• Nenhum resultado encontrado

The spinorial formalism, with applications in physics

N/A
N/A
Protected

Academic year: 2021

Share "The spinorial formalism, with applications in physics"

Copied!
113
0
0

Texto

(1)

DEPARTAMENTO DE FÍSICA

PROGRAMA DE PÓS-GRADUAÇÃO EM FÍSICA

JOÁS DA SILVA VENÂNCIO

THE SPINORIAL FORMALISM, WITH

APPLICATIONS IN PHYSICS

Recife 2017

(2)

THE SPINORIAL FORMALISM, WITH APPLICATIONS IN PHYSICS

Dissertation presented to the graduation program of the Physics Department of Uni-versidade Federal de Pernambuco as part of the duties to obtain the degree of Master of Philosophy in Physics.

Supervisor: Prof. Dr. Carlos Alberto Batista da Silva Filho

Recife 2017

(3)

V448s Venâncio, Joás da Silva.

The spinorial formalism, with appllications in physics / . – 2017. 112 f.

Orientador: Carlos Alberto Batista da Silva Filho.

Dissertação (Mestrado) – Universidade Federal de Pernambuco. CCEN. Física, Recife, 2017.

Inclui referências e índice.

1. Física matemática. 2. Formalismo espinorial. 3. Relatividade geral. 4. Geometria diferencial. I. Silva Filho, Carlos Alberto Batista da

(Orientador). II. Titulo.

(4)

THE SPINORIAL FORMALISM, WITH APPLICATIONS IN PHYSICS

Dissertation presented to the graduation pro-gram of the Physics Department of Universi-dade Federal de Pernambuco as part of the duties to obtain the degree of Master of Phi-losophy in Physics.

Accepted on: 23/02/2017.

EXAMINING BOARD:

Prof. Dr. Carlos Alberto Batista da Silva Filho Supervisor

Universidade Federal de Pernambuco

Prof. Dr. Bruno Geraldo Carneiro da Cunha Internal Examiner

Universidade Federal de Pernambuco

Prof. Dr. Marco Cariglia External Examiner

(5)

It is well-known that the rotation symmetries play a central role in the development of all physics. In this dissertation, the material is presented in a way which sets the scene for the introduction of spinors which are objects that provide the least-dimensional faithful representation for the group Spin(n), the group that is the universal coverage of the group SO(n), the group of rotations in n dimensions. With that goal in mind, much of this dissertation is devoted to studying the Clifford algebra, a special kind of algebra defined on vector spaces endowed with inner products. At the heart of the Clifford algebra lies the idea of a spinor. With these tools at our disposal, we studied the basic elements of dif-ferential geometry which enabled us to emphasise the more geometrical origin of spinors. In particular, we construct the spinor bundle which immediately lead to the notion of a spinor field which represents spin 1/2 particles, such as protons, electrons, and neutrons. A higher-dimensional generalization of the so-called monogenic multivector functions is also investigated. In particular, we solved the monogenic equations for spinor fields on conformally flat spaces in arbitrary dimension. Particularly, the massless Dirac field is a type of monogenic. Finally, the spinorial formalism is used to show that the Dirac equation minimally coupled to an electromagnetic field is separable in spaces that are the direct product of bidimensional spaces. In particular, we applied these results on the background of black holes whose horizons have topology R × S2× . . . × S2.

(6)

É bem conhecido que as simetrias de rotação desempenham um papel central no de-senvolvimento de toda a física. Nesta dissertação, apresentamos o conteúdo de forma a estabelecer o cenário para a introdução dos chamados spinors, os quais são objetos que fornecem as representações fiéis de menor dimensão para o grupo Spin(n), o grupo que é a cobertura universal do grupo SO(n), o grupo das rotações em n dimensões. Para este fim, grande parte desta dissertação é dedicada ao estudo da álgebra de Clifford, um tipo especial de álgebra definida em espaços munidos de um produto interno. No coração da álgebra de Clifford está precisa definição de um espinor. Com estas ferramentas à nossa disposição, estudamos os elementos básicos de geometria diferêncial, o que nos permi-tiu entender sobre a origem mais geométrica de espinores. Em particular, construimos o fibrado espinorial, o qual conduziu imediatamente a noção de um campo espinorial que, por sua vez, representa com precição as partículas com spin 1/2 tais como: pró-tons, elétrons e neutrons. Uma generalização para dimensões mais altas do conceito de multivetores monogênicos também é investigada. Em particular, resolvemos a equação dos monogênicos para campos espinoriais em espaços conformemente planos em dimen-são arbitrária. Particularmente, o campo de Dirac sem massa é um tipo de monogênico. Finalmente, o formalismo espinorial foi usado para mostrar que a equação de Dirac com massa minimamente acoplada ao campo eletromagnético é separável em espaços que são produtos diretos de espaços bidimensionais. Em particular, aplicamos estes resultados a buracos negros com horizontes topológicos R × S2× . . . × S2.

Palavras chaves: Álgebra de Clifford. Espinores. Monogenicos. Equação de Dirac. Sepa-rabilidade.

(7)

V Vector space over a field F: Page 11.

V∗ Dual vector space of V: Page 11.

F Field: Page 11.

R Field of the real numbers: Page 11. C Field of the complex numbers: Page 11. ∂µ Differential operators

∂xµ : Page 49.

T[a1a2...ap] Skew-symmetric part of the tensor Ta1a2...ap : Page 12.

∧pV Space of p-vectors: Page 12. ∧V Space of multivectors: Page 12.

h ip, hAip Set of projection operators: Page 12.

∧, A ∧ B Exterior product: Page 13.

< , >, < A, B > Non-degenerate inner product: Page 14.

C`(V) Clifford algebra of V: Page 14.

y, AyB Left contraction: Page 15.

b,Ab Degree involution: Page 16. e,Ae Reversion: Page 17.

,A Clifford conjugation: Page 18.

M(2, F) Algebra of 2 × 2 matrices over F : Page 21.

?, ?A Hodge dual:Page 23.

ˆ

S Reflection operator: Page 26.

ˆ

R Rotation operator: Page 27.

D Division algebra: Page 38.

( , ), (ψ, φ) Inner product between spinors: Page 38.

g The metric of the manifold:Page 50.

ωab, ωabc, ωabc Connection 1-form and its components:Page 57. Ta Torsion 2-form:Page 57.

Ra

b Curvature 2-form: Page 57.

£K Lie derivative along of K:Page 59.

Γ(SM ) Space of the local sections of spinorial bundle:Pages 74.

(8)

1 Introduction 9

2 Clifford Algebra and Spinors 11

2.1 The Exterior Algebra . . . 11

2.2 The Clifford Algebra . . . 13

2.2.1 Involutions . . . 16

2.2.2 Pseudoscalar, Duality Transformation and Hodge Dual . . . 21

2.2.3 Periodicity . . . 24

2.3 The Spin Groups . . . 26

2.4 Spinors . . . 29

2.4.1 Minimal Left Ideal . . . 30

2.4.2 Pure Spinors . . . 34

2.4.3 SPin-Invariant Inner Products . . . 37

2.5 Spinors, Gamma Matrices and their Symmetries . . . 41

3 Curved Spaces and Spinors 47 3.1 Manifolds and Tensors Fields . . . 47

3.2 The Curvature Tensor . . . 54

3.3 Non-Coordinate Frame and Cartan’s Structure Equations . . . 56

3.4 Symmetries, Killing Vectors and Hidden Symmetries . . . 58

3.4.1 Maximally Symmetric Manifolds . . . 61

3.4.2 Symmetries of Euclidean Space Rn . . . 67

3.5 Fiber Bundles . . . 70

3.5.1 Spinorial Bundles and Connections . . . 73

4 Monogenics and Spinors in Curved Spaces 78 4.1 Monogenics in the Literature . . . 78

4.2 Monogenic Multivector Functions : Definitions and Operator Equalities . . 80

4.3 Monogenics in Curved Spaces and Some Results . . . 82

4.4 Monogenic Equation for Spinor Fields . . . 85

4.5 Direct Product Spaces and the Separability of the Dirac Equation . . . 90

4.6 Black Hole Spacetimes . . . 96

4.6.1 The angular part of Dirac’s Equation . . . 100

(9)
(10)

1 Introduction

The so-called rotational symmetries play a fundamental role in the development of physics. For instance, in classical mechanics and classical field theory the invariance by rotations gives rise to conserved quantities that allow the analytical integration of the equations of motion, whereas in quantum mechanics the irreducible repre-sentations of the rotation group are used to label quantum states. Moreover, and foremost, the Lorentz group can be seen as the group of rotations in a space with a metric of Lorentzian signature. It is worth recalling that the invariance under Lorentz transformation is the foundation of the standard particle model, the most successful physical theory of the second half of the twentieth century. In this disser-tation, the material is presented in a way which sets the scene for the introduction of spinors which are objects that provide the least-dimensional faithful representa-tions for the group Spin(n), the universal coverage of the group SO(n), the group of rotations in n dimensions. To this end, we introduce a special kind of algebra defined by vector spaces endowed with inner products: the Clifford algebra, also known as geometric algebra. This Algebra was created by the English mathemati-cian William Kingdon Clifford around 1880 building on the earlier work of Hamilton on quaternions and Grassmann about exterior algebra. Although the definition of a spinor lies at the heart of the Clifford algebra, the spinors were discovered in 1913 by Élie Cartan as objects related to linear representations of simple groups; they provides a linear representation of the SO(n) group. Hence, spinors are of fundamental importance in several branches of physics and mathematics and this dissertation sheds light on the role played by spinors on physics and mathematics. This dissertation splits in three chapters. In chapter 1 we introduce the Clifford algebras intimately related to the orthogonal transformations, the rotations. These algebras play a central role in the construction of the groups P in(n) and Spin(n), which are the universal covering groups of the orthogonal groups O(n) and SO(n) respectively. Finally, at the end of chapter, we define the so-called spinors. In chap-ter 2 we study the basic elements of differential geometry, such as the curvature tensor, the Killing vectors and the fiber bundles which enabled us to emphasize on more geometrical origin of these objects. In particular, we construct the spinor bun-dle which immediately lead to the notion of a spinor field which represents spin 1/2 particles, such as protons, electrons, and neutrons. Finally, in chapter 3 the

(11)

mono-genic functions which are multivector functions that are annihilated from the left by the Dirac operator are reviewed and a higher-dimensional generalization these multivetor functions is also investigated. In particular, we solved the monogenic equations for spinor fields on conformally flat space in arbitrary dimension. More-over, in chapter 3 it is present the main results of this dissertation. It is shown that the Dirac Equation coupled to a gauge field can be decoupled in even-dimensional manifolds that are the direct product of bidimensional spaces.

