Home Mathematics Structures of spinors fiber bundles with special relativity of Dirac operator using the Clifford algebra
Article Open Access

Structures of spinors fiber bundles with special relativity of Dirac operator using the Clifford algebra

  • Yousif Atyeib Ibrahim Hassan EMAIL logo
Published/Copyright: December 3, 2021
Become an author with De Gruyter Brill

Abstract

The purpose of this article is to demonstrate how to use the mathematics of spinor bundles and their category. We have used the methods of principle fiber bundles obey thorough solid harmonic treatment of pseudo-Riemannian manifolds and spinor structures with Clifford algebras, which couple with Dirac operator to study important applications in cohomology theory.

MSC 2010: 55N08; 55R05; 55R20

1 Introduction

The interplay between physics and mathematics has spurred each of the disciplines to great heights. It has led to numerous discoveries, not the least of which is the Dirac operator and the related concept of the spinor. The first groundwork for these concepts was laid down by Clifford in the middle of the nineteenth century as a generalization of the quaternions of Hamilton and the exterior algebra [1] of Grassmann. In 1913 [1], Elie Cartan wrote down the general theory of spinors. Spinors were first applied to mathematical physics by Wolfgang Pauli in 1927 [1], when he introduced his spin matrices. On the physics, Dirac introduced his famed operator in 1928 [1,2] but made no mention of the connection with spinors, and this was only done much later. During the years 1940–1970, the spinors and Clifford algebra of bundle became a fundamental tool of particle physics and came back later, at the forefront of differential geometry and of mathematics in general, with the recognition of the importance of the Dirac operator and theory of spinors in differential geometry.

Connections on fiber bundles and their relation with covariant derivation is discussed. In the final section, spin structures are introduced as nontrivial coverings of SO bundles, followed by a proof of necessary and the sufficient condition for their existence in terms of SO bundle and Dirac operator.

2 Clifford algebras

A Clifford algebra is a type of associative algebra, which can be considered a generalization of the usual associative fields such as ℝ, ℂ or ℍ. In fact, these fields can be seen as particular examples of Clifford algebras [1]. It is defined over a vector space V over a field k denoted by V k equipped with a quadratic form q defined as follows:

Definition 1

A quadratic form q is an operator from a vector space V k to its fields k, such that if αk and δ, ɳV k , then q(αɳ) = α 2 q(ɳ) and 2q(δ, ɳ) = q(δ + ɳ) − q(δ) − q(ɳ) is a symmetric bilinear form on V k [2]. A nondegenerate quadratic form is a quadratic form with the extra condition that q(ɳ) = 0 ⇔ ɳ = 0 [1].

A Clifford algebra is a generalization of the exterior or Grassmann algebra precisely.

Definition 2

Let V k be a vector field over a field k and take the tensor algebra T(V k ) = i = 0 i V k = kV k V k V k ⊕ … and let q be a quadratic form [1]. A Clifford algebra Cℓ(V k , q) on V k is then Cℓ(V k , q) = T(V k )/I(V k , q) with I(V k , q), the two-sided ideal generated by ɳɳ + q(ɳ) [1,2]. This is an associative algebra with unit, the exterior algebra can be related with the Clifford algebra with quadratic form 0. Defining the Clifford algebra and deriving its basic properties, we follow [2], but we could have equivalently started with the universal property proven below and worked our way back from there.

Lemma 1

Every vector space with a quadratic form has a q-orthogonal basis {e 1,…, e n}, that is, q(e i , e j ) = 0 if ij [1].

Definition 3

(Pin and Spin groups)

The group Pin is the following subgroup of the Clifford group:

Pin ( V k , q ) = { η P ̌ ( V k , q ) | q ( η ) = ± 1 }

i.e., we define P(V k , q) ⊂ (V k , q) (where (V k , q) calls Lipschitz group which is “largest” spinor group) to be the subgroup of Cℓ(V k , q) generated by the elements ɳV k with q(ɳ) ≠ 0. And that there is a representation

P ( V k , q ) Ad O ( V k , q )

The group spin is the even subgroup of Pin: Spin(V k , q) = Pin(V k , q) ∩ Cℓ0(V k , q) [1,2].

Theorem 2

Let k be a spin field. Then, Pin(V k , q) is the kernel of N : (V k , q) → k ×, (where k × is nonzero multiples of 1) and the twisted adjoint Ad ˜ |Pin(V k,  q) is a surjection of Pin(V k , q) on O(V k , q). Then, the sequence:

1 μ 4 ( k ) Pin ( V k , q ) Ad ˜ O ( V k , q ) 1

with µ 4(k) the fourth roots of 1 in the field. For example, µ 4(ℝ) = {−1, 1} ≈ ℤ2. And µ 4(ℂ) = {−1, 1, −i, i} ≈ ℤ4 (1,10).

Proof

See [1].□

We can also consider the covering group of SO(V k , q) = {xO(V k , q) | det(x) = 1}. Then, we have the following:

Corollary 1

The group Spin(V k , q) is a cover of SO(V k , q), and we have the exact sequence:

1 μ 4 ( k ) Spin ( V k , q ) Ad ˜ SO ( V k , q ) 1

with µ 4(k) as in Theorem 2 [1,3].

Proof

Let x ∈ Pin(V k , q). Then, Ad ˜ x is equal to the composition of a number of reflections due to Theorem 2. We have that for all vV k , det ( Ad ˜ η ) = 1 . To see this, take a q-orthogonal basis with ɳ 1 = ɳ and q(ɳ, ɳ j ) = 0 for all j > 1. Then, Ad ˜ η ( η 1 ) = η by definition and Ad ˜ η ( η j ) = η j with j > 1 by definition, so det ( Ad ˜ η ) = 1 . So any element of SO(V k , q) is generated by an even number of reflection, and so Spin(V k , q) must be generated by an even number of vectors, so by the properties of the ℤ2 grading we get that it is also an element of Cℓ 0 (V k , q). The exact sequence follows immediately from Theorem 2 [1].□

Proposition 3

For any x in Spin(n), there exists an even integer p and elements f 1,…, f p of norm 1 such that x = f 1,…, f p . The reverse statement also holds [4].

Theorem 4

For any n ≥ 2, Spin(n) is connected. For n ≥ 3, Spin(n) is simply connected; hence, Spin(n) is the universal covering Lie group of SO(n).

