ(θ)= qθM where q H and M Mat2(H). (5.2.1.6) ∈ ∈ Now, inserting this definition into the condition for a symmetric endomorphism, we obtain the following two cases: 1. q R and M = M ∈ † 2. q Im(H) and M =−M†. ∈ 2 The first instance gives us our so(3) scalar modules in s1¯; therefore, these will map to either H or Z in s¯ to ensure we have so(3)-equivariance. The condition on M states that it must be 0 J of the form a b + m 1 0 0 1 0 0 0 1 M = = a + b + c + m , (5.2.1.7) b − m c 0 0 1 0 0 1 −1 0           where a, b, c R and m Im(H). We can make sense of this result using the decompo- ∈ ∈ ∼ 2 ∼ 2 sition of the odd part of the superalgebra, s1¯ = S = S R , and our knowledge of the 2 ⊗ maps in Hom( S, s0¯) derived from Section 4.1.1. Symmetrising the decomposed s1¯, we get 2 S2 =∼ 2 S 2 R2 2 S 2 R2. Recall that 2 S =∼ R 3V, where the scalar compo- J⊗ ⊕ ⊗ ⊕ nent is equivalent to the real span of the endomorphism L1R1, and the vector components are J J J V V J equivalent to the real span of the endomorphisms LqRq′ , where q, q′ Im(H). Notice, the scalar ∈ component is the only map that requires multiplication on the left by a real scalar. Since case 1 demands that the spinor module is multiplied on the left by a real scalar, we would expect that the terms in M which are associated with the 2 S 2 R2 component of 2 S2 would ⊗ also multiply the spinor module on the right by a real scalar. Indeed, we find that the first 2 J2 J J three terms of M correspond to maps inside S R2: they all take the form of the scalar 2 ⊗ 2 2 component of S, L1R1, multiplied by a basis element of R . J J J J To complete our decomposition of M for this case, note that Λ2S =∼ 3R V, where the three ⊕ copies of R correspond to the endomorphisms L1Ri, L1Rj, and L1Rk and V corresponds to spanR LiR1, LjR1, LkR1 . Again, case 1 imposes that we only consider the maps containing L1, which can be succinctly written as an imaginary quaternion acting on the right. Thus, the  final term in our decomposition is part of a map inside Λ2S Λ2R2 and takes the form of an ⊗ imaginary quaternion multiplied by the basis element of Λ2R2.

Now, the second case gives us our so(3) vector modules in 2 S2; therefore, these will map to B, P, or J to ensure so(3)-equivariance. The condition on M in this case produces J n d + l 1 0 0 1 0 0 0 1 M = = n + l + r + d , (5.2.1.8) −d + l r 0 0 1 0 0 1 −1 0           where n, l, r Im(H) and d R. Again, we can understand this result through the decomposi- ∈ ∈ 2 2 ∼ 2 2 2 2 2 2 tion of s1¯ and its symmetrisation, S = S R S R . From Section 4.1.1, 2 ⊗ ⊕ ⊗ we know that the so(3) vectors in S come from simultaneous left and right quaternionic J J J V V multiplication by Im(H). Since case 2 imposes that we must consider the maps which multiply J s1¯ on the left by an imaginary quaternion, we may expect that the M associated with the 2 S 2 R2 component of 2 S2 will also multiply on the right by an imaginary quaternion. ⊗ 2 2 Indeed, the first three components of M correspond to maps inside S R2: they con- J J 2 J ⊗ 2 2 sist of a map in S of the form LqRq′ , for q, q′ Im(H), multiplied by a basis element of R . ∈ J J J 2 2 2 J The final term in our decomposition comes from the S R2 component of S2. Re- ⊗ call, the only anti-symmetric endomorphisms of S which involve multiplication on the left by V V J an imaginary quaternion are in spanR LiR1, LjR1, LkR1 , which transforms as an so(3) vec- tor module. Therefore, we would anticipate that the terms in M associated with 2 S 2 R2  ⊗ V V 139 would correspond to a real scalar multiplied by the Λ2R2 basis element, which is indeed the case.

Putting all this together, we may write

[Q(θ), Q(θ)] = θ, θN0 H + θ, θN1 Z + θ, JθN2 + θ, BθN3 + θ, PθN4 , (5.2.1.9) h i h i h i h i h i where N0, N1 are quaternion Hermitian, and N2, N3, N4 are quaternion skew-Hermitian, as stated above, and

J = J1i + J2j + J3k B = B1i + B2j + B3k P = P1i + P2j + P3k. (5.2.1.10)

Using the fact that Re(ω¯J)= J(ω) and N = N† for i {0,1}, we can write i i ∈

[Q(θ), Q(θ)] = Re(θN0θ†)H + Re(θN1θ†)Z − J(θN2θ†)− B(θN3θ†)− P(θN4θ†). (5.2.1.11)

This polarises to 1 [Q(θ), Q(θ ′)] = Re(θN θ ′† + θ ′N θ†)H + Re(θN θ ′† + θ ′N θ†)Z 2 0 0 1 1 − J(θN2θ ′† + θ ′N2θ†)− B(θN3θ ′† + θ ′N3θ†)− P(θN4θ ′† + θ ′N4θ†) . (5.2.1.12) 

Preliminary Results

As in the N = 1 case, we can form a number of universal results that will help us when inves- tigating the super-extensions of the generalised Bargmann algebras. The following subsections will cover the (s0¯, s0¯, s1¯), (s0¯, s1¯, s1¯), and (s1¯, s1¯, s1¯) components of the super-Jacobi identity, respectively.

(s0¯, s0¯, s1¯)

In the N = 1 case, h, z, b, p H, and in the N = 2 case H, Z, B, P Mat2(H). Since H ∈ ∈ and Mat2(H) are both associative, non-commutative algebras, the algebraic manipulations are the same in both cases. Therefore, the N = 1 results generalise to the N = 2 case; the only difference being that the variables are 2 2 H matrices rather than H elements. ×

Lemma 5.2.1. The following relations between H, Z, B, P Mat2(H) are implied by the corre- ∈ sponding k-brackets:

[H, Z]= λH + µZ = [Z, H]= λH + µZ ⇒ [H, B]= λB + µP = [B, H]= λB + µP ⇒ [H, P]= λB + µP = [P, H]= λB + µP ⇒ [Z, B]= λB + µP = [B, Z]= λB + µP ⇒ (5.2.1.13) [Z, P]= λB + µP = [P, Z]= λB + µP ⇒ [B, B]= λB + µP + νJ = B2 = 1 λB + 1 µP + 1 ν ⇒ 2 2 4 [P, P]= λB + µP + νJ = P2 = 1 λB + 1 µP + 1 ν ⇒ 2 2 4 [B, P]= λH + µZ = BP + PB = 0 and [B, P]= λH + µZ. ⇒ Proof. See the proof of Lemma 5.1.1 for the algebraic manipulations that produce the above results.

(s0¯, s1¯, s1¯)

As in the N = 1 case, we use the universal generalised Bargmann algebra to simplify our

140 analysis here. Recall the brackets for this algebra are

[B, P]= Z [H, B]= λB + µP [H, P]= ηB + εP, (5.2.1.14) where λ, µ, η, ε R. Using these brackets, we obtain the following result. ∈

Lemma 5.2.2. The [H, Q, Q] identity produces the conditions

0 = HN + N H† where i {0,1,2} i i ∈ λN3 + ηN4 = HN3 + N3H†

µN3 + εN4 = HN4 + N4H†. The [Z, Q, Q] identity produces the conditions

0 = ZN + N Z† where i {0,1,2,3,4}. i i ∈ The [B, Q, Q] identity produces the conditions

0 = BN0 − N0B†

N4 = BN1 − N1B†

0 = βθBN2θ† + θN2(βθB)† 1 λ Re(θN0θ†)β + 2 [β, θN2θ†]= βθBN3θ† + θN3(βθB)†

µ Re(θN0θ†)β = βθBN4θ† + θN4(βθB)†. The [P, Q, Q] identity produces the conditions

0 = PN0 − N0P†

−N3 = PN1 − N1P†

0 = πθPN2θ† + θN2(πθP)†

η Re(θN0θ†)π = πθPN3θ† + θN3(πθP)† 1 ε Re(θN0θ†)π + 2 [π, θN2θ†]= πθPN4θ† + θN4(πθP)†, (5.2.1.15) where β, π Im(H) and θ H2. ∈ ∈

Proof. Beginning with the [H, Q, Q] identity, we have

[H, [Q(s), Q(s)]] = 2[[H, Q(s)], Q(s)]. (5.2.1.16)

Focussing on the L.H.S., note the general form of the [Q, Q] bracket has components along each of the s0¯ basis elements; however, H only commutes with B and P. Therefore,

L.H.S. = −[H, B(θN3θ†)]−[H, P(θN4θ†)] (5.2.1.17) =−B(θ(λN3 + ηN4)θ†)− P(θ(µN3 + εN4)θ†).

Substituting [H, Q(θ)] = Q(θH) into the R.H.S. and using the polarised form of the [Q, Q] bracket, we find

R.H.S. = Re(θ(HN0 + N0H†)θ†)H + Re(θ(HN1 + N1H†)θ†)Z (5.2.1.18) − J(θ(HN2 + N2H†)θ†)− B(θ(HN3 + N3H†)θ†)− P(θ(HN4 + N4H†)θ†).

Comparing coefficients and using the injectivity and linearity of the maps J, B and P, we get the desired conditions. The [Z, Q, Q] result follows in an analogous manner. Consider the [B, Q, Q] Jacobi identity [B(β), [Q(θ), Q(θ)]] = 2[[B(β), Q(θ)], Q(θ)]. (5.2.1.19)

141 Since B commutes with Z and B, the L.H.S. takes the following form

L.H.S. =[B(β), Re(θN0θ†)H − J(θN2θ†)− P(θN4θ†)] (5.2.1.20) B P 1 B ¯ Z =− Re(θN0θ†)(λ (β)+ µ (β))+ 2 ([θN2θ†, β]) − Re(βθN4θ†) . Turning attention to the R.H.S., we find

R.H.S. = Re(βθBN0θ† + θN0B†θ†β¯)H + Re(βθBN1θ† + θN1B†θ†β¯)Z

− J(βθBN2θ† + θN2B†θ†β¯)− B(βθBN3θ† + θN3B†θ†β¯)− P(βθBN4θ† + θN4B†θ†β¯). (5.2.1.21)

Using the property β¯ =−β, since β Im(H), and the cyclic property of Re, the first two terms ∈ can have their coefficients written in the form

Re(βθBNiθ† + θNiB†θ†β¯)= Re(βθ(BNi − NiB†)θ†), (5.2.1.22) for i {0,1}. Again, comparing coefficients we obtain the desired results. The [P, Q, Q] case ∈ follows identically.

(s1¯, s1¯, s1¯)

The last super-Jacobi identity component to consider is the (s1¯, s1¯, s1¯) case, [Q, Q, Q].

Lemma 5.2.3. The [Q, Q, Q] identity produces the condition

1 Re(θN0θ†)θH + Re(θN1θ†)θZ = 2 θN2θ†θ + θN3θ†θB + θN4θ†θP. (5.2.1.23)

Proof. The [Q, Q, Q] identity is written

0 = [[Q(θ), Q(θ)], Q(θ)]. (5.2.1.24)

Substituting in the [Q, Q] bracket, this becomes

0 =[Re(θN0θ†)H + Re(θN1θ†)Z − J(θN2θ†)− B(θN3θ†)− P(θN4θ†), Q(θ)]. (5.2.1.25)

Finally, using the brackets in (5.2.1.3) and (5.2.1.4) and the injectivity of Q, we obtain the desired result.

Basis Transformations

We will investigate the subgroup G GL(s¯) GL(s¯) by first looking at the transformations ⊂ 0 × 1 induced by the adjoint action of the rotational subalgebra r =∼ so(3). We will then look at the so(3)-equivariant maps transforming the basis of the underlying vector space. These will act 2 via Lie algebra automorphisms in s0¯ and endomorphisms of the so(3) module S in s1¯. Note, in the former case, where the automorphism is induced by adJi , each so(3) module will transform into itself, while, in the latter case, when the transformation is some so(3)-equivariant map, the modules transform into one another. For completeness, at the end of the section, we determine the automorphisms of each generalised Bargmann algebra.

Recall that Sp(1) is the double-cover of Aut(H), and Aut(H) =∼ SO(3). We, therefore, write λ Aut(H) as λ(s)= usu¯ for some u Sp(1), which will act trivially on the real component of s ∈ ∈ and rotate the imaginary components. Using this result, we can represent the action of Aut(H) on the so(3) vector modules in s0¯ by pre-composing the linear maps J, B, and P with Adu, for u Sp(1). To preserve the kinematical brackets in [s¯, s¯], we must pre-compose with the ∈ 0 0 same u for each map. Note, so(3) acts trivially on H and Z, so these basis elements will be left invariant under these automorphisms. For s1¯, we restrict to the individual copies of S through diagonal matrices. To preserve the [J, Q] bracket, we must pre-compose with the same u as above. Therefore, we write Q˜(θ)= Q(uθu1¯ ). We can now investigate how these automorphisms

142 affect our brackets

[J(ω), Q(θ)] = 1 Q(ωθ) 2 [H, Q(θ)] = Q(θH) [B(β), Q(θ)] = Q(βθB) [Z, Q(θ)] = Q(θZ) [P(π), Q(θ)] = Q(πθP) (5.2.1.26)

[Q(θ), Q(θ)] = Re(θN0θ†)H + Re(θN1θ†)Z − J(θN2θ†)− B(θN3θ†)− P(θN4θ†).

Transforming the basis, we have

[˜J(ω), Q˜(θ)] = 1 Q˜(ωθ) 2 [H˜, Q˜(θ)] = Q˜(θH˜) [B˜(β), Q˜(θ)] = Q˜(βθB˜) [Z˜, Q˜(θ)] = Q˜(θZ˜) [P˜(π), Q˜(θ)] = Q˜(πθP˜) (5.2.1.27)

[Q˜(θ), Q˜(θ)] = Re(sθN0θ†)H˜ + Re(θN1θ†)Z˜ − ˜J(θN2θ†)− B˜(θN3θ†)− P˜(θN4θ†), with H˜ = H, Z˜ = Z, ˜J f= J Adu, B˜ =fB Adu, P˜ f= P Adu,f and Q˜ = Q fAdu, where it is ◦ ◦ ◦ ◦ understood that Adu acts diagonally on the s1¯ basis, Q. The transformed matrices are

H˜ = DHD−1 B˜ = DBD−1 Ni = DNiD†, (5.2.1.28) Z˜ = DZD−1 P˜ = DPD−1

f −1 where D = u1 for u Sp(1) and i {0, 1, ..., 4}. Therefore, D = D† = u1¯ . These automor- ∈ ∈ phisms simultaneously rotate all quaternions, all the components of the matrices H, Z, B, P and Ni, by the same Sp(1) element.

Next, we want to consider the so(3)-equivariant linear maps which leave the rotational sub- algebra invariant: (J, B, P, H, Z, Q) (J, B˜, P˜, H˜, Z˜, Q˜). These take the general form → H˜ = aH + bZ Z˜ = cH + dZ B˜(β)= eB(β)+ fP(β)+ gJ(β) (5.2.1.29) P˜(π)= hB(π)+ iP(π)+ jJ(π) Q˜(θ)= Q(θM), where a, ..., j R and M GL(H2). Crucially, ∈ ∈ e f g a b A = GL(2, R) and C = h i j GL(3, R), (5.2.1.30) c d ∈   ∈   0 0 1   act on (H, Z)T and (B, P, J)T , respectively. Each of the generalised Bargmann algebra allows different transformations of this type; however, there are some important general results. There- fore, we will begin by working through the analysis of these maps with the universal generalised Bargmann algebra before focussing on each algebra separately.

As in the N = 1 case, the checking of brackets that include J is really verifying that the above maps are so(3)-equivariant, so this does not give us any information not already pre- sented. The first bracket we will consider is [B, P]= Z. Substituting in the maps of (5.2.1.29), we find the following important results:

d = ei − fh, c = 0, and g = j = 0. (5.2.1.31)

143 The vanishing of c tells us that d = 0 if we are to have A GL(2, R). Also, the vanishing of g 6 ∈ and j shows that we can reduce C to an element of GL(2, R),

e f C = , (5.2.1.32) h i   T acting on (B, P) . The remaining [s0¯, s0¯] brackets are [H, B] and [H, P], which produce

0 = λe(a − 1)+ ηaf − µh 0 = η(e − ai)+ εh − λah and (5.2.1.33) 0 = λf − εaf + µ(i − ea) 0 = ηf + εi(1 − a)− µah, respectively. Clearly, these conditions are dependent on the exact choice of generalised Bargmann algebra, so we will leave these results in this form for now.

Now, since the [s0¯, s1¯] and [s1¯, s1¯] brackets are so far independent of the chosen algebra, the following results will hold for all the generalised Bargmann algebras. Reusing (5.2.1.27), in this instance we find

H˜ = M(aH + bZ)M−1 Z˜ = dMZM−1 B˜ = M(eB + fP)M−1 P˜ = M(hB + iP)M−1

N = MN M 1 2 2 † N0 = MN0M† 1 a Nf3 = M(iN3 − hN4)M† (5.2.1.34) 1 ie − fh Nf1 = M(aN1 − bN0)M† 1 ad Nf = M(eN − fN )M†. 4 ie − fh 4 3 f Putting the two types of transformation inf G together, we have

J J Adu 7→ ◦ B eB Adu +fP Adu 7→ ◦ ◦ P hB Adu +iP Adu 7→ ◦ ◦ (5.2.1.35) H aH + bZ 7→ Z dZ 7→ Q Q Adu R . 7→ ◦ ◦ M a b e f 2 2 These transformations may be summarised by (A = 0 d , C = h i , M, u) GL(R ) GL(R ) 2 ∈ × × GL(H ) H×. Now that we have the most general element of the subgroup G GL(s¯) GL(s¯) ×   ⊂ 0 × 1 for s0¯ = k the universal generalised Bargmann algebra, we can restrict ourselves to the auto- morphisms of s¯ and set the parameters λ, µ, η, ε R to determine the automorphism group for 0 ∈ each of the generalised Bargmann algebras. The results of this investigation are presented in Table 5.5.

ˆa

In this instance, all the conditions vanish as λ = µ = η = ε = 0; therefore, the matrices A and C are left as stated above.

ˆn−

Having λ =−ε = 1 and µ = η = 0, the conditions in (5.2.1.33) become

0 = e(a − 1) 0 = h(1 + a) and (5.2.1.36) 0 = f(1 + a) 0 = i(1 − a).

144 Notice, if a / { 1} then C must vanish, which cannot happen if we are to retain the basis ∈ ± elements B and P. Therefore, we are left with two cases: a = 1 and a = −1. In the former instance, we have automorphisms with

1 b e 0 A = and C = . (5.2.1.37) 0 ei 0 i     In the latter instance, we have

−1 b 0 h A = and C = . (5.2.1.38) 0 −hf f 0    

ˆn+

In this case, λ = ε = 0 and µ =−η = 1. Therefore, our constraints become

0 = h + af 0 = e − ai and (5.2.1.39) 0 = i − ae 0 = f + ah.