(12)

2 Clifford Algebra and Spinors

A vector space V over a field F is a set of vectors with an operation of addition and a rule of scalar multiplication, which assigns a vector to the product of a vector with an element of the field. Elements in F will be called scalars. An algebra is a vector space in which an associative multiplication between the vectors is defined. In this chapter we introduce the Clifford algebra, a special kind of algebra defined on vector spaces endowed with inner products. These algebras are intimately related to the orthogonal transformations, namely the rotations. This special algebra plays a central role in the construction of the groups P in(V) and Spin(V), which are the universal covering groups of the orthogonal groups O(V) and SO(V) respectively, and also define the so-called spinors, which are elements of the a vector space on which a fundamental representation of these groups act.

2.1

The Exterior Algebra

Let V be an n-dimensional vector space over a field F (R or C) and {ea} an arbitrary basis for V, where a ∈ {1, 2, ..., n}. We can expand any vector V ∈ V in this basis as

V = Vaea, (2.1)

where we have started to employ the Einstein summation convention on which repeated indices are summed. Associated to V is the dual space of V, denoted by V∗ whose elements are linear functionals, also called co-vectors,

ω : V → F

V 7→ ω(V ) , (2.2)

which obey the rule of linearity ω(λ V + δ U ) = λ ω(V ) + δ ω(U ) ∀ λ , δ ∈ F. If the addition of co-vectors and their multiplication by an element of the field is trivially defined it is straightforward to prove that the dual of V is also a vector space.

We can define the co-vectors eb ∈ V(b = 1, 2, ..., n) by their action on elements of V as:

ea(eb) = δab =

1 , if a = b

(13)

hence the co-vectors {ea} provide a basis for V∗. So, any co-vector ω can be expanded in this basis and written as:

ω = ωaea, (2.4)

where ωa = ω(ea). It follows that if ω ∈ V∗ is co-vector and V ∈ V is a vector then ω(V ) = ωaVa ∈ F. We also may regard elements of V as linear functions on its dual V∗ by defining V (ω) ≡ ω(V ), since V∗∗∼ V.

The tensor product refers to way of constructing a big vector space out of two or more smaller vector spaces. For example, the tensor product of V ∈ V with ω ∈ V∗, denoted as V ⊗ ω, gives rise a new tensor t belonging to a vector space V ⊗ V∗ on which the n2 elements of the form e

a⊗ eb provide a basis, so that the most general element of this space can be written as t = tabea⊗ eb, where tab = t(ea, eb) ∈ F. Likewise, an arbitrary tensor T has the following abstract representation:

T = Ta1...ap b1... bqea1 ⊗ . . . ⊗ eap ⊗ e b1 ⊗ . . . ⊗ ebq, (2.5) where Ta1...ap b1... bq ≡ T (e a1, . . . , eap, e

b1, . . . , ebq). The space of such tensors is

de-noted Tp q(V).

The so-called p-vectors are totally skew-symmetric tensors of degree p and the vector space generated by all them, namely Aa1...ap = A[a1...ap], is denoted by ∧

pV, where we identify the field F with ∧0V and V with ∧1V. We must assume that all the indices inside square brackets take different values to ensure that the p-vector A ∈ ∧pV given by

A = A[a1...ap]e

a1 ⊗ . . . ⊗ eap, (2.6)

be non-null. Because the antisymmetry, it is a simple matter to prove that the dimension of ∧pV is: dim(∧pV) = n p  =  n! p!(n−p)! if 0 ≤ p ≤ n 0 if p > n , (2.7)

from which we can note that p-vectors of degree greater than n are zero. By taking the direct sum of the spaces ∧pV, we obtain the 2n-dimensional multivector space, that is ∧V = n M p=0 ∧pV . (2.8)

Elements in ∧V are called multivectors and we denote them by A. Because of de-composition of ∧V in p-vector subspaces, we can define a set of projection operators, denoted by h ip, whose action is:

h ip : ∧V → ∧pV

A 7→ hAip = Ap.

(14)

Notice that h ip is in fact a projector, since that it maps an arbitrary multivector to its p-vector component each of the form (2.6). Thus, an arbitrary multivector A can be decomposed into a sum of pure degree terms

A = hAi0+ hAi1+ . . . + hAin = A0+ A1 + . . . + An=

n X p=0

Ap. (2.10)

Multivectors containing terms of only one degree are called homogeneous.

An interesting feature of the space of multivectors is that it carries naturally a product, denoted by ∧, which maps two tensors of degrees p and q to a totally antisymmetric tensor of degree (p + q), this is the so-called exterior product. Therefore, if A is a p-vector and B is a q-vector, the exterior product ∧ is a map such that:

∧ : ∧pV × ∧qV → ∧p+qV

A , B 7→ A ∧ B, (2.11)

where A ∧ B = (−1)pqB ∧ A is defined as: A ∧ B = (p + q)!

p! q! A

[a1...apBb1...bq]e

a1 ⊗ . . . ⊗ eap⊗ eb1 ⊗ . . . ⊗ ebq. (2.12)

In n dimensions the set {1, ea1, ea1∧ ea2, ea1∧ ea2∧ ea3, . . . , ea1 ∧ . . . ∧ ean}, which

contains 2n elements furnish a basis for the space of multivectors ∧V. Therefore, a multivector A ∈ ∧V may be written in this basis as

A = A |{z} 0−vector + Aaea | {z } 1−vector + Aabea∧ eb | {z } 2−vector + Aabcea∧ eb∧ ec | {z } 3−vector . . . + Aa1...ane a1 ∧ . . . ∧ ean | {z } n−vector . (2.13) For instance, in V = R3 there exist eight elements that generate this basis. These are: one 0-vector which is denoted by 1, three 1-vectors ea, three 2-vectors ea∧ eb each of the form ea∧ eb = ea⊗ eb− eb⊗ ea for each choice of a 6= b = 1, 2, 3 and a single 3-vector. In particular, note that this 3-vector is written as

ea∧ eb ∧ ec = ea⊗ eb⊗ ec+ eb⊗ ec⊗ ea+ ec⊗ ea⊗ eb+ − eb ⊗ ea⊗ ec− ec⊗ eb⊗ ea− ea⊗ ec⊗ eb.

Definition 1. The exterior algebra is an associative algebra formed from ∧V and the exterior product on p-vectors.

2.2

The Clifford Algebra

The definitions of the exterior product, and of the multivectors do not depend on any inner product. However, Clifford algebras are a generalization of exterior

(15)

algebras, defined in the presence of an inner product. In what follows, we will assume an inner product in our vector space. Let V be a vector space endowed with a non-degenerate inner product < , >, this means that for V ∈ V, < V , U > = 0 for all U ∈ V if and only if V = 0. The pair (V, < , >) is called an orthogonal space. Then the Clifford product, assumed to be associative and denoted by juxtaposition, of a pair of vectors is defined to be such that its symmetric part gives the inner product:

V U + U V = 2 < V , U > ∀ V , U ∈ V . (2.14) Since the base used is completely arbitrary, it is convenient to adopt {ea} = {ˆe1, ˆe2, . . . , ˆen} as an orthonormal basis, since in this case we have < ˆea, ˆeb > = ±δab and therefore ˆeaeˆb = −ˆebeˆa, if a 6= b. Analogously, ˆeaˆebˆec is totally skew-symmetric if a 6= b 6= c 6= a. Thus, any element of C`(V, < , >), the Clifford algebra of V, can be written as a linear combination of p-vectors ˆea1. . . ˆeap with 1 ≤ p ≤ n.

In other words, the vector space of the Clifford algebra associated to V is ∧V, then it can be written as a direct sum:

C`(V) = n M

p=0

∧pV , (2.15)

hence the dimension dim(C`(V, < , >)) = 2n. This decomposition introduces a multivector structure into the Clifford algebra C`(V). The multivector structure is unique, that is, an arbitrary element A ∈ C`(V) can be uniquely decomposed into a sum of p-vectors A = A0+ A1+ . . . + An. In what follows, for simplicity, instead of the previous notation it will be denoted by C`(V) the Clifford algebra associated to V. It follows that the set {ˆ1, ˆea1, ˆea1eˆa2, . . . , ˆea1. . . ˆean} with the identity element

ˆ

1 ∈ F, forms a basis for C`(V). Therefore, a general element A ∈ C`(V) can be written as: A = A |{z} scalar + Aaeˆa | {z } vector + Aabˆeaeˆb | {z } 2−vector + Aabceˆaˆebˆec | {z } 3−vector + . . . + Aa1...anˆe a1. . . ˆean | {z } n−vector , (2.16)

Now, in the same way that (2.12) was defined in terms of the tensorial product, it is natural to define the exterior product in terms of the Clifford product. So, given a permutation σ : {1, 2, . . . , n} → {σ(1), σ(2), . . . , σ(2)} we define the exterior product as the totally antisymmetric part of the Clifford product:

V1∧ V2∧ . . . ∧ Vp := 1 p! X σ∈Sp (σ)Vσ(1)Vσ(2). . . Vσ(p), (2.17)

where Sp is the symmetric group whose elements are all the permutations and (σ) is +1 on even permutations, −1 otherwise. So, for example, given two vectors V and U we find that:

V ∧ U = 1

(16)

The equations (2.14) and (2.18) combine to give the Clifford product of two vectors: V U = < V , U > + V ∧ U , (2.19) also called the geometric product.

