• Keine Ergebnisse gefunden

Spinors and the Lorentz group

The SL(2,C) group is the universal cover of the group L+ of proper orthochronous Lorentz transformations. We can best understand the intertwining mapΛrelating the first with the second when studying the vector space of anti-Hermitian*2×2matrices.

*That we do not work here with Hermitian matrices but rather use anti-Hermitian elements of C2C¯2is related to the choice of signature(−,+,+,+)that we have agreed on earlier.

Given such a matrix we denote its row indices by roman capitalsA, B, C,· · · ∈ {0,1}, while indices referring to the column should carry a macron, i.e. an “overbar” such that we write A,¯ B,¯ C,¯ · · · ∈ {¯0,¯1}. The unmarked indices A, B, C refer to C2 and we call them left-handed, while their brothers A,¯ B,¯ C, . . .¯ belong to the complex conjugate vector space C¯2, and we call them right-handed or of opposite chirality. The logic behind this terminology should become clear in a moment. Using this index-notation we can express the anti-Hermiticity of a matrixX by saying:

XAB¯ is anti-Hermitian⇔XAB¯ ≡X¯AB¯ =−XBA¯. (A.15) The soldering matrices

σ0 = 1 0

0 1

, σ1 = 0 1

1 0

, σ2=

0 −i i 0

, σ3=

1 0 0 −1

, (A.16) form a basis in the vector space of all Hermitian 2×2 matrices. We introduce the component vectorXµ∈R4 of XAA¯ with respect to this basis and set:

XAA¯= i

√2σAA¯αXα. (A.17)

We can write this more abstractly by equating the component-vector with the matrix itself. i.e.:

XAA¯=Xα. (A.18)

The isomorphism (A.17) allows us to identify Minkowski indices with pairs of spinorial idices, for a generic world tensorTαβ... we can thus write:

Tαβ... =TAAB¯ B...¯ (A.19)

While the individual ordering of ordinary (right-handed) indicesA, B, C, . . . and their left-handed sistersA,¯ B,¯ C¯ matters, the relative ordering has no significance. With our choice of signature it turns out to be useful to declare the exchange of two adjacent indices of opposite chirality to result in an overall minus sign, e.g.:

XAA¯ =−XAA¯ . (A.20)

A short moment of reflection reveals now that the determinant of the matrix (A.17) turns into the Minkowski inner product of its components, for anyXµ∈R4:

2 det(X) =−(X0)2+ (X1)2+ (X2)2+ (X3)2αβXαXβ. (A.21) Let us now understand how Lorentz transformations can act on these spinorial in-dices. For any elementg∈SL(2,C) the definition

∀g∈SL(2,C) : Λ(g)XAA¯

=gABXBB¯A¯B¯ ≡gXg (A.22) defines a representation ofSL(2,C)on the vector spaceR4 formed by the components Xµ. This representation defines a linear map

Λ(g) :R4→R4, Xµ7→ Λ(g)Xµ

= Λ(g)µνXν. (A.23)

The component matrix Λµν must be a Lorentz transformation. This can be seen by looking at the Minkowski norm as defined by equation (A.21), and noting that the definition (A.22) cannot affect this norm simply because the determinant of a product of matrices equals the product of the determinant of the individual constituents, that is:

ηµνΛµαXαΛνβXβ = 2 det(gXg) = 2 det(g) det(X)det(g) =

= 2 det(X) =XΛXα. (A.24)

Where we used that for any g ∈ SL(2,C) we have, ipso facto, det(g) = 1. Since (A.2) holds for all Xµ ∈ R4, we can see Λ(g) must be a Lorentz transformation.

Since SL(2,C) is simply connected and the map g7→ Λ(g) is continuous, we can also continuously connect the image Λµν(g) with the identity δνµ. Therefore, Λ(g) must be a proper orthochronous Lorentz transformation, i.e. an element of L+. In fact, Λ(g) can cover all of L+. This can be seen by e.g. studying the corresponding Lie algebras so(1,3) and sl(2,C), establishing the isomorphism between the two, and recognising that for both L+ and SL(2,C) the exponential map can reach any element of the two respective groups. The homomorphism Λ :SL(2,C) → L+ so established is however not invertible; for every λ∈L+ there are exactly two elements g, g ∈SL(2,C), equal up to a signg=−g, that are both mapped towards the sameλ= Λ(g) = Λ(−g). We thus see, SL(2,C) is the universal cover of the group of proper orthochronous Lorentz transformation, and thus plays the same role thatSU(2)has for SO(3).

