• Keine Ergebnisse gefunden

Local SU(2) symmetry

Im Dokument Non-Abelian Gauge Theories (Seite 56-63)

Local Non-Abelian (Gauge) Symmetries

13.1 Local SU(2) symmetry

13.1.1 The covariant derivative and interactions with matter In this section we shall introduce the main ideas of the non-Abelian SU(2) gauge theory which results from the demand of invariance, or covariance,

13.1. Local SU(2) symmetry 41

under transformations such as (13.2). We shall generally use the language of isospin when referring to the physical states and operators, bearing in mind that this will eventually mean weak isospin.

We shall mimic as literally as possible the discussion of electromagnetic gauge covariance in sections 2.4 and 2.5 of volume 1. As in that case, no free particle wave equation can be covariant under the transformation (13.2) (taking the isospinor example for definiteness), since the gradient terms in the equation will act on the phase factor α(x). However, wave equations with a suitably defined covariant derivative can be covariant under (13.2); physically this means that, just as for electromagnetism, covariance under local non-Abelian phase transformations requires the introduction of a definite force field.

In the electromagnetic case the covariant derivative is

Dμ = μ + iqAμ(x). (13.4)

For convenience we recall here the crucial property of Dμ . Under a local U(1) phase transformation, a wavefunction transforms as (cf (13.3))

ψ(x)→ ψ'(x) = exp(iqχ(x))ψ(x), (13.5) from which it easily follows that the derivative (gradient) of ψ transforms as

μψ(x)→μψ'(x) = exp(iqχ(x))∂μψ(x) + iq∂μχ(x)exp(iqχ(x))ψ(x). (13.6) Comparing (13.6) with (13.5), we see that, in addition to the expected first term on the hand side of (13.6), which has the same form as the right-hand side of (13.5), there is anextra term in (13.6). By contrast, the covariant derivative of ψ transforms as (see section 2.4 of volume 1)

Dμψ(x)→ D'μψ'(x) = exp(iqχ(x))Dμψ(x) (13.7) exactly as in (13.5), with no additional term on the right-hand side. Note that Dμ has to carry a prime also, since it contains Aμ which transforms to A'μ = Aμ −∂μχ(x) whenψ transforms by (13.5). The property (13.7) ensured the gauge covariance of wave equations in the U(1) case; the similar property in the quantum field case meant that a globally U(1)-invariant Lagrangian could be converted immediately to a locally U(1)-invariant one by replacing

μ by Dˆμ (section 7.4).

In appendix D of volume 1 we introduced the idea of ‘covariance’ in the context of coordinate transformations of 3- and 4-vectors. The essential notion was of something ‘maintaining the same form’, or ‘transforming the same way’. The transformations being considered here are gauge transformations rather than coordinate ones; nevertheless it is true that, under them, Dμψ transforms in the same way as ψ, while ∂μψ does not. Thus the term covariant derivative seems appropriate. In fact, there is a much closer analogy between the ‘coordinate’ and the ‘gauge’ cases, which we did not present in volume 1, but give now in appendix N, for the interested reader.

13.1. Local SU(2)symmetry 43 where we have retained only the terms linear inEfrom an expansion of (13.8) withα→E. We have now dropped the x-dependence of theψ(12)’s, but kept that ofE(x), and we have used the simple ‘1’ for the unit matrix in the two-dimensional isospace. Equation (13.14) exhibits again an ‘extra piece’ on the right-hand side, as compared to (13.13). On the other hand, inserting (13.10) and (13.13) into our covariant derivative requirement (13.9) yields, for the left-hand side in the infinitesimal case,

D'μψ(12)'= (∂μ+ igτ·W/2)[1 + igτ·E(x)/2]ψ(12) (13.15) while the right-hand side is

[1 + igτ·E(x)/2](∂μ+ igτ·Wμ/2)ψ(12). (13.16) In order to verify that these are the same, however, we would need to know W'μ – that is, the transformation law for the three Wμ fields. Instead, we shall proceed ‘in reverse’, and use the imposed equality between (13.15) and (13.16) to determine the transformation law of Wμ.