The decomposition of the Clifford product of two vectors into a scalar term and a 2-vector term has a natural extension to general multivectors. This may be done through the use of the projection operator in terms of which we can conveniently express the inner and exterior products. So, given Ap ∈ ∧pV and Bq ∈ ∧qV the exterior product is written as:

Ap∧ Bq := hApBqip+q, and the inner product as:

< Ap, Bq > := hApBqi|p−q|. (2.20) Another important operation is the left contraction, denoted byy, defined as follows:

ApyBq :=

< Ap, Bq > if p ≤ q

0 if p > q . (2.21)

The products defined above can be extended by bilinearity for the whole algebra. Using the equations (2.18) and (2.17) the Clifford product of a vector and an arbi-trary multivetor is given by:

V A = V yA + V ∧ A ∀ V ∈ V, A ∈ C`(V) . (2.22) Moreover, one can prove that the contration by a vector satisfies Leibniz’s rule, this means that such a contraction is a derivation of the Clifford algebra C`(V). Indeed, a result that is extremely useful in practice is the following:

V y(V1∧ V2∧ . . . ∧ Vp) = (V yV1)V2∧ V3∧ V4∧ . . . ∧ Vp+ (2.23) − (V yV2)V1∧ V3∧ V4∧ . . . ∧ Vp+ + (V yV3)V1∧ V2∧ V4∧ . . . ∧ Vp+ . . . = p X i=1 (−1)i+1 < V , Vi > V1∧ . . . ∧ ˇVi∧ . . . ∧ Vp, (2.24) where the check on ˇVi denotes that the term should be withdrawn from the series. For instance, when p = 2 we find that:

(17)

Using the above result, by means of the equation (2.20), that in its turn ensures V yλ = 0 for any scalar λ, the equation (2.22) lead us at the following formula for the Clifford product of three vectors:

V U X = < V , U > X + < U , X > V − < V , X > U + V ∧ U ∧ X . (2.25) Note that in (2.22), as expected, multiplication by a vector raises and lowers the degree of a multivector by 1. But, in general, if Ap is a p-vector and Bqis a q-vector

ApBq 6= ApyBp + Ap∧ Bq,

where ApyBp ∈ ∧p−q(V) and Ap∧ Bp ∈ ∧p+q(V). Using (2.22) it is not so hard to prove that the products of two homogeneous multivector decompose as:

ApBq = hApBqi|p−q| + hApBqi|p−q|+2+ . . . + hApBqip+q. (2.26)

Definition 2. The Clifford algebra associated to V, denoted by C`(V), consists of the vector space ∧V together with the Clifford product.

2.2.1

Involutions

It is well-known that the conjugate of the conjugate of a complex number is the complex number itself. An operation, which, when applied to itself, returns the original object is called involution. In particular, the complex conjugation is a simple example of an involution. The Clifford algebra has three involutions similar to complex conjugation. The direct sum decomposition (2.15) gives C`(V) the structure of a Z-graded algebra. This induces in C`(V) the first involution, denoted by b , called the degree involution which is a linear map whose the action is defined on homogeneous multivectors by:

b : ∧V → ∧V

Ap 7→ Acp = (−1)pAp.

(2.27)

The degree involution is an automorphism1 such that: d

AB = bA bB ∀ A, B ∈ C`(V) . (2.28)

1An automorphism is an an invertible mapping from a set in itself. In particular, an automorphism s of V is said to be orthogonal with respect to < , > if s leaves < , > invariant, i.e, < sV , sV > = < V , V > ∀V ∈ V, see [14].

(18)

Since the map b applied to itself is identity map, c c

Ap = Ap , the eigenvalues of b are ±1, then the degree involution induces a Z2-grading on C`(V). Under degree involution the multivectors corresponding to eigenvalue +1 will be called even and the space of even multivectors will be denoted C`+(V), while the multivectors corresponding to eigenvalue −1 will be called odd, and belong to the subspace C`−(V). Thus, given A ∈ C`(V) we have

b

A = A0− A1+ A2− A3 + . . . . (2.29) In particular, we can write the inner and exterior products in terms of the degree involution. Indeed, we can prove that for an arbitrary multivector A ∈ C`(V) the following relations are satisfied:

V yA = 1

2(V A − bAV ) (2.30)

V ∧ A = 1

2(V A + bAV ) , (2.31)

where V ∈ V. Moreover, one can prove that the contration by a vector V ∈ V obeys the following version of of the Leibniz’s rule

V y(AB) = V yAB + bAV yB ∀ A, B ∈ ∧V . (2.32) The second involution in Clifford algebra is the called reversion, denoted by e , which reverses the order of vectors in any product. For instance, for a p-vector the reverse can be formed by a series of swaps of anticommuting vectors, each resulting in a minus sign. The first vector has to swap past p − 1 vectors, the second past p − 2, and so on. Using this it is easy matter to see that

^

V1∧ V2 ∧ . . . ∧ Vp = Vp∧ Vp−1∧ . . . ∧ V2∧ V1 = (−1)p(p−1)/2V1∧ V2∧ . . . ∧ Vp (2.33) It follows that the reversion, on a homogeneous vector, is a linear mapping such that:

e : ∧V → ∧V

Ap 7→ Afp = (−1)p(p−1)/2Ap

(2.34) From the above the degree involution is an anti-automorphism, that is,

g

AB = eB eA ∀ A, B ∈ C`(V), (2.35)

thus, we see that the reverse of A ∈ C`(V) is: e

(19)

The composition of the two previous involutions is called the Clifford conjuga-tion, denoted by A = g( bA) = d( eA), which is expressed by

: ∧V → ∧V

Ap 7→ Ap = (−1)p(p+1)/2Ap.

(2.37) and it is also an anti-automorphism,

AB = B A ∀ A, B ∈ C`(V). (2.38)

Thus,

A = A0− A1− A2− A3+ . . . . (2.39) The Clifford conjugation can be used to determine the inverse. Indeed, if A ∈ C`(V), then its inverse is given by

A−1 = A

AA. (2.40)

In particular, using the Clifford conjugation of a vector U ∈ V which is given by the following relation

U−1 = U

U2 , (2.41)

we can show that the associativity property of the Clifford product ensures that it is now possible to divide by vectors. In fact, defining B = V U ∀ V , U ∈ V we have that

BU = (V U )U = V (U U ) = V U2,

Now, from (2.41) we can recover V and we eventually arrive at the following ex-pression:

V = BU−1.

This ability to divide by vectors gives the Clifford algebra considerable power. The reversion can be used to extend the concept of norm. Now, we are going to define the norm of an arbitrary multivector A ∈ C`(V) as follows:

|A|2 = DAAe E 0 = D A eAE 0. (2.42)

For example, if V = R3 the most general element A ∈ C`(R3) is given by

A = a |{z} scalar + a1ˆe1+ a2ˆe2+ a3ˆe3 | {z } vector + a12ˆe1ˆe2+ a13eˆ1eˆ3+ a23ˆe2ˆe3 | {z } 2−vector + + a123ˆe1ˆe2ˆe3 | {z } 3−vector , (2.43)

then, using (2.43) and (2.14) it is straightforward to prove that:

(20)

Since we can split a multivector into those components that, under degree involu-tion, are even and those that are odd, every A ∈ C`(V) can be written as:

A = A++ A−, (2.45)

where A+ = A0 + A2+ A4+ . . . and A− = A1+ A3 + A5+ . . . . It follows that we can write C`(V) = C`+(V) ⊕ C`(V) where

C`±(V) = {A ∈ C`(V) | bA = ±A } . (2.46) This Z2-grading of the Clifford algebra ensures that elements of even degree form a subalgebra, the called even subalgebra. Indeed, note that, AB ∈ C`+(V) if A ∈ C`±(V) and B ∈ C`±(V) while AB ∈ C`(V) if A ∈ C`±(V) and B ∈ C`(V) or A ∈ C`∓(V) and B ∈ C`±(V).

The center Z of an algebra A consists of all those elements z of A such that za = az for all a in A. We can easily prove that the center of C`(V), denoted by Z(C`(V)), depends on the dimension of space as follows:

Z(C`(V)) = 

∧0V , if n is even ∧0V ⊕ ∧nV, if n is odd

. (2.47)

Before proceeding let us make as example.

Example 1: The C`(R3) algebra.