Let us now delve more into the structure of C2 and its complex conjugate vector space C¯2. Complex conjugation of the components relates one with the other:

· :C2→C¯2, ωA7→ωA= ¯ωA¯, (A.25) and analogously for the dual vector spaces(C2), and( ¯C2), the elements of which we can write as ωA and ω¯A¯ respectively. The spinor indices transform under the funda-mental or defining transformation of SL(2,C), that is just by matrix multiplication:

(gω)A=gABωB. (A.26)

All finite dimensional representations of SL(2,C) are labelled by spins (j, k) ∈ 12N0

×12N0 and can be constructed by simply tensoring the fundamental representation.

Hj,k := sym O2j

C2

⊗sym O2k

2

, (A.27)

where sym denotes total symmetrisation of the respective tensor product. For an elementΨ∈ Hj,k the irreducible group action is simply given by:

(gΨ)A1...A2jA¯1...A¯j =gA1B1· · ·gA2jB2jA¯1B¯1· · ·g¯A¯jB¯2kΨB1...B2jB¯1...B¯2k. (A.28) Comparing this equation with (A.17) we see Minkwoski vectors belong to the (12,12) representation of SL(2,C).

We have seen complex conjugation relatesC2 withC¯2, what we now need is an object allowing us to move the A, B, C, . . . and A,¯ B,¯ C¯ indices, that is a map from e.g. C2

to its algebraic dual (C2). This should be done respecting the symmetry group in question. To rise and lower Minkowski indices, we use the metric ηµν and not any other non-degenerate two-index tensor, simply because ηµν keeps unchanged if going from one inertial frame to another, while for a generic tensor this would not be true anymore. We have seen, in e.g. (A.22, A.2) that SL(2,C) linearly acts onto these indices when performing a Lorentz transformation. But there is an invariant object for thisSL(2,C)action, easily identified as the anti-symmetric ǫ-tensorǫABBA by observing:

∀g∈SL(2,C) :ǫCDgCAgDB = det(g)ǫABAB. (A.29) Since, again, for anyg∈SL(2,C),det(g) = 1. We define its contravariant versionǫAB by demanding:

ǫACǫBC ≡ǫAB =! δAB, ¯ǫA¯B¯ :=ǫAB, ¯ǫA¯B¯ :=ǫAB. (A.30) We can now use theǫ-tensor to establish the natural isomorphism between C2 and its dual vector space:

C2∋ωA7−→ωABAωA∈ C2

, (A.31)

C2

∋ωA7−→ωAABωB ∈C2. (A.32) and equally for the complex conjugate vector space. Here one has to be careful with index positions, particularly illustrated by the identity:

πAωA=−πAωA. (A.33)

We finally fix our conventions by choosing the matrix elements of theǫ-tensor as:

ǫ01= 1 =−ǫ10, ǫ00= 0 =ǫ11 (A.34) That theǫ-tensor plays the role of a metric is particularly well ilustrated when writing the Minkowski metric according to (A.19). Equation implies:

ηαβAAB¯ B¯AB¯ǫA¯B¯. (A.35) We close this section by studying the Lie algebra ofSL(2,C), and giving the relation to so(1,3). A possible basis in sl(2,C) is given by the wedge product of the soldering forms:

ΣABαβ =−1

AC¯σ¯CBβ]¯ , (A.36) where σ¯AAα¯ is nothing but σA¯ , and the anti-symmetrisation has to be taken over the pair[α, β] of indices only. These matrices obey the commutation relations of the Lorentz algebra. Suppressing the spinor indicesA, B, . . . we have in fact:

Σαβµν

= 4δαδβ]βηβµδµδν]νΣµν. (A.37) Using this basis we can write the isomorphism induced by (A.2) betweenso(1,3) and sl(2,C) by stating:

Λ :sl(2,C)∋ 1

αβωαβ 7−→ωαβ ∈so(1,3). (A.38)

These generators correspond to the self-dual sector of the Lorentz algebra, e.g.

ΣαβPαβµν = Σµν. (A.39)

Where we have introduced the self-dual projector Pαβµν = 1

2 δµδνβ]− i 2ǫαβµν

. (A.40)

Furthermore for any ω∈so(1,3)we find that:

1

αβωαβi1

minωmn+ iωio

=:τiωi, (A.41) whereσi = 2iτi are the Pauli spin matrices, and for anyω ∈so(1,3)

ωi = 1

minωmn+ iωio (A.42)

denote its self-dual components (A.13). Choosing these complex coordinates onsl(2,C), we can simplify calculations, in fact the commutation relations between the self-dual generators are nothing but

τi, τj

ijmτm. (A.43)

The differential map, that is the push forward Λ : sl(2,C) → so(1,3), induces an isomorphism between the two respective Lie algebras.

Λτi=−iΠi, Λτ¯i =−i ¯Πi, (A.44) whereΠi are the self-dual generators as defined by (A.10).