• Nie Znaleziono Wyników

Poincare group and quantum fields

N/A
N/A
Protected

Academic year: 2021

Share "Poincare group and quantum fields"

Copied!
47
0
0

Pełen tekst

(1)

Krzysztof Golec–Biernat

IFJ PAN

(25 października 2020)

Wersja robocza nie do dystrybucji

Kraków

2020

(2)
(3)

1 Poincare group 4

1.1 Definition . . . 4

1.2 Tensor notation . . . 5

1.3 Co and contravariant tensors . . . 7

1.4 Scrap . . . 7

1.5 Infinitesimal Lorentz transformations . . . 8

1.6 Finite Lorentz transformations . . . 9

1.7 Finite dimensional representations . . . 12

1.8 SL(2, C) and Lorentz group . . . . 13

1.9 Irreducible representations of SL(2, C) . . . . 15

2 UIR of the Poincare group 18 2.1 Unitary representations . . . 18

2.2 Generators . . . 19

2.3 Poincare group algebra . . . 20

2.4 Pauli-Lubanski vector . . . 22

2.5 Casimir operators . . . 25

2.6 Unitary irreducible representations . . . 26

2.7 Orbit and little group . . . 28

2.8 Wigner’s construction . . . 30

3 Massive particles 32 3.1 Standard momentum and little group . . . 32

3.2 Consistency condition . . . 33

3.3 Standard hermitian boosts . . . 34

3.4 Casimir operators . . . 37

3.5 Irreducible representations . . . 37 3

(4)

4 Massless particles 40 4.0.1 Standard momentum and little group . . . 40 4.0.2 Casimir operators . . . 43 4.0.3 Irreducible representations . . . 44

(5)

Poincare group

1.1 Definition

Let

g = (gµν) =

1 0 0 0

0 −1 0 0

0 0 −1 0

0 0 0 −1

(1.1)

be the matrix of the metric tensor, where µ, ν = 0, 1, 2, 3. The proper Lorentz group SO(1, 3) is defined by a set of nonsingular, 4-dimensional real matrices Λ which obey the relation

ΛTg Λ = g , detΛ = 1 (1.2)

The Poincare group P is defined as a set of pairs

(Λ, a) , a ∈ R4 (1.3)

which obey the following composition law

1, a1)(Λ2, a2) = (Λ1Λ2, Λ1a2+ a1) (1.4) The composition law is non-commutative since in general

1Λ2, Λ1a2+ a1) 6= (Λ2Λ1, Λ2a1+ a2) (1.5) The neutral element is given by (1, 0) since for any element (Λ, a) ∈ P

(1, 0)(Λ, a) = (Λ, a) (1.6)

while the inverse to this element is given by (Λ−1, −Λ−1a) since

−1, −Λ−1a)(Λ, a) = (1, 0) (1.7)

5

(6)

Notice that the set of elements (Λ, 0) ∈ P form a subgroup of P which is isomorphic to the Lorentz group since

1, 0)(Λ2, 0) = (Λ1Λ2, 0) (1.8) Considering the elements (1, a) ∈ P, we obtain the subgroup of P with the composition law

(1, a1)(1, a2) = (1, a1+ a2) (1.9) which is the abelian group of 4-dimensional translations.

The Poincare transformation is a composition of the Lorentz transformation and translation (in the indicated below order)

(Λ, a) = (1, a)(Λ, 0) (1.10)

1.2 Tensor notation

Let V be a real 4-dimensional vector space. Choosing the canonical base vectors eµ, where µ = 0, 1, 2, 3, any vector x ∈ V can be written as a linear combination of the base vectors

x = x0e0+ x1e1+ x2e2+ x3e3≡ xµeµ (1.11) where we apply the Einstein summation convention in which the upper and lower identical indices are summed. The four numbers

xµ= (x0, x1, x2, x3) (1.12) are coordinates of the vectors x in the base eµ. The vector space V is equiped with a real symmetric two-form, i.e. for any x, y, z ∈ V and a, b ∈ R we have

g(ax + by, z) = ag(x, z) + bg(y, z) , g(x, y) = g(y, x) ∈ R (1.13) which is not positive definite, i.e. there exists vector x 6= 0 such that

g(x, x) < 0 (1.14)