Taking the expressions for h and i from the conditions on the left and substituting them into the conditions on the right, we find

0 =(1 − a2)f 0 =(1 − a2)e. (5.2.1.40)

If a2 = 1, we would need both f and e to vanish, which contradicts our assumption that 6 C GL(2, R). Therefore, we need a2 = 1, which presents two cases: a = 1 and a =−1. In the ∈ former instance, we find automorphisms of the form

1 b e h A = and C = . (5.2.1.41) 0 e2 + h2 −h e     In the latter instance, we get

−1 b e h A = and C = . (5.2.1.42) 0 −e2 − h2 h −e     ˆg

Finally, we have λ = η = ε = 0 and µ = −1, which, when substituted into (5.2.1.33), pro- duces 0 = h and i = ae. (5.2.1.43) Therefore, automorphisms for the Bargmann algebra take the form

a b e f A = and C = . (5.2.1.44) 0 ae2 0 ae    

5.2.2 Establishing Branches Before proceeding to the discussion in which the non-empty sub-branches of S are identified, we first establish the possible [s0¯, s1¯] brackets. More specifically, we establish the possible forms for Z, H, B, P Mat2(H). In this section, we focus solely on the results of Lemma 5.2.1 con- ∈ cerning the (s0¯, s0¯, s1¯) component of the super-Jacobi identity. Using the universal generalised Bargmann algebra, we find that B, P Mat2(H), which encode the brackets [B, Q] and [P, Q], ∈ respectively, form a double complex. Analysing this structure, we identify four possible cases: 1. B = 0 and P = 0 2. B = 0 and P = 0 6

145 Table 5.5: Automorphisms of the Generalised Bargmann Algebras

k General (A, C) GL(R2) GL(R2) ∈ × a b e f ˆa , 0 d! h i!! 1 b e 0 −1 b 0 h ˆn− , , 0 ei! 0 i!! ∪ 0 −hf! f 0!! 1 b e h −1 b e h ˆn+ , , 0 e2 + h2! −h e!! ∪ 0 −e2 − h2! h −e!! a b e f ˆg , 0 ie! 0 ae!!

3. B = 0 and P = 0 6 4. B = 0 and P = 0. 6 6 Taking each of these cases in turn, we find forms for Z and H to establish four branches in S which may contain generalised Bargmann superalgebras. These branches will form the basis for our investigations into the possible super-extensions for each of the generalised Bargmann algebras in Section 5.2.3.

Using the results of Lemma 5.2.1, we notice that B2 = P2 = 0 and BP + PB = 0; therefore, B and P are the differentials of a double complex in which the modules are s1¯. What does this mean for the form of B and P? Notice that we could simply set B and P to zero. However, assuming at least one component of these matrices is non-vanishing, we find the following cases. Take P as our example and let p p P = 1 2 . (5.2.2.1) p p  3 4 The fact that this squares to zero tells us

2 2 p1 + p2p3 = 0 p1p2 + p2p4 = 0 p3p1 + p4p3 = 0 p3p2 + p4 = 0. (5.2.2.2)

There are two cases, p3 = 0 and p3 = 0, which we shall now consider in turn. 6

In the p3 = 0 case, the constraints in (5.2.2.2) become

2 2 p1 = 0 p1p2 + p2p4 = 0 p4 = 0. (5.2.2.3)

Therefore, p1 = p4 = 0 and p2 is unconstrained, leaving the matrix

0 p P = 2 . (5.2.2.4) 0 0   2 −1 In the p3 = 0 case, we can use the first and third constraints of (5.2.2.2) to get p2 =−p1p3 and 6 −1 p4 =−p3p1p3 , respectively. These choices trivially satisfy the second and fourth constraints such that we arrive at p −p2p−1 P = 1 1 3 . (5.2.2.5) p −p p p−1  3 3 1 3  In a completely analogous manner, we find

0 b b −b2b−1 B = 2 and B = 1 1 3 . (5.2.2.6) 0 0 b −b b b−1    3 3 1 3  Now, what does the anti-commuting condition, BP + PB = 0, tell us about the non-vanishing matrices? Assuming, for now, that B and P are non-vanishing, we have four options:

146 1. p3 = 0, b3 = 0, 6 6 2. p3 = 0, b3 = 0, 6 3. p3 = 0, b3 = 0, and 6 4. p3 = 0, b3 = 0.

Option 1 Here, we will find three distinct sub-options. Interestingly, these three sub-options are equivalent to options 2, 3, and 4 above. Substituting the matrices associated with p3 = 0 6 and b3 = 0 into BP + PB = 0 gives us 6 2 −1 2 −1 0 = b1p1 − b1b3 p3 + p1b1 − p1p3 b3 2 −1 2 −1 −1 2 −1 2 −1 −1 0 =−b1p1p3 + b1b3 p3p1p3 − p1b1b3 + p1p3 b3b1b3 −1 −1 (5.2.2.7) 0 = b3p1 − b3b1b3 p3 + p3b1 − p3p1p3 b3 2 −1 −1 −1 2 −1 −1 −1 0 =−b3p1p3 + b3b1b3 p3p1p3 − p3b1b3 + p3p1p3 b3b1b3 .

−1 Multiplying the first of these conditions on the right by b1b3 and adding it to the second condition, we obtain −1 −1 −1 0 = b1(p1 − b1b3 p3)(b1b3 − p1p3 ). (5.2.2.8) Since the quaternions have no zero-divisors, one of these terms must vanish. The vanishing of the second is equivalent to the vanishing of the third, so we have two sub-options:

1.1 b1 = 0, and

−1 −1 1.2 b1b3 = p1p3 . In the latter case, the third and fourth conditions of (5.2.2.7) are trivially satisfied, but in the former case, a little more work is required. Setting b1 = 0, we obtain

2 −1 −1 0 = b3p1p3 and 0 = b3p1 − p3p1p3 b3. (5.2.2.9)

Again, using the fact the quaternions have no zero-divisors, these conditions mean this sub- option further divides into two:

1.1.1 b3 = 0, and

1.1.2 p1 = 0, with p3 left free. Recall that to arrive at these options we first made a choice to multiply the −1 −1 first condition of (5.2.2.7) by b1b3 . We could equally have multiplied by p1p3 such that case 1.1 above read p1 = 0. (Notice, the second case is symmetric, so would remain the same in this instance.) Analogous subsequent calculations would lead to sub-options p3 = 0 and b1 = 0. Putting all of this together, we have four sub-options to consider:

p −p2p−1 Sub-option 1: B = 0 P = 1 1 3 p −p p p−1  3 3 1 3  0 0 0 0 Sub-option 2: B = P = b 0 p 0  3   3  b −b2b−1 Sub-option 3: B = 1 1 3 P = 0 b −b b b−1  3 3 1 3  b −b2b−1 p −p2p−1 Sub-option 4: B = 1 1 3 P = 1 1 3 where b b−1 = p p−1. b −b b b−1 p −p p p−1 1 3 1 3  3 3 1 3   3 3 1 3  (5.2.2.10) In fact, this list can be simplified further. For all generalised Bargmann algebras, we can choose a transformation (1, 1, M,1), where, M takes the form

1 −b b−1 M = 1 3 , (5.2.2.11) 0 1  

147 such that sub-option 4 becomes sub-option 2. In summary, the p3 = 0 and b3 = 0 assumption 6 6 lead to three separate sub-options.

p −p2p−1 Sub-option 1: B = 0 P = 1 1 3 p −p p p−1  3 3 1 3  0 0 0 0 Sub-option 2: B = P = (5.2.2.12) b 0 p 0  3   3  b −b2b−1 Sub-option 3: B = 1 1 3 P = 0. b −b b b−1  3 3 1 3 

Option 2 Letting p3 = 0 and b3 = 0, the anti-commuting condition tells us 6 −1 0 = b2p3 and p1b2 = b2p3p1p3 . (5.2.2.13)

Using the first condition, b2 = 0, and, with b2 = 0, we are left with sub-option 1 above.

Option 3 Now, consider p3 = 0 and b3 = 0. Substituting the relevant forms of P and B into 6 the anti-commuting condition, BP + PB = 0, we find

−1 0 = p2b3 and b1p2 = p2b3b1b3 . (5.2.2.14)

This is identical to option 2 only b and p have been swapped. Therefore, we have a similar result: p = 0 such that we have sub-option 3 above.

Option 4 The final case to consider is p3 = 0 and b3 = 0, where

0 p 0 b P = 2 and B = 2 . (5.2.2.15) 0 0 0 0     These strictly upper-triangular matrices are equivalent to the strictly lower-triangular matrices of sub-option 2 above. Thus, again, we find no new cases to carry forward.

To simplify the rest of the calculations, we will choose to use the transformation in (5.2.2.11) for all generalised Bargmann algebras and all options. Combining the case in which both B and P vanish with the non-vanishing options, we find

Case 1: B = 0 P = 0 0 0 Case 2: B = 0 P = p 0  3  0 0 0 0 (5.2.2.16) Case 3: B = P = b 0 p 0  3   3  0 0 Case 4: B = P = 0. b 0  3  In all cases, it is a straight-forward computation to show that [B, P] = Z tells us that Z = 0. Therefore, we are left with only H to determine. From the results in Lemma 5.2.1, the conditions we have including B, P and H are

[B, H]= λB + µP and [P, H]= ηB + εP. (5.2.2.17)

Case 1 The vanishing of B and P in this instance, when substituted into (5.2.2.17), means we do not obtain any conditions on H. Thus, we find a branch with matrices

B = P = Z = 0 and H unconstrained. (5.2.2.18)

148 Case 2 Notice that the vanishing of B means that the second condition in (5.2.2.17) becomes

0 0 0 0 h h h h 0 0 ε = 1 2 − 1 2 . (5.2.2.19) p 0 p 0 h h h h p 0  3   3   3 4  3 4  3  This gives us two constraints

0 = h2p3 and εp3 = p3h1 − h4p3. (5.2.2.20)

The first constraint here tells us that either h2 or p3 must vanish. In the latter instance, we recover the matrices from case 1. In the former instance, we can use the second constraint to write h4 in terms of h1. Thus, we find a branch with matrices

0 0 h 0 B = Z = 0 P = H = 1 . (5.2.2.21) p 0 h p h p−1 − ε  3   3 3 1 3  The first condition in (5.2.2.17) does not add any new branches to those already given as, with B = 0, it reduces to 0 = µP. Therefore, for those generalised Bargmann algebras with µ = 0, it 6 gives the branch identified in Case 1, and, for those with µ = 0, it leaves p3 free to fix h4 as prescribed for the branch presented in this case.

Case 3 Substituting the B and P associated with this case into (5.2.2.17), we get the following constraints

0 = h2b3 0 = h2p3 λb3 + µp3 = b3h1 − h4b3 ηb3 + εp3 = p3h1 − h4p3. (5.2.2.22)

The first two constraints above tell us that if h2 = 0, then we again arrive at the branch with 6 B = P = Z = 0 and H unconstrained. Letting h2 = 0, we focus on the second two constraints. Notice, for this branch to be distinct from the other two, we require b3 = 0 and p3 = 0. 6 6 These assumptions allow us to take inverses of both b3 and p3 in the following calculations. −1 Multiplying the third constraint on the right by b3 , we can rearrange for h4 and substitute this into the fourth constraint to get

−1 −1 ηb3 + εp3 = p3h1 − b3h1b3 p3 + µp3b3 p3 + λp3. (5.2.2.23)

−1 Multiplying this expression by b3 on the left and rearranging, we find

2 [u, h1]=−µu +(ε − λ)u + η, (5.2.2.24)

−1 where u = b3 p3. Alternatively, we could have chosen to multiply the fourth condition on −1 the right by p3 to get our expression for h4 and substituted this into the third constraint. −1 Multiplying this on the left by p3 produces the similar condition

2 [v, h1]= ηv +(λ − ε)v + µ, (5.2.2.25)

−1 where v = p3 b3. Depending on the generalised Bargmann algebra in question, one of these will prove more useful than the other. We will leave these constraints in this form to be analysed separately for each generalised Bargmann algebra. Note, this analysis show that, depending on the algebra in question, we may find super-extensions for which B = 0 and P = 0. Thus, we can 6 6 think of promoting this case to a branch.

Case 4 The calculations for this case are nearly identical to those for Case 2. The vanishing of P means that the first constraint in (5.2.2.17) produces

0 = h2b3 and λb3 = b3h1 − h4b3. (5.2.2.26)

149 From the first expression above, if h2 = 0, we recover the branch presented in Case 1. However, 6 setting h2 = 0, b3 is general, and we can use the second constraint to write h4 in terms of h1 and b3. Thus, we find a branch with matrices

0 0 h 0 P = Z = 0 B = H = 1 . (5.2.2.27) b 0 h b h b−1 − λ  3   3 3 1 3  The second constraint in (5.2.2.17) does not produce any new branches for B, P, and H. Substi- tuting in P = 0, it becomes 0 = ηB. Therefore, if η = 0, B must vanish leaving the branch from 6 Case 1; and, if η = 0, b3 is left free so we can write h4 as prescribed for the branch presented here.

In summary, we have the following three branches for all generalised Bargmann algebras 1. B = P = Z = 0 and H unconstrained 0 0 h 0 2. B = Z = 0 P = H = 1 p 0 h p h p−1 − ε  3   3 3 1 3  0 0 h 0 3. P = Z = 0 B = H = 1 . b 0 h b h b−1 − λ  3   3 3 1 3  There is also a possible fourth branch depending on the generalised Bargmann algebra:

h 0 0 0 0 0 Z = 0 H = 1 B = P = , (5.2.2.28) h h b 0 p 0  3 4  3   3  subject to

2 2 [u, h1]=−µu +(ε − λ)u + η and [v, h1]= ηv +(λ − ε)v + µ (5.2.2.29)

−1 −1 where u = b3 p3 and v = p3 b3.

5.2.3 Classification In this section, we complete the story started in Section 5.2.2. Each branch we identified in Section 5.2.2 encodes the possible [s0¯, s1¯] brackets for a generalised Bargmann superalgebra s. Here, we take each branch in turn and find corresponding [Q, Q] brackets. Since our interests are in supersymmetry, we will always impose the condition that [Q, Q] = 0; therefore, we are 6 only interested in branches for which at least one of the Ni matrices does not vanish. Note, the imposition of this condition means that the various branches identified here belong to the sub-variety S of the real algebraic variety cut out by the super-Jacobi identity J V . ⊂ We will begin our investigation into each branch by stating the associated matrices, B, P, H, Z ∈ Mat2(H). These matrices are then substituted into the conditions from Lemmas 5.2.2 and 5.2.3, which use the Lie brackets of the universal generalised Bargmann algebra. This process pro- duces a system of equations containing B, P, H, Z encoding the [s0¯, s1¯] components of the bracket, the matrices N for i {0, 1, ..., 4} encoding the [s¯, s¯] components of the bracket, and the four i ∈ 1 1 parameters of the universal generalised Bargmann algebra, λ, µ, η, ε R. Any conditions which ∈ do not contain one of the parameters λ, µ, η, ε are analysed and possible dependencies among the Ni matrices are found. Once these dependencies have been established, we start setting parameters to consider the various generalised Bargmann algebras. In branches 1 and 2, we will see that multiple generalised Bargmann algebras produce the same set of conditions. In these instances, we will highlight the relevant algebras but only analyse the system once to avoid repetition.

In branches 2, 3 and 4, we find that the vanishing of certain matrices Ni imposes the van- ishing of other Ni. Thus, we end up with a chain of dependencies, which lead to different sub-branches. These sub-branches will be labelled such that sub-branches with a larger branch number will have more non-vanishing matrices Ni. For example, sub-branch 2.2 may have non-vanishing N0 and N1, but sub-branch 2.3 may additionally have non-vanishing N3. Within

150 each sub-branch, we regularly find two options: one in which N0 vanishes, leaving H free, and one in which H = 0 such that N0 is unconstrained. Using sub-branch 2.2 as our example, the former instance, with N0 = 0, will be labelled 2.2.i, and the latter instance will be labelled 2.2.ii. In branch 4, we will find some instances in which both N0 and H can be non-vanishing. Using sub-branch 4.3 as an example, we will label these cases as 4.3.iii.

Each sub-branch is designed to have a unique set of non-vanishing matrices. However, the components within the matrices are not completely fixed by the super-Jacobi identity. There- fore, each sub-branch is given as a tuple (Mk, X, Ck, X), where k labels the underlying generalised Bargmann algebra, and X will be the branch number. This tuple consists of M, the subset of non-vanishing matrices in {B, P, H, Z, N0, N1, N2, N3, N4} describing the branch, and C, the set of constraints on the components of the matrices. After stating (M, C) for a given sub-branch, we proceed to a discussion on possible parameterisations of the super-extensions in the sub-branch. In particular, the aim of these discussions is to highlight the existence of super-extensions in the sub-branch. First we set as many of the parameters to zero as possible. In general, this will involve setting H to zero along with a small number of components in the matrices Ni. Then, using any residual transformations in the group G GL(s¯) GL(s¯), we fix the remaining ⊂ 0 × 1 parameters. Once the existence of super-extensions has been established, we introduce some other parameters to produce further examples of generalised Bargmann superalgebras contained within the sub-branch.

Recall, we build the [Q, Q] bracket from the Ni matrices as follows

[Q(θ), Q(θ)] = Re(θN0θ†)H + Re(θN1θ†)Z − J(θN2θ†)− B(θN3θ†)− P(θN4θ†), (5.2.3.1) where N0 and N1 are quaternion Hermitian, Ni† = Ni, and N2, N3 and N4 are quaternion skew-Hermitian, Nj† = −Nj. Throughout this section, we will use the following forms for the quaternion Hermitian matrices:

a q c r N = and N = , (5.2.3.2) 0 ¯q b 1 r¯ d     where a, b, c, d R, and q, r H. The quaternion skew-Hermitian matrices will be defined ∈ ∈ e f n m N = and N = , (5.2.3.3) 3 −¯f g 4 −m¯ l     where e, g, n, l Im(H) and f, m H. We will only briefly need to consider parts of the N2 ∈ ∈ matrix explicitly; therefore, we will define its components as necessary.

Branch 1

B = P = Z = 0 and H unconstrained. (5.2.3.4)

Using the remaining conditions from the (s0¯, s1¯, s1¯) and (s1¯, s1¯, s1¯) components of the super- Jacobi identity, we can look to find some expressions for the matrices Ni. The conditions derived from the [B, Q, Q] identity in Lemma 5.2.2 immediately give N4 = 0 due to the vanishing of B. Similarly, the [P, Q, Q] conditions give us N3 = 0 due to the vanishing of P. We are thus left with

0 = µ Re(θN0θ†)

0 = HNi + NiH† i {0,1,2} 0 = η Re(θN0θ†) ∈ β, π Im(H), θ H2. 1 1 ∀ ∈ ∀ ∈ 0 = Re(θN0θ†)θH − 2 θN2θ†θ 0 = λ Re(θN0θ†)β + 2 [β, θN2θ†] 1 0 = ε Re(θN0θ†)π + 2 [π, θN2θ†] (5.2.3.5)

151 Since [c, d] is perpendicular to both c and d for c, d H, the final two conditions can be reduced ∈ to

0 = λ Re(θN0θ†) 0 =[β, θN2θ†] 0 = ε Re(θN0θ†) 0 =[π, θN2θ†]. (5.2.3.6)

Substituting θ =(1,0), θ =(0,1), and θ =(1,1) into the N2 conditions above, we find that

0 e N = , (5.2.3.7) 2 −e 0   where e R. Now substituting θ =(1, i) into the N2 conditions, we find ∈ 0 =−2e[β, i]. (5.2.3.8)

We can choose any β Im(H); therefore, we may choose β = j. Thus we find that e must ∈ vanish, making N2 = 0. This result reduces the conditions further:

0 = µ Re(θN0θ†)

0 = HNi + NiH† i {0,1} 0 = η Re(θN0θ†) ∈ β, π Im(H), θ H2. (5.2.3.9) 0 = Re(θN0θ†)θH 0 = λ Re(θN0θ†)β ∀ ∈ ∀ ∈

0 = ε Re(θN0θ†)π

Focussing on the conditions common to all generalised Bargmann algebras, i.e. those conditions which do not contain λ, µ, η, or ε, we have only

0 = HNi + NiH† i {0,1} ∈ (5.2.3.10) 0 = Re(θN0θ†)θH.