Let us work out in the vector space V = R3, whose Clifford algebra C`(R3) is gener-ated by {1, ˆe1, ˆe2, ˆe3, ˆe1eˆ2, ˆe1eˆ3, ˆe2ˆe3, ˆe1ˆe2ˆe3}, which contains 23 = 8 elements, where ˆe1eˆ1 = ˆe2ˆe2 = ˆe3eˆ3 = 1 and ˆeaeˆb = −ˆebˆea if a 6= b. A priori just for the sake of simplicity in notation, let us denote the product ˆe1ˆe2ˆe3 ≡ I and in the next section we discuss this important symbol. The reversion this symbol is e

I = −I and their square is equal to

I2 = eˆ1eˆ2eˆ3eˆ1eˆ2eˆ3 = −ˆe1ˆe3ˆe2ˆe1ˆe2eˆ3 = ˆe3ˆe1eˆ2ˆe1ˆe2ˆe3 = −ˆe3eˆ2eˆ1eˆ1eˆ2eˆ3 = −ˆe3eˆ2eˆ2eˆ3 = −ˆe3ˆe3

=⇒ I2 = −1 . (2.48)

The symbol I commutes with all vectors in three dimensions, using this one can prove that IA = AI ∀ A ∈ C`(R3) given by (2.43). C`+(R3), the the even subalgebra of C`(R3), is generated by {1, ˆe1ˆe2, ˆe1eˆ3, ˆe2eˆ3} which contains 1223 = 4 elements in terms of which a multivector A+ ∈ C`+(R3) can be written as:

A+ = a |{z} scalar + a12ˆe1ˆe2 | {z } 2−vector + a13eˆ1eˆ3 | {z } 2−vector + a23eˆ2eˆ3 | {z } 2−vector .

(21)

Defining i = ˆe2ˆe3, j = ˆe1ˆe3, k = ij = ˆe1eˆ2, we find that i2 = j2 = k2 = −1 , i,e., is the quaternion algebra H. Thus, the even subalgebra of C`(R3) is isomorphic to the quaternion algebra, as can be seen by the following correspondences:

H C`+(R3) 1 1 i eˆ2eˆ3 j eˆ1eˆ3 k eˆ1eˆ2 (2.49)

The basis vectors ˆe1, ˆe2 and ˆe3 can also be represented by matrices σ1, σ2 and σ3 given by σ1 =  0 1 1 0  ; σ2 =  0 −i i 0  ; σ3 =  1 0 0 −1  ,

the so-called Pauli matrices, since these matrices in accordance with this represen-tation satisfy the same algebra, namely σ1σ1 = σ2σ2 = σ3σ3 = I , σaσb = −σbσa if a 6= b. The quaternion algebra admits the following matrix representation

1 ' I ; i = ˆe2eˆ3 ' iσ1; j = ˆe1eˆ3 ' iσ2; k = ij = ˆe1eˆ2 ' iσ3,

where I is the 2 × 2 identity matrix. Equivalently, the multivetor A ∈ C`(R3) can be represented as: [A] = a 1 0 0 1  + a1  0 1 1 0  + a2  0 −i i 0  + a3  1 0 0 −1  + + a12  i 0 0 −i  + a13  0 1 −1 0  + a23  0 i i 0  + a123  i 0 0 i  =⇒ [A] =  z1 z3 z2 z4  , (2.50)

where z1 = (a + a3) + i(a12+ a123) , z2 = (a1 + a13) + i(a2 + a23) , z3 = (a1 − a13) − i(a2− a23), and z4 = (a − a3) − i(a12− a123) . Note also that A+ ∈ C`+(R3) is represented by [A] = w1 −w2 ∗ w2 w1∗  , (2.51)

where w1 = a + ia12, w2 = a13+ ia23 and w∗i is the complex conjugate of wi (i = 1, 2). By equation (2.47), the center of C`(R3) is the subalgebra of scalars and 3-vectors, namely ∧0V ⊕ ∧3V = {a + a123I}. Note that σ1σ2σ3 = iI and the symbol I can therefore be viewed as the unit imaginary2 i and the combination of a scalar 2The symbol i is an element which commutes with all others, which is not necessarily a property of I. But, in this case it commutes with all elements and squares to −1. It is therefore a further candidate for a unit imaginary.

(22)

and a pseudoscalar as a complex number. This implies the center of C`(R3) is isomorphic to the complex field C, that is,

Z(C`(R3)) = ∧0V ⊕ ∧3V ' C . (2.52) The correspondences ˆe1 ' σ1, ˆe2 ' σ2 and ˆe3 ' σ3 establish the isomorphism C`(R3

) ' M(2, C)3 with the following correspondences of the basis elements: M(2, C) C`(R3) I 1 σ1, σ2, σ3 eˆ1, ˆe2, ˆe3 σ1σ2, σ1σ3, σ2σ3 ˆe1ˆe2, ˆe1ˆe3, ˆe2eˆ3 iI I (2.53)

According to with what was seen above, the Clifford algebra C`(R3) contains two subalgebras: the center Z(C`(R3)) which is isomorphic to the complex field C and C`+

(R3) isomorphic to the quaternion algebra H. Now, since the elements of C and H commute, zq = qz for z ∈ C , q ∈ H and that, as real algebra, C`(R3) is generated by C and H, noting that (dimC)(dimH) = dimC`(R3), we are left with the following conclude:

C`(R3

) ' C ⊗ H. (2.54)



2.2.2

Pseudoscalar, Duality Transformation and Hodge Dual

The object I mentioned in the previous example is an important element of C`(V). This is the highest degree element in a given algebra, the so-called pseudoscalar4. For a given vector space the highest degree element exists and is unique up to a multiplicative scalar. The exterior product of n vectors is therefore a multiple of the unique pseudoscalar for C`(V). Given an orthonormal basis {ea} (a = 1, 2, . . . , n), we shall define the pseudoscalar to be I = ˆe1eˆ2. . . ˆen. This convention is well-defined if we do not change the orientation of the basis. Let us consider that the space V = Rp,q has dimension n = p + q with p vectors whose square is equal to 1 and q vectors whose square is equal to −1 and s = |p − q| the signature of the inner product. Noting that the p + 1 vector has square −1, the norm of I depends

3M(2, C) denotes the algebra of 2 × 2 matrices over C

(23)

on the dimension of space and the signature in the following form: |I|2 = DeII E 0 = hˆeqeˆq−1. . . ˆep+1eˆp. . . ˆe2eˆ1eˆ1eˆ2. . . ˆepeˆp+1. . . ˆeq−1ˆeqi0 = ˆe21ˆe22. . . ˆe2pˆe2p+1. . . ˆe2q−1ˆe2q = (−1)q =⇒ |I|2 = (−1)12(n−s). (2.55)

In particular |I|2 = 1 if the signature is Euclidean, s = n, and |I|2 = −1 if Lorentzian signature, s = (n − 2) . When a multivector A is homogeneuous |A|2 = eAA. So, immediately we see that the sign of I2 is specified by:

I2 = (−1)12n(n−1)eII = (−1) 1

2[n(n−1)+(n−s)]. (2.56)

Another property of the pseudoscalar is that it defines an orientation. Apart from a sign, the choice of I is independent of the choice of orthonormal basis. Choosing a sign amounts to choosing an orientation for V which is swapped by exchange of any pair of vectors. Since the dimension of λn is 1, given any vectors V1, V2, . . . , Vn it follows that

V1∧ V2∧ . . . ∧ Vn = λnI, (2.57) where5 λn ∈ F. Given the independent vectors V1, V2, . . . , and Vn their exterior product will either have the same sign as I, or the opposite sign. Those with the same sign are said to have a positive orientation, otherwise have a negative orientation. In particular, the pseudoscalar I = ˆe1ˆe2ˆe3 of C`(R3) is always chosen to be positive orientation.

An important property is that the pseudoscalar commutes with all vectors in odd dimension while in even dimension it anticomutes with all vectors. Using this, it is an immediate consequence that in three dimensions I ˆe1 = ˆe2eˆ3. Then, I(ˆe1∧ˆe2) = I ˆe1ˆe2 = ˆe2eˆ3eˆ2 = −ˆe3 = − ˆe1× ˆe2 where × denote the vector cross product. It follows that for any two vectors V , U ∈ R3 the vector cross product is defined in Clifford algebra as:

V × U = −I(V ∧ U ) . (2.58)

Note that, in this case, the result of the product of I ∈ ∧3V with the 2-vector ∧2V is a vector ∧1V, but this results is not valid in any dimension. This product only exists in three dimensions. In general we have the product of the pseudoscalar I ∈ ∧nV with a homogeneous multivector Ap ∈ ∧pV is another homogeneous multivector IAp ∈ ∧n−pV. This operation is called a duality transformation. In this language, the equation (2.58) means that the 2-vector was mapped to a vector by a duality transformation. Note that this is valid just on n = 3 and by this it is commom refer to the 2-vector V ∧ U as a pseudovector in three dimensions.

5One can prove that |λ

n| is the analogous of the volume of a parallelepiped which is generated by the vectors V1, V2, . . . , Vn−1and Vn, see [16, 1, 4].

(24)

All properties relative to commutation and anticomutation of I are contained in the following equation:

IAp = (−1)p(n−1)ApI . (2.59)

The pseudoscalar can be used to define an important operation, denoted by ?, called Hodge dual. This is a linear map from the space of p-vectors to the space of (n − p)-vectors:

? : ∧pV → ∧n−pV Ap 7→ ?Ap = fApI .

(2.60) Note that the Hodge map depends of the choice of orientation. By linearity one can extend this definition to inhomogeneuos multivectors of the C`(V) [1]

?A = eAI , (2.61)

and since the dimension dim(∧pV) = np 

= n−pn 

= dim(∧n−pV) we have an isomorphism between ∧pV and ∧n−pV. In particular, in three dimensions, using (2.61) one finds by anticommutation that:

?1 = I = ˆe1eˆ2eˆ3 ; ?ˆe1 = ˆe2ˆe3 ; ?ˆe2 = ˆe3eˆ1 ; ?ˆe3 = ˆe1ˆe2 ;

?(ˆe1eˆ2) = ˆe3 ; ?(ˆe1eˆ3) = ˆe2 ; ?(ˆe2eˆ3) = ˆe1 ; ?I = 1 . (2.62) By equation (2.56) it follows the (2.61) can be written as:

?A = (−1)12[n(n−1)+(n−s)]AIf (2.63)

with this form the above equation makes clear that the Hodge dual map depends on the signature of the inner product and of the choice of orientation. In particular, the dual of the dual of a p-vector A ∈ ∧pV is proportional to identity

? ? A = (−1)[p(n−p)+12(n−s)]A (2.64)

Example 2: Maxwell’s equations .

