Добавил:
Опубликованный материал нарушает ваши авторские права? Сообщите нам.
Вуз: Предмет: Файл:

Supersymmetry. Theory, Experiment, and Cosmology

.pdf
Скачиваний:
79
Добавлен:
01.05.2014
Размер:
12.6 Mб
Скачать

Symmetries 369

Note that a mass term for the vector field is forbidden by the gauge symmetry (A.31).

A.1.4 Nonabelian local gauge transformations: Yang–Mills theory

Following Yang and Mills [384], we now try to make the SU (2) isospin symmetry local. For future reference, we will adopt slightly more general notations than in Section A.1.2

and consider a general doublet Ψ(x) = ψ1(x) which transforms under SU (2) as

ψ2(x)

Ψ(x)

Ψ (x) = e−iαa(x)ta

Ψ(x) = U (αa(x)) Ψ(x)

(A.37)

 

 

 

 

where ta = σa/2 and the three parameters αa(x) a = 1, 2, 3, are now position dependent. Everything which follows in this section will be equally valid for a general symmetry group G: the ta are the generators of the group (a = 1, . . . , n = dimG) in an N -dimensional matrix representation; Ψ(x) is then an N -dimensional vector. In the case of SU (2) specifically considered here, n = 3 and N = 2.

Since µΨ(x) → U (α)µΨ(x) + [µU (α)]ψ(x), we must complete the spacetime derivative to make it covariant. By analogy with the abelian case, we introduce n vector fields Aaµ(x) (because there are n independent transformations associated with the n parameters αa(x)) and we define the derivative as

DµΨ(x) = µΨ(x) − igAµa (x)taΨ(x).

(A.38)

It is a covariant derivative if it transforms like Ψ itself, i.e. if

(DµΨ(x)) = U (α) DµΨ(x).

(A.39)

It is straightforward to check that this corresponds to transforming the gauge fields as follows (beware of noncommuting matrices!)

Aµata = U (α)AaµtaU (α)1 gi [µU (α)] U (α)1.

From now on, we will introduce the N × N matrix:

Aµ ≡ Aaµta.

More specifically, in our SU (2) example,

 

 

 

σ

a

 

 

Aµ3

 

Aµ1 −iAµ2

 

Aµ

Aµa

 

=

A1

2

 

 

 

2

 

.

 

 

 

 

 

 

2

+iA2

 

 

A3

 

 

 

 

 

 

µ

 

µ

 

µ

 

 

 

 

 

 

 

 

 

2

 

 

2

Then the transformation law (A.40) reads

Aµ = U AµU 1 gi µU U 1.

(A.40)

(A.41)

(A.42)

(A.43)

370 A review of the Standard Model and of various notions of quantum field theory

Comparison with the abelian case is more transparent in the case of infinitesimal transformations:

U (α) 1 − iαa(x)ta + · · ·

(A.44)

In order to expand (A.43) we need the commutator of two generators. In the general case, we have

ta, tb = iCabctc,

(A.45)

where the Cabc are the structure constants of the group G. Comparison with (A.26) shows that, in the case of SU (2), Cabc = abc. One then easily obtains from expanding (A.43) to linear order in αa,

a

a

abc

b

c

1

 

a

 

 

Aµ

= Aµ + C

 

α

Aµ

 

µα

 

+ · · ·

(A.46)

 

g

 

The last term is similar to the abelian case in (A.31). The second term is specific to the nonabelian situation: it involves the nonvanishing structure constants Cabc of the nonabelian gauge group. Introducing the generators T a in the adjoint representation ((T a)bc = −iCabc), one may write the infinitesimal transformation (A.46) as

 

1

&µαa − igAµc

(T c)ab αb'

1

 

 

δAµa

=

 

≡ −

 

Dµαa.

(A.47)

g

g

It remains to write a dynamics for the vector fields, and thus to define a field strength. By analogy with the abelian case, we may consider µAν − ∂ν Aµ. Using (see exercise 2)

µU 1 = −U 1 [µU ] U 1,

 

 

 

 

 

(A.48)

we obtain from (A.43)

 

 

 

 

 

 

 

 

 

 

 

µAν − ∂ν Aµ = U (µAν − ∂ν Aµ) U 1

 

 

 

 

 

 

 

 

 

U

1

 

U A

U

 

1

U U

 

1

+ iµU Aν 1

 

 

1 ν

 

 

µ

 

 

+

 

ν U U µU U

(µ ↔ ν) .

g

From the form of the first term on the right-hand side, it is clearly hopeless to expect a gauge invariant field strength as in the abelian case. We may at best expect that Fµν transforms as U Fµν U 1: a covariant field strength. Indeed, because

 

µAν

=

ig U [A

µAν ]

1

1

1 i

1

1

 

 

ig A

 

 

 

U

 

 

 

 

 

 

 

 

 

 

+ − ∂µU U , U Aν U +

 

µU U ν U U (µ ↔ ν) ,

 

 

 

 

g

we check that

 

 

Fµν = µAν − ∂ν Aµ − ig [AµAν ]

 

 

 

 

 

 

 

 

(A.49)

transforms covariantly:

 

 

= U Fµν U 1.

 

 

 

 

 

 

 

 

 

 

Fµν

 

 

(A.50)

Symmetries 371

We could have obtained directly the field strength from the commutation of two covariant derivatives (as in the abelian case):

[Dµ, Dν ] = −igFµν .