Since the second condition must hold for all θ H2, we find that either ∈

(i) N0 = 0 and H = 0, or 6

(ii) N0 = 0 and H = 0. 6 We can now split this analysis in two depending on the generalised Bargmann algebra of in- terest. First, we will discuss the algebras in which at least one of the parameters λ, µ, η, ε are non-vanishing. Subsequently, we will consider the algebras in which all of these parameters vanish. The former instance encapsulates ˆn and ˆg, and the latter encapsulates ˆa. ±

ˆn and ˆg ± All of these algebras have non-vanishing values for at least one of the parameters, λ, µ, η, ε. Therefore, all have the conditions for Branch 1 reduce to

0 = HN + N H† i {0,1} i i ∈ 0 = Re(θN0θ†) (5.2.3.11)

0 = Re(θN0θ†)θH.

Substituting θ =(1,0), θ =(0,1), and θ =(1,1) into the second condition above, we find that

0 Im(q) N = . (5.2.3.12) 0 − Im(q) 0  

Now substitute θ =(1, i) into this condition, using the convention that q = q1i + q2j + q3k, to find 0 = Re(i¯q)= q1. (5.2.3.13)

152 Using θ = (1, j) and θ = (1, k), we get analogous expressions for q2 and q3, so q = 0. There- fore, N0 = 0, and we cannot produce a super-extension in sub-branch 1.ii for these generalised Bargmann algebras.

The only remaining matrices are H and N1, such that

0 = HNi + NiH†, (5.2.3.14) with no constraints on H and N1 = N1† . So far, we have not used any basis transformations for this branch; therefore, we can choose N1 to be the canonical quaternion Hermitian form, 1. The above condition then states that H† =−H. Thus, this branch produces one non-empty sub-branch for ˆn and ˆg, with the set of non-vanishing matrices given by ±

h1 h2 1 0 Mnˆ± and gˆ, 1.i = H = , N1 = . (5.2.3.15) −h¯2 h3 0 1     Although already explicit in the forms of H and N1, we note that the set of constraints for this sub-branch is

Cnˆ± and gˆ, 1.i = {H† =−H, N1 = N1† }. (5.2.3.16) Our only comment on H going into the analysis of this branch was that it was unconstrained; therefore, we may choose to have H = 0. Thus there is certainly a super-extension in this sub-branch, one with only N1 = 1 non-vanishing. However, wanting to introduce some more parameters, we may let h1, h2 and h3 be non-vanishing. These quaternions can be fixed using the group of basis transformations G GL(s¯) GL(s¯) by noticing that H =−H tells us that ⊂ 0 × 1 † H sp(2). Therefore, the residual Sp(2) GL(s¯) which fixes N1 = 1 acts on H via the adjoint ∈ ⊂ 1 action of Sp(2) on its Lie algebra. Thus, we can make H diagonal and choose the two imaginary quaternions parameterising it, arriving at

i 0 1 0 H = and N = . (5.2.3.17) 0 j 1 0 1    

ˆa

Since ˆa has λ = µ = η = ε = 0, the conditions in (5.2.3.9) become

0 = HNi + NiH† i {0,1} ∈ (5.2.3.18) 0 = Re(θN0θ†)θH.

Unlike the ˆn and ˆg case, these conditions do not instantly set N0 = 0; therefore, we may have ± super-extensions with either (i) N0 = 0 and H = 0, or (ii) N0 = 0 and H = 0. First, setting H = 0, 6 6 6 we know this imposes N0 = 0, and, as in the ˆn and ˆg case, we may use the basis transformations ± to set N1 = 1, such that H† = −H. Therefore, one of the possible super-extensions for ˆa has non-vanishing matrices

h1 h2 1 0 Maˆ, 1.i = H = , N1 = . (5.2.3.19) −h¯2 h3 0 1     As before, the set of conditions for this super-extension is

Caˆ, 1.i = {H† =−H, N1† = N1}, (5.2.3.20) and we can use G to fix the quaternions in h. Alternatively, setting N0 = 0, we need H = 0. 6 Thus the second possible super-extension in this branch has

a q c r M = N = , N = and C = {N† = N , N† = N }. (5.2.3.21) aˆ, 1.ii 0 ¯q b 1 r¯ d aˆ, 1.ii 0 0 1 1     153 Since the primary constraint on these matrices is that both be non-vanishing, we can choose to have b, q, c and r vanish. Using the scaling symmetry of the s0¯ basis elements present in G GL(s¯) GL(s¯), we can write down the super-extension ⊂ 0 × 1 1 0 0 0 N = N = . (5.2.3.22) 0 0 0 1 0 1     Therefore, this sub-branch is not empty. Additionally, we may choose to keep all the parameters in the matrices of Maˆ, 1.ii and use the basis transformations to fix them. In particular, we can let N0 = 1. This choice leaves us with a residual Sp(2) action with which to fix the parameters of N1, which may give us N1 = 1.

Branch 2

0 0 h 0 B = Z = 0 P = H = 1 . (5.2.3.23) p 0 h p h p−1 − ε  3   3 3 1 3  As above, it is useful to exploit the vanishing matrices of the branch to simplify the conditions from Lemmas 5.2.2 and 5.2.3. In particular, the [B, Q, Q] super-Jacobi identity tells us N4 = 0 due to the vanishing of B. The rest of the [B, Q, Q] conditions tells us that

1 0 = λ Re(θN0θ†)β + [β, θN2θ†] 2 (5.2.3.24) 0 = µ Re(θN0θ†)β.

The [P, Q, Q] conditions become

0 = PN0 − N0P†

−N3 = PN1 − N1P†

0 = πθPN2θ† + θN2(πθP)† (5.2.3.25)

η Re(θN0θ†)π = πθPN3θ† + θN3(πθP)† 1 0 = ε Re(θN0θ†)π + 2 [π, θN2θ†]. Since the conditions from the [Z, Q, Q] identity are all satisfied due to Z = 0, the final conditions are

0 = HN + N H† where i {0,1,2} i i ∈ λN3 = HN3 + N3H† (5.2.3.26)

0 = µN3, from [H, Q, Q]. The result from Lemma 5.2.3 then gives us

1 Re(θN0θ†)θH = 2 θN2θ†θ. (5.2.3.27) As in Branch 1, the conditions

1 1 0 = λ Re(θN0θ†)β + 2 [β, θN2θ†] and 0 = ε Re(θN0θ†)π + 2 [π, θN2θ†], (5.2.3.28)

154 tell us that N2 = 0. Therefore, the conditions reduce further to

0 = µN3

0 = HN + N H† where i {0,1} i i ∈ 0 = PN0 − N0P† λN3 = HN3 + N3H† 2 0 = λ Re(θN0θ†)β −N3 = PN1 − N1P† β, π Im(H), θ H . ∀ ∈ ∀ ∈ 0 = µ Re(θN0θ†)β η Re(θN0θ†)π = πθPN3θ† + θN3(πθP)†

0 = ε Re(θN0θ†)π

0 = Re(θN0θ†)θH (5.2.3.29)

We can now use the following two conditions common to all generalised Bargmann algebras to highlight the possible sub-branches:

− N3 = PN1 − N1P† and 0 = Re(θN0θ†)θH. (5.2.3.30)

Substituting the N1 from (5.2.3.2) and the N3 from (5.2.3.3) into the first condition here, we can write 0 cp¯ N = 3 . (5.2.3.31) 3 −cp r¯p¯ − p r  3 3 3  This result tells us that N3 is dependent on N1: if N1 = 0 then N3 = 0. Therefore, we may organise our investigation into the possible super-extensions by considering each of the following sub-branches in turn

1. N1 = 0 and N3 = 0,

2. N1 = 0 and N3 = 0, 6

3. N1 = 0 and N3 = 0. 6 6 Next, consider the condition from the [Q, Q, Q] identity:

0 = Re(θN0θ†)θH. (5.2.3.32)

Notice, this is identical to the condition from the [Q, Q, Q] identity we found in Branch 1. Therefore, as before, we have two cases to consider in each sub-branch:

(i) N0 = 0 and H = 0, and 6

(ii) N0 = 0 and H = 0. 6 We will now consider each generalised Bargmann algebra in turn to determine whether they have super-extensions associated to these sub-branches.

ˆa

In addition to the conditions already discussed in producing the possible sub-branches,

− N3 = PN1 − N1P† and 0 = Re(θN0θ†)θH, (5.2.3.33) substituting λ = µ = η = ε = 0 into (5.2.3.29) leaves us with

0 = HN + N H† where i {0,1,3} i i ∈ 0 = PN0 − N0P† (5.2.3.34)

0 = πθpN3θ† + θN3(πθP)†.

155 None of these conditions force the vanishing of any more Ni; therefore, a priori we may find super-extensions in each of the sub-branches. The only restriction to the matrices so far has been the re-writing of N3: 0 cp¯ N = 3 . (5.2.3.35) 3 −cp r¯p¯ − p r  3 3 3 

Sub-Branch 2.1 Setting N1 = N3 = 0, we are left with only N0, subject to

0 = HN0 + N0H† and 0 = PN0 − N0P†. (5.2.3.36)

We know that we may have two possible cases for this sub-branch: either (i) N0 = 0 and H = 0, 6 or (ii) N0 = 0 and H = 0. Since we need N0 = 0 for a supersymmetric extension, we must have 6 6 the latter case. This leaves only the second condition above with which to restrict the form of N0. Since p3 = 0, this tells us 6

0 = a and 0 = p3q − ¯qp¯3. (5.2.3.37)

Thus the sub-branch is given by

0 0 0 q Maˆ, 2.1.ii = P = , N0 = and Caˆ, 2.1.ii = {0 = p3q − ¯qp¯3}. (5.2.3.38) p3 0 ¯q b     This sub-branch is parameterised by two collinear quaternions p3 and q, and a single real scalar b, such that it defines an 8-dimensional space in the sub-variety S . Notice that we can choose either q = 0 or b = 0 and this sub-branch remains supersymmetric. Choosing the former case, we can use the endomorphisms of s1¯ to set p3 = i and the scaling symmetry of H to produce

0 0 0 0 P = and N = . (5.2.3.39) i 0 0 0 1    

In the latter case, we can still choose p3 = i, and the condition in Caˆ, 2.1.ii will impose that q must also lie along i. Again using the scaling symmetry of H in G GL(s¯) GL(s¯), we arrive ⊂ 0 × 1 at 0 0 0 i P = and N = . (5.2.3.40) i 0 0 −i 0     These two examples turn out to be the only super-extensions in this sub-branch. Keeping both b and q at the outset, we can use the endomorphisms of s1¯ to set b = 0 while imposing that p3 and q lie along i. Thus, in this case, we could always retrieve the second example above.

Sub-Branch 2.2 Setting N1 = 0 but keeping N3 = 0, the conditions in (5.2.3.33) and 6 (5.2.3.34) become

0 = PNi − NiP† where i {0,1}. (5.2.3.41) 0 = HNi + NiH† ∈

Importantly, we can now have super-extensions in either of the two cases: (i) N0 = 0 and H = 0, 6 or (ii) N0 = 0 and H = 0. In the former case, in which N0 = 0, (5.2.3.33) and (5.2.3.34) become 6

0 = PN1 − N1P† and 0 = HN1 + N1H†. (5.2.3.42)

The first of these conditions tells us that 0 r N = , (5.2.3.43) 1 r¯ d  

156 such that 0 = p3r − r¯p¯3. Substituting this N1 into the latter condition, we find

0 = h r + rp h p−1 1 3 1 3 (5.2.3.44) 0 = Re(h3r)+ d Re(h1).

Assuming r = 0 and h1 = 0, take the real part of the first constraint to get Re(h1) = 0. 6 6 Alternatively, with r = 0, d = 0 for N1 = 0; therefore, the second constraint would also impose 6 6 Re(h1)= 0. This result allows us to simply the constraints to

0 = Re(h1) 0 =[h1, rp3] 0 = Re(h3r). (5.2.3.45)

In fact, the second constraint above is satisfied by

0 = p3r − r¯p¯3, (5.2.3.46) so the set of constraints on this sub-branch becomes

Caˆ, 2.2.i = {0 = Re(h1), 0 = Re(h3r), 0 = p3r − r¯p¯3}. (5.2.3.47)

Subject to these constraints, we have the following non-vanishing matrices

0 0 h1 0 0 r Maˆ, 2.2.i = P = , H = −1 , N1 = . (5.2.3.48) p3 0 h3 p3h1p3 r¯ d       This sub-branch consists of two collinear quaternions p3 and r, one quaternion h3 that is perpendicular to these two in Im(H), and one imaginary quaternion h1. In addition, there is a single real scalar, d. Notice that if H vanishes, we produce a system that is equivalent to the one found in Sub-Branch 2.1.ii; therefore, this sub-branch is certainly non-empty. However, to investigate the role of H, we will require at least one of its components to be non-vanishing. To simplify H as far as possible, let h1 = 0. Now we can choose either r or d to vanish while maintaining supersymmetry. Letting r = 0, we can use the endomorphisms of s1¯ on p3 and h3, and employ the scaling of Z on N1 to arrive at

0 0 0 0 0 0 P = , H = , N = . (5.2.3.49) i 0 i 0 1 0 1       Thus, there exist super-extensions in this sub-branch for which H = 0. Wanting to be more be 6 a little more general, we can choose for only h3 to vanish. Then, using the endomorphisms in GL(s1¯), we can set r = di such that p3 also lies along i. Utilising the scaling symmetry of P and Z in GL(s0¯), we can remove the constants from the matrices P and N1 to get

0 0 0 i P = and N = . (5.2.3.50) i 0 1 −i 1    

Employing the residual endomorphisms of s1¯, we can now choose h1 to lie along i. This change allows us to use the scaling symmetry of H in GL(s0¯) such that H becomes

i 0 H = . (5.2.3.51) 0 i  

Now, returning to the latter case, in which H vanishes and N0 = 0, we have only 6

0 = PN − N P† where i {0,1}, (5.2.3.52) i i ∈ which tells us that 0 q 0 r N = and N = , (5.2.3.53) 0 ¯q b 1 r¯ d     where 0 = p3q − ¯qp¯3 and 0 = p3r − r¯p¯3. (5.2.3.54)

157 Therefore, the set of non-vanishing matrices is given by

0 0 0 q 0 r Maˆ, 2.2.ii = P = , N0 = , N1 = , (5.2.3.55) p3 0 ¯q b r¯ d       subject to Caˆ, 2.2.ii = {0 = p3q − ¯qp¯3 and 0 = p3r − r¯p¯3}. (5.2.3.56)

Notice that the matrices N0 and N1 and the constraints on their components take the same form as the matrix N0 and its constraints in Sub-Branch 2.1.ii. However, this sub-branch is distinct. Notice that, using the endomorphisms of s1¯ and the conditions in Caˆ, 2.2.ii, we can make all the quaternions parameterising this sub-branch of S lie along i. The scaling symmetry of P may then be employed to set p3 = i, leaving only b and d unfixed. The last of the endomorphisms of s1¯ may set one of these parameters to zero, but not both; therefore, we cannot have N0 = N1, which would be a necessary condition for this sub-branch to be equivalent to (Maˆ, 2.1.ii, Caˆ, 2.1.ii). However, we can fix all the parameters of this sub-branch. Had we chosen q = bi with the initial s1¯ endomorphism and set d = 0, we could scale H and Z to find

0 0 0 i 0 i P = , N = , N = . (5.2.3.57) i 0 0 −i 1 1 −i 0       Thus, this sub-branch is non-empty and we can fix all parameters in each super-extension it contains.

Sub-Branch 2.3 Finally, with N1 = 0 and N3 = 0, we can substitute θ =(0,1) into 6 6

0 = πθPN3θ† + θN3(πθP)† (5.2.3.58) to find c = 0. Therefore, N1 and N3 are reduced to

0 r 0 0 N = and N = . (5.2.3.59) 1 r¯ d 3 0 r¯p¯ − p r    3 3  Recall, the condition 0 = Re(θN0θ†)θH (5.2.3.60) tells us that either (i) N0 = 0 and H = 0, or (ii) N0 = 0 and H = 0. Letting N0 = 0, the final 6 6 conditions for this sub-branch are

0 = HN + N H† where i {1, 3}. (5.2.3.61) i i ∈

From the discussion in Sub-Branch 2.2.i, we know that the N1 case produces the constraints

0 = Re(h1), 0 =[h1, rp3], and 0 = Re(h3r). (5.2.3.62)

Interestingly, the N3 condition adds no new constraints to this set; therefore, we have

Caˆ, 2.3.i = {0 = Re(h1), 0 =[h1, rp3], 0 = Re(h3r)}. (5.2.3.63)

The corresponding matrices for this sub-branch are given by

0 0 h 0 0 r 0 0 M = P = , H = 1 , N = , N = . aˆ, 2.3.i p 0 h p h p−1 1 r¯ d 3 0 r¯p¯ − p r  3   3 3 1 3     3 3  (5.2.3.64) To establish the existence of super-extensions in this sub-branch, begin by setting H = 0 and d = 0. The endomorphisms of s1¯ may be used to set p3 to lie along i and scale r such that r Sp(1). We can then utilise the automorphisms of H and the scaling symmetry of P and B ∈ in GL(s0¯) to set r and fix the parameters in P and N3. This leaves us with a super-extension

158 whose matrices are written 0 0 0 1 + j 0 0 P = , N = , N = . (5.2.3.65) i 0 1 1 − j 0 3 0 k      

Using this parameterisation, we can also introduce h1. Substituting p3 = i and r = 1 + j into the constraints of Caˆ, 2.3.i, we find

0 0 i − k 0 0 1 + j 0 0 P = , H = , N = , N = . (5.2.3.66) i 0 0 i + k 1 1 − j 0 3 0 k        

Looking to include h3 or d leads to the introduction of parameters that cannot be fixed using the basis transformations G GL(s¯) GL(s¯) and the constraints. ⊂ 0 × 1

In the latter case, for which H = 0 and N0 = 0, the only remaining condition is 6

0 = PN0 − N0P†, (5.2.3.67) which we know from the previous sub-branches, tells us that p3 and q are collinear, and that a = 0. Therefore, the non-vanishing matrices for this sub-branch are

0 0 0 q 0 r 0 0 M = P = , N = , N = , N = , aˆ, 2.3.ii p 0 0 ¯q b 1 r¯ d 3 0 r¯p¯ − p r  3       3 3  (5.2.3.68) and the constraints are given by

Caˆ, 2.3.ii = {0 = p3q − ¯qp¯3}. (5.2.3.69)

This sub-branch of S has 13 real parameters, being parameterised by two collinear quaternions p3 and q, an additional quaternion r and two real scalars, b and d. Letting d = 0 and q = 0, we can use the same transformations as in Sub-Branch 2.3.i to fix 0 0 0 1 + j 0 0 P = , N = , N = . (5.2.3.70) i 0 1 1 − j 0 3 0 k      

Subsequently employing the scaling symmetry of H in GL(s0¯), we can fix b such that

0 0 N = . (5.2.3.71) 0 0 1   Therefore, there are certainly super-extensions in this sub-branch. We can introduce either q or d while continuing to fix all the parameters of the super-extension; however, attempting to include both leads to the inclusion of a parameter that we cannot fix with the constraints of Caˆ, 2.3.ii and basis transformations in G.