The Clifford algebra treatment is useful in several branches of the physics, since it provides a more compact formulation, for example, the Eletromagnetism. The four Maxwell’s equations can be united into a single equation. Let us define the multivector operator D by

D = ∂

∂t + ∇, (2.65)

where ∇ = ˆe1∂x1 + ˆe2∂x2 + ˆe3∂x3 is the usual gradient. The electromagnetic field strength is represented by the multivetor F and the two scalar densities charge ρ and vectorial charge current J are combined into a single multivector J given by

(25)

where the vector fields E and B are the eletric field and the magnetic induction, respectively. Since the pseudoscalar commutes with all vectors in three dimensios, using the equations (2.19) and (2.58), the Clifford action of D on F is:

DF = ( ∂ ∂t + ∇)(E + IB) = ∂E ∂t + ∂(IB) ∂t + ∇E + ∇(IB) = ∂E ∂t + (< ∇, E > + ∇ ∧ E) + I  ∂B ∂t + (< ∇, B > + ∇ ∧ B)  = < ∇, E > | {z } scalar +  ∂E ∂t − ∇ × B  | {z } vector + I ∂B ∂t + ∇ × E  | {z } 2−vector + + I < ∇, B > | {z } 3−vector . (2.67)

From above identity, the Maxwell equations can be written as the following compact formula

DF = J . (2.68)

Indeed, by comparing the both sides we immediately find that: < ∇, E > = ρ , ∂E ∂t − ∇ × B = J , ∂B ∂t + ∇ × E = 0 , < ∇, B > = 0 . (2.69) 

2.2.3

Periodicity

Now it is useful to introduce the periodicity of the Clifford algebras from which we can relate any Clifford algebra to a number of low-dimensional Clifford algebras. The structure of a real Clifford algebra is determined by the dimension of the vector space and the signature of the metric. Then, we take the field F to be the real numbers R and let us denote the Clifford algebra of the vector space Rp,q by C`p,q, i.e, C`p,q = C`(Rp,q). The Clifford algebra C`0,1 is generated by {1, ˆe2} where ˆ

e22 = −1, it is therefore isomorphic to the algebra of complex numbers

(26)

but the Clifford algebra C`1,0 with basis {1, ˆe1} where now ˆe21 = 1 is not related to any known algebra. We shall note that the elements f1 = 12(1 + ˆe1) and f2 =

1

2(1 − ˆe1) form a new basis {f1, f2}, with f 2

1 = f1, f22 = f2 and f1f2 = f2f1 = 0, it follows that f1 and f2 each span mutually orthogonal one-dimensional subalgebras, each of which is isomorphic to the field R. Then,

C`1,0 ' R ⊕ R . (2.71)

By multiplication of the elements of {1, ˆe1} and {1, ˆe2} we can construct a basis for Clifford algebra C`1,1 for which a possibility is {1, ˆe1, ˆe2, ˆe1eˆ2}. Then, the algebra C`1,1 can be naturally associated to algebra C`2,0 which, in its turn, is isomorphic to M(2, R) [2]. So,

C`1,1 ' M(2, R) . (2.72)

In particular, one can easily prove that [12, 1, 2] C`0,2 ' H .

Actually, these are some general features of associative algebras [1, 14].

The following theorem, called Periodicity theorem and demonstrated in [1], establish important isomorphisms between different Clifford algebras.

Theorem 1. Let C`p,q be the Clifford algebra for the vector space Rp,q:

C`p+1,q+1 ' C`1,1⊗ C`p,q ; C`q+2,p ' C`2,0⊗ C`p,q ; C`q,p+2 ' C`0,2⊗ C`p,q ,

(2.73) where p > 0 or q > 0.

This theorem immediately implies the following corollary:

Corollary 1. Any Clifford algebra C`p,q can be determined from the algebras C`0,1, C`1,0, C`1,1, C`0,2 and C`2,0.

Clifford algebras admit a periodicity of dimension 8 over the real numbers. Indeed, using the latter theorem, we obtain the following relations

C`0,4 ' C`0,2⊗ C`2,0 , C`4,0 ' C`2,0⊗ C`0,2 , C`0,8 ' C`0,4⊗ C`0,4. From above identity, we are left with the following final relation

C`p,q+8 ' C`0,4⊗ C`0,4⊗ C`p,q' C`0,8⊗ C`p,q. (2.74) thus achieving the periodicity that we were looking for.

(27)

2.3

The Spin Groups

In this section we establish the connection between the Clifford algebra and rota-tions, which is the most relevant application of the Clifford algebra. Indeed, this algebra provides a very clear and compact method for performing rotations, which is considerably more powerful than working with matrices. Let n ∈ V be a non-null vector, n2 = < n, n >6= 0. This elements are invertible in C`(V),

n−1 = n

< n, n >. (2.75)

The set of all the non-null vectors generates a group under Clifford product, this is called Clifford group

Γ =  np. . . n2n1 ∈ C`(V) | ni is non-null . (2.76) Note that n−1 = ±n when n is a normalized vector. Any vector V ∈ V can be decomposed as:

V = V nn−1

= (< V , n > + V ∧ n)n−1 = < V , n > n−1 + (V ∧ n)n−1

=⇒ V = Vk + V⊥ (2.77)

where Vk is the component of V along of n and V⊥ is the component of V orthogonal to n given by

Vk = < V , n > n−1 , V⊥ = (V ∧ n)n−1. (2.78) One can easily verify that the above relation has the properties that we desire. In order to see this, note that the expression for Vk is, in fact, the projection of V onto vector n, since < V , n > is a constant and the remaining term must be the perpendicular component. Moreover, from the following inner product

< n, V⊥> = n(V ∧ n)n−1

0 =V ∧ n −1

0 = 0 , we check that V⊥ is perpendicular to n. We can rewrite (2.77) as:

V − 2Vk = V⊥ − Vk

V − 2 < V , n > n−1 = (V ∧ n)n−1− < V , n > n−1

(2.79) If the linear operator ˆS : V 7→ V is the reflection in the plane orthogonal to n, we have that:

ˆ

S(V⊥ + Vk) = V⊥ − Vk

(28)

and by means of (2.14) and (2.18) we easily obtain that: ˆ S(V ) = ( V ∧ n − < V , n > )n−1 =  V U − U V 2 − V U + U V 2  n−1 =⇒ S(V ) = −nV nˆ −1 (2.81)

which is valid on spaces of any dimension and ˆS is called reflection operator. It follows that for V ∈ V the operation ˆS(V ) ∈ V and it is the reflection of the vector V with respect to the the plane orthogonal to n. Note that for a normalized vector n2 = ±1 the action of ˆS with itself is ˆS( ˆS(V )) = ˆS ◦ ˆS(V ) = (−1) ˆS(nV n−1) = nnV nn = V , i.e., ˆS ◦ ˆS = I. We should check that the expression for the reflection has the desired property of leaving the inner product invariant. A simple proof is given by:

< ˆS(V ), ˆS(U ) > = S(V ) ˆˆ S(U ) + ˆS(U ) ˆS(V )

2 =

n(V U + U V )n−1 2

= < V , U > . (2.82)

The reflection operator is therefore a linear transformation that preserves the inner product. Moreover, it can be proved that this transformation has determinant −1. In fact, defining the action of ˆS on an arbitrary homogeneous multivector by

ˆ

S(V1V2. . . Vp) = ˆS(V1) ˆS(V2) . . . ˆS(Vp), using the following relation between the determinant and the pseudoscalar ˆS(I) = det( ˆS)I, see [4], and the equation (2.59) we find that:

det( ˆS)I = S(I) = ˆˆ S(ˆe1eˆ2. . . ˆen) = ˆS(ˆe1) ˆS(ˆe2) . . . ˆS(ˆen) = (−1)nnIn−1 = det( ˆS)I = (−1)n(−1)1−nI = −I

=⇒ det( ˆS) = −1 . (2.83)

We say thus that a linear transformation which leaves the inner product invariant and it has determinant −1 is called reflection.

Now, suppose that the vector U is the reflection of the vector V , ˆS1(V ) = U , with respect to plane orthogonal to normalized vector n1 and the vector X obtained by reflection of U , ˆS2(U ) = X, with respect to plane orthogonal to n2. Thus,

X = ( ˆS2◦ ˆS1)(V ) = −n2U n−12 = n2n1V n−11 n −1 2 . Defining R = n2n1 we can now write the result of the rotation as:

ˆ

R(V ) = RV R−1, (2.84)

where ˆR = ˆS2◦ ˆS1 is a linear operator ˆR : V 7→ V called rotation operator which represent the rotations. We shall note that in this derivation the dimension of the

(29)

vector space was never specified, so that it must work in all spaces, whatever their dimension. Analogously equations (2.82) and (2.83) it is simple matter to prove that the rotation operator preserves the inner product and it has determinant 1. Although this demonstration is completely analogous to the equation (2.82), let us do it again using the projection operator. Suppose that ˆR(V ) = RV R−1 and

ˆ

R(U ) = RU R−1, then

< ˆR(V ), ˆR(U ) > = RV R−1

RU R−1 0 = hV U i0 = < V , U > .