Proof

Then, the short exact sequence of Lie groups.

1 µ 4 ( k ) Spin ( V k , q ) Ad ˜ SO ( V k , q ) 1 .

Since π 1SO(n) = ℤ2 for any n ≥ 3, any connected couple covering of SO(n) is simply connected for n ≥ 3. Hence, it suffices to show that Spin(n) is connected for any n ≥ 2 [3].

First, let us show that 1 and −1 are path connected in Spin(n). Since the dimension is at least 2, there are elements e 1 and e 2 in V k , such that ‖e 1‖ = ‖e 2‖ = 1 and ⟨e 1, e 2⟩ = 0. For 0 ≤ tπ, we define γ(t) = cos t + e 1 e 2 sin t = e 1(−e 1 cos t + e 2 sin t). It is trivial that the norm (−e 1 cos t + e 2 sin t) is 1. By Proposition 3, γ(t) is an element of Spin(n) for any t. We have γ(0) = 1 and γ(π) = −1.

Second, note that any element y of Spin(n) can be connected with y by the path (t).

Finally, let x and y be in Spin(n). Since SO(n) is connected, there exists a path from ρ(x) to ρ(y) in SO(n). We may lift this path to a path starting in x and ending in a point y′ of Spin(n). Since the kernel of ρ equals {±1}, it holds that y= y or y=y. In the first case, we are done. In the second case, we connect x with −y and then connect −y with y [4].□

We now can construct any real Clifford algebra we are interested in. For example, the physically interesting Clifford algebras Cℓ1,3 and Cℓ3,1 are now constructed as Cℓ1,3 ≃ Cℓ0,2 ⊗ Cℓ1,1 = ℍ ⊗ ℝ(2) ≃ ℍ(2) and Cℓ3,1 ≃ Cℓ2,0 ⊗ Cℓ1,1 = ℝ(2) ⊗ ℝ(2) ≃ ℝ(4) [1].

Theorem 5

There are isomorphisms

C n , 0 C 0 , 2 C 0 , n + 2 , C 0 , n C 2 , 0 C n + 2 , 0 , C r , s C 1 , 1 C r + 1 , s + 1 ,

or all n, r, s ≥ 0 [2].

Note that here we are using the ungraded tensor product.

Proof

See [1].□

Remark 1

It is standard notation to write: q r,s ≡ q, O r.s ≡ O(V k , q) and SO r,s ≡ SO(V k , q). In accordance, we write Pin r,s ≡ Pin(V k , q) and Spin r,s ≡ Spin(V k , q). Similarly, it is conventional to write O n ≡ O n,0 O 0,n and SO n SO n,0 SO 0,n . Thus, we set Pin n = Pin n,0 , and Spin n = Spin n,0 .

We also write P r.s P(V k , q) and r,s (V k , q) [1].

Definition 4

A section of Clifford Cℓ(V k , q) is called Clifford field [5].

3 Spinors

We will build representations with m-dimensions of spinor on a complex vector, and it will become clear that the complex Clifford algebra has a much simpler structure than the real one, a period of degree 2 instead of degree 8 as in the real case [3]. Now if m is even, m = 2n, we have Welly representation, which restricts to Cℓ0(V k , q).

Definition 5

Clifford’s complex algebra Cℓ(V k , Q) is a Clifford algebra that is constructed by starting with a complex vector space V k ℂ, and Q extends through the complex linearity and then using the definition as real case. By starting with a real vector space V k of dimension n, then this is denoted by Cℓ(n). One can easily see that Cℓ(n) = Cℓ(n) ⊗ Cℓ. Build a spin representation as being reversible elements in Cℓ(n) are complexified, producing a structure of Spin(n, ℂ). (The complexification of Spin(n)) is a reversible element in Cℓ(n).)

By an inductive argument. The algebras Cℓ (n) constructed to begin

C ( 1 ) = , C ( 2 ) = ,

so

C ( 1 ) = C ( 1 ) =

and

C ( 2 ) = C ( 2 ) = = M ( 2 , ) .

Theorem 6

Cℓ(n + 2) = Cℓ(n) ⊗ Cℓ(2) = Cℓ(n) ⊗ M(2, ℂ).

To start the induction, we put n = 1, 2.

Corollary 2

If n = 2k, Cℓ(2k) = M(2, ℂ) ⊗⋯⊗ M(2, ℂ) = M(2 K , ℂ), where the product has k factors, and if n = 2k + 1, Cℓ(2k + 1) = Cℓ(1) ⊗ M(2 k , ℂ) = M(2 K , ℂ) ⊕ M(2 K , ℂ) [3].

Proof

Choose generators g 1, g 2 of Cℓ(2), f 1,…, f n of Cℓ(n) and ĕ 1,…, ĕ n+2 of Cℓ(n + 2). Then, we get the symmetry by the following map:

e ˘ 1 1 g 1 e ˘ 2 1 g 2 e ˘ 3 if 1 g 1 g 2 e n + 2 if n g 1 g 2

If n = 2k (even case). In this case, Clifford’s complex algebra is the algebra of 2 k by 2 k complex matrices.□

Definition 6

A spin structure on a pseudo-Riemannian vector bundle with signature (r, s) E is a principal Spin(r, s) bundle P Spin(E) together with a two-sheeted covering ξ: P Spin(E) → P SO(E), such that ξ(pg) = ξ(p)ξ 0(g) for all pP Spin and g ∈ Spin(r, s) and ξ0 the covering map Spin(r, s) → SO(r, s) [2,3].

Two spin structures P Spin(E) and P Spin ' (E) are called equivalent if there is a mapping F, such that F(gh) = F(g)h for gP Spin and h ∈ Spin and the following diagram commutes:

This means that if they are equivalent as spin structures, they are equivalent as principal fiber bundles [1,2].

From now on, we will only consider r, s such that π 1(SO(r, s)) = ℤ2. This makes many proofs much simpler since it makes Spin(r,s) simply connected. Also, this situation is the one where most examples interesting to physics are, when r ≥ 3 and s = 0 or 1, or vice-versa. The definition of the spin structure gives the following commutative diagram:

Since restriction to fibers gives the covering map ξ 0, this diagram can be extended [1,2]:

In the above figure, we see that the vertical lines show the inclusions of fibers in the fiber bundle. Now we can consider whether given a two-sheeted covering φ: C 2 (E) → P SO(E) of P SO(E) gives a spin structure. It certainly gives a fiber bundle over E since we can set π′ = π ξ.