ˆn−

Setting µ = η = 0, λ = 1 and ε =−1, the conditions in (5.2.3.29) reduce to

0 = HN + N H† where i {0,1} i i ∈ 0 = PN0 − N0P† 0 = πθPN3θ† + θN3(πθP)†

0 = Re(θN0θ†)β N3 = HN3 + N3H† (5.2.3.72)

0 =− Re(θN0θ†)π −N3 = PN1 − N1P†.

0 = Re(θN0θ†)θH

159 From the third and fourth condition, we instantly get N0 = 0. Therefore, we cannot have any solutions along sub-branches satisfying case (ii) for ˆn−. We are left with

0 = HN1 + N1H† N3 = HN3 + N3H† (5.2.3.73) 0 = πθPN3θ† + θN3(πθP)† −N3 = PN1 − N1P†.

Sub-Branch 2.1 Since N1 and N3 are the only possible non-vanishing matrices encoding the [Q, Q] bracket, we cannot have a super-extension in this branch.

Sub-Branch 2.2 With N3 = 0, we are left with

0 = HN1 + N1H† and 0 = PN1 − N1P†. (5.2.3.74)

The latter condition tells us that p3 and r are collinear and 0 = cp3. Since we must have p3 = 0 6 in this branch, we have c = 0. Using this result, the first condition above tells us

0 = h r + rp h p−1 + 1 1 3 1 3 (5.2.3.75) 0 = Re(h3r)+ d(Re(h1)+ 1).

In fact, utilising the collinearity of p3 and r, the first of these constraints becomes

0 =(2 Re(h1)+ 1) Re(p3r). (5.2.3.76)

Thus, we have

Cnˆ−, 2.2.i = {0 = r¯p¯3 − p3r, 0 =(2 Re(h1)+ 1) Re(p3r), 0 = Re(h3r)+ d(Re(h1)+ 1)}. (5.2.3.77) The non-vanishing matrices in this instance are

0 0 h1 0 0 r Mnˆ−, 2.2.i = P = , H = −1 , N1 = . (5.2.3.78) p3 0 h3 p3h1p3 + 1 r¯ d       Therefore, the sub-branch in S for these super-extensions of ˆn− is parameterised by two collinear quaternions p3 and r, two quaternions encoding the action of H on s1¯, h1 and h3, and one real scalar d. Notice, this is the first instance in which setting some parameters to zero imposes particular values for other parameters in the extension. In particular, the vanishing of r imposes Re(h1)=−1 by the third constraint in Cˆ , 2.2.i, since d = 0 in this instance. n− 6 However, if r = 0, the second constraint implies 2 Re(h1)=−1. In the former case, we can 6 set h3 and the imaginary part of h1 to zero. Using the endomorphisms of s1¯ to set p3 = i, we can subsequently employ the scaling symmetry of H and Z to obtain a super-extension with matrices 0 0 1 0 0 0 P = , H = and N = . (5.2.3.79) i 0 0 0 1 0 1       Therefore, there exist super-extensions in this sub-branch for which r = 0. Letting r = 0, we 6 may again use the endomorphisms of s1¯ to impose that p3 lies along i; however, due to the first constraint in Cnˆ−, 2.2.i, this also means that r lies along i. Utilising the scaling symmetry of the s0¯ basis elements, we may write down the matrices

0 0 1 0 0 i P = , H = and N = . (5.2.3.80) i 0 0 −1 1 −i 0      

Thus, super-extensions for which r = 0 exist in this sub-branch. In both cases, residual s¯ endo- 6 1 morphisms may be used to set h3 and the imaginary part of h1 should we choose to include them.

160 Sub-Branch 2.3 Setting N3 = 0, we must now consider 6

0 = πθPN3θ† + θN3(πθP)†, (5.2.3.81) which, on substituting in θ = (0,1), tells us that c = 0. Therefore, as in sub-branch 2.2, the first condition of (5.2.3.73) tells us

0 = h r + rp h p−1 + 1 1 3 1 3 (5.2.3.82) 0 = Re(h3r)+ d(Re(h1)+ 1).

However, unlike sub-branch 2.2, r and p3 are not collinear since the imaginary part of p3r makes up the only non-vanishing component of N3:

0 0 N = . (5.2.3.83) 3 0 r¯p¯ − p r  3 3 

Substituting this N3 into its condition from the [H, Q, Q] identity, we find

(1 − 2 Re(h4)) Im(l)=[Im(h4), Im(l)], (5.2.3.84)

−1 where h4 = p3h1p3 + 1 and l = r¯p¯3 − p3r. Since Im(l) is perpendicular to [Im(h4), Im(l)], both sides of this expression must vanish separately. Substituting h4 and l into the above expressions, we find 0 =(1 + 2 Re(h1)) Im(p3r) and 0 =[h1, rp3]. (5.2.3.85)

As stated above, r and p3 are not collinear; therefore, the first constraint here tells us that

2 Re(h1)=−1. (5.2.3.86)

Substituting this result into the second constraint in (5.2.3.82), we find

2 Re(h3r)=−d. (5.2.3.87)

Putting all these results together, the constraints are

Cnˆ−, 2.3.i = {2 Re(h1)=−1, 2Re(h3r)=−d, 0 =[h1, rp3]}, (5.2.3.88) for the non-vanishing matrices

0 0 h 0 0 r 0 0 M = P = , H = 1 , N = , N = . nˆ−, 2.3.i p 0 h p h p−1 + 1 1 r¯ d 3 0 r¯p¯ − p r  3   3 3 1 3     3 3  (5.2.3.89) Notice, the sub-branch in S describing these super-extensions of ˆn− is parameterised by four quaternions p3, h1, h3 and r, and one real scalar d. Wanting to establish the existence of super-extensions in this sub-branch, we can choose to set h3, d, and the imaginary part of h1 to zero. Then, utilising the endomorphisms of s1¯, we can impose that p3 must lie along i and that r must have unit norm. Subsequently employing Aut(H) to fix r, we can finally scale H, Z, P, and B to get the super-extension

0 0 1 0 0 1 + j 0 0 P = , H = , N = , N = . (5.2.3.90) i 0 0 −1 1 1 − j 0 3 0 k         Having established that this sub-branch is not empty, we may look to introduce the components we have set to zero for this example. Notably, we may introduce the imaginary part of h1 while still fixing all parameters using the basis transformations G GL(s¯) GL(s¯). However, the ⊂ 0 × 1 inclusion of either h3 or d will introduce parameters that cannot be fixed.

ˆn+ and ˆg

161 4 Substituting λ = ε = 0, µ = 1 into the conditions of (5.2.3.29), we instantly have N3 = 0 and ±

0 = HN + N H† where i {0,1} i i ∈ 0 = PNi − NiP† where i {0,1} ∈ (5.2.3.92) 0 = Re(θN0θ†)θH

0 = Re(θN0θ†)β. ±

The final condition here states that N0 = 0; therefore, N1 is the only possible non-vanishing matrix of those encoding [Q, Q]. This result tells us there will be no sub-branch 2.1 or 2.3 for these algebras and no sub-branch satisfying case (ii), in which N0 = 0. Therefore, the conditions 6 reduce to 0 = HN1 + N1H† and 0 = PN1 − N1P†. (5.2.3.93)

Under the assumption that p3 = 0 for this branch of super-extensions, the latter condition tells 6 us that c = 0 and that p3 and r are collinear:

0 = r¯p¯3 − p3r. (5.2.3.94)

Substituting these results into the first condition, we find

0 = Re(h1), 0 =[h1, rp3] and 0 = Re(h3r). (5.2.3.95)

Notice that, since p3 and r are collinear, the second constraint is instantly satisfied. Thus, our constraints reduce to

Cnˆ+ and gˆ, 2.2.i = {0 = Re(h1), 0 = Re(h3r), 0 = r¯p¯3 − p3r}. (5.2.3.96)

The non-vanishing matrices in this instance are

0 0 h1 0 0 r Mnˆ+ and gˆ, 2.2.i = P = , H = −1 , N1 = . (5.2.3.97) p3 0 h3 p3h1p3 r¯ d       This sub-branch has identical (M, C) to sub-branch 2.2.i for ˆa. Therefore, for a discussion on the existence of such super-extensions, we refer the reader to the discussion found there.

Branch 3

0 0 h 0 P = Z = 0 B = H = 1 . (5.2.3.98) b 0 h b h b−1 − λ  3   3 3 1 3  Exploiting the vanishing of Z and P, we can reduce the conditions from Lemmas 5.2.2 and 5.2.3. In particular, the vanishing of P, when substituted into the conditions from the [P, Q, Q] super- Jacobi identity, tells us that N3 = 0 and

0 = η Re(θN0θ†)π 1 (5.2.3.99) 0 = ε Re(θN0θ†)π + 2 [π, θN2θ†].

4 Whether we are in the nˆ+ or gˆ case makes no difference: the distinction between the two is the value of η, which, if non-vanishing, would add the condition

0 = Re(θN0θ†)π. (5.2.3.91)

This condition sets N0 = 0, but we already have this result from another condition. Therefore, the super- extensions are the same for both of these generalised Bargmann algebras.

162 The [B, Q, Q] identity then produce

0 = BN0 − N0B†

N4 = BN1 − N1B†

0 = βθBN2θ† + θN2(βθB)† (5.2.3.100) 1 0 = λ Re(θN0θ†)β + 2 [β, θN2θ†]

µ Re(θN0θ†)β = βθBN4θ† + θN4(βθB)†.

The conditions from the [Z, Q, Q] identity are satisfied since Z = 0, and, lastly, the [H, Q, Q] super-Jacobi identity produces

0 = HN + N H† where i {0,1,2} i i ∈ 0 = ηN4 (5.2.3.101)

εN4 = HN4 + N4H†.

From Lemma 5.2.3, we get 1 Re(θN0θ†)θH = 2 θN2θ†θ. (5.2.3.102) As in both previous branches, the conditions

1 0 = λ Re(θN0θ†)β + 2 [β, θN2θ†] 1 (5.2.3.103) 0 = ε Re(θN0θ†)π + 2 [π, θN2θ†], tell us N2 = 0, such that, putting everything together, we have

0 = ηN4

0 = HN + N H† where i {0,1} i i ∈ 0 = BN0 − N0B† εN4 = HN4 + N4H†.

0 = η Re(θN0θ†)π N4 = BN1 − N1B† β, π Im(H), θ H. ∀ ∈ ∀ ∈ 0 = λ Re(θN0θ†)β µ Re(θN0θ†)β = βθBN4θ† + θN4(βθB)†

0 = ε Re(θN0θ†)π

0 = Re(θN0θ†)θH (5.2.3.104)

We can now use some of the conditions common to all generalised Bargmann algebras to identify possible sub-branches with which we can organise our investigations. Substituting the N1 from (5.2.3.2) and the N4 from (5.2.3.3) into the condition

N4 = BN1 − N1B†, (5.2.3.105) we can write N4 in terms of the parameters in N1 and B:

0 −cb¯ N = 3 . (5.2.3.106) 4 cb b r − r¯b¯  3 3 3

Notice that this means N4 is completely dependent on N1: if N1 = 0 then N4 = 0. Therefore, in general, we have the following sub-branches:

1. N1 = 0 and N4 = 0,

2. N1 = 0 and N4 = 0, 6 3. N1 = 0 and N4 = 0. 6 6 Also, as in Branches 1 and 2, the condition derived from the [Q, Q, Q] identity tells us that either N0 or H vanishes. We will consider both of these cases within each sub-branch, identifying them as

163 (i) N0 = 0 and H = 0, and 6

(ii) N0 = 0 and H = 0. 6

ˆa

Setting λ = µ = η = ε = 0, the conditions in (5.2.3.104) reduce to

0 = HN + N H† where i {0,1,4} i i ∈ 0 = BN0 − N0B†

0 = βθBN4θ† + θN4(βθB)† (5.2.3.107)

0 = Re(θN0θ†)θH

N4 = BN1 − N1B†.

As in Branch 2, none of these conditions force the vanishing of any more Ni; therefore, super- extensions may be found in each of the sub-branches. In fact, because of the symmetry of the generators B and P in this generalised Bargmann algebra, we may use automorphisms to transform the above conditions into those in (5.2.3.33) and (5.2.3.34), which describe the super-extensions of ˆa in Branch 2. More explicitly, substitute the transformation with matrices

1 0 0 −1 1 0 A = , C = , and M = , (5.2.3.108) 0 1 1 0 0 1       and the quaternion u = 1, into (5.2.1.34). Putting the transformed matrices into the condi- tions of (5.2.3.107), we recover the conditions of (5.2.3.33) and (5.2.3.34). Therefore, all the super-extensions of ˆa in this branch are equivalent to the super-extensions of Branch 2. Thus, for this particular generalised Bargmann algebra, this branch produces no new super-extensions.

ˆn−

Setting µ = η = 0, λ = 1 and ε =−1, the conditions of (5.2.3.104) become

0 = HN + N H† where i {0,1} i i ∈ 0 = BN0 − N0B† −N4 = HN4 + N4H†.

0 = Re(θN0θ†)β N4 = BN1 − N1B† (5.2.3.109)

0 =− Re(θN0θ†)π 0 = βθBN4θ† + θN4(βθB)†.

0 = Re(θN0θ†)θH

The conditions 0 = Re(θN0θ†)β and 0 =− Re(θN0θ†)π (5.2.3.110) tell us that N0 must vanish, leaving only

0 = HN1 + N1H† −N4 = HN4 + N4H† (5.2.3.111) 0 = βθBN4θ† + θN4(βθB)† N4 = BN1 − N1B†.

Notice, this result tells us that we cannot have any sub-branches satisfying case (ii); therefore, all sub-branches (M, C) discussed below will have a subscript ending in i. Like the ˆa case, this generalised Bargmann algebra allows for an automorphism which transforms the conditions for this branch into the conditions for Branch 2. However, in this instance, this branch will produce some distinct super-extensions. This result is a consequence of the parameters ε and λ and their appearance in H. In Branch 2, the matrix H is written as

h 0 H = 1 , (5.2.3.112) h p h p−1 − ε  3 3 1 3 

164 and in this branch, it is written

h 0 H = 1 . (5.2.3.113) h p h p−1 − λ  3 3 1 3 

Since ˆn− has ε = −1 and λ = 1, this matrix differs in these branches, if only be a sign. Thus, although the investigations into the super-extensions of ˆn− in this branch will be very similar to those in the previous branch, we will give a partial presentation of them here to demonstrate any consequences of this change in sign. In particular, we will omit the discussions on the ex- istence of super-extensions and parameter fixing as these require only trivial adjustments from the discussions found in Branch 2.

Sub-Branch 3.1 As N0 = 0, we cannot have both N1 and N4 vanish; therefore, there is no super-extension in this sub-branch.

Sub-Branch 3.2 Letting N1 = 0 and N4 = 0, we are left with only the conditions 6

0 = HN1 + N1H† and 0 = BN1 − N1B†. (5.2.3.114)

The second condition above tells us that

0 = cb3 and 0 = b3r − r¯b¯3. (5.2.3.115)

As b3 = 0 by assumption, c = 0. Substituting this result into the first condition above, we find 6 0 = h r + rb h b−1 − 1 1 3 1 3 (5.2.3.116) 0 = Re(h3r)+ d(Re(h1)− 1).

Using the collinearity of b3 and r, the first of these constraints tells us that

0 =(2 Re(h1)− 1) Re(b3r). (5.2.3.117)

Therefore, the constraints in this instance are given by ¯ Cnˆ−, 3.2.i = {0 = b3r − r¯b3, 0 =(2 Re(h1)− 1) Re(b3r), 0 = Re(h3r)+ d(Re(h1)− 1)}. (5.2.3.118) The non-vanishing matrices in this instance are

0 0 h1 0 0 r Mnˆ−, 3.2.i = B = , H = −1 , N1 = . (5.2.3.119) b3 0 h3 b3h1b3 − 1 r¯ d       This sub-branch of S is parameterised by two collinear quaternions b3 and r, two quater- nions encoding the action of H on s1¯, h1 and h3, and one real scalar d. Notice that the real component of h1 varies depending on whether r vanishes. Together with the super-extensions in Sub-Branch 2.2.i for ˆn−, these are the only super-extensions that demonstrate this type of dependency. If r = 0, the first two constraints of Cnˆ−, 3.2.i are trivial, and the third condition tell us that Re(h1) = 1, since d = 0 for N1 = 0. However, if r = 0, the second constraint 6 6 6 requires 2Re(h1) = 1. In this instance, the third constraint then becomes 2 Re(h3r) = d. As the matrices and conditions for this sub-branch are so similar to those in 2.2.i, we refer the reader to the discussion on existence of super-extensions and parameter fixing presented there.

Sub-Branch 3.3 Finally, let N1 = 0 and N4 = 0. The condition 6 6

0 = βθBN4θ† + θN4(βθB)† (5.2.3.120)

165 imposes c = 0, such that

0 r 0 0 N = and N = . (5.2.3.121) 1 r¯ d 4 0 b r − r¯b¯    3 3 This result reduces the conditions in (5.2.3.111) to

0 = HN1 + N1H† (5.2.3.122) −N4 = HN4 + N4H†.

Substituting the N4 from (5.2.3.121) into the second condition above, we have

− l = h4l + lh¯4, (5.2.3.123)

¯ −1 where l = b3r − r¯b3 and h4 = b3h1b3 − 1. We can rewrite this condition as

(1 + 2 Re(h4))l =[l, h4]. (5.2.3.124)

Notice that the R.H.S. of this expression must lie in Im(H) and be orthogonal to l, which is imaginary by construction. Therefore, both sides of this expression must vanish independently:

0 =(1 + 2 Re(h4))l 0 =[l, h4]. (5.2.3.125)

Substituting l and h4 into these constraints, we find

0 =(2 Re(h1)− 1)(b3r − r¯b¯3) and 0 =[h1, rb3], (5.2.3.126) respectively. For N4 to not vanish, we must have Im(b3r) = 0, so, by the first constraint above, 6 we need 2 Re(h1) = 1. The first condition in (5.2.3.122) produces the same constraints as in Sub-Branch 3.2; namely,

0 = h r + rb h b−1 − 1 1 3 1 3 (5.2.3.127) 0 = Re(h3r)+ d(Re(h1)− 1).