Now, let ˆR2 be a rotation operator such that ˆR1(V ) = R1V R−11 = U and ˆR2 another rotation operator such that ˆR2(U ) = R2U R−12 = X. It follows that:

X = Rˆ2(U ) = ˆR2( ˆR1(V )) = ( ˆR2◦ ˆR2)(V ) = ˆR(V ) = R2U R−12 = R2R1V R−11 R −1 2 = RV R −1 , (2.85)

where ˆR = ˆR2◦ ˆR1 and R = R2R1. In general, for any inhomogeneous multivec-tor A ∈ C`(V) we have that:

ˆ

R(A) = RAR−1. (2.86)

We say thus that a linear transformation which leave the inner product invariant and it has determinant 1 is called rotation. In particular, by what was seen above, a rotation in the plane generated by two unit vectors n2 and n1 is achieved by successive reflections with respect to the planes perpendicular n2 and n1. So, we can construct all rotations and reflections by application of an even number or an odd number of successive reflection operators. Moreover, it can be proved that in n dimensions any rotation can be decomposed as a product of at most n reflections [3].

The orthogonal group, denoted by O(V), is the group of linear transformations on (V, < , >) that preserve the inner product and the special orthogonal group, denoted by SO(V), is the subgroup of O(V) restrict to determinant 1. The theorem below summarizes the results obtained previously

Theorem 2. Any orthogonal transformation T ∈ O(V) can be written as a compo-sition of reflections with respect to the hyperplanes orthogonal to non-null vectors. The set of all normalized vectors form a group under Clifford product. Denoted by P in(V), this is called pin group

P in(V) = {R ∈ C`(V)|R = np. . . n2n1, ni ∈ V and n2i = ±1} . (2.87) Due to the Z2-grading, the elements of C`(V) split into those of even degree and those of odd degree. The even degree elements of the Pin group form a subgroup, denoted by SP in(V), called the spin group

(30)

Note that SP in(V) = P in(V) ∩ C`+(V). These groups are naturally defined on the Clifford algebra and can be used to implement reflections and rotations in arbitrary vectors V ∈ V. This can be seen very clear and compact below [12].

Rotation + Ref lection : (−1)pRV R−1, R ∈ P in(V)

P ure Rotation : RV R−1, R ∈ SP in(V) (2.89) where p is the degree of R. Since the action of the elements of Γ on vectors is quadratic, R and −R generate the same transformation. So there is a two-to-one map between elements of P in(V) and rotations and reflections. Mathematically, P in(V) and SP in(V) providing a double cover representation of the orthogonal groups O(V) and SO(V), respectively. Then we establish the following relations

O(V) = P in(V) Z2

; SO(V) = SP in(V) Z2

(2.90)

Note that the group Spin(V) is a subgroup of P in(V) formed by elements of even degree. One can also define P in(V) as being the group of the elements R ∈ Γ such that eRR = ± 1 , and SP in(V) as being a subgroup of P in(V) formed by elements R ∈ Γ+ such that eRR = ± 1 . It follows that the group P in(V) has a subgroup, denoted by P in+(V), whose elements are R ∈ Γ such eRR = 1 . Analogously, the group SP in(V) has an even subgroup, denoted by Spin+(V), whose elements are R ∈ Γ+ such eRR = 1 . Other subgroups can be obtained by restriction to certain subgroups of P in(V) and Spin(V). In fact, the group Spinb+(V) has as its elements R ∈ Γ such that its norm defined by c|R| = RR 0 is equal to 1 [8, 2]. By equation (2.89), an element of the group Spin(V) act in the vector space V and yields an element on the same vector space. This implies that the vector space V furnish a representation for the spin groups. But since the action is quadratic, it follows it is not faithful [12, 14, 1].

2.4

Spinors

In the previous section it was established that the vector space V provides a repre-sentation for the spins groups, since the action of elements of the spins groups in elements of V results in element of V. However, since elements of the spin group which differ by a sign produce the same orthogonal transformation, it follows that this representation is not faithful. In this section we introduce the so-called spinors, that generate a vector space that provides a faithful representation for the spin group. For simplicity, we will only consider the Clifford algebra of an orthogonal space (V, < , >) whose dimension is even, n = 2r with r ∈ N.

(31)

2.4.1

Minimal Left Ideal

There is an important class of subspaces which are called left ideals. A left ideal L ⊂ C`(V) of an algebra C`(V) is a subalgebra of C`(V) such that:

Aφ ∈ L ∀ φ ∈ L , A ∈ C`(V) , (2.91)

i.e, L is invariant under left multiplication of the whole algebra. Since C`(V) has a Z2-grading, we can use its even subalgebra C`+(V) as the representation space of C`(V) and define ρ : C`(V) 7→ End(C`+(V)). By means of (2.27) we can split any multivector A ∈ C`(V) as A = A++ A−, where A± = 12(A ± bA) ∈ C`±(V). We should establish in what conditions ρ = ρ++ ρ− such that ρ(A) = ρ+(A+) + ρ−(A−) will be a representation of C`(V). Note that A+B ∈ C`+(V) if B ∈ C`+(V), this implies that ρ+(A+)(B) = A+B ∀ B ∈ C`+(V), while A−B ∈ C`−(V) if B ∈ C`+(V), this requires that ρ

−(A−)(B) 6= A−B when B ∈ C`+(V). If, however, we can take an odd element C ∈ C`−(V) and define

ρ−(A−)(B) = A−BC ∈ ∀B ∈ C`+(V) (2.92) such that C2 = 1, then ρ = ρ

++ ρ− is a representation in the space C`+(V). One might then wonder, how can we know if such representation is reducible? Suppose that there exists an element C1 ∈ C`+(V) such that

C2

1 = 1 ; C1C = CC1. (2.93)

Thus we can write C`+(V) = +C`+(V) ⊕ −C`+(V), where

±C`+(V) = C`+(V) 1

2(1 ± C1) . (2.94)

so that, for B± ∈ ±C`+(V) we have B±C1 = ±B±. Each of these spaces ±C`+(V) is invariant under the action of ρ as we can see by the equation (2.93) and it is also a subalgebra of C`+(V). If another element C2 ∈ C`+(V) such that C22 = 1 , C2C1 = C1C2 and C2C = CC2 can be found, then the subspaces±C`+(V) carry no irreducible representation. So, we define another four subspaces

±C`+(V) = ±C`+(V) 1

2(1 ± C1)(1 ± C2), (2.95) each of which invariant under the action of ρ. We can proceed with this construction in an analogous fashion if there are other elements Cj commuting with the previous ones.

The regular representation ρ : C`(V) 7→ End(C`(V)) given by ρ(A)B = AB pre-serves certain vector subspaces. The minimal left ideals are just the invariant sub-spaces S ⊂ C`(V) for which the map ρ : C`(V) 7→ End(S) defined by ρ(A)ψ = Aψ

(32)

is an irreducible representation. Note that the minimal left ideal contains no left ideals apart from itself and zero, it provides the least-dimensional faithful represen-tation of C`(V), the so-called spinorial represenrepresen-tation of the Clifford algebra. The minimal left ideal S is called, as a vector space, the spinor space, and its elements, denoted by ψ, are called spinors. Note also that these minimal left ideals are al-gebras, the subalgebras of C`(V). Thus, under Clifford product, the space carrying a such representation of C`(V) will be called spinorial algebra and the choice of a different minimal left ideal gives other equivalent representations [1, 11, 14]. We will therefore establish that the spinor representation of the Clifford algebra induces a representation of any subset by restricting to left multiplication on the ideal by elements of that set. In particular it induces a representation of the Clifford group [1]. When the dimension of V is n = 2r one can prove that the dimension of the spinor space is 2r. When C`(V) is thought of as a matrix algebra, an example of a minimal left ideal is the subalgebra of matrices with all columns but the first vanishing [2, 11, 30]. In particular, C`(V), P in(V), SP in(V), O(V) and SO(V) can be faithfully represented by 2r× 2r matrices. In this case spinors are represented by the column vectors on which these matrices act.

By what was seen above, the minimal left ideals are of great relevance in the study of the spinors. These can be obtained by action of the whole algebra on the so-called primitive idempotent of C`(V), this makes clear the fundamental importance of the primitive idempotents. An element ξ of an any algebra A is said to be an idempotent if ξ2 = ξ and ξ 6= 0. Particularly, such an idempotent generates a subalgebra ξAξ. Now, if A is a division algebra, then the single idempotent is the identity, since if ξ 6= 0, then every ξ has an inverse ξ−1, it implies that ξ2 = ξ ∵ ξ−1ξ2 = ξ−1ξ = 1A ∴ ξ = 1A. We can split the element ξ into a sum of other two elements ξ1 and ξ2, ξ = ξ1 + ξ2. Hence, ξ will be idempotent if the following relations are satisfied: ξ21 = ξ1, ξ22 = ξ2 and ξ1ξ2 = −ξ2ξ1. If such elements satisfying the conditions ξ1ξ2 = 0 and ξ = ξ12cannot be found, we call ξ a primitive idempotent. Therefore, a idempotent ξ is primitive if and only if it is the single idempotent on ξAξ [2, 14, 17]. Let ξ be a generic primitive idempotent in C`(V), then any minimal left ideal S. has the form:

S = C`(V)ξ = {Aξ | A ∈ C`(V)} . (2.96) When the dimension of V is even the square of I depends only on the signature of the inner product. So, equation (2.56) is reduced to:

I2 = (−1)12[n(n−1)+(n−s)] ∀ n

=⇒ I2 = (−1)−s2 , n = 2r . (2.97)

(33)

the action of I on S gives a decomposition [12]

S = S+ + S− ; S± = {ψ ∈ S | Iψ = ±εψ} , (2.98) where S± are subspaces of dimension 2r−1 whose elements are called semi-spinors. The semi-spinors are also called Weyl spinors of positive and negative chirality. These subspaces are invariant under the action of the even subalgebra. Indeed, by means of the equation (2.59) we have that IA+ = A+I for A+ ∈ C`+(V) and n even. Then, due to (2.98) we have:

IA+ψ = A+

=⇒ IA+ψ = ±εA+ψ ∀ A+∈ C`+(V) . (2.99) This means that the spinorial representation of C`+(V) on S is reducible while the spinorial representation of C`+(V) on S± is irreducible [12, 1]. In paricular, the spinorial representation of C`+(V) splits in two blocks of dimension 2r−1 × 2r−1. It happens that, by equation (2.88), elements of Spin(V) are of degree even. So, by restriction, the spinorial representation of C`+(V) induce a representation of dimension 2r−1 for Spin(V) [14].