Now we see that this bundle gives a spin structure if the covering is nontrivial on the fibers.

Theorem 7

If π 1(SO(r, s)) = ℤ2, then the spin structures are in one-to-one correspondence with two-sheeted coverings of P SO(E), which are nontrivial on the fibers [1,2].

Now consider a spin structure ξ: P Spin → P SO. Define α F π 1(SO (r,s)), the nontrivial element. The spin structure ξ induces a group homomorphism ξ : π 1(P) → π 1(Q). This subgroup of π 1(Q) is a subgroup of index 2 because of covering morphism, and P Spin is a double covering of P SO fiber wise.

Lemma 8

α F ξ ( π 1 ( P Spin ) )

Proof

Suppose α F ξ (π 1(P Spin)). Then, the inclusion map i: SO → P SO lifts to a continuous map.

I: SO → P Spin such that commutes. I (SO) ⊂ P Spin is contained in one fiber Spin of P Spin, so I: SO(r, s) → Spin(r,s) with ξ 0 ◦ I = IdSO(r, s) = Id 2 . Then, ξ 0∗ I = Idπ 1(SO(r, s)) and π 1(SO(r, s)) = ℤ2 and π 1(Spin(r, s)) = 1, so we get a contradiction, so α F ξ (π 1(P Spin)) [6].

This lemma is then used to prove a classification theorem of spin structures, following and using the classification of covering spaces.□

Theorem 9

A SO(r, s)-principal fiber bundle P SO over a manifold M has a spin structure if and only if there is a short split exact sequence.

0 Z 2 i π 1 ( P SO ) π π 1 ( M ) 0 ,

meaning that π 1(P SO) is isomorphic to K × ℤ2 and π , the map induced by projection map, maps K isomorphically to π 1(M).

To prove this theorem, we need a few lemmas. First, we prove the necessity of the conditions, we assume P SO has a spin structure P Spin.

Lemma 10

The map π: π 1(P Spin) → π 1(M) is an isomorphism.

Proof

First, we prove surjectivity. Take an element [g] ∈ π 1(M) represented by a loop g: [0, 1] → M based at m 0M. We can then look at a path : [0, 1] → P Spin such that πg̃(t) = g(t). This path exists because P Spin is a fiber bundle over M. This path does not have to be a loop. However, π((0)) = π((1)) = m 0 and the fibers of P Spin are path connected, so we can find a path δ in spin such that δ(0) = 1 and δ(1)(1) = g(0). The product path δg̃ is now a loop and π(δg̃) = g. This proves surjectivity.

As for injectivity, consider a loop : [0, 1] → P Spin and assume π()(t) = h(t) is homotopic to trivial loop, say at the point x 0 = h(0). There is then a homotopy from the loop h to the point x 0. This homotopy can be covered by a homotopy of into a new loop lying in the fiber π 1(x 0), or its generalization to fiber bundles over paracompact spaces [7]. Since spin is simply connected, there is then a homotopy to a single point in this fiber; hence, is homotopic to the trivial loop, so π is injective.□

Lemma 11

The map π : π 1(P SO) → π 1(M) restricted to the image of ξ (π 1(P Spin)) is an isomorphism.

Proof

Call the image of ξ K. We have the following commutative diagram:

The map ξ is injective, and according to the covering space, morphism is injective. We also know that π′ is an isomorphism, so π |k is injective. Furthermore, π |k◦ξ = π′ is an isomorphism, so π |k must be surjective, hence an isomorphism [1].□

Lemma 12

The map i : π 1(SO) = ℤ2π 1(P SO) is injective, so it is an isomorphism onto its image.

Proof

Suppose [g] ∈ π 1(SO) and i [g] = [α] ∈ π 1(P SO). If [α] is trivial, then we can lift α to a path α ˆ in P Spin. The loop α lies within a single fiber, so the loop α ˆ also does. Because Spin is simply connected, there is a homotopy between α ˆ and the trivial loop within the fiber. Applying the covering map to of P SO, SO, so [g] is trivial and i is injective.□

Lemma 13

The sequence of Lemma 11 is exact, in particular: ker(π ) = img(i ) [1,2].

Proof

Take a [g] ∈ π 1(SO). Then, i [g] has a representative lying within a single fiber. Then, π i [g] = [e] and hence, im(i ) ⊆ ker(π ). Take a loop α ⊆ P SO. We will construct a loop α ˆ i(P Spin) such that α = α ˆ g with g a loop in SO. First, let α′ = π α be a loop in M. Then, we lift this loop one in P Spin by lifting and multiplying with a path δ with δ(0) = 1 and α′(1)δ(1) = α(0). Now define α ˆ = ξ(αδ). We can now conclude that [ α ˆ ] = ξ π ' 1 π ( [ α ] ) and that π(α) = π( α ˆ ). Thus, we have that α = α ˆ g for some loop g ⊆ SO; hence, [α] = [ α ˆ g] = [ α ˆ ]i [g], so if [α] ∈ ker(π) we have that because π is an isomorphism on the image of ξ, that [α] = [ α ˆ ]i [g] = [e]i [g], so ker(π ) ⊆ im(i ) [1].□

If the fundamental group of SO(r, s) is not ℤ2, this result can be generalized, but the double covering space of SO(r, s) will not be a universal covering space.

Instead, one must look at either the universal covering space of SO(r, s), which is then not equal to Spin(r,s), or one must look at not simply connected Spin(r,s) [1,6].

The condition of Lemma 11 for the existence of a spin structure can be shown to be equivalent with the usual condition that the second Stiefel-Whitney class w 2 of M vanishes [1,6,8]. For the special case of SO0(1, 3)-principal fiber bundles over a noncompact 4-manifold M, the case in general relativity, it has been shown that any existing spin structures are trivial, M × Spin(1, 3) and so it has a spin structure, if and only if the SO0(1, 3)-bundle is parallelizable, meaning that there is a global section of the SO0(1,3) bundle [1].

Definition 7

We call global sections of P SO(1,3)(M) Lorentz frames and global sections of P Spin(1,3)(M) spin frames [5].

Remark 2

When we will use the concept of the spin structure in physics, the space–time is fourdimensional manifold with a metric of signature (3, 1) or (1, 3), we will naturally use SL(2, ℂ) as group instead of Spin(p, q).