Notice that the requirement of setting 2 Re(h1)= 1 makes the second constraint here 2 Re(h3r)= d, and says that the first constraint is equivalent to 0 =[h1, rb3]. Therefore, the constraints on this sub-branch are given by

Cnˆ−, 3.3.i = {d = 2 Re(h3r), 1 = 2 Re(h1) and 0 =[h1, rb3]}, (5.2.3.128) and the non-vanishing matrices are

0 0 h 0 0 r 0 0 M = B = , H = 1 , N = , N = . nˆ−, 3.3.i b 0 h b h b−1 − 1 1 r¯ d 4 0 b r − r¯b¯  3   3 3 1 3     3 3 (5.2.3.129) For the discussion on existence of super-extensions and how to fix the parameters of the matri- ces describing this sub-branch of S , we refer the reader to Sub-Branch 2.3.i. The application of the discussion to the present case requires only minor adjustments.

ˆn+

Substituting λ = ε = 0, µ = 1, and η = −1 into the results for Lemmas 5.2.2 and 5.2.3,

166 we find

0 =−N4

0 = HNi + NiH† where i {0,1,4} ∈ N4 = BN1 − N1B† 0 = BN0 − N0B† (5.2.3.130) Re(θN0θ†)β = βθBN4θ† + θN4(βθB)†. 0 =− Re(θN0θ†)π

0 = Re(θN0θ†)θH

Therefore, N4 vanishes, and N0 vanishes by 0 =− Re(θN0θ†)π. This leaves us with

0 = HN1 + N1H† and 0 = BN1 − N1B†. (5.2.3.131)

Notice that these conditions are similar to those of (5.2.3.93), which describe the super- extensions of ˆn+ in Branch 2. In fact, we can utilise the automorphisms of ˆn+ to transform the above conditions into those in (5.2.3.93). Unlike the ˆn− case, since ˆn+ has vanishing ε and λ, there is no discrepancy between the transformed matrices and those of Branch 2; therefore, the super-extensions of ˆn+ in Branches 2 and 3 are equivalent. Thus, we have no new super- extensions here.

ˆg

Substituting λ = η = ε = 0 and µ =−1 into (5.2.3.104), we have

0 = HNi + NiH† where i {0,1,4} ∈ N4 = BN1 − N1B† 0 = BN0 − N0B† (5.2.3.132) − Re(θN0θ†)β = βθBN4θ† + θN4(βθB)†. 0 = Re(θN0θ†)θH

With these conditions, we can now investigate the three sub-branches.

Sub-Branch 3.1 We cannot have N1 = N4 = 0, since the vanishing of N4 means N0 = 0 through − Re(θN0θ†)β = βθBN4θ† + θN4(βθB)†. (5.2.3.133)

This would cause all Ni to vanish such that [Q, Q]= 0. Therefore, there is no super-extension in this sub-branch.

Sub-Branch 3.2 With only N1 = 0, the conditions reduce to 6

0 = HN1 + N1H† and 0 = BN1 − N1B†. (5.2.3.134)

Notice that this is the same set of conditions as the ˆn+ case above. Therefore, we may expect the analysis for this generalised Bargmann algebra to be analogous. However, there is a very important distinction. In the ˆn+ case, we were able to use the automorphisms to transform the conditions into those of Sub-Branch 2.2. This automorphism is not permitted by the generalised Bargmann algebra ˆg. Therefore, although the analysis will be the same mutatis mutandis as that of Sub-Branch 2.2, the resulting super-extensions will be distinct.

Now, since N1 is the only possible non-vanishing matrix in the [Q, Q] bracket, it must have non-zero components. The latter condition above tells us that c = 0 and b3 and r are collinear quaternions, while the former condition imposes

0 = h r + rb h b−1 1 3 1 3 (5.2.3.135) 0 = Re(h3r)+ d Re(h1).

167 Notice that if r = 0, we need d = 0 for the existence of a super-extension; therefore, the final 6 constraint above would impose Re(h1) = 0. Similarly, if r = 0, the first constraint would also 6 enforce Re(h1) = 0. Thus, in all super-extensions, we require Re(h1) = 0. Using this result, these two constraints simplify to

0 =[h1, rb3] and 0 = Re(h3r). (5.2.3.136)

However, since b3 and r are collinear and it is only the imaginary part of rb3 that will con- tribute to [h1, rb3], the first of these constraints is already satisfied. Therefore, the final set of constraints on this sub-branch is

Cgˆ, 3.2.i = {0 = b3r − r¯b¯3, 0 = Re(h1), 0 = Re(h3r)}. (5.2.3.137)

Subject to these constraints, we have non-vanishing matrices are

0 0 h1 0 0 r Mgˆ, 3.2.i = B = , H = −1 , N1 = . (5.2.3.138) b3 0 h3 b3h1b3 r¯ d       Since this (M, C) is analogous to the one found in Branch 2 for ˆn+ and ˆg, we will omit the discussion on existence of super-extensions and parameter fixing.

Sub-Branch 3.3 Finally, with N4 = 0, we can think of setting N0 = 0 and H = 0. But first, 6 6 try setting N0 = 0 to allow H = 0. The conditions in (5.2.3.132) become 6

0 = HN + N H† where i {1,4} i i ∈ N4 = BN1 − N1B† (5.2.3.139)

0 = βθBN4θ† + θN4(βθB)†.

Notice that the second condition above allows us to write N4 in terms of B and N1:

0 −cb¯ N = 3 . (5.2.3.140) 4 cb b r − r¯b¯  3 3 3

The third condition then imposes c = 0, since b3 = 0, leaving us with 6 0 r 0 0 N = and N = . (5.2.3.141) 1 r¯ d 4 0 b r − r¯b¯    3 3 Using these matrices in the final conditions,

0 = HN + N H† where i {1,4}, (5.2.3.142) i i ∈ produces 0 =[h1, rb3], (5.2.3.143) when i = 4, and, when i = 1, we obtain

0 = h r + rb h b−1 1 3 1 3 (5.2.3.144) 0 = Re(h3r)+ d Re(h1).

Since r = 0 for N4 = 0, the first condition here states that Re(h1)= 0. Therefore, the constraints 6 6 on the parameters of this super-extension are given by

Cgˆ, 3.3.i = {0 = Re(h1), 0 =[h1, rb3], 0 = Re(h3r)}. (5.2.3.145)

168 The non-vanishing matrices associated with this sub-branch are

0 0 h 0 0 r 0 0 M = B = , H = 1 , N = , N = . gˆ, 3.3.i b 0 h b h b−1 1 r¯ d 4 0 b r − r¯b¯  3   3 3 1 3     3 3 (5.2.3.146) The (M, C) of this sub-branch is the same mutatis mutandis as that of Sub-Branch 2.3.i for ˆa; therefore, we refer the reader to the discussion found there on existence of super-extensions and parameter fixing.

Finally, let N0 = 0 such that H = 0. The conditions remaining from (5.2.3.132) are 6

0 = BN0 − N0B†

N4 = BN1 − N1B† (5.2.3.147)

− Re(θN0θ†)β = βθBN4θ† + θN4(βθB)†.

We know how the second condition acts from the discussion at the beginning of this branch. The first of these conditions tells us

0 = a and 0 = b3q − ¯qb¯3, (5.2.3.148) and the third, substituting in θ =(0,1), produces

2 − b = 2c|b3| . (5.2.3.149)

Now, substituting θ =(1, s) into the third condition, we find

2 2 2 − 2 Re(s¯q)− b|s| = 2c|s| |b3| . (5.2.3.150)

Therefore, using the previous result and letting s = 1, s = i, s = j and s = k, we see that all components of q must vanish. We thus have non-vanishing matrices

0 0 0 0 c r 0 −cb¯3 Mgˆ, 3.3.ii = B = , N0 = 2 , N1 = , N4 = . b3 0 0 −2c|b3| r¯ d cb b r − r¯b¯        3 3 3 (5.2.3.151) Interestingly, there are no additional constraints to the parameters of this sub-branch; therefore, Cgˆ, 3.3.ii is empty. Notice the sub-branch of S for this type of super-extension is parameterised by two quaternions b3 and r, and two real scalars c and d. To demonstrate that this sub-branch is not empty, we begin by setting both r and d to zero. This choice allows us to utilise the endomorphisms of s1¯ to set b3 = i and c = 1. Employing the scaling symmetry of the basis elements, we arrive at

0 0 0 0 1 0 0 i B = , N = , N = , N = . (5.2.3.152) i 0 0 0 1 1 0 0 4 i 0        

We may now look to introduce r and d. Again, using the endomorphisms of s1¯, we can impose 2 that b3 must lie along i, set |r| = 1, and choose √2c = 1. This choice for r imposes that r Sp(1), and we may utilise Aut(H) to fix √2r = 1 + i. Having chosen r = 0, we can always ∈ 6 employ the residual endomorphisms of s1¯ to set d = 0. Using the only remaining symmetry, the scaling of H, Z, B, and P, we find

0 0 0 0 1 1 + i 0 i B = , N = , N = , N = . (5.2.3.153) √2i 0 0 0 1 1 1 − i 0 4 i 2i        

Branch 4

h 0 0 0 0 0 Z = 0 H = 1 B = P = , (5.2.3.154) h h b 0 p 0  3 4  3   3 

169 subject to

2 2 [u, h1]=−µu +(λ − ε)u + η or [v, h1]= ηv +(λ − ε)v + µ, (5.2.3.155)

−1 −1 where 0 = u = b p3 and 0 = v = p b3. Recall, we keep both of these constraints as, de- 6 3 6 3 pending on the generalised Bargmann algebra under investigation, one of them will prove more useful than the other. We still need to determine the generalised Bargmann algebras for which this branch could provide a super-extension. Therefore, we will consider each algebra in turn, and analyse those for which the above constraints may hold.

ˆa

Setting λ = µ = η = ε = 0 in (5.2.3.155), we could still get a super-extension, as long as we impose 0 =[u, h1]. (5.2.3.156)

Throughout this section, we will choose to write parameters in terms of b3; therefore, we write −1 p3 = b3u and h4 = b3h1b , where u H. Notice that the significance of u is only manifest 3 ∈ when h1 = 0: when h1 vanishes, we are simply replacing p3 with u. However, since u will be 6 important is several instances, we will always use this notation.

Since neither B nor P vanish, there are no immediate results as in the three previous branches: all the conditions of Lemmas 5.2.2 and 5.2.3 must be taken into consideration. However, as with Branches 2 and 3, we can organise our investigations based on dependencies. In particular, the conditions

N4 = BN1 − N1B†

−N3 = PN1 − N1P† 1 (5.2.3.157) 2 [β, θN2θ†]= βθBN3θ† + θN3(βθB)† 1 2 [π, θN2θ†]= πθPN4θ† + θN4(πθP)†, show us that if N1 vanishes, so must N2, N3, and N4. Additionally, the vanishing of either N3 or N4 means we must have N2 = 0. Therefore, we can divide our investigations into the following sub-branches.

1. N1 = N2 = N3 = N4 = 0

2. N1 = 0 and N2 = N3 = N4 = 0 6 3. N1 = 0, N3 = 0, and N2 = N4 = 0 6 6 4. N1 = 0, N4 = 0, and N2 = N3 = 0 6 6 5. N1 = 0, N3 = 0, N4 = 0, and N2 = 0 6 6 6 6. N1 = 0, N2 = 0, N3 = 0, and N4 = 0. 6 6 6 6 Unlike Branches 1, 2 and 3, the [Q, Q, Q] super-Jacobi identity will not always result in the cases (i), in which N0 = 0 and H = 0, or (ii), in which N0 = 0 and H = 0. There are instances 6 6 in which both N0 and H may not vanish. These cases, will be labelled (iii).

Sub-Branch 4.1 With only N0 left available, it cannot vanish for a supersymmetric extension to exist. Therefore, the [Q, Q, Q] identity,

Re(θN0θ†)θH = 0, (5.2.3.158) tells us we must have H = 0. The remaining conditions are then

0 = BN0 − N0B† and 0 = PN0 − N0P†, (5.2.3.159)

170 which tell us 0 = a, 0 = b3q − ¯qb¯3 and 0 = b3uq − ¯qu¯b¯3. (5.2.3.160) This sub-branch thus has non-vanishing matrices

0 0 0 0 0 q B = , P = , N = , (5.2.3.161) b 0 b u 0 0 ¯q b  3   3    subject to the constraints 0 = b3q − ¯qb¯3, 0 = b3uq − ¯qu¯b¯3. (5.2.3.162)

Notice that these matrices and constraints are very similar to (Maˆ, 2.1.ii, Caˆ, 2.1.ii). In fact, em- ploying the automorphisms of ˆa, we can show that the above system is equivalent to Sub-Branch 2.1.ii. Using the endomorphisms of s1¯ and the constraints above, we can set b3, b3u and q to lie along i, and set b = 0. In particular, this means that u R. Scaling B, P, and H, we find ∈ the matrices 0 0 0 0 0 i B = , P = , N = , (5.2.3.163) i 0 i 0 0 −i 0       which under the basis transformation with 1 −1 C = , (5.2.3.164) 0 1   recovers the maximal super-extension of Sub-Branch 2.1.ii. Thus, this sub-branch does not contribute any new super-extensions to ˆa.

Sub-Branch 4.2 The [Q, Q, Q] identity still imposes that either N0 or H must vanish in this sub-branch; however, we can now consider the case where N0 = 0 as we have N1 = 0. First, 6 consider case (i), with N0 = 0 such that H = 0. The conditions remaining are 6

0 = HN1 + N1H†

0 = BN1 − N1B† (5.2.3.165)

0 = PN1 − N1P†.

The latter two conditions tell us that c = 0 and b3 is collinear with b3u and r. Substituting these results into the first condition, we find

0 = h r + rb h b−1 1 3 1 3 (5.2.3.166) 0 = Re(h3r)+ d Re(h1).

We know from the analysis of Branch 3 that demanding N1 = 0 under these conditions imposes 6 Re(h1)= 0; and, that having the condition

0 = b3r − r¯b¯3 (5.2.3.167) means we always satisfy the imaginary part of

−1 0 = h1r + rb3h1b3 . (5.2.3.168)

Putting all this together, we find the constraints on this sub-branch to be

0 = Re(h1), 0 = Re(h3r), 0 = b3r − r¯b¯3, 0 = b3ur − r¯u¯b¯3, 0 =[u, h1]. (5.2.3.169)

The non-vanishing matrices are then

0 0 0 0 h 0 0 r B = , P = , H = 1 and N = . (5.2.3.170) b 0 b u 0 h b h b−1 1 r¯ d  3   3   3 3 1 3   

171 Notice that b3, r and b3u all being collinear implies that u R. Thus the final constraint is sat- ∈ isfied, and, as in Sub-Branch 4.1, we can use the endomorphisms of s1¯ and the automorphisms of ˆa to rotate B and P such that we only have the matrix P, in which p3 = i. The resulting matrices and constraints are then equivalent to those found in Sub-Branch 2.2.i, and, therefore, this sub-branch does not produce any new super-extensions for ˆa.

Now, considering case (ii), let H = 0. The remaining conditions are

0 = BNi − NiB† where i {0,1} ∈ (5.2.3.171) 0 = PN − N P† where i {0,1}. i i ∈

Therefore, N0 and N1 take the same form in this instance: both a and c vanish, with q and r being collinear to both b3 and b3u. In summary, the constraints are

0 = b3q − ¯qb¯3, 0 = b3uq − ¯qu¯b¯3, 0 = b3r − r¯b¯3, 0 = b3ur − r¯u¯b¯3, (5.2.3.172) and the non-vanishing matrices are

0 0 0 0 0 q 0 r B = , P = , N = , and N = . (5.2.3.173) b 0 b u 0 0 ¯q b 1 r¯ d  3   3     

Through the same use of the subgroup G GL(s¯) GL(s¯) as discussed for case (i), we find ⊂ 0 × 1 that this sub-branch is equivalent to 2.2.ii for ˆa.

Sub-Branch 4.3 Now with N3 = 0, we can use 6

− N3 = PN1 − N1P† and 0 = πθPN3θ† + θN3(πθP)† (5.2.3.174) to first write N3 in terms of P and N1 before setting c = 0 by substituting θ = (0,1) into the latter condition. This produces the matrix

0 0 N = . (5.2.3.175) 3 0 r¯u¯b¯ − b ur  3 3 

Since N3 and B are non-vanishing, the condition from the [Q, Q, Q] identity no longer states that we must set either N0 or H to zero. We have

Re(θN0θ†)θH = θN3θ†θB. (5.2.3.176)

Substituting θ =(0,1) into the above condition, we find

bh3 =(r¯u¯b¯3 − b3ur)b3 and bh1 = 0. (5.2.3.177)

By assumption N3 = 0; therefore, both b and h3 cannot vanish. Using this result, the second 6 constraint tells us that h1 = 0. Thus H is reduced to a strictly lower-diagonal matrix. As in Sub-Branch 4.2, we have

0 = BNi − NiB† where i {0,1} ∈ (5.2.3.178) 0 = PN0 − N0P†, which tell us a and c vanish, and

0 = b3q − ¯qb¯3, 0 = b3uq − ¯qu¯b¯3, and 0 = b3r − r¯b¯3. (5.2.3.179)

Using these results and the rewriting of h3 in (5.2.3.177), the conditions from the [H, Q, Q] identity are instantly satisfied. Therefore, the constraints on the parameters of this sub-branch

172 are

0 = b3q − ¯qb¯3, 0 = b3uq − ¯qu¯b¯3, 0 = b3r − r¯b¯3, and bh3 =(r¯u¯b¯3 − b3ur)b3. (5.2.3.180)

Notice that the first three constraints here tell us that b3 is collinear with both q and r and that b3u is collinear with q. In particular, were we to use the endomorphisms of s1¯ to set q to lie along i, b3, b3u and r would all lie along i as well. Thus, b3ur R, such that N3 = 0. ∈ Therefore, this sub-branch is empty.

Sub-Branch 4.4 This sub-branch will be very similar to the one above due to the similarity in the conditions the super-Jacobi identity imposes on N3 and N4. Using

N4 = BN1 − N1B† and 0 = βθBN4θ† + θN4(βθB)†, (5.2.3.181) we know N4 may be written 0 0 N = . (5.2.3.182) 4 0 b r − r¯b¯  3 3 Lemma 5.2.3 then tells us that

Re(θN0θ†)θH = θN4θ†θP. (5.2.3.183)

Substituting θ =(0,1) into this condition produces

bh3 =(b3r − r¯b¯3)b3u and bh1 = 0. (5.2.3.184)

Since b3u = 0 and b3r − r¯b¯3 = 0 by assumption, b cannot vanish; therefore, h1 = 0. The 6 6 conditions

0 = PNi − NiP† where i {0,1} ∈ (5.2.3.185) 0 = BN0 − N0B†, tell us that both a and c vanish, and

0 = b3q − ¯qb¯3, 0 = b3uq − ¯qu¯b¯3, and 0 = b3ur − r¯u¯b¯3. (5.2.3.186)

Finally, we have the conditions from the [H, Q, Q] identity, which impose

0 = Re(h3q) and 0 = Re(h3r). (5.2.3.187)

However, using the form of h3 in (5.2.3.184) and the collinearity of b3u with q and r, both of these constraints are already satisfied. Therefore, the final set of constraints on this sub-branch is

0 = b3q − ¯qb¯3, 0 = b3uq − ¯qu¯b¯3, 0 = b3ur − r¯u¯b¯3, bh3 =(b3r − r¯b¯3)b3u. (5.2.3.188)

Notice that the first three constraints tell us that b3, b3u, q, and r are collinear. This tells us that b3r R; therefore, significantly, N4 = 0. Thus this sub-branch is empty. ∈

Sub-Branch 4.5 Now with non-vanishing N3 and N4, we can begin by using

−N3 = PN1 − N1P† N4 = BN1 + N1B† and (5.2.3.189) 0 = πθPN3θ† + θN3(πθP)† 0 = βθBN4θ† + θN4(βθB)†, to write 0 r 0 0 0 0 N = , N = and N = . (5.2.3.190) 1 r¯ d 3 0 r¯u¯b¯ − b ur 4 0 b r − r¯b¯    3 3   3 3

173 Using these results, substitute θ =(0,1) into the condition from the [Q, Q, Q] identity to find

bh3 =(r¯u¯b¯3 − b3ur)b3 +(b3r − r¯b¯3)b3u and bh1 = 0. (5.2.3.191)

As in all previous sub-branches, the [P, Q, Q] and [B, Q, Q] conditions on N0 tell us

0 = a, 0 = b3q − ¯qb¯3 and 0 = b3uq − ¯qu¯b¯3. (5.2.3.192)

Finally, the [H, Q, Q] identities tell us

0 = Re(h3q)+ b Re(h1) 0 = Re(h3r)+ d Re(h1) and (5.2.3.193) −1 −1 0 = h1q + qb3h1b3 0 = h1r + rb3h1b3 .