Though we have restricted ourselves to even-dimensional spaces, let us clarify these concepts with an example in three dimensions following the line of the previous examples. We will assume that the dimension of S is 2[n/2], where [ ] denotes the integer part of the number inside it.

Example 3: Spinors in the minimal left ideal of C`(R3).

Let {ˆe1, ˆe2, ˆe2} be an orthonormal basis for vector space V = R3 where ˆe1ˆe1 = ˆ

e2eˆ2 = ˆe3ˆe3 = 1 , ˆeaeˆb = −ˆebeˆaif a 6= b. The set {1, ˆe1, ˆe2, ˆe3, ˆe1ˆe2, ˆe1ˆe3, ˆe2ˆe3, I} span the Clifford algebra C`(R3) whose the most general element is:

A = a + a1ˆe1+ a2ˆe2+ a3ˆe3 + a12ˆe1eˆ2+ a13eˆ1eˆ3+ a23ˆe2ˆe3 + a123I , where I is the pseudoscalar. In three dimensions, we must find a spinor space S of dimension dim(S) = 2[3/2] = 2. Now, consider the element

ξ1 = 1

2(1 + ˆe3).

Note that ξ1, ξ12 = 12(1 + ˆe3)12(1 + ˆe3) = ξ1, is a primitive idempotent. It follows that the set of the form

(34)

is a minimal left ideal of C`(R3). Using the Clifford algebra, one finds after some manipulations that

1 = [(a + a3) + I(a12+ a123)]ξ1 + [(a1+ a13) + I(a2+ a23)]ˆe1ξ1. (2.100) Defining ξ2 = ˆe1ξ1, we obtain that:

1 = [(a + a3) + I(a12+ a123)]ξ1 + [(a1 + a13) + I(a2+ a23)]ξ2. (2.101) In the same way, we obtain

2 = [(a1 − a13) − I(a2+ a23)]ξ1 + [(a − a3) − I(a12+ a123)]ξ2. (2.102) Note that that the pseudoscalar I commutes with all elements and squares to −1, then it can be viewed as the unit imaginary i in C`(R3), Iξ1 = iξ1. Since the action of A on ξ1 and ξ2 are both linear combination of the set {ξ1, ξ2}, we see that:

S = {ψ ∈ C`(R3) | ψ = ψ1ξ1 + ψ2ξ2 ∀ ψ1, ψ2 ∈ C} , (2.103) where ψ1 = α1 + iβ1 and ψ2 = α2 + iβ2. This space admits no proper left ideal, so the elements ψ = ψAξA are called spinors and {ξA} where A ∈ {1, 2} form a basis for S, the called spinor basis. It is simple to prove that S is invariant by the action of C`(R3). In order to prove this explicitly, we only need to act the elements that span V which are ˆe1, ˆe2 and ˆe3 on the elements of S and this implies in the following spinorial representation for the vectors of the basis:

ˆ e1ψ = ˆe1(ψ1ξ1 + ψ2ξ2) = ψ1ξ2 + ψ2ξ1 =⇒ ˆe1 ' 0 1 1 0  (2.104) ˆ e2ψ = ˆe2(ψ1ξ1 + ψ2ξ2) = −iψ1ξ2 + i ψ2ξ1 =⇒ ˆe2 ' 0 −i i 0  (2.105) ˆ e3ψ = ˆe3(ψ1ξ1 + ψ2ξ2) = ψ1ξ1 − ψ2ξ2 =⇒ ˆe3 ' 1 0 0 −1  , (2.106) i.e., the action of V on S yield elements on S. Thus, as a matrix algebra, using these three last equations we can see that the spinor can be represented by:

ψ = ψ1 ψ2  ∈ C2 or ψ = ψ1 0 ψ2 0  ∈ M(2, C)ξ1, (2.107)

(35)

where ξ1 = 1 0 0 0 

. If we multiply ψ on the left by an arbitrary matrix ∈ M(2, C) we are left with a completely analogous relation [2]. Using these matrices we arrive on the matricial representation of the multivetor A obtained on Example 1.

Defining the normalized vectors n1 = ˆe1, n2 = cosθ ˆe1 + sinθ ˆe2, then we can construct the following element of the group SP in(V): R1 = n2n1 = cosθ − sinθˆe1eˆ2. It follows that an element R1 of Spin(R3) acts on the elements of R3 as R1ˆeaR−11 (a = 1, 2, 3). Thus, ˆ e1 7→ ˆR1(ˆe1) = n2n1ˆe1n1n2 = cos(2θ)ˆe1 + sin(2θ)ˆe2 ˆ e2 7→ ˆR1(ˆe2) = n2n1ˆe2n1n2 = −sin(2θ)ˆe1 + cos(2θ)ˆe2 ˆ e3 7→ ˆR1(ˆe3) = n2n1ˆe3n1n2 = ˆe3 . (2.108) The result of the action of ˆR1 is therefore a rotation of 2θ on the plane generated by the two unit vectors ˆe1 and ˆe2. Note that from the expression of R1 it can be represented by R1 = e−θˆe1eˆ2. But we can implement rotation in any one of the ˆ

eaeˆb planes. So, we can perform others rotation by R = e−θˆeaˆeb whose action yields a rotation of 2θ on the ˆeaeˆb plane. In particular, taking θ = π, we have a rotation through 2θ on the ˆeaeˆb plane. In this case, the vectors remain unchanged while the spinors change of sign.



2.4.2

Pure Spinors

From the geometrical point of view, there exists an important class of spinors which are associated with maximal isotropic vector subspaces, observed by Cartan [3], these are the so-called pure spinors. For example, let V be a null vector, namely < V , V > = 0. In three dimensions we can write the complex components (V1, V2, V3) of this vector in terms of two elements ψ1, ψ2 as:

V1 = ψ12 − ψ22 V2 = i(ψ12 + ψ22)

V3 = −2ψ1ψ2, (2.109)

which automatically guarantees the isotropic character of the vector. Indeed, the latter equation immediately implies that V12 + V22 + V32 = 0. Note that these equations are solved for

ψ12 = V1 − iV2 2 ; ψ 2 2 = − V1 + iV2 2 . (2.110)

(36)

So, using the above equation and that V3 = −2ψ1ψ2 we obtain that V3ψ1 + (V1 − iV2)ψ2 = 0 (V1 + iV2)ψ1 − V3ψ2 = 0 =⇒  V3 V1 − iV2 V1 + iV2 −V3  ψ1 ψ1  = 0 , (2.111) which is the matrix representation of V ψ = 0, with

 V3 V1 − iV2 V1 + iV2 −V3  V3 V1 − iV2 V1 + iV2 −V3  = 1 0 0 1  < V , V > = 0 and < V , V > = 0. Note that the rotation (2.108) on the vector V implies,

V10 = V1cos(2θ) − V2sin(2θ) V20 = V1sin(2θ) + V2cos(2θ)

V30 = V3. (2.112)

Using this, it is simple matter to prove that: ψ1 ψ2  =⇒ψ 0 1 ψ02  = e −iθ ψ1 eiθψ 2  . (2.113)

Note that taking θ = π the vector remains unchanged by the action of the rotation operator, while the elements ψ1, ψ2 change the sign. Thus, the pair ψ1, ψ2 consti-tutes the components of a spinor ψ which is associated a null subspace spanned by the vectors that annihilate it, V ψ = 0, whose representation is given by (2.111). The spinor ψ is an example of a pure spinor.

Now, let us make a rigorous definition of a pure spinor. Formally, given a spinor ψ ∈ S one can construct a subspace of V, denoted by Nψ, defined by:

Nψ = {V ∈ V | V ψ = 0 } . (2.114)

Note that if ψ in non-null then for all V , U ∈ Nψ we have 2 < V , U > ψ = (V U + U V )ψ = 0

=⇒ < V , U > = 0 . (2.115) This subspace is clearly totally null, which means that all vectors belonging to it are orthogonal to each other including to itself.

In what follows, we will only consider the complexified Clifford algebra of an even dimensional orthogonal space. The complexification of a real orthogonal space (V, < , >) is the space (VC, < , >C) whose elements are of the form V + iU for some V , U ∈ V and i the unit imaginary. We can define the sum and multiplication by a complex scalar (λ + iδ) as:

(V1 + iU1) + (V2 + iU2) = (V1 + V2) + i(U1 + U2)

(37)

The complex conjugate of V = V1 + iU1 ∈ VC for V1, U1 ∈ V is defined by V∗ = V1− iU1. Since VCcan be obtained by the complexification V, VC = C⊗V, we see that the complex dimension of VCis n and the real dimension 2n. The inner product < , > is extended to < , >C on VC by assuming bilinearity of the inner product in the complex field. For simplicity, from now on we shall assume that the complex Clifford algebra associated to VC endowed is simply C`(VC) and the complexification of C`(V, < , >) is C`C(V). From this, we say that theses algebras are isomorphic, namely:

C`(VC) ' C`C(V) , (2.117)

where C`C(V) = C ⊗ C`(V).