The isomorphism SL(2, ℂ) ≃ Spin(1, 3) [9]. Here, the fundamental interactions of spinors, Dirac, Weyl spinors (half-spinors), Majorana spinors, and bi-quaternions are used to describe the most fundamental particles in same dimension.

4 Clifford algebras and spinor bundles

There are two equivalent ways of defining a Clifford bundle of a pseudo-Riemannian vector bundle E over a manifold X. One is the obvious generalization of Cℓ( n ):

C ( E ) = x X C ( E x , q x ) .

where q is a smooth quadratic form on E and q x is the restriction of that form to the fiber over x. This definition emphasizes that the Clifford bundle is a bundle of Clifford algebra’s over X. The other definition uses associated bundles and can be used to determine the topology, as follows: An orthogonal transformation with respect to an inner product of signature (r, s) in (r+s) induces an orthogonal transformation in Cℓ r,s , since it preserves the ideal. This induced map preserves the multiplication in Cℓ r,s , so if we take an orthogonal transformation ρ r,s from SO(r, s), we get a map cl(ρ r,s ): SO r,s → Aut(Cℓ r,s ). The bundle associated to this bundle is called the Clifford bundle:

C ( E ) = P SO ( E ) × c l ( ρ r , s ) C r , s

So Cℓ(E) as the quotient bundle:

C ( E ) = r = 0 r E / I ( E ) ,

where I(E) is the bundle of ideals [2].

It is also clear that all fundamental concepts on Clifford algebras carries over to Clifford bundles. For example, Cℓ(E) = Cℓ0(E) ⊕ Cℓ1(E) corresponding to the even–odd decomposition of the algebras.

These are the +1 and −1 eigen bundles of the bundle automorphism

A : C ( E ) C ( E ) [10] .

These two definitions are the same since the fiber at xE of P SO (E) × cl(ρ r,s ) Cℓ r,s is Cℓ r,s [10]. All notions familiar from Clifford algebras over real vector spaces carry over to Clifford bundles over manifolds. If X is a pseudo-Riemannian manifold, we can construct the Clifford bundle Cℓ(TX) associated with the pseudo-Riemannian form on the tangent bundle TX. We will also call this bundle Cℓ(X), in case there is no confusion possible [10].

Definition 8

A smooth manifold endowed with a spin structure will be called a spin manifold [5].

Definition 9

(Spinor bundles). Let X be a smooth manifold with a spin structure ξ: P Spin(X) → P SO(X). A real or complex spinor bundle of X is a bundle of the form [13]

S ( X ) = P Spin ( X ) × μ L , P Spin ( X ) × μ L c ,

where L is a left module of Cℓ r,s and µ: Spin(r, s) → End (L) is left-multiplication of elements in Spin(r, s), and where L is a complex left module for Cℓ(ℝ n ) ⊗ ℂ and µ: Spin(r, s) → End (L ) is the left-multiplication of elements in Spin(r, s). When (r, s) = (1, 3) defined as certain equivalence classes in appropriate sets, and a first definition for field of these object on living Minkowski space time was given [5].

Remark 3

The complexified left spin Clifford bundle denoted by

C Spin 1 , 3 l ( X ) = P Spin 1 , 3 ( X ) × l 1 , 3 P Spin 1 , 3 ( X ) × l 4 , 1

and the complexified right Clifford bundle by

C Spin 1 , 3 r ( X ) = P Spin 1 , 3 ( X ) × r 1 , 3 P Spin 1 , 3 ( X ) × r 4 , 1 [ 5 ] .

Remark 4

Taking, e.g., L = ℂ4 and µ the D (1\2,0)   D (O,1\2) of Spin1,3 ≌ SL(2, ℂ) in End(ℂ4) and L is covariant spinor bundle [5] (where D is Dirac operator see the last sections).

If the module L(or L ) is ℤ2 graded, the corresponding bundle is said to be ℤ2 graded.

Example 1

Consider Cℓ(ℝ n ) as a module over itself by left multiplication l. The corresponding real spinor bundle CℓSpin(X) = P Spin(X) × l Cℓ(ℝ n ), then:

  1. CℓSpin(X) is a “principal Cℓ(ℝ n )-bundle,” i.e., it admits a free action of Cℓ(ℝ n ) on the right.

  2. There is a natural embedding P Spin(X) ⊂ CℓSpin(X), which comes from the embedding Spin n ⊂ Cℓ(ℝ n ). Hence, every real spinor bundle for X can be captured from this one [2].

A similar remark holds for the complex case.

Of course, the bundle CℓSpin(X) differs from the Clifford bundle Cℓ(X). They can be compared as follows.

Ad : Spin n Aut ( C ( n ) )

given by Ad g (ϕ) = gφg −l for g ∈ Spin n ⊂ Cℓ(ℝ n ). Clearly Ad−1 = identity, and this representation come to acting Ad′ of SO n . One easily checks that Ad′ is just the representation Cℓ( ρ n ) given by

C ( X ) = P Spin ( X ) × Ad C ( n ) .

Two spinor bundles of X are equivalent if they are equivalent as bundles of Cℓ(X)-modules.

Real bundle, complex bundle, graded, and ungraded bundle of Cℓ(X) modules is called irreducible if at each x fiber is irreducible as a (real or complex, graded or ungraded) module over Cℓ(X x ) [2].

Then, when X is a spin manifold;

  1. the components of Cℓ (X) = P Spin(X) ×Ad Cℓ(ℝ n ) are homomorphism classes [(β, α)] of pairs (β, α), where βP Spin(X), α ∈ Cℓ(ℝ n ) and (β, α) ∼ (β′, α′) ⇔ β′ = βu −1, α′ = uαu −1, for some u ∈ Spin.

  2. the components of C l Spin(X) are homomorphism classes of pairs (β, α) where βP Spin(X), α C ( n ) and ( β , α ) ( β , α ) β = β u 1 , α = α u 1 , for some u Spin .

  3. the components of C r Spin(X) are homomorphism classes of pairs (β, α), where βP Spin(X), α C ( n ) and ( β , α ) ( β ' , α ' ) β ' = β u 1 , α ' = α u 1 , for some u Spin .

Proposition 14

There is a natural pairing secCℓ l Spin(X) × secCℓ r Spin(X) = secCℓ(X).