This two form allows to define the scalar product of two vectors x, y ∈ V

x · y = g(x, y) (1.15)

In a given base eµ, we have

g(x, y) = g(xµeµ, yνeν) = xµyνg(eµ, eν) (1.16) and introducing

gµν= g(eµ, eν) (1.17)

we may write

x · y = gµνxµyν (1.18)

(7)

Thus, for the Minkowski space with gµν given by (1.1), we have for the canonical base vectors e0· e0= −e1· e1= −e2· e2= −e3· e3= 1 (1.19) while for µ 6= ν

eµ· eν = 0 (1.20)

In the active interpretation, the Poincare transformation (Λ, a) is the vector space trans- formation V → V such that for any x, x0, a ∈ V

x0≡ (Λ, a) x = Λx + a (1.21)

where Λ is a Lorentz operator which obeys the linearity conditions

Λ(αx + βy) = α Λx + β Λy (1.22)

for any x, y ∈ V and α, β ∈ R, and preserves the scalar product

g(x, y) = g(Λx, Λy) (1.23)

With such a definition, we find the composition law (1.5)

x0= (Λ1, a1)(Λ2, a2) x = (Λ1, a1)(Λ2x + a2) = Λ12x + a2) + a1

= Λ1Λ2x + (Λ1a2+ a1) = (Λ1Λ2, Λ1a2+ a1)x (1.24) In a given canonical base eµ, the transformation (1.21) reads

x0= xeµ= Λ(xνeν) + aµeµ= xν(Λeν) + aµeµ (1.25) Defining the matrix Λµν of the Lorentz operator Λ through the equation

Λeν = eµΛµν (1.26)

we find the following coordinate form of the transformation (1.21)

x= Λµνxν+ aµ (1.27)

which could be interpreted as matrix multiplication with the coordinates forming columns,

y0 y1 y2 y3

=

Λ00 Λ01 Λ02 Λ03 Λ10 Λ11 Λ12 Λ13 Λ20 Λ21 Λ22 Λ23 Λ30 Λ31 Λ32 Λ33

x0 x1 x2 x3

+

a0 a1 a2 a3

(1.28)

The Lorentz transformation condition (1.23) can also be written in the coordinate form since g(eα, eβ) = g(Λeα, Λeβ) = g(eµΛµα, eνΛνβ) = ΛµαΛνβg(eµ, eν) (1.29) which gives

gαβ = gµνΛµαΛνβ (1.30)

(8)

1.3 Co and contravariant tensors

Introducing the symmetric contravariant metric tensor

g−1= (gµν) =

1 0 0 0

0 −1 0 0

0 0 −1 0

0 0 0 −1

(1.31)

such that

gαµgαν = δνµ (1.32)

we can lower and raise tensor component indices, e.g.

xµ= gµνxν, xµ= gµνxν (1.33)

Thus, the scalar product (1.18) can also be written as follows

x · y = xµyµ= xµyµ (1.34)

Introducing the inverse Lorentz matrix (Λ−1)µν such that

Λµα−1)αν= (Λ−1)µαΛαν= δµν (1.35) we can write (1.29) in the following form

gµρΛµα= gαβ−1)βρ (1.36) Using symmetry of the metric tensors, we can lower the indices to obtain

Λρα= (Λ−1)αρ (1.37)

and raising α we find

Λρα= (Λ−1)αρ (1.38)

1.4 Scrap

In the passive interpretation, the Poincare transformation is due to the change of the base vectors and the point to which they are fixed to, i.e.