Suppose that, under this infinitesimal transformation,

Wμ→W'μ=Wμ+δWμ. (13.17) Then the condition of equality is

[∂μ+ igτ/2·(Wμ+δWμ)][1 + igτ·E(x)/2]ψ(12)

= [1 + igτ·E(x)/2](∂μ+ igτ·Wμ/2)ψ(12). (13.18) Multiplying out the terms, neglecting the term of second order involving the product ofδWμ andE and noting that

μ(Eψ) = (∂μE)ψ+E(∂μψ) (13.19) we see that many terms cancel and we are left with

igτ·δWμ Using the identity for Pauli matrices (see problem 3.4(b))

σ·aσ·b=a·b+ iσ·a×b (13.21) this yields

τ·δWμ =−τ·∂μE(x)−gτ·(E(x)×Wμ). (13.22)

42 13. Local Non-Abelian (Gauge) Symmetries

We need the local SU(2) generalization of (13.4), appropriate to the local SU(2) transformation (13.2). Just as in the U(1) case (13.6), the ordinary

and differentiating term by term. By analogy with (13.7), the key property

1 us ‘decode’ the desired property (13.9), for the algebraically simpler case of an infinitesimal local SU(2) transformation with parameters E(x), which are

42 13. Local Non-Abelian (Gauge) Symmetries We need the local SU(2) generalization of (13.4), appropriate to the local SU(2) transformation (13.2). Just as in the U(1) case (13.6), the ordinary gradient acting on ψ(12)(x) does not transform in the same way as ψ(12)(x):

takingμ of (13.2) leads to

μψ(12)1(x) = exp[igτ·α(x)/2]∂μψ(12)(x)

+ igτ·∂μα(x)/2 exp[igτ·α(x)/2]ψ(12)(x) (13.8) as can be checked by writing the matrix exponential exp[A] as the series

exp[A] = c n=0

An/n!

and differentiating term by term. By analogy with (13.7), the key property we demand for ourSU(2) covariant derivative Dμψ(12) is that this quantity should transform likeψ(12) – i.e. without the second term in (13.8). So we require

(D1μψ(12)1(x)) = exp[igτ·α(x)/2](Dμψ(12)(x)). (13.9) The definition ofDμ which generalizes (13.4) so as to fulfil this requirement is

Dμ(acting on an isospinor) =μ+ igτ·Wμ(x)/2. (13.10) The definition (13.10), as indicated on the left-hand side, is only appropri-ate for isospinorsψ(12); it has to be suitably generalized for otherψ(t)’s (see (13.44)).

We now discuss (13.9) and (13.10) in detail. Theμis multiplied implicitly by the unit 2 matrix, and the τ’s act on the two-component space ofψ(12). TheWμ(x) arethreeindependent gauge fields

Wμ= (W1μ, W2μ, W3μ), (13.11) generalizing the single electromagnetic gauge fieldAμ. They are called SU(2) gauge fields, or more generallyYang-Mills fields. The term τ·Wμ is then the 2×2 matrix

τ·Wμ=

( W3μ W1μiW2μ W1μ+ iW2μ −W3μ

)

(13.12) using the τ’s of (12.25); the x-dependence of the Wμ’s is understood. Let us ‘decode’ the desired property (13.9), for the algebraically simpler case of an infinitesimal local SU(2) transformation with parametersE(x), which are of course functions ofx since the transformation is local. In this case,ψ(12) transforms by

ψ(12)1= (1 + igτ·E(x)/2)ψ(12) (13.13) and the ‘uncovariant’ derivativeμψ(12)transforms by

μψ(12)1= (1 + igτ·E(x)/2)∂μψ(12)+ igτ·∂μE(x)/2ψ(12), (13.14)

13.1. Local SU(2) symmetry 43

where we have retained only the terms linear in E from an expansion of (13.8)

1

with α → E. We have now dropped the x-dependence of the ψ( )2 ’s, but kept that of E(x), and we have used the simple ‘1’ for the unit matrix in the two-dimensional isospace. Equation (13.14) exhibits again an ‘extra piece’ on the right-hand side, as compared to (13.13). On the other hand, inserting (13.10) and (13.13) into our covariant derivative requirement (13.9) yields, for the left-hand side in the infinitesimal case,

1 2

1

· W /2)[1 + igτ 2

D'μψ( )' = (∂μ + igτ · E(x)/2]ψ( ) (13.15) while the right-hand side is

1

· E(x)/2](∂μ + igτ · W μ/2)ψ( )2

[1 + igτ . (13.16)

In order to verify that these are the same, however, we would need to know W 'μ – that is, the transformation law for the three W μ fields. Instead, we shall proceed ‘in reverse’, and use the imposed equality between (13.15) and

W μ (13.16) to determine the transformation law of .