Proof

Given δ ∈ secCℓ l Spin(X), ρ ∈ secCℓ r Spin(X) select representatives (β, α) for δ(x) and (β, b) for ρ(x), (with β ∈ π 1(x)) and define (δρ)(x) ≔ [(β, αb)] ∈ Cℓ (X). If alternative representatives (βu −1 , uα) and (βu −1 , bu −1 ) are chosen for δ(x) and ρ(x), we have (βu −1 , uαbu −1 ) ∼ (β, αb) and thus, (δρ)(x) is well define component of Cℓ(X) [5].

Let us now say a word about the ℤ2-graded case. There is irreducible bundle of ℤ2-graded modules over Cℓ(X) = Cℓ0(X) Cℓ1(X) and classes irreducible bundle of modules over Cℓ0(X). Given a bundle S(X) = S 0(X) S 1(X) of the first kind, S 0(X) is of the second. Given an S 0(X) of the second kind, the bundle

S ( X ) = C ( X ) ) C 0 ( X ) S O ( X ) is of the first .

Suppose now that n = 2 m and S (X) is the irreducible complex spinor bundle of X. We will show clearly how to split S (X) into a direct sum:

S ( X ) = S + ( X ) S ( X )

of Cℓ0(X) modules. Interpreting S + ( X ) as S 0 ( X ) and S ( X ) as S 1 ( X ) , or the other way around, gives a ℤ2-graded module structure to S ( X ) .

There is a similar construction in the real case.

S ( X ) = S + ( X ) S ( X ) [ 2 ] .

Recall that every module for Cℓ(ℝ n ) is a direct sum of irreducible ones, and there are at most two homomorphism classes of irreducible modules [2].□

Proposition 15

If S(X) is a real spinor bundle of X, then S(X) is a bundle of modules over the bundle of algebras Cℓ(X). In particular, the sections of the spinor bundle are a module over the sections of the Clifford bundle [5].

Proof

See [2].□

Remark 5

The corresponding fact holds in the complex and ℤ2-graded cases and Sections of S(X) are called spinors.

Definition 10

The dual spinor bundle S (X) is a real or a complex spinor bundle

S ( X ) = P Spin ( X ) × μ L , P Spin ( X ) × μ L c ,

where L is a right module of Cℓ r,s and µ: Spin(r, s) → End(L) is the representation given by right-multiplication of (inverse) elements in Spin(r, s), and where L is a complex right module for Cℓ(ℝ n ) ⊗ ℂ and µ: Spin(r, s) → End(L ) is the representation given by right-multiplication of inverse elements in Spin(r, s) [5].

Definition 11

If U is a normal bundle of the Grassmannian. Then, if s U = S (X) the spinor bundle on Q 2k+1. Its rank is 2 k . We call s U = S ( X ) and s U S ( X ) the two spinor bundles on Q 2k . Their rank is 2 k−1.

If f is an automorphism of Q 2k that exchanges the two families of k-planes, we have

f S ( X ) S ( X ) and f S ( X ) S ( X ) .

Clear that S (X) (spinor bundles) on all quadrics Q are homogeneous, i.e., f S (X) S (X) for all f ∈ Aut(Q)0, where Aut(Q)0 is the connected component of the identity in Aut(Q).

Theorem 16

  1. Let S ( X ) , S(X) are spinor bundles on Q 2k , and let i: Q 2k-1Q 2k be a smooth hyper plane section. Then i S ( X ) i S ( X ) S (X), where S (X) is the spinor bundle on Q 2k−1.

  2. Let S (X) is spinor bundle on Q 2k+i , and let i: Q 2k Q 2k+i be a smooth hyperplane section.

Then, i S (X) S ( X ) S ( X ) , where S ( X ) and S ( X ) are the spinor bundles on Q 2k [11].

Example 2

Two embeddings s′: Q4 → Gr(l, 3) and S″: Q 4 → Gr(l, 3) are isomorphisms from the definition of spinor bundles. So S (X) on Q 4 are the normal bundle and the dual of the quotient bundle.

The embedding S: Q 3 → Gr(l, 3) corresponds to a hyperplane section. If S (X) is the spinor bundle on Q3, then S 2 (X) S (X) = TQ3. In fact, TQ4/Q3S (X)S (X)S 2 (X) S (X) ⊕ (1) and the exact sequence splits

0 T Q 3 T Q 4 / Q 3 ( 1 ) 0 .

On Q 2 the two of S (X) are the duals of the two line bundles corresponding to two skew-lines on Q 2. On Q 1 ≃ ℙ1, then the spinor bundle can defined to be 1 (−1) [11].

Corollary 3

Let lQ n (n > 3) be a line and let S (X) on Q n [11]. Then

S X | l = O l 2 ( n 3 ) 2 O l ( 1 ) 2 ( n 3 ) 2 .

5 Relations between spinor bundles and spinor structures

For every xM, the spinor representation τ x : C ( g x ) End Σ x defines the real line a(τ x ) and the circle c(τ x ) [12]. The set

α ( τ ) = x M α ( τ x ) Hom ( Σ , Σ ¯ ) = Σ Σ ¯

has the structure of a real line bundle over M. The set

c ( τ ) = x M c ( τ x ) Hom ( Σ , Σ ¯ ) = Σ Σ ¯

has the structure of a bundle of circles over M: It is a principal U(1)-bundle. If this bundle is trivial, i.e., if it has a (global) section C: Mc(τ), then the real line bundle over M defined as follows:

b ( τ , C ) = x M b ( τ x , C ( x ) ) Hom ( Σ , Σ ) = Σ Σ .

Proposition 17

Let (M, g) be an smooth Riemannian manifold with (V, h) as the local model.

  1. There corresponds a Clifford c structure to every spinor bundle such that the associated spinor bundle is isomorphic to τ: Cℓ(g) → End Σ [3].

  2. The Clifford c structure can be reduced to a spin structure iff the line bundle above is trivial.

  3. The Clifford c structure can be reduced to a Clifford structure iff the bundle of circles above is trivial.

  4. The resulting Clifford structure can be reduced to a spin structure iff the real line bundle above is trivial [13,14].