(Λ, a) : {0, eµ} → {O0, eµ0} (1.39) Under the Lorentz transformation any vector x ∈ V remains the same while its coordinates are different due to the change of the basis,

x = xµeµ= xν0eν0 (1.40)

(9)

Assuming (1.29) for the change of the base vectors

eν0= eµΛµν0 (1.41)

we find the following transformation of the coordinates

xµ= Λµν0xν0 (1.42)

or introducing the inverse matrix Λ−1

xµ0 = (Λ−1)µν0xν (1.43)

This should be compared to (1.27) for aµ= 0,

x= Λµνxν, (1.44)

assuming that numerically

Λµν = Λµν0,−1)µν= (Λ−1)µν0 (1.45) With such an assumption xν0 are coordinates of the vector x in the new base eµ0, while x are coordinates of the vector x0= Λx in the base eµ.

The change of the reference points by the vector a = −−−→

OO0 amounts to the vector shift x → x0 such that

x0= x + a (1.46)

which leads to the following transformation of the coordinates in the basis eµ

x= xµ+ aµ= (Λ−1)µν(xν− aµ) (1.47) The composition law (1.5) is also obeyed since in the matrix interpretation, we find

x= (Λ1, a1)(Λ2, a2) x = (Λ1, a1)(Λ2x + a2) = Λ12x + a2) + a1

= Λ1Λ2x + (Λ1a2+ a1) = (Λ1Λ2, Λ1a2+ a1)x (1.48)

1.5 Infinitesimal Lorentz transformations

Let us consider infinitesimal Lorentz transformation

Λµν = δνµ+ µν, µν 1 (1.49) where from (1.29)

gµναµ+ µα)(δβν+ νβ) = gαβ (1.50) Up to linear terms in , we obtain

gµβµα+ gαννβ= 0 (1.51)

which can be written as

αβ= −βα (1.52)

where we lowered the first index. Thus, we found six infinitesimal parameters of the proper Lorentz transformation: 0iand ij where i, j = 1, 2, 3 and i < j.

(10)

1.6 Finite Lorentz transformations

Let us write the proper Lorentz transformation in the form

Λ = (Λµν) = e, (1.53)

gdzie the matrix Ω = (Ωµν) will be found from the Lorentz transformation condition

ΛTg Λ = g (1.54)

Multiplying from the right by the inverse Lorentz transformation matrix

Λ−1= e−Ω (1.55)

we find

eTg = g e−Ω. (1.56)

Expanding this expression in powers of Ω



1 + ΩT + 1

2!(ΩT)2 + ...

 g = g



1 − Ω + 1

2! Ω2 + ... ,



, (1.57)

and comparing the same order terms on both sides, we find that

Tg = −g Ω . (1.58)

Introducing the matrix

ω = g Ω ,ωµν= gµααν (1.59)

we find that ω is antisymmetric

ωT = (gΩ)T = ΩTg = −g Ω = −ω . (1.60) Hence the six finite parameters of the Lorentz transformation

β = (β1, β2, β3) , φ = (φ1, φ2, φ3) (1.61) which can be organized in the following way

ω = (ωµν) =

0 β1 β2 β3

−β1 0 φ3 −φ2

−β2 −φ3 0 φ1

−β3 φ2 −φ1 0

(1.62)

and

βi= ω0i, φi=12ijkωjk (1.63)

The matrix Ω for these parameters is given by

Ω = g−1ω ↔ Ωµν= gµαωαν (1.64)

(11)

and we find

Ω = (Ωµν) =

0 β1 β2 β3

β1 0 −φ3 φ2 β2 φ3 0 −φ1 β3 −φ2 φ1 0

(1.65)

Let us write Ω in the form

Ω = iβ · K − iφ · J (1.66)

which defines six generators of the Lorentz group,

K = (K1, K2, K3) , J = (J1, J2, J3) (1.67) By the comparison with (1.65), we find their matrix representation

K1=

0 −i 0 0

−i 0 0 0

0 0 0 0

0 0 0 0

K2=

0 0 −i 0

0 0 0 0

−i 0 0 0

0 0 0 0

K3=

0 0 0 −i

0 0 0 0

0 0 0 0

−i 0 0 0

(1.68)

and

J1=

0 0 0 0

0 0 0 0

0 0 0 −i

0 0 i 0

J2=

0 0 0 0

0 0 0 i

0 0 0 0

0 −i 0 0

J3=

0 0 0 0

0 0 −i 0

0 i 0 0

0 0 0 0

(1.69)