Suppose that, under this infinitesimal transformation,

W μ W 'μ = W μ + δW μ . (13.17) Then the condition of equality is

1

[∂μ + igτ /2· (W μ + δW μ)][1 + igτ · E(x)/2]ψ( ) 2 1

= [1 + igτ · E(x)/2](∂μ + igτ · W μ/2)ψ( )2 . (13.18) Multiplying out the terms, neglecting the term of second order involving the product of δW μ and E and noting that

μ(Eψ) = (∂μE)ψ + E(∂μψ) (13.19) we see that many terms cancel and we are left with

τ · δW μ τ ·μE(x)

ig = ig

2 |(2 τ · E(x)) (τ · W μ ) (τ · W μ ) (τ · E(x))|

+ (ig)2 .

2 2 2 2

(13.20) Using the identity for Pauli matrices (see problem 3.4(b))

σ ·· b = a · b + iσ · a × b (13.21)

this yields

τ · δW μ = −τ ·μE(x)−· (E(x)× W μ). (13.22)

13.1. Local SU(2)symmetry 45 so thatψ(12)transforms by

ψ(12)'=U(α(x))ψ(12). (13.28) Then we require

Dψ(12)'=U(α(x))Dμψ(12). (13.29) The left-hand side is

(∂μ+ igτ·W'μ/2)U(α(x))ψ(12)

= (∂μU)ψ(12)+U∂μψ(12)+ igτ·W'μ/2(12), (13.30) while the right-hand side is

U(∂μ+ igτ·Wμ/2)ψ(12). (13.31) TheU∂μψ(12)terms cancel leaving

(∂μU)ψ(12)+ igτ·W'μ/2(12)=Uigτ·Wμ/2ψ(12). (13.32) Since this has to be true for all (two-component)ψ(12)’s, we can treat it as an operator equation acting in the space ofψ(12)’s to give

μU+ igτ·W'μ/2U=Uigτ·Wμ/2, (13.33) or equivalently

1

2τ·W= i

g(∂μU)U1+U1

2τ·WμU1, (13.34) which defines the (finite) transformation law for SU(2) gauge fields. Problem 13.1 verifies that (13.34) reduces to (13.26) in the infinitesimal case α(x)→ E(x).

Suppose now that we consider a Dirac equation forψ(12):

(iγμμ−m)ψ(12)= 0 (13.35) where both the ‘isospinor’ components of ψ(12) are four-component Dirac spinors. We assert that we can ensure local SU(2) gauge covariance by re-placing μ in this equation by the covariant derivative of(13.10). Indeed, we have

U(α(x))[iγμDμ−m]ψ(12) = iγμU(α(x))[Dμψ(12))−mU(α(x)]ψ(12)

= iγμD'μψ(12)'−mψ(12)' (13.36) using equations (13.9) and (13.28). Thus if

(iγμDμ−m)ψ(12)= 0 (13.37)

)'

44 13. Local Non-Abelian (Gauge) Symmetries

Equating components of τ on both sides, we deduce

δW μ = −∂μE(x)− g[E(x)× W μ]. (13.23)

The reader may note the close similarity between these manipulations and those encountered in section 12.1.3.

W μ Equation (13.23) defines the way in which the SU(2) gauge fields transform under an infinitesimal SU(2) gauge transformation. If it were not for the presence of the first term μE(x) on the right-hand side, (13.23) would be simply the (infinitesimal) transformation law for the T = 1 triplet repre-sentation of SU(2) – see (12.64) and (12.65) in section 12.1.3. As mentioned at the end of section 12.2, the T = 1 representation is the ‘adjoint’, or ‘regular’, representation of SU(2), and this is the one to which gauge fields belong, in general. But there is the extra term −∂μE(x). Clearly this is directly analo-gous to the −∂μχ(x) term in the transformation of the U(1) gauge field Aμ; here, an independent infinitesimal function Ei(x) is required for each compo-nent W i μ (x). If the E’s were independent of x, then μE(x) would of course vanish and the transformation law (13.23) would indeed be just that of an SU(2) triplet. Thus we can say that under global SU(2) transformations, the W μ behave as a normal triplet. But under local SU(2) transformations they acquire the additional −∂μE(x) piece, and thus no longer transform

‘prop-1

erly’, as an SU(2) triplet. In exactly the same way, μψ( ) 2 did not transform

‘properly’ as an SU(2) doublet, under a local SU(2) transformation, because of the second term in (13.14), which also involves μE(x). The remarkable

re-1

sult behind the fact that Dμψ( ) 2 does transform ‘properly’ under local SU(2) transformations, is that the extra term in (13.23) precisely cancels that in (13.14)!