Proof

We already know that Hom C M ( S , S ¯ ) is a complex line bundle: each fiber contains a Cℓ(T p M )-linear isomorphism τ p : S p p and by irreducibility of S p and p and Schur’s Lemma any Cℓ(T p M )-linear map τ p : p S p satisfies τ p τ p = λ/S p for some λ . In case of a spin structure τ defines a non-vanishing section in Hom C M ( S , S ¯ ) , hence

Hom C M ( S , S ¯ ) is trivial. Conversely, if this bundle is trivial and if τ̃′ is a nonvanishing section, then τ̃2 = λ/ S for some nonvanishing map λ (M, ) . But λ(p) τ̃ p (v) = τ̃ p τ̃ p τ̃ p (v) = τ ˜ ( λ ( p ) v ) = λ ( p ) ¯ τ ˜ ( v ) , i.e., λ (M) is real valued, and replacing τ̃ by τ = |λ|−1/2 τ̃, we obtain a structural map.

Now given an irreducible complex spinor bundle S and a structural map τ, we can choose a Riemannian structure compatible with Clifford multiplication and such that τ is an isometry.□

Definition 12

Suppose E is a smooth Riemannian vector bundle over a manifold X and that ξ: P Spin(E) → P SO(E) is a spin structure on E. Then, of course, any connection on P SO(E) can be lifted via ξ to a connection on P Spin(E), and this, in turn, defines a connection on the associated spinor bundles [2].

We give two equivalent definitions of a connection on a principal G-bundle with projection P π X.

First, we define the vertical tangent space T p v P at p as the subspace of T p P , which is tangent to the fiber of the projection P π X. A connection on P π X is a smoothly varying family of linear subspaces ( T p H ) pP of the tangent bundle T P, which is everywhere complementary to the vertical distribution ( T p v V ) pP and which is invariant under the action of the group G. The distribution ( T p H ) pP is called the horizontal distribution. Note that

T P = T V P T H P [ 12 ] .

Second, to give a connection on P π X is to give a connection 1-form ω on P with values in g satisfying two conditions.

It transforms by the adjoint action, i.e., for any ɡ in G, p in P and any γ in T p P, we have

ω p g ( γ g ) = g 1 ω p ( γ ) g .

For any a in g, the associated vector field Va on P defined by the tangent vector of the curve pe ta at any pP. It holds that ω(V a ) = a.

Note that the horizontal distribution can be recaptured by taking the kernel of the connection 1-form assigned to it.

Lemma 18

Given a Euclidean connection on a real vector bundle E, there is a canonical orthogonal connection (i.e., the decomposition T P = T V P T H P is Euclidean) on its orthonormal frame bundle P SO ( n ) X . Reversely, any orthogonal connection on the frame bundle P S O ( n ) X induces a Euclidean connection on E. Furthermore, these operations are inverse to each other [4].

Lemma 19

Given a Lie group G and a connection on a principal G-bundle P π X, there is an induced connection on any vector bundle P × G V coming from a linear representation GGL(V) [4].

The curvature g-valued 2-form Ω is

Ω = d ω + [ ω , ω ] .

Example 3

(Orthogonal connections). Let P = P SO(E), where E is a smooth, oriented Riemannian vector bundle. The Lie algebra SO n of real, skew-symmetric n × n-matrices. Hence, a connection 1-form ω can be considered as an n × n-matrix of 1-forms ω = (ω ij ), where ω i j = ω j i .

The corresponding curvature is a matrix of 2-forms Ω = Ω ij, where

Ω i j = d ω i j + k = 1 n ω i k Λ ω k j .

Suppose that μ = (e 1,…, e n ) is just a section of P SO(E) over U X , and it can be lifted to a section μ ˜ of P Spin(E) over U. There are two possible such liftings. They satisfy the relation:

ξ ە μ ˜ = μ

The connection 1-form on P Spin(E) is just the lift ξ ω (the pull down) of the connection 1-form ω on P SO(E) [2].

Theorem 20

Let ω be the connection i-form on P SO(E) and let S(E) be any spinor bundle associated to E. Then, the covariant derivative s on S(E) is given locally by the formula [2]:

s = 1 2 i < j ω ˜ i j e i e j .

where μ = (e 1,…, e n ) is a local section of P SO(E), ω = μ (ω), and where E = (σ 1,…, σ n ) is a local section of P SO(S(E)) determined by ω.

Theorem 21

Let Ω be the curvature 2-form on P SO(E) and let S(E) be any spinor bundle associated to E. Then, the curvature by the formula S of S(E) is locally given by

R s = 1 2 i < j Ω ˜ i j e i e j ,

where μ = (e 1,…, en ) is a section of P SO(E), Ω ˜ = μ ( Ω ) and σ is any section of S(E) [2].

Then, any two tangent vectors V and W at x X , the curvature transformation V , W s : S(E x ) → S(E x ) is

R V , W s = 1 2 i < j R V , W ( e i , e j ) e i e j ,

where v , w is the curvature transformation of E x .

Dirac operator D acting on sections of spinor bundle (S (X)) Σ → M is globally defined as follows.

Definition 13

(Dirac operator). Let U ι be an open subset of M and let e = (e µ ) µ=1,…,m be a field of (not necessarily orthonormal) frames on U ι . For every pU ι , the components of the metric tensor g with respect to e at p are g µν (p) = g(e µ (p), e ν (p)) and there is the inverse g µν (p) of g µν (p). The restriction of the Dirac operator to U ι is [2] expressed as follows:

D = g μ ν τ ( e μ ) e ν .

The Dirac operator on M well defined by its restrictions to the sets (U) providing an open cover of M.

TM is connection on the spinor bundle to be metric, but may have torsion.

Definition 14

A Dirac bundle is a bundle S over a Riemannian manifold X of left modules over Cℓ(X) together with a Riemannian metric and connection on S with properties:

e 1 σ 1 , e 2 σ 2 = σ 1 , σ 2 and ( φ σ ) = ( φ ) σ + φ ( σ ) for all φ Γ ( C ( X )   ) , σ Γ ( S ) .

The operator D is elliptic if the linear map σ ξ (D): E x → E x is an isomorphism for all ξ ≠ 0.

Lemma 22

If D is the Dirac operator of the bundle S defined above. Then, for any ξ T ( X ) T ( X ) , we have that

σ ξ ( D ) = i ξ ,

σ ξ ( D 2 ) = ξ 2 ,

where the symbol on the right denotes Clifford multiplication by the vector ξ and the scalar ξ 2 . In particular, both D and D 2 are elliptic operators.