Notice that with such definition, the parameters β and φ are real while the generators K are antihermitian while J are hermitian matrices

K= −K , J= J (1.70)

The generators K cannot be hermitian since the Lorentz group is not compact and cannot be represented by finite size unitary matrices, which would be the case if K were hermitian. It can be easily checked that these generators obey the commutation relations which define the Lie algebra of the Lorentz group SO(1, 3)

[ Ji, Jj] = i ijkJk (1.71)

[ Ji, Kj] = i ijkKk (1.72)

[ Ki, Kj] = −i ijkJk, (1.73) where ijk is totally antisymmetric symbol with 123= 1.

We see that J are generators of the rotation group in three dimensional position space, SO(3). Computing the Casimir operator

J2= (J1)2+ (J2)2+ (J3)2=

0 0 0 0 0 2 0 0 0 0 2 0 0 0 0 2

(1.74)

(12)

we find that in the position space

J2= j(j + 1) = 2 (1.75)

which means that the representation (1.53) of the Lorentz group corresponds to spin j = 1.

The direction of the vector φ gives the rotation axis, while its length, φ = |φ|, the rotation angle. For rotation along the axis ˆ1, when φ = (φ, 0, 0), from the relation J21= 1 we find the following matrix

Λ(φ) = e−iφJ1 = cos φ − iJ1sin φ =

1 0 0 0

0 1 0 0

0 0 cos φ − sin φ 0 0 sin φ cos φ

(1.76)

which corresponds to anticlockwise rotation in the (2, 3) plane for φ > 0.

The generators K are vectorial operators and generate Lorentz boosts in direction given by vector β. For example, for the boost along the axis ˆ1, when β = (β, 0, 0), we find from the relation K21= −1 the following matrix

Λ(β) = eiβK1 = cosh β + iK1sinh β =

cosh β sinh β 0 0 sinh β cosh β 0 0

0 0 1 0

0 0 0 1

(1.77)

which correspond to the boost along the axis ˆ1 in the negative direction for β > 0.

Organizing the six generators K and J in the antisymmetric matrix

Jµν = −Jνµ=

0 K1 K2 K3

−K1 0 J3 −J2

−K2 −J3 0 J1

−K3 J2 −J1 0

(1.78)

such that

Ki= J0i, Jk= 12klmJlm (1.79)

and the six parameters β and φ in the tensor

ωµν=

0 −β1 −β2 −β3 β1 0 φ3 −φ2 β2 −φ3 0 φ1

β3 φ2 −φ1 0

(1.80)

we find the Lorentz transformation in the following form

Λ = expn2iωµνJµνo= exp{iβ · K − iφ · J} (1.81)

(13)

1.7 Finite dimensional representations

Let us define the following operators which are built out the SO(1, 3) generators J and K J±=1

2(J ± iK) (1.82)

In this way, we have two sets of operators

J+= (J1+, J2+, J3+) , J= (J1, J2, J3) (1.83) which from (1.71)-(1.73) obey

hJi+, Jj+i= i ijkJk+ hJi, Jji= i ijkJk h

Ji+, Jji= 0 (1.84)

Notice that J± are hermitian operators, (J±)=1

2(J ± iK)=1

2(J∓ iK) =1

2(J ± iK) = J± (1.85) Thus, we constructed two independent generators of group SU (2). Therefore, we obtain the following isomorphism for Lie algebras of the groups below

SO(1, 3) ∼= SU (2) ⊗ SU (2) (1.86)

This identification means that we can construct finite dimension representations of SO(1, 3) as a tensor product of finite dimension irreducible representations of SU (2), which are eumera- ted by spin j = 0,12, 1, . . .. The SO(1, 3) representations will be enumerated by the pairs (j+, j) and the resulting representations will have dimension equal to N = (2j++ 1)(2j+ 1). In ge- neral, it will be reducible representations. The subsequent representations in ascending order of their dimensionality are shown in Table 1.7.