Since, by assumption, N1 = 0, these constraints mean we must have Re(h1) = 0. If r = 0, we 6 would need d = 0, which, when substituted into 0 = d Re(h1), mean Re(h1)= 0. Alternatively, 6 if r = 0, we multiply 6 −1 0 = h1r + rb3h1b3 (5.2.3.194) −1 on the right by r and take the real part to obtain Re(h1)= 0. Knowing this, we can use the −1 fact b3h1b Im(H) to rewrite the remaining imaginary part of this constraint as 3 ∈

0 =[h1, rb3]. (5.2.3.195)

Additionally, since Re(h1)= 0, we can use 0 = b3q − ¯qb¯3 to instantly satisfy the condition

−1 0 = h1q + qb3h1b3 . (5.2.3.196)

These results leave us with

Caˆ, 4.5.iii = {0 = b3q − ¯qb¯3, 0 = b3uq − ¯qu¯b¯3,

0 = Re(h3q), 0 = Re(h3r), 0 = Re(h1), (5.2.3.197)

0 =[h1, rb3], 0 = bh1, bh3 =−2 Im(b3ur)b3 + 2 Im(b3r)b3u}.

Subject to these constraints, the non-vanishing matrices are

0 0 0 0 h1 0 Maˆ, 4.5.iii = B = , P = , H = −1 , b3 0 b3u 0 h3 b3h1b3       0 q 0 r 0 0 0 0 N = , N = , N = , N = . 0 ¯q b 1 r¯ d 3 0 r¯u¯b¯ − b ur 4 0 b r − r¯b¯      3 3   3 3 (5.2.3.198)

The wealth of parameters describing this sub-branch mean we will only highlight one parame- terisation of these super-extensions here, though many more may exist. In particular, we will choose to set b, d and h3 to zero. We will also utilise the subgroup G GL(s¯) GL(s¯) to ⊂ 0 × 1 impose that q, b3 and b3u, lie along i. The residual endomorphisms of s1¯ may then scale r such that its norm becomes 1. Employing Aut(H), we can set h1 to lie along i as well. Having made these choices, the constraint 0 =[h1, rb3] (5.2.3.199) tells us r R 1, i . Notice that for N3 and N4 to be non-vanishing r must have a real component; ∈ h i therefore, to simplify the form of the matrices in our example, we will choose r = 1. The remaining constraints in Caˆ, 4.5.iii are then satisfied, and we can use the scaling symmetry of the s0¯ basis elements to produce

0 0 0 0 i 0 B = , P = , H = , i 0 i 0 0 i       (5.2.3.200) 0 i 0 1 0 0 0 0 N = , N = , N = and N = . 0 −i 0 1 1 0 3 0 i 4 0 i        

174 Sub-Branch 4.6 We find that this sub-branch is empty using the analysis from the previous sub-branch. Again, we use

−N3 = PN1 − N1P† N4 = BN1 + N1B† and (5.2.3.201) 0 = πθPN3θ† + θN3(πθP)† 0 = βθBN4θ† + θN4(βθB)†, to write 0 0 0 0 N = N = . (5.2.3.202) 3 0 r¯u¯b¯ − b ur 4 0 b r − r¯b¯  3 3   3 3 Substituting these matrices into

1 2 [β, θN2θ†]= βθBN3θ† + θN3(βθB)† 1 (5.2.3.203) 2 [π, θN2θ†]= πθPN4θ† + θN4(πθP)†, the R.H.S. of both of these constraints vanishes, setting N2 = 0. Therefore, this sub-branch is empty.

ˆn−

Setting µ = η = 0, λ = 1 and ε =−1, the first condition in (5.2.3.155) becomes

[u, h1]= 2u. (5.2.3.204)

Since [u, h1] is perpendicular to u in Im(H) this branch cannot provide a super-extension for ˆn−.

ˆn+

In this case, for which λ = ε = 0, µ = 1, and η = −1, the first constraint in (5.2.3.155) gives us 2 [h1, u]= u + 1. (5.2.3.205) Taking the real part of (5.2.3.205) produces

Re(u2)=−1, (5.2.3.206) therefore, u Im(H), such that |u|2 = 1, i.e. it is a unit-norm vector quaternion, or right . ∈ The imaginary part of (5.2.3.205) imposes

[u, h1]= 0. (5.2.3.207)

Thus, we could get a super-extension of ˆn+ in this branch. Wishing to write our parameters in −1 2 terms of b3, we have p3 = b3u and h4 = b3(h1 − u)b , where u Im(H), such that u =−1. 3 ∈ As with the ˆa case above, all of the conditions of Lemmas 5.2.2 and 5.2.3 must be taken into account. The conditions

N4 = BN1 − N1B† and − N3 = PN1 − N1P† (5.2.3.208) tell us that if N1 = 0, N3 = 0 and N4 = 0. Substituting these results into

− Re(θN0θ†)π = πθPN3θ† + θN3(πθP)† Re(θN0θ†)β = βθBN4θ† + θN4(βθB)† 1 1 2 [β, θN2θ†]= βθBN3θ† + θN3(βθB)† 2 [π, θN2θ†]= πθPN4θ† + sN4(πθP)†, (5.2.3.209)

175 we see that if N3 or N4 vanish, so must N0 and N2. Equally, if N3 vanishes N4 necessarily vanishes and vice versa due to

− N4 = HN3 + N3H† and N3 = HN4 + N4H†. (5.2.3.210)

Therefore, based on these dependencies, our investigation into this branch of possible super- extensions of ˆn+ divides into the following sub-branches.

1. N1 = 0, and N0 = N2 = N3 = N4 = 0 6

2. N1 = 0, N3 = 0, N4 = 0, and N0 = N2 = 0 6 6 6

3. N1 = 0, N3 = 0, N4 = 0, N0 = 0 and N2 = 0 6 6 6 6

4. N1 = 0, N3 = 0, N4 = 0, N0 = 0 and N2 = 0 6 6 6 6

5. N1 = 0, N3 = 0, N4 = 0, N0 = 0 and N2 = 0 6 6 6 6 6

Unlike the super-extensions of ˆn+ found in Branches 1, 2 and 3, the [Q, Q, Q] identity will not impose that either N0 or H must vanish. In the first sub-branch above, we instantly see that N0 = 0; therefore, the super-extensions found here are extensions satisfying (i). However, all other sub-branches have either non-vanishing N3 or non-vanishing N4. Since B = 0 and P = 0, 6 6 the [Q, Q, Q] identity will now form relationships between N0, N3 and N4, with, in general, H = 0. Therefore, these super-extensions, for which N0 = 0 and H = 0, will be labelled (iii) to 6 6 6 distinguish them from the cases (i) and (ii).

Sub-Branch 4.1 With only N1 = 0, the conditions from Lemmas 5.2.2 and 5.2.3 reduce to 6

0 = HN1 + N1H†

0 = PN1 − N1P† (5.2.3.211)

0 = BN1 − N1B†.

The latter two conditions tell us

0 = c, 0 = b3r − r¯b¯3 and 0 = b3ur − r¯u¯b¯3, (5.2.3.212) which, when substituted into the first conditions, produce

0 = Re(h1) and 0 = Re(h3r). (5.2.3.213)

Therefore, the non-vanishing matrices for this super-extension are

0 0 0 0 h 0 0 r B = , P = , H = 1 , N = , (5.2.3.214) b 0 b u 0 h b (h − u)b−1 1 r¯ d  3   3   3 3 1 3    subject to the constraints

2 0 =[u, h1], 0 = Re(h1), 0 = Re(h3r), 0 = b3r − r¯b¯3, 0 = b3ur − r¯u¯b¯3, u =−1. (5.2.3.215) However, notice that the final three constraints listed above require one of b3, u or r to vanish. Since neither b3 or u can vanish in this sub-branch, it must be that r = 0. Therefore, the final set of matrices is 0 0 0 0 h 0 0 0 M = B = , P = , H = 1 , N = , nˆ+, 4.1.i b 0 b u 0 h b (h − u)b−1 1 0 d  3   3   3 3 1 3    (5.2.3.216) and the final set of constraints is

2 Cnˆ+, 4.1.i = {0 =[u, h1], 0 = Re(h1), u =−1}. (5.2.3.217)

176 To demonstrate that this sub-branch of S is not empty, choose to set h1 and h3 to zero. Using the endomorphisms of s1¯ and Aut(H) on b3 and u, respectively, we may write b3 = i and u = j. Employing the scaling symmetry of Z, we arrive at the super-extension

0 0 0 0 0 0 0 0 B = , P = , H = , N = . (5.2.3.218) i 0 k 0 0 j 1 0 1        

Thus, this sub-branch is not empty. We may then introduce h1 while continuing to fix all the parameters of the super-extension; however, this cannot be achieved on introducing h3.

Sub-Branch 4.2 Now with N3 = 0 and N4 = 0 as well as N1 = 0, we can use the conditions 6 6 6

−N3 = PN1 − N1P† 0 = βθBN3θ† + θN3(βθB)† 0 = βθBN4θ† + θN4(βθB)†

N4 = BN1 − N1B† 0 = πθPN3θ† + θN3(πθP)† 0 = πθPN4θ† + θN4(πθP)†, (5.2.3.219) and the analysis of Branches 2 and 3 to write

0 r 0 0 0 0 N = N = N = . (5.2.3.220) 1 r¯ d 3 0 r¯u¯b¯ − b ur 4 0 b r − r¯b¯    3 3   3 3 This leaves only the [H, Q, Q] conditions:

0 = HN1 + N1H†

−N4 = HN3 + N3H† (5.2.3.221)

N3 = HN4 + N4H†.

We know from Sub-Branch 4.1.i that, since c = 0, the first of these produces

0 = Re(h1) and 0 = Re(h3r). (5.2.3.222)

Writingr ¯u¯b¯3 −b3ur =−2 Im(b3ur) and b3r−r¯b¯3 = 2 Im(b3r) to simplify our expressions, the second and third conditions give us

Im(b3r)= h4 Im(b3ur)+ Im(b3ur)h¯4 (5.2.3.223) − Im(b3ur)= h4 Im(b3r)+ Im(b3r)h¯4,

−1 respectively, where h4 = b3(h1 − u)b . Notice that since h1, u Im(H), and h4 is written in 3 ∈ terms of the adjoint action of b3 H, h4 Im(H). Therefore, using h¯4 =−h4, we find ∈ ∈

Im(b3r)=−[h4, [h4, Im(b3r)]] and Im(b3ur)=−[h4, [h4, Im(b3ur)]]. (5.2.3.224)

This imposes the constraint that Im(b3r) and Im(b3ur) must be perpendicular to h4 in Im(H). The constraints for this sub-branch are summarised as follows.

2 Cnˆ , 4.2.iii = {0 =[h1, u], −1 = u 0 = Re(h1), 0 = Re(h3r), + (5.2.3.225) Im(b3r)=−[h4, [h4, Im(b3r)]], Im(b3ur)=−[h4, [h4, Im(b3ur)]]}.

The non-vanishing matrices are then

0 0 0 0 h1 0 Mnˆ+, 4.2.iii = B = , P = , H = −1 , b3 0 b3u 0 h3 b3(h1 − u)b3       0 r 0 0 0 0 N = , N = , N = . 1 r¯ d 3 0 −2 Im(b ur) 4 0 2 Im(b r)    3   3  (5.2.3.226)

177 To demonstrate the existence of super-extensions in this sub-branch, we will begin by simplifying our parameter set as much as possible. In particular, we begin by setting both h3 and d to zero. We then utilise Aut(H) and the endomorphisms of s1¯ to set u = j and impose that b3 lies along i. Notice that with u along j, the first constraint in Cnˆ+, 4.2.iii tells us that h1 must −1 also lie along j, as must h4 = b3(h1 − u)b3 . With these choices, the two constraints involving 1 3 h4 impose r R 1, j , and that |h1| = or |h1| = . Residual endomorphisms then allow us to ∈ h i 2 2 scale r such that it has unit norm. Finally, we can scale the s0¯ basis elements to arrive at

0 0 0 0 j 0 B = , P = , H = , i 0 k 0 0 j       (5.2.3.227) 0 1 + j 0 0 0 0 N = , N = , N = . 1 1 − j 0 3 0 k − i 4 0 k + i      

Sub-Branch 4.3 The beginning of the investigation of this sub-branch is identical to that of the previous sub-branch. The [P, Q, Q] and [B, Q, Q] identities produce

−N3 = PN1 − N1P† 0 = βθBN3θ† + θN3(βθB)† (5.2.3.228) N4 = BN1 − N1B† 0 = πθPN4θ† + θN4(πθP)†, where the first two conditions give N3 and N4 the form

0 cb u 0 −cb¯ N = 3 and N = 3 . (5.2.3.229) 3 −cb u r¯u¯b¯ − b ur 4 cb b r − r¯b¯  3 3 3   3 3 3

Substituting this N3 with θ =(0,1) into

0 = βθBN3θ† + θN3(βθB)†, (5.2.3.230) we acquire 2 0 = 2c|b3| Im(βu) β Im(H). (5.2.3.231) ∀ ∈ As, by assumption, b3 = 0 and u = 0, this imposes c = 0, such that 6 6 0 r 0 0 0 0 N = N = N = . (5.2.3.232) 1 r¯ d 3 0 r¯u¯b¯ − b ur 4 0 b r − r¯b¯    3 3   3 3

With this form of N3 and N4,

− Re(θN0θ†)π = πθPN3θ† + θN3(πθP)† (5.2.3.233) Re(θN0θ†)β = βθBN4θ† + θN4(βθB)†, have a vanishing R.H.S., showing that N0 = 0. This result contradicts our assumption that N0 = 0; therefore, this sub-branch does not contain any super-extensions. 6

Sub-Branch 4.4 Letting N0 = 0 and N2 = 0, we can use 6

−N3 = PN1 − N1P† 0 = βθBN4θ† + θN4(βθB)† (5.2.3.234) N4 = BN1 − N1B† 0 = πθPN3θ† + θN3(πθP)†, to again write

0 r 0 0 0 0 N = N = N = . (5.2.3.235) 1 r¯ d 3 0 r¯u¯b¯ − b ur 4 0 b r − r¯b¯    3 3   3 3

178 Substituting these Ni into

1 2 [β, θN2θ†]= βθBN3θ† + θN3(βθB)† 1 (5.2.3.236) 2 [π, θN2θ†]= πθPN4θ† + θN4(πθP)†, the R.H.S. vanishes for both, showing N2 = 0, contradicting our initial assumption in this sub- branch.

Sub-Branch 4.5 With none of the Ni vanishing, we start again by writing N3 and N4 in terms of N1 using conditions from the [P, Q, Q] and [B, Q, Q] identities:

0 cb u 0 −cb¯ N = 3 and N = 3 . (5.2.3.237) 3 −cb u r¯u¯b¯ − b ur 4 cb b r − r¯b¯  3 3 3   3 3 3 Letting n m N2 = where n, l Im(H), m H, (5.2.3.238) −m¯ l ∈ ∈   we can use 1 2 [β, θN2θ†]= βθBN3θ† + θN3(βθB)† (5.2.3.239) to write N2 in terms of b3 and u. First let θ =(0,1) to find

1 [β, l]=−c[β, b3ub¯3] 0 = β Im(H). (5.2.3.240) 2 ∀ 6 ∈ Therefore, l =−2cb3ub¯3. (5.2.3.241) Next, substitute in θ =(1,1) to get

1 ¯ [β,2Im(m)] + 2 [β, l]=−c[β, b3ub3]. (5.2.3.242) Using the previous result, this tells us that Im(m) = 0. Analogous calculations with θ = (0, i) and θ = (1, i) show that, in fact, all of m must vanish. Finally, substituting in θ = (1,0) into (5.2.3.239), we find n = 0 since the R.H.S. vanishes. Therefore, we are left with

0 0 N = . (5.2.3.243) 2 0 −2cb ub¯  3 3

We would have arrived at the same expression had we used N4 and

1 2 [π, θN2θ†]= πθPN4θ† + θN4(πθP)†. (5.2.3.244)

This form of N2 automatically satisfies all other conditions it is involved in from both the [B, Q, Q] and [P, Q, Q] identities. Finally, we can put this N2 into

0 = HN2 + N2H† (5.2.3.245) to get 0 = h4l + lh¯4, (5.2.3.246) −1 ¯ 2 where h4 = b3(h1 − u)b3 and l =−2cb3ub3. Working through some algebra, noting Re(u )= −1 and the fact c = 0 for N2 = 0, we arrive at h1u = h1u. Since u Im(H), this forces 6 6 ∈ h1 Im(H) such that u and h1 are collinear. ∈

Now turn to N0 and consider

− Re(θN0θ†)π = πθPN3θ† + θN3(πθP)† (5.2.3.247) Re(θN0θ†)β = βθBN4θ† + θN4(βθB)†.

179 Letting θ =(1,0) in either of these conditions, we find that a = 0. Next, substituting θ =(0,1) into the second condition produces 2 − b = 2c|b3| . (5.2.3.248) We would have arrived at the same result had we substituted into the first condition and used the fact |u|2 = 1. Now substituting θ =(1, s) into the second condition, we find

2 2 2 − 2 Re(s¯q)− b|s| = 2c|s| |b3| . (5.2.3.249)

Therefore, using the previous result and letting s = 1, s = i, s = j and s = k, we see that all components of q must vanish. All other conditions on N0 are now automatically satisfied, leaving 0 0 N = . (5.2.3.250) 0 0 −2c|b |2  3  Equipped with these Ni, we can now analyse the condition from Lemma 5.2.3:

1 Re(θN0θ†)θH = 2 θN2θ†θ + θN3θ†θB + θN4s†θP. (5.2.3.251) Letting θ =(0,1):

2 −1 ¯ − 2c|b3| h3 b3(h1 − u)b3 =−cb3ub3 0 1 − Im(b3ur) b3 0 + Im(b3r) b3u 0 . (5.2.3.252)     Concentrating on the second component, we have

2 −1 ¯ − 2c|b3| b3(h1 − u)b3 =−cb3ub3. (5.2.3.253)

2 Using the fact |b3| b3 = b¯3 and cancelling relevant terms leaves

2b3h1b¯3 = 0. (5.2.3.254)

Since, by assumption b3 = 0, we get h1 = 0. The first component of (5.2.3.252) gives us a 6 prescription for h3, 2 − 2c|b3| h3 =−2 Im(b3ur)b3 + 2 Im(b3r)b3u, (5.2.3.255) therefore, we can fully describe H in terms of B, P, and N1.