To summarize progress so far: we have shown that, for infinitesimal trans-formations, the relation

1

2 1

(D'μψ( )') = [1 + igτ · E(x)/2](Dμψ( )2 ) (13.24) (where Dμ is given by (13.10)) holds true if in addition to the infinitesimal

1

local SU(2) phase transformation on ψ( ) 2 1

2 1

ψ( = [1 + igτ · E(x)/2]ψ( )2 (13.25) the gauge fields transform according to

W 'μ = W μ μE(x)− g[E(x)× W μ]. (13.26)

In obtaining these results, the form (13.10) for the covariant derivative has been assumed, and only the infinitesimal version of (13.2) has been treated explicitly. It turns out that (13.10) is still appropriate for the finite (non-infinitesimal) transformation (13.2), but the associated transformation law for the gauge fields is then slightly more complicated than (13.26). Let us write

U(α(x)) exp[igτ · α(x)/2] (13.27)

44 13. Local Non-Abelian (Gauge) Symmetries Equating components ofτ on both sides, we deduce

δWμ=−∂μE(x)−g[E(x)×Wμ]. (13.23) The reader may note the close similarity between these manipulations and those encountered in section 12.1.3.

Equation (13.23) defines the way in which the SU(2) gauge fields Wμ transform under an infinitesimal SU(2) gauge transformation. If it were not for the presence of the first termμE(x) on the right-hand side, (13.23) would be simply the (infinitesimal) transformation law for theT = 1 triplet repre-sentation of SU(2) – see (12.64) and (12.65) in section 12.1.3. As mentioned at the end of section 12.2, theT = 1 representation is the ‘adjoint’, or ‘regular’, representation of SU(2), and this is the one to which gauge fields belong, in general. But there is the extra term−∂μE(x). Clearly this is directly analo-gous to the−∂μχ(x) term in the transformation of the U(1) gauge fieldAμ; here, an independent infinitesimal functionEi(x) is required for each compo-nentWiμ(x). If the E’s were independent ofx, then μE(x) would of course vanish and the transformation law (13.23) would indeed be just that of an SU(2) triplet. Thus we can say that under global SU(2) transformations, the Wμ behave as a normal triplet. But underlocalSU(2) transformations they acquire the additional −∂μE(x) piece, and thus no longer transform ‘prop-erly’, as an SU(2) triplet. In exactly the same way,μψ(12)did not transform

‘properly’ as an SU(2) doublet, under a local SU(2) transformation, because of the second term in (13.14), which also involvesμE(x). The remarkable re-sult behind the fact thatDμψ(12) doestransform ‘properly’ under local SU(2) transformations, is that the extra term in (13.23) precisely cancels that in (13.14)!

To summarize progress so far: we have shown that, for infinitesimal trans-formations, the relation

(D'μψ(12)') = [1 + igτ·E(x)/2](Dμψ(12)) (13.24) (where Dμ is given by (13.10)) holds true if in addition to the infinitesimal local SU(2) phase transformation onψ(12)

ψ(12)' = [1 + igτ·E(x)/2]ψ(12) (13.25) the gauge fields transform according to

W'μ=Wμ−∂μE(x)−g[E(x)×Wμ]. (13.26) In obtaining these results, the form (13.10) for the covariant derivative has been assumed, and only the infinitesimal version of (13.2) has been treated explicitly. It turns out that (13.10) is still appropriate for the finite (non-infinitesimal) transformation (13.2), but the associated transformation law for the gauge fields is then slightly more complicated than (13.26). Let us write

which defines the (finite) transformation law for SU(2) gauge fields. Problem 13.1 verifies that (13.34) reduces to (13.26) in the infinitesimal case α(x)→

13.1. Local SU(2)symmetry 47 quantum number, so as to emphasize that it is not the hadronic isospin, for which we retain T; t will be the symbol used for the weak isospin to be introduced in chapter 20. The general local SU(2) transformation for a t-multiplet is then