The one dimensional representation (0, 0) is given by 1 and its generators Ji±= 0. The two dimensional representations are constructed from generators which are proportional to Pauli matrices σi. For (12, 0) we have

Ji+=12σi, Ji= 0 (1.87)

and the generators

Ji= (Ji++ Ji) =1

2σi, Ki= −i(Ji+− Ji) = −i

2σi (1.88)

(14)

(j1, j2) N = (2j1+ 1)(2j2+ 1)

(0, 0) 1

(12, 0) (0,12) 2

(1, 0) (0, 1) 3

(12,12) (32, 0) (0,32) 4

(2, 0) (0, 2) 5

(1,12) (12, 1) 6

Tablica 1.1: Finite dimension representations of SO(1, 3)

give the representation

A(1/2,0)= exp{ −12β · σ +2iφ · σ} (1.89) Similarly, for (0,12) we have

Ji+= 0, , Ji=12σi (1.90)

which gives

Ji= (Ji++ Ji) = 1

2σi, Ki= −i(Ji+− Ji) = i

2σi (1.91)

and the representation

A(0,1/2)= exp{12β · σ +2iφ · σ} (1.92) These are the spinor representations which act in complex two dimensional space of vectors, called spinors. Notice that from

det M = exp{Tr ln M } (1.93)

we have

det A(1/2,0)= det A(0,1/2)= 1 (1.94) since Pauli matrices are traceless. Therefore, the two spinor representations are SL(2, C) ma- trices which form group. Their tensor product gives vector Lorentz group representation

A(1/2,0)⊗ A(0,1/2)= Λ (1.95)

1.8 SL(2, C) and Lorentz group

The explicit relation between the Lorentz group SO(1, 3) and the group SL(2, C) of 2 × 2 complex matrices with determinant equal one is the following. For any momentum pµ, define the hermitean matrix

P = pµσµ= p0+ p3 p1− ip2 p1+ ip2 p0− p3

!

(1.96) with the determinant

det(P ) = p20− p21− p22− p23= p2 (1.97)

(15)

where σµ= (1, σ) and σ = (σ1, σ2, σ3) denotes three Pauli matrices. Consider the transformation

P0= AP A (1.98)

where A ∈ SL(2, C). The matrix P0 is hemitian and has the same determinant as P

det(P0) = det(A) det(P ) det(A) = det(P ) (1.99) Thus, the transformation (1.98) induces Lorentz transformation: p0µ= Λµνpν. Both A and −A give the same Λ since SL(2, C) is simple connected double covering of the Lorentz group.

Nevertheless, SL(2, C) has six independent parameters like the Lorentz group.

Introducing

σ¯µ= (1, −σ) (1.100)

which obeys

Tr(σνσ¯µ) = 2 δνµ (1.101)

and computing

TrA(pνσν)Aσ¯µ= Tr(pν0σνσ¯µ) = 2p0µ= 2Λµνpν (1.102) we obtain the inverse relation

Λµν =12Tr(¯σµA σνA) (1.103) We can also write the relation (1.98) with the help of ¯σµ using the relations

σ¯µ= (σ2σµσ2)T, σµ= (σ2σ¯µσ2)T (1.104) Then

p0µ2σ¯µσ2)T = A(pµσ2σ¯µσ2)TA (1.105) and taking the transposition of both sides, we obtain

p0µ2σ¯µσ2) = A(pµσ2σ¯µσ2)AT (1.106) Multiplying from the left and right by σ2, we find

p0µσ¯µ= (σ2Aσ2)(pµσ¯µ)(σ2ATσ2) (1.107) Thus, with

B = σ2Aσ2 (1.108)

we finally obtain

B(pµσ¯µ)B= p0µσ¯µ (1.109)

(16)

1.9 Irreducible representations of SL(2, C)

Since SL(2, C) is not compact group, its finite dimension, irreducible representations are not unitary. To construct them, let us introduce analytic functions of two complex variables f (z1, z2) and define the action of the representation D(A) of A ∈ SL(2, C) in the linear space H of such functions by

[D(A)f ](z1, z2) = f (zTA) (1.110) where

zTA = (z1, z2) α β γ δ

!