Proof

See [2].□

Theorem 23

Let X be a complete Riemannian manifold and if D is the Dirac operator of Dirac bundle S over X. Then, the closure of D in L 2(S) is a self-adjoint operator. Furthermore, ker(D) = ker(D 2) on L 2(S), where L is a canonical bundle map L: Cℓ(X) → Cℓ(X) defined by L ( φ ) = e j φ e j [2].

Lemma 24

Let D, S and X be as above. Then for any f C ( X ) and any φ Γ ( S ) C ( X ) , we have that

D ( f φ ) = ( grad f ) φ + f D φ .

Proof

D() = e j e j ( f φ )   = e j { ( e j f ) φ   +   f e j φ } = ( ( e j f )   e j φ )   +   f D φ )   = (grad f) ⋅ φ + fDφ [2].

Having discussed Dirac bundles in general terms, it is now time to look hard at some important examples. We begin with the basic ones.□

Example 4

(An historical case). Let X = R n , Euclidean n-space, and let S = R n × V, where V some module for Cℓ n . Here, D is a constant on V-valued functions [2]

D = k = 1 n γ k x k

and each γ k is a linear map γ k :V → V and γ j γ k + γ k γ j = 2 δ j k .

This particular operator has historical roots in physics. In the 1920s [2], the physicist P.A.M. Dirac was searching for a Lorentz-invariant first-order differential operator whose square would be the Klein-Gordon operator. Thus, he was essentially led to search for a first-order operator D of the form above, which satisfied the equation D 2 = ∆, where ∆ = −∑∂2/∂x2 is the positive Laplacian in ℝ n . Realizing that the γ k s must be matrices, he was led immediately by this equation to the aforementioned relations, which we recognize now as the generating relations of a representation of Cℓ n .

Let n = 1, so that C 1 = V = 2 . Then we have [3,8]

D = i x 1

the generator of a basic semi-group of unitary operators on L 2.

Let n = 2, so that C 2 = V = . The construction of into is natural and corresponds to the ℤ2-grading C 2 0 C 2 1 [3]

D = e 1 x 1 + e 2 x 2

has the form D = 0 z z ¯ 0 , where / z ¯ = / x 1 + i / x 2 .

Let n = 3, so that C 3 = and V = ℍ. Cℓ3 has two representations on ℍ given as Identify ℝ3 with Im(ℍ), by letting i, j, and k act on either the right or the left in ℍ. On the left, we get Dirac operator on ℍ:

D = i x 1 + j x 2 + k x 3 .

Let n = 4, so that Cℓ4 = ℍ(2) and V = 2 = . To describe the full Dirac operator, we consider first the following ℍ under the basis (1, i, j, k):

q ¯ = x 0 + i x 1 + j x 2 + k x 3 , q = x 0 + i x 1 + j x 2 + k x 3 ,

then D = 0 q q ¯ 0 .

Thus, left multiplication be represented by complex 2 × 2-matrices σ 0, σ 2 and σ 3 respectively, then the operator / q ¯ becomes

q ¯ = x 0 + σ 1 x 1 + σ 2 x 2 + σ 3 x 3 .

The matrices σ k can be chosen to be the classical Pauli matrices:

σ 1 = i 0 0 i , σ 2 = 0 1 1 0 , σ 3 = 0 i i 0 .

Note that these matrices generate the fundamental representation of Cℓ3 in complex form [2,3].

Example 5

(The Clifford bundle). Let S = Cℓ(X) with its canonical Riemannian connection, and view Cℓ(X) as a bundle of left modules over itself by left Clifford multiplication. The Dirac operator D in this case is a square root of the classical Hodge Laplacian [2].

Example 6

(The spinor bundles). Suppose X is a spin manifold with a spin structure on its tangent bundle. Let S (X) be any spinor bundle associated with T(X). Then, S (X) is a bundle of modules over Cℓ(X), and S (X) carries a canonical Riemannian connection, which has property of Definition 14. The Dirac operator in this case was first written down by Atiyah and Singer in their work on the Index theorem. Finding this operator was a major accomplishment, and for this reason, we shall call it the Atiyah-Singer operator [2,15].

Notation. For spin manifolds X of even dimension, we shall denote the (unique) irreducible complex spinor bundle by $ℂ; and when dim(X) ≢ 3(mod 4), we denote the irreducible real spinor bundle by $.

In both cases, the Atiyah-Singer operator will be written as .

These basic examples each generate large families of new examples by the following construction.

Let S and E be a given Dirac bundle with connection s and E over a Riemannian manifold X. Then, the tensor product S E is again a bundle of left modules over Cℓ(X), where for φ Cℓ(X), σ S, e E, φ . ( σ e ) = ( φ . σ ) e [2].

Furthermore, we can equip S E with the canonical tensor product connection, = s E , which is defined on sections of the form σ e by the formula:

( σ e ) = ( s σ ) e + σ ( E e ) .

6 Conclusion

The researches have explained how Riemannian geometry, with the theory of spinor fiber bundle, fits into the general meaning of principal fiber bundles. We see that the spinor bundles defined as vector bundles whose fibers carry spinor representations of the Clifford algebras Cℓ(g), spinor fields are sections of spinor bundles.

With this constructive classification tool, we have investigated the usefulness of spinor structures of Clifford algebra. We have also given an explicit description of an important physical applications of spinor bundle and used a spinor connection and the constant Dirac matrices of special relativity to define the Dirac operator using the Clifford algebra of space–time and Dirac operation.

Acknowledgments

The author would like to thank the Deanship of Scientific Research, Qassim University for funding publication of this project.