ψ(t)→ψ(t)' = exp[igα(x)·T(t)(t) (13.42) where the (2t+ 1)×(2t+ 1) matricesTi(t) (i= 1,2,3) satisfy (cf (12.47))

[Ti(t), Tj(t)] = iEijkTk(t). (13.43) The appropriate covariant derivative is

Dμ=μ+ igT(t)·Wμ (13.44) which is a (2t+ 1)×(2t+ 1) matrix acting on the (2t+ 1) components of ψ(t). The gauge fields interact with such ‘isomultiplets’ in auniversalway – only one g, the same for all the particles – which is prescribed by the local covariance requirement to be simply that interaction which is generated by the covariant derivatives. The fermion vertex corresponding to (13.44) is obtained by replacingτ/2 in (13.40) by T(t).

We end this section with some comments:

(i) It is a remarkable fact that only one constantg is needed. This isnotthe same as in electromagnetism. There, each charged field interacts with the gauge fieldAμvia a coupling whose strength is its charge (e,−e,2e,5e . . .).

The crucial point is the appearance of the quadratic g2 multiplying the commutatorof theτ’s, [τ·E,τ·W], in theWμ transformation (equation (13.20)). In the electromagnetic case, there is no such commutator – the associated U(1) phase group is Abelian. As signalled by the presence of g2, a commutator is a non-linear quantity, and the scale of quantities ap-pearing in such commutation relations is not arbitrary. It is an instructive exercise to check that, once δWμ is given by equation (13.23) – in the SU(2) case – then theg’s appearing in ψ(12)' (equation (13.13)) andψ(t)' (via the infinitesimal version of equation (13.42)) must be thesameas the one appearing inδWμ.

(ii) According to the foregoing argument, it is actually a mystery why electric charge should be quantized. Since it is the coupling constant of an Abelian group, each charged field could have an arbitrary charge from this point of view: there are no commutators to fix the scale. This is one of the motivations of attempts to ‘embed’ the electromagnetic gauge transfor-mations inside a larger non-Abelian group structure. Such is the case, for example, in ‘grand unified theories’ of strong, weak and electromagnetic interactions.

/

46 13. Local Non-Abelian (Gauge) Symmetries

FIGURE 13.1

Vertex for isospinor-W interaction.

then

1

(iγμD'μ −m)ψ( )2 ' = 0, (13.38) proving the asserted covariance. In the same way, any free particle wave

1

equation satisfied by an ‘isospinor’ ψ( ) 2 – the relevant equation is determined by the Lorentz spin of the particles involved – can be made locally covariant by the use of the covariant derivative Dμ, just as in the U(1) case.

The essential point here, of course, is that the locally covariant form

in-1

2)’s and the gauge fields W μ, which are cludes interactions between the ψ(

determined by the local phase invariance requirement (the ‘gauge principle’).

Indeed, we can already begin to find some of the Feynman rules appropriate to tree graphs for SU(2) gauge theories. Consider again the case of an SU(2)

1

isospinor fermion, ψ( )2 , obeying equation (13.38). This can be written as

1 2

1

∂ −m)ψ( ) = g(τ /2)· /W ψ( )2 . (13.39) (i

In lowest-order perturbation theory the one-W emission/absorption process is given by the amplitude (cf (8.39)) for the electromagnetic case)

ig { ψ¯(

f

1 2

1

) ( ) 2

·W μd4 x (13.40) (τ /2)γμψi

exactly as advertized (for the field-theoretic vertex) in (12.129). The ma-trix degree of freedom in the τ’s is sandwiched between the two-component

1

isospinors ψ( )2 ; the γ matrix acts on the four-component (Dirac) parts of

1

2 The external W μ field is now specified by a spin-1 polarization vector ψ( ).

Eμ, like a photon, and by an ‘SU(2) polarization vector’ ar(r = 1, 2, 3) which tells us which of the three SU(2) W-states is participating. The Feynman rule for figure 13.1 is therefore

ig(τr/2)γμ (13.41)

which is to be sandwiched between spinors/isospinors ui, u¯f and dotted into Eμ and ar. (13.41) is a very economical generalization of rule (ii) in Comment (3) of section 8.3.1.

The foregoing is easily generalized to SU(2) multiplets other than doublets.