= (αz1+ γz2, βz1+ δz2) (1.111) The operators D(A) form the representation since

[D(A2)D(A1)f ](z1, z2) = ˜f (zTA2) = [D(A1)f ](zTA2)

= f (zTA2A1) = [D(A2A1)f ](z1, z2) ‘ (1.112) and

D(1) = 1 , D(A−1) = D−1(A) (1.113)

where D−1 means the inverse to the operator D.

The finite dimension, irreducible representations act in the invariant subspaces Hj⊂ H of the functions f (z1, z2) which are homogenous of degree 2j,

f (λz1, λz2) = λ2jf (z1, z2) (1.114) where

j = 0,12, 1, . . . (1.115)

Indeed, Hj is invariant under the action of the representation element

[Dj(A)f ](z1, z2) = ¯f (z1, z2) = f (αz1+ γz2, βz1+ δz2) (1.116) since ¯f is homogeneous function of degree 2j

f (λz¯ 1, λz2) = f (λ(αz1+ γz2), λ(βz1+ δz2))

= λ2jf (αz1+ γz2, βz1+ δz2) = λ2jf (λz¯ 1, λz2) (1.117) From analyticity, any homogeneous function of degree 2j can be written as

f (z1, z2) =

j

X

m=−j

cjm zj+m1 zj−m2 p(j + m)!(j − m)!

j

X

m=−j

cjmφjm(z1, z2) (1.118)

where φjm are base functions in Hj. In particular, for j = 1/2 and m = ±, we find

φ+(z1, z2) = z1, φ(z1, z2) = z2 (1.119)

(17)

while for j = 1 and m = −1, 0, 1, we obtain φ+(z1, z2) = z12

√2, φ0(z1, z2) = z1z2, φ(z1, z2) = z22

√2 (1.120)

The matrix of the representation Dj(A) is defined from the action on the base functions [Dj(A) φjm](z1, z2) = φjm(zTA) =

j

X

m0=−j

φjm0(z1, z2) [ ˜Dj(A)]m0m (1.121)

In particular, for j = 1/2 we obtain

[D1/2(A) φ+](z1, z2) = φ+(zTA) = αz1+ γz2= φ+α + φγ

= φ+D˜+++ φD˜−+

[D1/2(A) φ](z1, z2) = φ(zTA) = βz1+ δz2= φ+β + φδ (1.122)

= φ+D˜+−+ φD˜−−

and the matrix representation

D˜1/2(A) = D˜++ D˜+−

D˜−+ D˜−−

!

= α β

γ δ

!

= A (1.123)

Similarly, it can be shown that for j = 1 the matrix representation reads

D˜1(A) =

α2

2αβ β2

2αγ αδ + γβ2βδ

γ2

2γδ δ2

(1.124)

Obviously these are not unitary matrices.

It can easily be shown that

Dj(A) , Dj((AT)−1) , Dj(A) , Dj((A)−1) (1.125) fulfil the representation conditions. However, for SL(2, C) there exists the matrix

C = iσ2= 0 1

−1 0

!

, C−1= −iσ2= 0 −1

1 0

!

(1.126) such that for any A ∈ SL(2, C)

CA C−1= (AT)−1, CAC−1= (A)−1 (1.127) which can be easily proven starting from the SL(2, C) matrices

A = α β γ δ

!

, A−1= δ −β

−γ α

!

(1.128)

(18)

where αδ − βγ = 1. Now, one can check that the first relation in (1.127) is fulfilled

0 1

−1 0

! α β γ δ

! 0 −1

1 0

!

= δ −γ

−β α

!

(1.129) while the second relation can be found by taking complex conjugation of the first one.