The final conditions to consider are those from the [H, Q, Q] super-Jacobi identity for N1, N3 and N4. First, the N1 condition tell us

¯ −1 0 = ch3 + rb3ub3 0 = Re(h3r). (5.2.3.256)

Notice that the second constraint here is automatically satisfied by the first, since c = 0 for a 6 non-vanishing N0. Substituting this expression for h3 into the previous prescription, we find

2 2 −1 |r| |b3| b3ub3 =[Im(b3r), Im(b3ur)] + Re(b3ur) Im(b3r)− Re(b3r) Im(b3ur). (5.2.3.257)

Now, the constraints that the N3 and N4 conditions produce are the ones given in Sub-Branch 4.2.iii:

Im(b3r)=−[h4, [h4, Im(b3r)]] and Im(b3ur)=−[h4, [h4, Im(b3ur)]]. (5.2.3.258)

These tell us that Im(b3r) and Im(b3ur) are perpendicular to h4 in Im(H). Therefore, the expression in (5.2.3.257) becomes

2 0 = Re(b3r), 0 = Re(b3ur) and |r| b3ub¯3 =[Im(b3r), Im(b3ur)]. (5.2.3.259)

180 Putting all of these constraints together, we have

2 Cnˆ+, 4.5.iii = {u =−1, Im(b3r)=−[h4, [h4, Im(b3r)]], Im(b3ur)=−[h4, [h4, Im(b3ur)]], 2 0 = Re(b3r), 0 = Re(b3ur), |r| b3ub¯3 =[Im(b3r), Im(b3ur)]}, (5.2.3.260) for non-vanishing matrices

0 0 0 0 0 0 Mnˆ+, 4.5.iii = B = , P = , H = −1 −1 −1 , b3 0 b3u 0 −c b3ub3 r¯ b3ub3       0 0 c r 0 0 N = , N = , N = , (5.2.3.261) 0 0 −2c|b |2 1 r¯ d 2 0 −2cb ub¯  3     3 3 0 cb3u 0 −cb¯3 N3 = , N4 = . −cb3u −2 Im(b3ur) cb3 2 Im(b3r)     To demonstrate that there are super-extensions in this sub-branch, we will first simplify this system by letting parameters vanish where possible. In particular, r and d in N1 may be set 2 to zero. This reduces Cnˆ+, 4.5.iii to contain only u = −1. Now we can use the endomorphisms of s1¯ to impose b3 = i, u = j and c = 1. With these choices, the matrices become

0 0 0 0 0 0 B = , P = , H = , i 0 k 0 0 j       0 0 1 0 0 0 N = , N = , N = , (5.2.3.262) 0 0 −2 1 0 0 2 0 2j       0 −k 0 i N = , N = . 3 −k 0 4 i 0     As there are no restrictions on the parameter d, we may introduce it without affecting our other parameter choices; however, this is not the case for r. There are several constraints in

Cnˆ+, 4.5.iii involving r; therefore, we need to examine these constraints to determine whether new parameters must be chosen. Interrogating

Im(b3r)=−[h4, [h4, Im(b3r)]] and Im(b3ur)=−[h4, [h4, Im(b3ur)]] (5.2.3.263) with the parameter choices stated above, we find that r must vanish. In particular, due to −1 h4 = b3ub3 having unit length, we cannot replicate the analysis of Sub-Branch 4.2.iii, where 1 1 the magnitude of h4 was necessarily either + 2 or − 2 . Thus, we cannot produce a super- extension in this sub-branch for which r = 0. This simplifies the above (M, C), such that the 6 remaining constraints are 2 Cnˆ+, 4.5.iii = {u =−1}, (5.2.3.264) and the non-vanishing matrices are now

0 0 0 0 0 0 Mnˆ+, 4.5.iii = B = , P = , H = −1 , b3 0 b3u 0 0 b3ub3       0 0 c 0 0 0 N = , N = , N = , (5.2.3.265) 0 0 −2c|b |2 1 0 d 2 0 −2cb ub¯  3     3 3 0 cb3u 0 −cb¯3 N3 = , N4 = . −cb3u 0 cb3 0    

ˆg

Finally, substitute λ = η = ε = 0 and µ = −1 into the second constraint in (5.2.3.155) to investigate the ˆg case. We find [v, h1]=−1, (5.2.3.266)

181 which, since [v, h1] Im(H), is inconsistent. Therefore, we cannot get a super-extension of ˆg in ∈ this branch.

5.2.4 Summary Table 5.6 lists all the sub-branches of S we found that contain N = 2 generalised Bargmann superalgebras. Each Lie superalgebra in one of these branches is an N = 2 super-extension of one of the generalised Bargmann algebras given in Table 5.2, taken from [3]. It is interesting that Z only appears in [B, P]= Z and [Q, Q]= Z. (5.2.4.1) Therefore, in all instances, Z remains central after the super-extension. In particular, this means that we may always find a kinematical Lie superalgebra (without a central-extension) by taking the quotient of our generalised Bargmann superalgebra s by the R-span of Z, s/ Z . h i Table 5.6: Sub-Branches of N = 2 Generalised Bargmann Superalgebras (with [Q, Q] = 0) 6

(S)B k H Z B P [Q, Q] 1.i ˆa X Z 1.ii ˆa H + Z 2.1.ii ˆa X H 2.2.i ˆa X X Z 2.3.i ˆa X X Z + B 2.3.ii ˆa X H + Z + B 4.5.iii ˆa X X X H + Z + B + P

1.i ˆn− X Z

2.2.i ˆn− X X Z

2.3.i ˆn− X X Z + P

3.2.i ˆn− X X Z

3.3.i ˆn− X X Z + P

1.i ˆn+ X Z

2.2.i ˆn+ X X Z

4.1.i ˆn+ X X X Z

4.2.iii ˆn+ X X X Z + B + P

4.5.iii ˆn+ X X X H + Z + B + P 1.i ˆg X Z 2.2.i ˆg X X Z 3.2.i ˆg X X Z 3.3.i ˆg X X Z + P 3.3.ii ˆg X H + Z + P

The first column indicates the sub-branch of generalised Bargmann superalgebras, so that the reader may navigate back to find the conditions on the non-vanishing parameters of these superalgberas. The second column then tells us the underlying generalised Bargmann algebra k. The next four columns tells us which of the s0¯ generators H,Z, B, and P act on Q. Recall, J necessarily acts on Q, so we do not need to state this explicitly. The final column shows which s0¯ generators appear in the [Q, Q] bracket.

Unpacking the Notation Although the formalism employed in this classification was useful for our purposes, it may be unfamiliar to the reader. Therefore, in this section, we will convert one of the N = 2 super- extensions of the Bargmann algebra in sub-branch 3.3.ii into a more standard notation. The

182 brackets for this algebra, excluding the s0¯ brackets, which are shown in Table 5.2, take the form

[B(β), Q(θ)] = Q(βθB) and [Q(θ), Q(θ)] = Re(θN0θ†)H+Re(θN1θ†)Z−P(θN4θ†), (5.2.4.2) where

0 0 0 0 c r 0 −cb¯3 B = N0 = 2 N1 = N4 = . (5.2.4.3) b3 0 0 −2c|b3| r¯ d cb b r − r¯b¯        3 3 3 1 1 2 Let {Qa} be a real basis for the first so(3) spinor module in s1¯ = S S where a {1,2,3,4} , 2 ⊕ ∈ and {Qa} be a basis for the second so(3) spinor module. Letting θ =(θ1, θ2), and substituting the above matrices into the [s0¯, s1¯] bracket, we get

1 2 1 [B(β), Q (θ1)] = 0 and [B(β), Q (θ2)] = Q (βθ2b3). (5.2.4.4)

Substituting θ = θ ′ = (θ1,0), θ =(θ1,0) and θ ′ =(0, θ2), and θ = θ ′ =(0, θ2) into the [s1¯, s1¯] bracket we find

1 1 2 [Q (θ1), Q (θ1)] = c|θ1| Z Q1 Q2 ¯ c P ¯ ¯ ¯ [ (θ1), (θ2)] = Re(θ1rθ2)Z − 2 (θ2b3θ1 − θ1b3θ2) (5.2.4.5) 2 2 2 2 2 [Q (θ2), Q (θ2)]=−2c|b3| |θ2| H + d|θ2| Z − P(θ2(b3r − r¯b¯3)θ¯2).

For the purposes of the present example, we will set the parameters of this super-extension as specified in (5.2.3.152); therefore, we have [s0¯, s1¯] brackets

1 2 2 [B(β), Q (θ1)] = Q (βθ2i) and [B(β), Q (θ2)] = 0, (5.2.4.6) and [s1¯, s1¯] brackets 1 1 2 1 2 1 ¯ ¯ 2 2 2 [Q (θ1), Q (θ1)] = |θ1| Z, [Q (θ1), Q (θ2)]=− 2 P(θ2iθ1+θ1iθ2) and [Q (θ2), Q (θ2)] = |θ2| H. (5.2.4.7) Now, we can write

4 3 2 1 b 1 1 1 2 i 2 2 [Bi,Qa]= Qbβi a [Qa,Qb]= δabZ [Qa,Qb]= PiΓab, [Qa,Qb]= δabH. b=1 i=1 X X (5.2.4.8) Our brackets then produce the βi matrices

−1 0 0 −σ 0 σ β = β = 1 β = 3 , (5.2.4.9) 1 0 1 2 −σ 0 3 σ 0    1   3  and the symmetric Γ i matrices

−1 0 0 −σ 0 σ Γ 1 = Γ 2 = 1 Γ 3 = 3 , (5.2.4.10) 0 1 −σ 0 σ 0    1   3  where σ1 and σ3 are the first and third Pauli matrix, respectively. This N = 2 Bargmann superalgebra takes the same form as the (2 + 1)-dimensional Bargmann superalgebra utilised in [109].

5.3 Conclusion

In this chapter, we classified the N = 1 super-extensions of the generalised Bargmann alge- bras with three-dimensional spatial isotropy up to isomorphism. We also presented the non- empty sub-branches of the variety S describing the N = 2 super-extensions of the generalised Bargmann algebras. To simplify this classification problem, we utilised a quaternionic formal- ism such that so(3) scalar modules were described by copies of R, so(3) vector modules were described by copies of Im(H), and so(3) spinor modules were described by copies of H. We began

183 by defining a universal generalised Bargmann algebra, which, under the appropriate setting of some parameters, may be reduced to the centrally-extended static kinematical Lie algebra ˆa, the centrally-extended Newton-Hooke algebras ˆn , or the Bargmann algebra ˆg. The most general ± form for the [s0¯, s1¯] and [s1¯, s1¯] bracket components were found before substituting them into the super-Jacobi identity and finding the constraints on the parameters for these maps. Because of the formalism in use, solving these constraints amounted to some linear algebra over the quaternions, and paying attention to the allowed basis transformations G GL(s¯) GL(s¯). ⊂ 0 × 1 Since we are only interested in supersymmetric extensions of these algebras, we limited our- selves to those branches which allow for non-vanishing [Q, Q]. The results of the N = 1 and N = 2 analyses are in Tables 5.3 and 5.6, respectively. We found 9 isomorphism classes in the N = 1 case, and 22 non-empty sub-branches in the N = 2 case.

184 Chapter 6

Conclusion

In this thesis, we presented a framework to explore kinematical symmetries beyond the standard Lorentzian case. This framework consisted of an algebraic classification, a geometric classifi- cation, and a derivation of the geometric properties required to define physical theories on the classified spacetime geometries. We will now briefly run through the main results from each of the primary chapters.

In Chapter 3, each step in this framework was discussed in detail for the case of kinemati- cal symmetries. Section 3.1 reviewed the classification of kinematical Lie algebras with spatial dimension D = 3 (up to isomorphism). These Lie algebras assumed spatial isotropy as well as homogeneity in both time and space. Known and named Lie algebras were identified, and 18 isomorphism classes were found. In Section 3.2, the integration of these Lie algebras to spatially-isotropic simply-connected homogeneous spacetime was discussed, and the classifi- cation of these geometries was reviewed. Five kinematical spacetime classes were identified; namely, Lorentzian, Riemannian, Galilean, Carrollian, and Aristotelian. This classification was followed by identifying the characteristic Lie brackets of each spacetime class and identifying a procedure for connecting the spacetimes via geometric limits. Section 3.4 then derived various geometric properties for each kinematical spacetime including the fundamental vector fields, soldering forms, vielbeins, invariant structures, and the space of invariant affine connections.

In Chapter 4, the algebraic and geometric classifications of our framework were carried out in the super-kinematical case. Section 4.1 gave the classification of the N = 1 kinematical and Aristotelian Lie superalgebras in three spatial dimensions, up to isomorphism. We found 43 isomorphism classes, some with parameters. In Section 4.2, we then classified the corresponding simply-connected homogeneous (4|4)-dimensional kinematical superspaces, finding 27 isomor- phism classes. It was then shown how these superspaces might be connected via geometric limits and the low-rank invariants of each superspace were determined.

Chapter 5 presented the algebraic classification of our framework for the super-Bargmann case. In Section 5.1, the classification of N = 1 generalised Bargmann superalgebras was presented, identifying 9 isomorphism classes of Lie superalgebra. Section 5.2 then discussed the generali- sation of this classification to the N = 2 case. Here, owing to the increased complexity of the problem, we only identified non-empty branches in the real algebraic variety S describing the possible generalised Bargmann superalgebra structures. In particular, we found 22 non-empty branches in S .

Outside the classifications and the derivations mentioned above, some other key results in- clude the proof that boosts act with non-compact orbits in Lorentzian, Galilean, and Carrollian spacetimes. The property of having non-compact boosts is an essential physical requirement, as discussed in [20]. If this property does not hold, then a sufficiently large boost would no longer be a boost; we would arrive back at our starting point. Thus, compact boosts are deemed unphysical.

185 Another significant result presented in this thesis was the demonstration that kinematical and Aristotelian Lie algebras may admit several non-isomorphic super-extensions. These non- isomorphic super-extensions may then integrate into inequivalent kinematical and Aristotelian superspaces. Interestingly, due to geometric realisability and effectivity being, respectively, in- dependent and dependent of the super-extension, Aristotelian Lie algebras may form effective Lie super pairs (s, h) where the underlying Aristotelian Lie pair (k, h) is not effective. In these cases, the “boost” generators act as R-symmetries, which transform the odd dimensions, but not the even dimensions.

This thesis’s results suggest several directions for future study, which we will present here in no particular order. Given that the geometric classifications contain homogeneous (super)spaces, which may also be called Klein geometries, a natural generalisation would be to build Cartan geometries modelled on these spaces. This construction is described at length in [92]; how- ever, for our purposes, the crucial point is that we can view our classifications as describing the possible local geometries for a spacetime manifold. By modelling a Cartan geometry on these spaces, we allow for the introduction of curvature in the same way Riemannian geom- etry introduces curvature to Euclidean geometry. This process was developed by Cartan in his rewriting of Newtonian gravity in [13, 14]. Additionally, for a discussion on this process in 2 + 1-dimensions, see [57]. Note, the addition of Cartan geometries would constitute a new step in the presented framework, taking us closer to a complete set of kinematical spacetime theories.

More immediate work, which is currently underway, is to complete the presented framework for all the given cases. In particular, this would involve deriving the geometric properties for the kinematical superspaces, building the superspaces corresponding to the generalised Bargmann superalgebras, and calculating the necessary geometric properties in these instances. Work on constructing the non-supersymmetric Bargmann spacetimes is also in progress and should ap- pear soon.

In addition to further classifications, we may look to build theories using some of the novel (super)spaces found here. In particular, we may look to utilise some of these spaces in holo- graphic theories, as in [33, 35, 110, 111] or we may gauge some of the Bargmann superalgebras to arrive at novel non-relativistic supergravity theories. Additionally, in a similar spirit to the Cartan generalisation mentioned above, we could look to generate interesting gravity theories by putting the classified Lie (super)algebras through the procedure which leads to MacDowell- Mansouri gravity [112–114].

186 Bibliography

[1] J. Figueroa-O’Farrill, R. Grassie, and S. Prohazka, “Geometry and BMS Lie algebras of spatially isotropic homogeneous spacetimes,” arXiv:1905.00034 [hep-th].

[2] J. Figueroa-O’Farrill and R. Grassie, “Kinematical superspaces,” arXiv:1908.11278 [hep-th].

[3] J. M. Figueroa-O’Farrill, “Kinematical Lie algebras via deformation theory,” J. Math. Phys. 59 (2018), no. 6, 061701, arXiv:1711.06111 [hep-th].

[4] J. M. Figueroa-O’Farrill, “Higher-dimensional kinematical Lie algebras via deformation theory,” J. Math. Phys. 59 (2018), no. 6, 061702, arXiv:1711.07363 [hep-th].

[5] J. Figueroa-O’Farrill and S. Prohazka, “Spatially isotropic homogeneous spacetimes,” arXiv:1809.01224 [hep-th].

[6] K. Brading and E. Castellani, “Symmetries in physics: Philosophical reflections,” arXiv:quant-ph/0301097.

[7] E. P. Wigner, Symmetries and reflections: Scientific essays.

[8] D. J. Gross, “The role of symmetry in fundamental physics,” Proceedings of the National Academy of Sciences 93 (1996), no. 25, 14256–14259, https://www.pnas.org/content/93/25/14256.full.pdf.

[9] J. D. Logan, The Physical Origins of Partial Differential Equations, pp. 1–49. Springer US, New York, NY, 1998.

[10] H. A. Kastrup, “The contributions of emmy noether, felix klein and sophus lie to the modern concept of symmetries in physical systems,” tech. rep., PRE-27220, 1984.

[11] C. N. Yang, “Symmetry and physics,” Proceedings of the American Philosophical Society 140 (1996), no. 3, 267–288.

[12] T. Hawkins, “The erlanger programm of felix klein: Reflections on its place in the history of mathematics,” Historia Mathematica 11 (1984), no. 4, 442–470.

[13] É. Cartan, “Sur les variétés à connexion affine et la théorie de la relativité généralisée (première partie),” in Annales scientifiques de l’École normale supérieure, vol. 40, pp. 325–412. 1923.

[14] E. Cartan, “Sur les variétés à connexion affine, et la théorie de la relativité généralisée (première partie)(suite),” in Annales scientifiques de l’École Normale Supérieure, vol. 41, pp. 1–25. 1924.

[15] T. Koetsier, “The case of kinematics, the genesis of a discipline,” in Explorations in the History of Machines and Mechanisms, T. Koetsier and M. Ceccarelli, eds., pp. 491–504. Springer Netherlands, Dordrecht, 2012.

[16] A. A. Martínez, Kinematics: The lost origins of Einstein’s relativity, vol. 256. Johns Hopkins University Press Baltimore, 2009.