We shall change the notation slightly to use t instead of T for the ‘isospin’

46 13. Local Non-Abelian (Gauge) Symmetries

FIGURE 13.1

Vertex for isospinor-W interaction.

then

(iγμD'μ−m)ψ(12)'= 0, (13.38) proving the asserted covariance. In the same way, any free particle wave equation satisfied by an ‘isospinor’ψ(12)– the relevant equation is determined by the Lorentz spin of the particles involved – can be made locally covariant by the use of the covariant derivativeDμ, just as in the U(1) case.

The essential point here, of course, is that the locally covariant form in-cludes interactions between the ψ(12)’s and the gauge fields Wμ, which are determined by the local phase invariance requirement (the ‘gauge principle’).

Indeed, we can already begin to find some of the Feynman rules appropriate to tree graphs for SU(2) gauge theories. Consider again the case of an SU(2) isospinor fermion,ψ(12), obeying equation (13.38). This can be written as

(i/∂−m)ψ(12)=g(τ/2)· /Wψ(12). (13.39) In lowest-order perturbation theory the one-W emission/absorption process is given by the amplitude (cf (8.39)) for the electromagnetic case)

ig

{ ψ¯f(12)/2)γμψi(12)·Wμd4x (13.40)

exactly as advertized (for the field-theoretic vertex) in (12.129). The ma-trix degree of freedom in theτ’s is sandwiched between the two-component isospinors ψ(12); the γ matrix acts on the four-component (Dirac) parts of ψ(12). The externalWμ field is now specified by a spin-1 polarization vector Eμ, like a photon, and by an ‘SU(2) polarization vector’ar(r= 1,2,3) which tells us which of the three SU(2) W-states is participating. The Feynman rule for figure 13.1 is therefore

ig(τr/2)γμ (13.41)

which is to be sandwiched between spinors/isospinorsui,u¯f and dotted into Eμandar. (13.41) is a very economical generalization of rule (ii) in Comment (3) of section 8.3.1.

The foregoing is easily generalized to SU(2) multiplets other than doublets.

We shall change the notation slightly to use t instead of T for the ‘isospin’

13.1. Local SU(2) symmetry 47

quantum number, so as to emphasize that it is not the hadronic isospin, for which we retain T; t will be the symbol used for the weak isospin to be introduced in chapter 20. The general local SU(2) transformation for a t-multiplet is then

ψ(t) ψ(t)' = exp[igα(x)· T(t)(t) (13.42) where the (2t+ 1) × (2t+ 1) matrices Ti (t) (i= 1,2,3) satisfy (cf (12.47))

(t) (t) (t)

[Ti , Tj ] = iEijk Tk . (13.43) The appropriate covariant derivative is

Dμ = μ + igT(t) · W μ (13.44) which is a (2t+ 1) × (2t+ 1) matrix acting on the (2t+ 1) components of ψ(t). The gauge fields interact with such ‘isomultiplets’ in a universal way – only one g, the same for all the particles – which is prescribed by the local covariance requirement to be simply that interaction which is generated by the covariant derivatives. The fermion vertex corresponding to (13.44) is obtained by replacing τ /2 in (13.40) by T (t) .

We end this section with some comments:

(i) It is a remarkable fact that only one constant g is needed. This is not the same as in electromagnetism. There, each charged field interacts with the gauge field Aμ via a coupling whose strength is its charge (e,−e,2e,5e . . .).

The crucial point is the appearance of the quadratic g2 multiplying the commutator of the τ ’s, [τ · E,τ · W ], in the W μ transformation (equation (13.20)). In the electromagnetic case, there is no such commutator – the associated U(1) phase group is Abelian. As signalled by the presence of g2, a commutator is a non-linear quantity, and the scale of quantities ap-pearing in such commutation relations is not arbitrary. It is an instructive exercise to check that, once δW μ is given by equation (13.23) – in the SU(2) case – then the g’s appearing in ψ( 1 2 )' (equation (13.13)) and ψ(t)' (via the infinitesimal version of equation (13.42)) must be the same as the one appearing in δW μ .

(ii) According to the foregoing argument, it is actually a mystery why electric charge should be quantized. Since it is the coupling constant of an Abelian group, each charged field could have an arbitrary charge from this point of view: there are no commutators to fix the scale. This is one of the motivations of attempts to ‘embed’ the electromagnetic gauge transfor-mations inside a larger non-Abelian group structure. Such is the case, for example, in ‘grand unified theories’ of strong, weak and electromagnetic interactions.

Im Dokument Non-Abelian Gauge Theories (Seite 56-63)