Therefore, the representations (1.125) are pairwise equivalent

Dj(A) ∼ Dj((AT)−1) , Dj(A) ∼ Dj((A)−1) (1.130) and there exists only two inequivalent representations, e.g.

D(j,0)(A) ≡ Dj(A) , D(0,j)(A) ≡ Dj((A)−1) (1.131) In particular, for j = 1/2 we have

D˜(1/2,0)(A) = A = α β γ δ

!

D˜(0,1/2)(A) = (A)−1= δ −γ

−β α

!

(1.132) Let us notice that for A ∈ SU (2) ⊂ SL(2, C)

A−1= A => A = (A)−1 (1.133)

and the two representations (1.131) are identical

D(j,0)(A) = D(0,j)(A) = D(j)(A) (1.134)

Thus, D(j)(A) in this case is unitary since SU (2) is compact group.

By definition, the representation D(j,k) is a tensor product of irreducible representations D(j,k)(A) = D(j,0)(A) ⊗ D(0,k)(A) (1.135) which matrix has double indices

D˜aa(j,k)0,bb0(A) = ˜Dab(j,0)(A) ˜D(0,k)a0b0 (A) (1.136) The tensor product representation is in general reducible. In particular

D(1/2,0)⊗ D(0,1/2)= D(1,0)⊕ D(0,0) (1.137) where D(0,0)= 1 and D(1,0) is given by (1.124). It can be shown that such representation is unitary equivalent to the vectorial Lorentz group representations.

(19)

UIR of the Poincare group

2.1 Unitary representations

Topologically, the Poincare group P is not compact. Thus, their unitary and irreducible re- presentations (UIR) are infinite dimensional. It means that they are realized in an infinite dimensional Hilbert space with the inner product h .|.i by linear unitary operators U

(Λ, a) → U (Λ, a) (2.1)

such that the following homeomorphism is valid

U (Λ1, a1) U (Λ2, a2) = U (Λ1Λ2, Λ1a2+ a1) (2.2) With this composition law, the unitary operators U form group with the neutral element

U (1, 0) = 1 (2.3)

and the inverse element for each U (Λ, a)

U (Λ−1, −Λ−1a)) = [U (Λ, a)]−1 (2.4) Since we consider the unitary representations

[U (Λ, a)]−1= [U (Λ, a)]≡ U(Λ, a) (2.5) and we have

U (Λ, a) U(Λ, a) = U(Λ, a) U (Λ, a) = 1 (2.6) From (1.10) we find

U (Λ, a) = U (1, a) U (Λ, 0) (2.7)

Introducing the shorthand notation

U (a) = U (1, a) , U (Λ) = U (Λ, 0) (2.8)

19

(20)

we obtain in the indicated order

U (Λ, a) = U (a) U (Λ) (2.9)

Thus, we may consider two subgroups. The subgroup of the Lorentz transformations for which U (Λ1) U (Λ2) = U (Λ1Λ2) (2.10) and the abelian subgroup of four dimensional translations for which

U (a) U (a0) = U (a0) U (a) = U (a + a0) (2.11)

2.2 Generators

From (1.10), the general form of the unitary representation is given by

U (Λ, a) = U (1, a) U (Λ, 0) = exp{ iaµPµ} exp{ −2iωµνJµν} (2.12) where aµ and ωµν= −ωνµ are ten parameters of the Poincare group while Pµ and Jµν are the corresponding hermitian generators

Pµ= Pµ, Jµν = Jµν= −Jνµ (2.13)

For the Poincare transformation with infinitesimal parameters aµ= µand ωµν, we have up to the linear order in these parameters

U (1 + ω, ) = 1 + iµPµ12µνJµν (2.14) Let us find the transformation laws for these generators under the Poincare transformations.