We say that a spinor is pure if the null subspace associated to it is maximal [3, 14], namely when the null subspace has the maximum dimension possible. But, what is maximum dimension of a null subspace? When the complex dimension is n = 2r, the maximal dimension that a null subspace can be have is r. For example, given the vectors e1, e2, . . . , er, er+1, . . . , e2r, if we define na = ea+ i ea+r, then we have a null subspace of dimension r. In this case we say that it is maximal or maximally null. Therefore, a spinor associated to a null subspace with this dimension is said to be pure. Apart from a multiplicative factor, there is a one-to-one association between ψ and Nψ, thus a pure spinor ψ is a representative for Nψ if and only if V ψ = 0 for all V ∈ Nψ.

We should recall that a vector V transforms under rotation as RV R−1 while the spinor ψ transform as Rψ. So, let ψ be a pure spinor and V a null vector that belongs to Nψ, then V ψ = 0. If V0 = RV R−1 and ψ0 = Rψ. It follows that

V0ψ0 = RV R−1Rψ = RV ψ

=⇒ V0ψ0 = 0 . (2.118)

i.e., we can transform any null subspace Nψ into any another null subspace Nψ0 by a rotation or reflection [3]. It is worth stressing that the sum of two pure spinors is not a pure spinor in general, since the purity condition is quadratic on the spinor [11, 1]. In practice, given a spinor ψ when is it pure? In order to answer this question, let us workout an example in R3. Let S = {ψ ∈ C`(R3) | ψ = ψ1ξ1 + ψ2ξ2 ∀ ψ1, ψ2 ∈ C} be the minimal left ideal of the Clifford algebra C`(R3) where ξ

1 = (1 + ˆe3)/2, ξ2 = e1ξ1 and let V = V1eˆ1 + V2ˆe2 + V3ˆe3 be a null vector, i.e., < V , V > = V2

1 + V22 + V32 = 0. Here, the pseudoscalar I = ˆe1eˆ2eˆ3 commutes with all elements and squares equal to −1 and therefore be viewed as the unit imaginary i. Indeed Iξ1 = iξ1. Acting V on ψ

V ψ = (V1ˆe1 + V2ˆe2 + V3ˆe3)(ψ1ξ1 + ψ2ξ2)

(38)

we obtain the following system to be solved    V1ψ2 + V3ψ1 − iV2ψ2 = 0 V1ψ1 − V3ψ2 + iV2ψ1 = 0 V12 + V22 + V32 = 0 whose solution is V1 = ψ12 − ψ22 V2 = i(ψ12 + ψ22) V3 = −2ψ1ψ2. (2.120)

Therefore, if the components of the spinor ψ is characterized by a set of quadratic relations, then the spinor ψ is said to be pure. It turns out that pure spinors are necessarily semi-spinors. Let us prove this. Since the dimension of VC is 2r we can choose the first r vectors to be the set {e1, e2, . . . , er} such that < ea, eb > = 0. Thus, this set forms a basis for a maximal null subspace. To complete this basis we choose the other r as being the set of the vectors {er+1, er+2, . . . , e2r} such that < ea, eb > = 0. Theses two sets form a basis for the whole vector space VC and its elements are such that < ea, eb > = 12δba. The pseudoscalar of C`C(V) has the following proportionality

I ∝ (e1∧ e1)(e2∧ e2) . . . (er∧ er)

∝ [e1, e1][e2, e2] . . . [er, er] , (2.121) where [ea, eb] = eaeb − ebea denote the Clifford commutators. Since the vectors ea span Nψ, by definition that eaψ = 0 . Hence ebeaψ = 0. Thus we only need to know the action of eaeb on ψ, that is,

(eaeb)ψ = (eaeb + ebea)ψ = 2 < ea, eb > ψ = δabψ . (2.122) Thus, the action of commutator is

[ea, eb]ψ = (eaeb − ebea)ψ = δabψ . (2.123) Therefore, using (2.121) and (2.123) is trivial show that Iψ ∝ ψ. It follows that every pure spinor ψ must be a semi-spinor, or Weyl spinor. It turns out that in two, four and six dimensions all semi-spinors are pure, while in higher dimensions not all are semi-spinors.

2.4.3

SPin-Invariant Inner Products

We have referred to a minimal left ideal as the spinor space and its elements as spinors, we now examine some operations and properties that this space possess

(39)

with respect to products of two of its elements. We shall construct an invariant product under the action of the spin group. For example, the space of spinors S = C`(R3

1 given by

S = ψ ∈ C`(R3) | ψ = ψ1ξ1 + ψ2ξ2 ∀ ψ1, ψ2 ∈ C , (2.124) obtained in the Example 3 can be represented as follows:

S '  [ψ] ∈ M(2, C)ξ1| [ψ] = ψ1 0 ψ2 0  ∀ ψ1, ψ2 ∈ C  . (2.125) But since the product ξ1ξ2 = 0, the set

D = ξ1C`(R 3 ) ξ1 = {ψ1ξ1, ψ1 ∈ C} ' ψ1 0 0 0  | ψ1 ∈ C  (2.126) is a subalgebra of C`(R3) with unity ξ1 such that λξ1 = ξ1λ , for all λ ∈ D. According to equation (2.40), none of its elements is invertible as an element of C`(R3), however there is a unique λ

2 ∈ D with λ2λ1 = ξ1 if λ1 is a non-zero element of D. Thus, D is a division algebra and hence the idempotent ξ1 is primitive in C`(R3) as expected. The space S has a natural right D-linear structure defined by

S × D → S

ψ, λ 7→ ψλ . (2.127)

The space S endowed with this right D-linear structure is called, more formally, the spinor space. We must note that on spinorial representation an arbitrary multivec-tor A ∈ C`(R3) and its reverse are given by (see the Example 1):

[A] = z1 z3 z2 z4  ; [ eA] = z ∗ 1 z2∗ z3∗ z4∗  , zi ∈ C . (2.128) Then, given two spinors ψ, φ ∈ S the product

[ eψ][φ] = ψ ∗ 1 ψ∗2 0 0  φ1 0 φ2 0  = ψ ∗ 1φ1 + ψ∗2φ2 0 0 0  ∈ D (2.129)

falls in the division algebra D. Now, the inner product can be naturally introduced on the spinor space as

S × S 7→ D

ψ, φ 7→ (ψ, φ) = eψφ , (2.130)

if for ψ ∈ S and λ ∈ D the following relation holds: ]

(40)

where λ → eλ is the complex conjugation of the division algebra D and the reversion ψ 7→ eψ is a semi-linear map. Now, let A be an arbitrary element of C`(R3). Then,

(Aψ, Aφ) = gAψAφ = eψ eAAφ

This implies that the elements A ∈ C`(R3) such that eAA = 1 preserve the inner product, i.e., this inner product between spinors is invariant under the action of the group Spin+(R3). In the same way, we can use another involution to define a different inner product. Indeed, the spinorial representation of A ∈ C`(R3) furnish that its Clifford conjugate is:

[A] =  z4 −z2 −z3 z1



, zi ∈ C . (2.132)

We can use this Clifford conjugation to introduce an inner product for spinors. However, the product

[ψ][φ] =  0 0 −ψ2 ψ1  φ1 0 φ2 0  =  0 0 ψ1φ2 − ψ2φ1 0  / ∈ D (2.133)

does not fall in the division algebra D. But since ψφ = (ψ1φ2 − ψ2φ1)ξ2 its always possible to find an invertible element V = ˆe1 ∈ C`(R3), in this case, such that the product ˆe1ψφ = (ψ1φ2 − ψ2φ1)ˆe1ξ2 = (ψ1φ2 − ψ2φ1)ξ1. Thus, [V ][ψ][φ] = 0 1 1 0   0 0 ψ1φ2 − ψ2φ1 0  = ψ1φ2 − ψ2φ1 0 0 0  ∈ D . (2.134) This leads us to the conclusion that,

S × S 7→ D

ψ, φ 7→ (ψ, φ) = V ψφ (2.135)

also defines an inner product. One can easily prove that this latter inner product between spinors is invariant under the action of the group Spin(R3).

In what follows, we will deal with a general Clifford algebra on which every spinor space admits a real, complex or quaternion division algebra. Let η be some involu-tion. If ξ is any primitive idempotent then ξη = J ξJ−1 for some element J with Jη = εJ and ε = ±1 [1]. Then, if ψ, φ ∈ S, we define:

S × S → D

ψ, φ 7→ (ψ, φ) = J−1ψηφ . (2.136)

If A ∈ C`C(V) is arbitrary element we find that:

Referências

Documentos relacionados

The probability of attending school four our group of interest in this region increased by 6.5 percentage points after the expansion of the Bolsa Família program in 2007 and

Para tanto foi realizada uma pesquisa descritiva, utilizando-se da pesquisa documental, na Secretaria Nacional de Esporte de Alto Rendimento do Ministério do Esporte

Na própria aplicação (disponível no menu de aplicações), pode aceder à sua caixa de entrada, assim como realizar diferentes configurações tais como a forma de mostrar os

i) A condutividade da matriz vítrea diminui com o aumento do tempo de tratamento térmico (Fig.. 241 pequena quantidade de cristais existentes na amostra já provoca um efeito

Ousasse apontar algumas hipóteses para a solução desse problema público a partir do exposto dos autores usados como base para fundamentação teórica, da análise dos dados

didático e resolva as ​listas de exercícios (disponíveis no ​Classroom​) referentes às obras de Carlos Drummond de Andrade, João Guimarães Rosa, Machado de Assis,

This log must identify the roles of any sub-investigator and the person(s) who will be delegated other study- related tasks; such as CRF/EDC entry. Any changes to

Além disso, o Facebook também disponibiliza várias ferramentas exclusivas como a criação de eventos, de publici- dade, fornece aos seus utilizadores milhares de jogos que podem