  1. Conflict of interest: The author states no conflict of interest.

References

[1] J. J. Venselaar , Spinors and Dirac Operators, Caltech, California, USA, 2007.Search in Google Scholar

[2] H. B. Lawson and M.-L. Michelsohn , Spin Geometry, Princeton University, New Jersey, USA, 1989.Search in Google Scholar

[3] H. Schröder, On the definition of geometric Dirac operators, https://arxiv.org/abs/math/0005239, (2000).Search in Google Scholar

[4] J. M. Bismut , Introduction to the Index Theory, University of Paris, France, 1986.Search in Google Scholar

[5] A. R. Mosona and A. W. Rodrigues , The Fundle of Algebraic and Dirac–Hestenes Spinor Field, 13083-970 Campinas, SP Brazil, 2004.Search in Google Scholar

[6] S. Morrison , Classifying spinor structures, Master’s thesis, University of New South Wales, https://arxiv.org/abs/math-ph/0106007, (2001).Search in Google Scholar

[7] N. Steenrod , The Topology of Fibre Bundles, Princeton University, USA, 1999.Search in Google Scholar

[8] K. Wernli , Lecture notes on spin geometry, arXiv:1911.09766v1, (2019).Search in Google Scholar

[9] L. Claessens , Field theory from a bundle point of view, Belgium, 2011, https://www.yumpu.com/en/document/view/33648732/field-theory-from-a-bundle-point-of-view.Search in Google Scholar

[10] J. L. Koszul , Lectures on Fibre Bundles and Differential Geometry, Bombay, 1960.Search in Google Scholar

[11] G. Ottaviani , Spinor bundles on quadrics, Trans. Amer. Math. Soc. 307 (1988), no. 1, 301–316.10.1090/S0002-9947-1988-0936818-5Search in Google Scholar

[12] A. Trautman , Connections and the Dirac operator on spinor bundles, J. Geom. Phys. 58 (2007), 238–252.10.1016/j.geomphys.2007.11.001Search in Google Scholar

[13] C. Kassel , Homology and Cohomology of Associative Algebras, Notes in the Advanced School on Non-commutative Geometry at ICTP, Trieste, Italy, 2004.Search in Google Scholar

[14] T. Friederish and A. Trautman , Clifford Structures and Spinor Bundles, Berlin, Germany and Warsaw, Poland, 1997.Search in Google Scholar

[15] M. F. Atiyah , R. Bott , and A. Shapiro , Clifford modules, Topology 3 (1964), 3–38.10.1007/978-1-4612-5367-9_1Search in Google Scholar

Received: 2021-03-11
Revised: 2021-05-18
Accepted: 2021-06-06
Published Online: 2021-12-03

© 2021 Yousif Atyeib Ibrahim Hassan, published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Regular Articles
  2. Graded I-second submodules
  3. Corrigendum to the paper “Equivalence of the existence of best proximity points and best proximity pairs for cyclic and noncyclic nonexpansive mappings”
  4. Solving two-dimensional nonlinear fuzzy Volterra integral equations by homotopy analysis method
  5. Chandrasekhar quadratic and cubic integral equations via Volterra-Stieltjes quadratic integral equation
  6. On q-analogue of Janowski-type starlike functions with respect to symmetric points
  7. Inertial shrinking projection algorithm with self-adaptive step size for split generalized equilibrium and fixed point problems for a countable family of nonexpansive multivalued mappings
  8. On new stability results for composite functional equations in quasi-β-normed spaces
  9. Sampling and interpolation of cumulative distribution functions of Cantor sets in [0, 1]
  10. Meromorphic solutions of the (2 + 1)- and the (3 + 1)-dimensional BLMP equations and the (2 + 1)-dimensional KMN equation
  11. On the equivalence between weak BMO and the space of derivatives of the Zygmund class
  12. On some fixed point theorems for multivalued F-contractions in partial metric spaces
  13. On graded Jgr-classical 2-absorbing submodules of graded modules over graded commutative rings
  14. On almost e-ℐ-continuous functions
  15. Analytical properties of the two-variables Jacobi matrix polynomials with applications
  16. New soft separation axioms and fixed soft points with respect to total belong and total non-belong relations
  17. Pythagorean harmonic summability of Fourier series
  18. More on μ-semi-Lindelöf sets in μ-spaces
  19. Range-Kernel orthogonality and elementary operators on certain Banach spaces
  20. A Cauchy-type generalization of Flett's theorem
  21. A self-adaptive Tseng extragradient method for solving monotone variational inequality and fixed point problems in Banach spaces
  22. Robust numerical method for singularly perturbed differential equations with large delay
  23. Special Issue on Equilibrium Problems: Fixed-Point and Best Proximity-Point Approaches
  24. Strong convergence inertial projection algorithm with self-adaptive step size rule for pseudomonotone variational inequalities in Hilbert spaces
  25. Two strongly convergent self-adaptive iterative schemes for solving pseudo-monotone equilibrium problems with applications
  26. Some aspects of generalized Zbăganu and James constant in Banach spaces
  27. An iterative approximation of common solutions of split generalized vector mixed equilibrium problem and some certain optimization problems
  28. Generalized split null point of sum of monotone operators in Hilbert spaces
  29. Comparison of modified ADM and classical finite difference method for some third-order and fifth-order KdV equations
  30. Solving system of linear equations via bicomplex valued metric space
  31. Special Issue on Computational and Theoretical Studies of free Boundary Problems and their Applications
  32. Dynamical study of Lyapunov exponents for Hide’s coupled dynamo model
  33. A statistical study of COVID-19 pandemic in Egypt
  34. Global existence and dynamic structure of solutions for damped wave equation involving the fractional Laplacian
  35. New class of operators where the distance between the identity operator and the generalized Jordan ∗-derivation range is maximal
  36. Some results on generalized finite operators and range kernel orthogonality in Hilbert spaces
  37. Structures of spinors fiber bundles with special relativity of Dirac operator using the Clifford algebra
  38. A new iteration method for the solution of third-order BVP via Green's function
  39. Numerical treatment of the generalized time-fractional Huxley-Burgers’ equation and its stability examination
  40. L -error estimates of a finite element method for Hamilton-Jacobi-Bellman equations with nonlinear source terms with mixed boundary condition
  41. On shrinkage estimators improving the positive part of James-Stein estimator
  42. A revised model for the effect of nanoparticle mass flux on the thermal instability of a nanofluid layer
  43. On convergence of explicit finite volume scheme for one-dimensional three-component two-phase flow model in porous media
  44. An adjusted Grubbs' and generalized extreme studentized deviation
  45. Existence and uniqueness of the weak solution for Keller-Segel model coupled with Boussinesq equations
  46. Special Issue on Advanced Numerical Methods and Algorithms in Computational Physics
  47. Stability analysis of fractional order SEIR model for malaria disease in Khyber Pakhtunkhwa
Downloaded on 10.1.2026 from https://www.degruyterbrill.com/document/doi/10.1515/dema-2021-0035/html
Scroll to top button