To this end, let us compute using the composition law (2.2)

U (Λ, a) U (1 + ω, ) U(Λ, a) = U (Λ(1 + ω), Λ + a) U (Λ−1, −Λ−1a)

= U (1 + ΛωΛ−1, Λ − ΛωΛ−1a) (2.15) and from (2.14), we obtain

U (Λ, a) U (1 + ω, ) U(Λ, a) = 1 + i(Λ − ΛωΛ−1a)αPα12i(ΛωΛ−1)αβJαβ

= 1 + i(Λ)αPα12i(ΛωΛ−1)αβ(Jαβ+ 2Pαaβ) (2.16) On the other hand, from (2.14)

U (Λ, a) U (1 + ω, ) U(Λ, a) = 1 + iµU (Λ, a)PµU(Λ, a) −12µνU (Λ, a)JµνU(Λ, a) (2.17) To compare the corresponding expressions on the rhs of (2.16) and (2.17), we compute

(Λ)αPα= ΛαµµPα= µ(PαΛαµ)

(ΛωΛ−1)αβ= Λαµωµν−1)νβ = ΛαµΛβνωµν=12ωµνΛαµΛβν− ΛανΛβµ (2.18)

(21)

where we used antisymmetry of ωµν. Therefore, we find

U (Λ, a) PµU(Λ, a) = PαΛαµ (2.19) and

U (Λ, a)JµνU(Λ, a) = 12(Jαβ+ 2Pαaβ)ΛαµΛβν− ΛανΛβµ

= (Jαβ+ Pαaβ− Pβaα) ΛαµΛβν (2.20) Thus, Pµ and Jµν are vectorial and tensorial operators, respectively. Raising and lowering indices, we obtain with the help of relation (1.38)

U (Λ, a) PµU(Λ, a) = (Λ−1)µαPα

U (Λ, a) JµνU(Λ, a) = (Λ−1)µα−1)νβ(Jαβ+ Pαaβ− Pβaα) (2.21)

2.3 Poincare group algebra

The commutation relations between the generators Pµ and Jµν can be found with the help of the transformation laws (2.19) and (2.20). For pure translations Λµν = δµν and

U (a) PµU(a) = Pµ

U (a) JµνU(a) = Jµν+ Pµaν− Pνaµ (2.22) Thus, for infinitesimal translations U (a) we have up to linear terms in aµ

U (a)PµU(a) = (1 + iaνPν)Pµ(1 − iaνPν)

= Pµ+ iaν[Pν, Pµ] = Pµ (2.23) and from the comparison of both sides of the last equality, we find

[Pµ, Pν] = 0 (2.24)

We also consider up to linear terms in aµ

U (a)JµνU(a) = (1 + iaαPα)Jµν(1 − iaαPα) = Jµν+ iaα[Pα, Jµν]

= Jµν+ Pµaν− Pνaµ

= Jµν+ aα(gναPµ− gµαPν) (2.25) and by the comparison of the last expressions in the two lines, we obtain

[Jµν, Pα] = i (gναPµ− gµαPν) (2.26)

Cytaty

Powiązane dokumenty

It is well known that any complete metric space is isomet- ric with a subset of a Banach space, and any hyperconvex space is a non- expansive retract of any space in which it

Thus eigenfunctions of the Fourier transform defined by the negative definite form −x 2 in one variable are the same as eigenfunctions of the classical in- verse Fourier

We give a direct proof of this characterization and get stronger results, which allows us to obtain some other results on ω-limit sets, which previously were difficult to prove.. Let

The essential part of the paper is Section 3 in which we give a formula allowing to compute the scalar part of a given Clifford number.. As an application of this formula, we are

In fact, the proof of the preceding section can be viewed as a generalization of the fact that adding super-perfect real adds a minimal real in the sense that the structure of

The result of [LM-L] implies that an ultrametric space can be bi-Lipschitz embedded in R n if and only if its Assouad dimension is finite (see also [A]).. According to Semmes

a Artificially generated noisy image, b its magnitude gradient image calculated using a two-point operator, c watersheds of b, d watersheds significant at the 10−5 significance level,

3. Free zero-dimensional topological groups. Here is probably the most natural example of a topological group free in a class that fails to form a variety in a most spectacular