187 [17] C. W. F. Everitt, Experimental Tests of General Relativity: Past, Present and Future, pp. 529–555. Springer US, Boston, MA, 1980. [18] H. Bacry and J. Lévy-Leblond, “Possible kinematics,” Journal of Mathematical Physics 9 (1968), no. 10, 1605–1614, https://doi.org/10.1063/1.1664490. [19] H. Bacry and J. Nuyts, “Classification of ten-dimensional kinematical groups with space isotropy,” Journal of Mathematical Physics 27 (1986), no. 10, 2455–2457, https://doi.org/10.1063/1.527306. [20] H. Bacry and J. Levy-Leblond, “Possible kinematics,” J. Math. Phys. 9 (1968) 1605–1614. [21] H. Bacry and J. Nuyts, “Classification of Ten-dimensional Kinematical Groups With Space Isotropy,” J. Math. Phys. 27 (1986) 2455. [22] T. Andrzejewski and J. M. Figueroa-O’Farrill, “Kinematical lie algebras in 2 + 1 dimensions,” J. Math. Phys. 59 (2018), no. 6, 061703, arXiv:1802.04048 [hep-th]. [23] A. Einstein, “Zur elektrodynamik bewegter körper,” Annalen der physik 4 (1905). [24] M. Peskin, An introduction to quantum field theory. CRC press, 2018. [25] A. Zee, Quantum field theory in a nutshell, vol. 7. Princeton university press, 2010. [26] K. Osterwalder and R. Schrader, “Axioms for euclidean green’s functions,” Comm. Math. Phys. 31 (1973), no. 2, 83–112. [27] R. Banerjee, A. Mitra, and P. Mukherjee, “A new formulation of non-relativistic diffeomorphism invariance,” Phys. Lett. B 737 (2014) 369–373, arXiv:1404.4491 [gr-qc]. [28] D. T. Son, “Newton-Cartan Geometry and the Quantum Hall Effect,” arXiv:1306.0638 [cond-mat.mes-hall]. [29] M. Geracie, D. T. Son, C. Wu, and S.-F. Wu, “Spacetime Symmetries of the Quantum Hall Effect,” Phys. Rev. D 91 (2015) 045030, arXiv:1407.1252 [cond-mat.mes-hall]. [30] P. Salgado-Rebolledo and G. Palumbo, “Extended Nappi-Witten Geometry for the Fractional Quantum Hall Effect,” arXiv:2102.03886 [cond-mat.mes-hall]. [31] M. H. Christensen, J. Hartong, N. A. Obers, and B. Rollier, “Torsional Newton-Cartan Geometry and Lifshitz Holography,” Phys. Rev. D 89 (2014) 061901, arXiv:1311.4794 [hep-th]. [32] J. Hartong, Y. Lei, and N. A. Obers, “Nonrelativistic Chern-Simons theories and three-dimensional Horava-Lifshitz gravity,” Phys. Rev. D 94 (2016), no. 6, 065027, arXiv:1604.08054 [hep-th]. [33] J. Hartong and N. A. Obers, “Horava-Lifshitz gravity from dynamical Newton-Cartan geometry,” JHEP 07 (2015) 155, arXiv:1504.07461 [hep-th]. [34] E. A. Bergshoeff, J. Hartong, and J. Rosseel, “Torsional Newton–Cartan geometry and the Schrödinger algebra,” Class. Quant. Grav. 32 (2015), no. 13, 135017, arXiv:1409.5555 [hep-th]. [35] J. Hartong, E. Kiritsis, and N. A. Obers, “Lifshitz space–times for Schrödinger holography,” Phys. Lett. B 746 (2015) 318–324, arXiv:1409.1519 [hep-th]. [36] J. Kluson, “Remark About Non-Relativistic String in Newton-Cartan Background and Null Reduction,” JHEP 05 (2018) 041, arXiv:1803.07336 [hep-th]. [37] T. Harmark, J. Hartong, L. Menculini, N. A. Obers, and G. Oling, “Relating non-relativistic string theories,” JHEP 11 (2019) 071, arXiv:1907.01663 [hep-th].

188 [38] E. Bergshoeff, A. Chatzistavrakidis, J. Lahnsteiner, L. Romano, and J. Rosseel, “Non-Relativistic Supersymmetry on Curved Three-Manifolds,” arXiv:2005.09001 [hep-th]. [39] R. Auzzi, S. Baiguera, G. Nardelli, and S. Penati, “Renormalization properties of a Galilean Wess-Zumino model,” JHEP 06 (2019) 048, arXiv:1904.08404 [hep-th]. [40] C. D. Blair, “A worldsheet supersymmetric Newton-Cartan string,” JHEP 10 (2019) 266, arXiv:1908.00074 [hep-th]. [41] K. Morand and J.-H. Park, “Classification of non-Riemannian doubled-yet-gauged spacetime,” Eur. Phys. J. C 77 (2017), no. 10, 685, arXiv:1707.03713 [hep-th]. [Erratum: Eur.Phys.J.C 78, 901 (2018)]. [42] S. M. Ko, C. Melby-Thompson, R. Meyer, and J.-H. Park, “Dynamics of Perturbations in Double Field Theory & Non-Relativistic String Theory,” JHEP 12 (2015) 144, arXiv:1508.01121 [hep-th]. [43] D. S. Berman, C. D. Blair, and R. Otsuki, “Non-Riemannian geometry of M-theory,” JHEP 07 (2019) 175, arXiv:1902.01867 [hep-th]. [44] D. Grumiller, J. Hartong, S. Prohazka, and J. Salzer, “Limits of JT gravity,” arXiv:2011.13870 [hep-th]. [45] J. Gomis, D. Hidalgo, and P. Salgado-Rebolledo, “Non-relativistic and Carrollian limits of Jackiw-Teitelboim gravity,” arXiv:2011.15053 [hep-th]. [46] R. Andringa, E. Bergshoeff, S. Panda, and M. de Roo, “Newtonian Gravity and the Bargmann Algebra,” Class. Quant. Grav. 28 (2011) 105011, arXiv:1011.1145 [hep-th]. [47] E. Bergshoeff, J. Rosseel, and P. Townsend, “Non-relativistic limits and 3d massive gravity,” Journal of Physics: Conference Series 1194 (apr, 2019) 012013. [48] E. Bergshoeff, J. Gomis, M. Kovacevic, L. Parra, J. Rosseel, and T. Zojer, “Nonrelativistic superparticle in a curved background,” Phys. Rev. D 90 (2014), no. 6, 065006, arXiv:1406.7286 [hep-th]. [49] E. Bergshoeff, A. Chatzistavrakidis, L. Romano, and J. Rosseel, “Newton-Cartan Gravity and Torsion,” JHEP 10 (2017) 194, arXiv:1708.05414 [hep-th]. [50] D. Hansen, J. Hartong, and N. A. Obers, “Action Principle for Newtonian Gravity,” Phys. Rev. Lett. 122 (2019), no. 6, 061106, arXiv:1807.04765 [hep-th]. [51] D. Hansen, J. Hartong, and N. A. Obers, “Non-Relativistic Gravity and its Coupling to Matter,” JHEP 06 (2020) 145, arXiv:2001.10277 [gr-qc]. [52] J. Gomis, A. Kleinschmidt, and J. Palmkvist, “Symmetries of M-theory and free Lie superalgebras,” JHEP 03 (2019) 160, arXiv:1809.09171 [hep-th]. [53] J. Gomis, A. Kleinschmidt, J. Palmkvist, and P. Salgado-Rebolledo, “Newton-Hooke/Carrollian expansions of (A)dS and Chern-Simons gravity,” JHEP 20 (2020) 009, arXiv:1912.07564 [hep-th]. [54] E. Bergshoeff, J. Gomis, and P. Salgado-Rebolledo, “Non-relativistic limits and three-dimensional coadjoint Poincare gravity,” arXiv:2001.11790 [hep-th]. [55] J. A. de Azcárraga, D. Gútiez, and J. M. Izquierdo, “Extended D = 3 Bargmann supergravity from a Lie algebra expansion,” arXiv:1904.12786 [hep-th]. [56] J. A. de Azcarraga, J. M. Izquierdo, M. Picon, and O. Varela, “Generating Lie and gauge free differential (super)algebras by expanding Maurer-Cartan forms and Chern-Simons supergravity,” Nucl. Phys. B 662 (2003) 185–219, arXiv:hep-th/0212347.

189 [57] J. Matulich, S. Prohazka, and J. Salzer, “Limits of three-dimensional gravity and metric kinematical Lie algebras in any dimension,” JHEP 07 (2019) 118, arXiv:1903.09165 [hep-th]. [58] G. Papageorgiou and B. J. Schroers, “A Chern-Simons approach to Galilean quantum gravity in 2+1 dimensions,” JHEP 11 (2009) 009, arXiv:0907.2880 [hep-th]. [59] J. Gomis, A. Kleinschmidt, J. Palmkvist, and P. Salgado-Rebolledo, “Symmetries of post-Galilean expansions,” Phys. Rev. Lett. 124 (2020), no. 8, 081602, arXiv:1910.13560 [hep-th]. [60] E. Bergshoeff, J. M. Izquierdo, T. Ortín, and L. Romano, “Lie Algebra Expansions and Actions for Non-Relativistic Gravity,” JHEP 08 (2019) 048, arXiv:1904.08304 [hep-th]. [61] L. Romano, “Non-Relativistic Four Dimensional p-Brane Supersymmetric Theories and Lie Algebra Expansion,” arXiv:1906.08220 [hep-th]. [62] P. Concha, M. Ipinza, and E. Rodríguez, “Generalized Maxwellian exotic Bargmann gravity theory in three spacetime dimensions,” arXiv:2004.01203 [hep-th]. [63] P. Concha and E. Rodríguez, “Non-Relativistic Gravity Theory based on an Enlargement of the Extended Bargmann Algebra,” JHEP 07 (2019) 085, arXiv:1906.00086 [hep-th]. [64] L. Avilés, E. Frodden, J. Gomis, D. Hidalgo, and J. Zanelli, “Non-Relativistic Maxwell Chern-Simons Gravity,” JHEP 05 (2018) 047, arXiv:1802.08453 [hep-th]. [65] P. Concha, L. Ravera, E. Rodríguez, and G. Rubio, “Three-dimensional Maxwellian Extended Newtonian gravity and flat limit,” arXiv:2006.13128 [hep-th]. [66] A. Bagchi, R. Basu, A. Mehra, and P. Nandi, “Field Theories on Null Manifolds,” JHEP 02 (2020) 141, arXiv:1912.09388 [hep-th]. [67] L. Ciambelli, R. G. Leigh, C. Marteau, and P. M. Petropoulos, “Carroll Structures, Null Geometry and Conformal Isometries,” Phys. Rev. D 100 (2019), no. 4, 046010, arXiv:1905.02221 [hep-th]. [68] L. Donnay and C. Marteau, “Carrollian Physics at the Black Hole Horizon,” Class. Quant. Grav. 36 (2019), no. 16, 165002, arXiv:1903.09654 [hep-th]. [69] C. Duval, G. W. Gibbons, and P. A. Horvathy, “Conformal Carroll groups and BMS symmetry,” Class. Quant. Grav. 31 (2014) 092001, arXiv:1402.5894 [gr-qc]. [70] L. Ciambelli, C. Marteau, P. M. Petropoulos, and R. Ruzziconi, “Fefferman-Graham and Bondi Gauges in the Fluid/Gravity Correspondence,” PoS CORFU2019 (2020) 154, arXiv:2006.10083 [hep-th]. [71] L. Ciambelli, C. Marteau, P. M. Petropoulos, and R. Ruzziconi, “Gauges in Three-Dimensional Gravity and Holographic Fluids,” JHEP 11 (2020) 092, arXiv:2006.10082 [hep-th]. [72] Y. Herfray, “Asymptotic shear and the intrinsic conformal geometry of null-infinity,” J. Math. Phys. 61 (2020), no. 7, 072502, arXiv:2001.01281 [gr-qc]. [73] J. de Boer, J. Hartong, E. Have, N. A. Obers, and W. Sybesma, “Non-Boost Invariant Fluid Dynamics,” SciPost Phys. 9 (2020), no. 2, 018, arXiv:2004.10759 [hep-th]. [74] J. Armas and A. Jain, “Effective field theory for hydrodynamics without boosts,” arXiv:2010.15782 [hep-th]. [75] L. O’RAIFEARTAIGH, The Dawning of Gauge Theory. Princeton University Press, 1997.

190 [76] J. Wess and J. Bagger, Supersymmetry and Supergravity: Revised Edition. Princeton university press, 2020.

[77] S. Coleman and J. Mandula, “All possible symmetries of the s matrix,” Phys. Rev. 159 (Jul, 1967) 1251–1256.

[78] M. Bertolini, “Lectures on supersymmetry,” Lecture notes given at SISSA (2015).

[79] S. P. Martin, “A Supersymmetry primer,” Adv. Ser. Direct. High Energy Phys. 21 (2010) 1–153, arXiv:hep-ph/9709356.

[80] S. J. Gates, M. T. Grisaru, M. Rocek, and W. Siegel, Superspace Or One Thousand and One Lessons in Supersymmetry, vol. 58 of Frontiers in Physics. 1983. arXiv:hep-th/0108200.

[81] J. Rembieliński and W. Tybor, “Possible superkinematics,” Acta Phys. Polon. B 15 (1984), no. 7, 611–615.

[82] J. Rembielinski and W. Tybor, “Possible Superkinematics,” Acta Phys. Polon. B 15 (1984) 611–615.

[83] P. Kosinski, J. Rembielinski, and W. Tybor, “Extended Superkinematics,” Acta Phys. Polon. B 16 (1985) 827–830.

[84] V. Hussin, J. Negro, and M. A. del Olmo, “Kinematical superalgebras,” Journal of Physics A: Mathematical and General 32 (jan, 1999) 5097–5121.

[85] R. Campoamor-Stursberg and M. Rausch de Traubenberg, “Kinematical superalgebras and Lie algebras of order 3,” J. Math. Phys. 49 (2008) 063506, arXiv:0801.2630 [hep-th].

[86] C.-G. Huang and L. Li, “Possible Supersymmetric Kinematics,” Chin. Phys. C39 (2015), no. 9, 093103, arXiv:1409.5498 [hep-th].

[87] S. Chapman, Y. Oz, and A. Raviv-Moshe, “On Supersymmetric Lifshitz Field Theories,” JHEP 10 (2015) 162, arXiv:1508.03338 [hep-th].

[88] I. Arav, Y. Oz, and A. Raviv-Moshe, “Holomorphic Structure and Quantum Critical Points in Supersymmetric Lifshitz Field Theories,” JHEP 11 (2019) 064, arXiv:1908.03220 [hep-th].

[89] F. W. Warner, Foundations of differentiable manifolds and Lie groups, vol. 94. Springer Science & Business Media, 2013.

[90] S. Helgason, Differential geometry and symmetric spaces, vol. 341. American Mathematical Soc., 2001.

[91] A. Kirillov Jr, An introduction to Lie groups and Lie algebras, vol. 113. Cambridge University Press, 2008.

[92] R. W. Sharpe, Differential geometry: Cartan’s generalization of Klein’s , vol. 166. Springer Science & Business Media, 2000.

[93] Y. Voglaire, “Strongly exponential symmetric spaces,” Int. Math. Res. Not. IMRN (2014), no. 21, 5974–5993, arXiv:1303.5925 [math.DG].

[94] P. K. Rozanov, “On the exponentiality of affine symmetric spaces,” Funktsional. Anal. i Prilozhen. 43 (2009), no. 1, 68–80.

[95] K. Yagi et al., “Super lie groups,” in Progress in Differential Geometry, pp. 407–412, Mathematical Society of Japan. 1993.

191 [96] B. Kostant, “Graded manifolds, graded , and prequantization,” in Differential geometrical methods in mathematical physics (Proc. Sympos., Univ. Bonn, Bonn, 1975), pp. 177–306. Lecture Notes in Math., Vol. 570. Springer, , 1977. [97] J.-L. Koszul, “Graded manifolds and graded Lie algebras,” in Proceedings of the international meeting on geometry and physics (Florence, 1982), pp. 71–84. Pitagora, Bologna, 1983. [98] A. Santi, “Superization of homogeneous spin manifolds and geometry of homogeneous supermanifolds,” Abh. Math. Semin. Univ. Hambg. 80 (2010), no. 1, 87–144, arXiv:0905.3832 [math.DG]. [99] M. Batchelor, “The structure of supermanifolds,” Trans. Amer. Math. Soc. 253 (1979) 329–338. [100] W. Rossmann, Lie groups, vol. 5 of Oxford Graduate Texts in Mathematics. Oxford University Press, Oxford, 2002. An introduction through linear groups. [101] K. Nomizu, “Invariant affine connections on homogeneous spaces,” American Journal of Mathematics 76 (1954), no. 1, 33–65. [102] A. Farahmand Parsa, H. R. Safari, and M. M. Sheikh-Jabbari, “On Rigidity of 3d Asymptotic Symmetry Algebras,” arXiv:1809.08209 [hep-th].

[103] H. Safari and M. Sheikh-Jabbari, “BMS4 algebra, its stability and deformations,” JHEP 04 (2019) 068, arXiv:1902.03260 [hep-th]. [104] H. Safari, “Deformation of Asymptotic Symmetry Algebras and Their Physical Realizations,” other thesis, 11, 2020. [105] K. Nomizu, “Invariant affine connections on homogeneous spaces,” Amer. J. Math. 76 (1954) 33–65. [106] C. Duval, G. Gibbons, P. Horvathy, and P. Zhang, “Carroll versus Newton and Galilei: two dual non-Einsteinian concepts of time,” Class. Quant. Grav. 31 (2014) 085016, arXiv:1402.0657 [gr-qc]. [107] G. Hochschild and J.-P. Serre, “Cohomology of Lie algebras,” Ann. of Math. (2) 57 (1953) 591–603. [108] P. de Medeiros, J. Figueroa-O’Farrill, and A. Santi, “Killing superalgebras for Lorentzian four-manifolds,” J. High Energy Physics 2016 (2016), no. 6, 1–50, arXiv:1605.00881 [hep-th]. [109] R. Andringa, E. A. Bergshoeff, J. Rosseel, and E. Sezgin, “3D Newton–Cartan supergravity,” Class. Quant. Grav. 30 (2013) 205005, arXiv:1305.6737 [hep-th]. [110] M. H. Christensen, J. Hartong, N. A. Obers, and B. Rollier, “Boundary Stress-Energy Tensor and Newton-Cartan Geometry in Lifshitz Holography,” JHEP 01 (2014) 057, arXiv:1311.6471 [hep-th]. [111] J. Hartong, E. Kiritsis, and N. A. Obers, “Field Theory on Newton-Cartan Backgrounds and Symmetries of the Lifshitz Vacuum,” JHEP 08 (2015) 006, arXiv:1502.00228 [hep-th]. [112] D. K. Wise, “MacDowell-Mansouri gravity and Cartan geometry,” Class. Quant. Grav. 27 (2010) 155010, arXiv:gr-qc/0611154. [113] R. D’Auria, “Geometric supergravitty,” arXiv:2005.13593 [hep-th]. [114] L. Castellani, “Supergravity in the Group-Geometric Framework: A Primer,” Fortsch. Phys. 66 (2018), no. 4, 1800014, arXiv:1802.03407 [hep-th].

192