(A.51)

We also note that T r F µν Fµν is invariant. Since Fµν is a N × N matrix, we may write it Fµν = Fµνa ta where, from (A.45) and (A.49)

Fµνa = µAνa − ∂ν Aµa + gCabcAµb Aνc .

(A.52)

The complete Lagrangian reads

 

L = 21 Tr (F µν Fµν ) +

 

 

 

 

(A.53)

ψiγµDµψ − mψψ.

Normalizing the generators ta as

 

Tr tatb = 21 δab

(A.54)

as in (A.25), we have 12 Tr(F µν Fµν ) = 14 F aµν Fµνa . As in the abelian case, the fermion– gauge field coupling is of the form

 

 

 

Lint = gJAµa ,

(A.55)

a

¯

a

Ψ is the fermionic part of the Noether current.

 

where Jµ

= Ψγµt

 

 

The corresponding equation of motion is obtained by varying with respect to Aa .

It reads

µ

 

DµF aµν = µF aµν − gCbacAµb F cµν = −gJ,

(A.56)

where we have introduced the covariant derivative in the adjoint representation:

DρFµνb = ρFµνb − igAρa (T a)bc Fµνc , (T a)bc = −iCabc.

(A.57)

One important di erence between abelian and nonabelian gauge theories is the notion of charge or quantum number. In (A.29) and (A.32), we have included a real number q whose value is not fixed by the symmetry. If we try to introduce such a number in the nonabelian case and replace (A.37) and (A.38), respectively, by

Ψ (x) = e−iXαa(x)ta Ψ(x)

(A.58)

and

 

DµΨ(x) = µΨ(x) − ig X Aµa (x)taΨ(x),

(A.59)

then it is easy to show that DµΨ is covariant under an infinitesimal transformation only if X = 0 or 1. This shows that, in a nonabelian gauge theory, charges (i.e. quantum numbers) can only take quantized values fixed by group theory (i.e. matrix elements of generators ta). This will be one of the motivations for embedding the quantum electrodynamics U (1) symmetry into a nonabelian grand unified gauge symmetry group.

372 A review of the Standard Model and of various notions of quantum field theory

A.2 Spontaneous breaking of symmetry

A.2.1 Example of a global symmetry

We start with the simplest example of a global abelian symmetry and consider a complex scalar field φ(x) with a Lagrangian density

L = µφµφ − V (φφ).

(A.60)

This has a global phase invariance φ(x) → φ (x) = e−iθφ(x) ; and the associated conserved Noether current is Jµ = −iφ∂µφ+µφ. Requirement of renormalizability imposes to limit the interaction potential V to quartic terms:

V (φφ) = a φφ +

λ

(φφ)2,

(A.61)

2

 

 

 

where λ > 0 to avoid instability at large values of the field. The corresponding equation of motion reads:

(

 

+ a)φ =

λφ(φφ).

(A.62)

 

 

 

 

If a > 0, we may write a ≡ m2 and the equation of motion

( + m2)φ = j(φ)

(A.63)

where j(φ) ≡ −λφ(φφ) describes the self-interaction of the field φ. It is then possible to quantize the theory: first quantize the free field theory corresponding to vanishing self-interaction (the field obeys the Klein–Gordon equation ( + m2)φ = 0) and then use perturbation theory in λ.

We are interested in the vacuum structure of the theory, i.e. in identifying its ground state. Using Noether formalism, one may compute the energy of the system

P0 = d3x H(x)

H(x) = π∂0φ(x) + π0φ(x) − L

(A.64)

where π (resp. π) is the canonical conjugate momentum of φ(x) (resp. φ(x)) as in (A.16). This gives in the present case:

H(x) = 0φ(x)0φ(x) + iφ(x)iφ(x) + V (φφ).

(A.65)

Since the first two terms are positive definite, minimization of the energy density is obtained by setting them to zero, i.e. by taking a constant value for the field φ, and by minimizing the potential energy. In other words, we minimize the energy density by choosing a constant value φ0 which corresponds to the minimum of V . If a ≥ 0, the minimum is at φ = 0 (see Fig. A.1) i.e. for vanishing field. We are familiar with this situation: for example, the electromagnetic energy density 81π (E2 + B2) is minimized for E = B = 0. In this latter case, a nonvanishing electric or magnetic field in the vacuum (i.e. the ground state) would obviously break the isotropy of space. There is no such risk with a scalar field.

Spontaneous breaking of symmetry 373

V (φ)

Re φ

Im φ

Fig. A.1

Indeed, when a < 0, the ground state corresponds to a nonvanishing value for the scalar field: φ0φ0 = −a/λ. If we set a ≡ −m2, the equation of motion reads

( − m2)φ = j(φ).

(A.66)

In the limit of vanishing self-interaction (j →

0), the equation is not the standard

2

 

Klein–Gordon equation because of the sign of m and we thus do not know how to quantize the theory. Writing −m2 = (im)2, we may identify the origin of this sign as the instability of the system for small values of the field, i.e. around φ = 0 (see Fig. A.2). In other words, the problem might be that we evaluate the scalar field φ around an unstable value. It is then preferable to evaluate it around one of its stable values defined by φ0φ0 = m2.

We note that the potential has a rotation symmetry corresponding to φ → e−iθφ. On the other hand, the ground state of the system (which corresponds to the choice of one point on the set of degenerate minima) necessarily breaks this symmetry. We are in a situation of spontaneous symmetry breaking:

the Lagrangian (the Hamiltonian) is invariant under the symmetry;

the ground state (the vacuum) is not left invariant by the symmetry.

Such a situation is often encountered in classical physics with macroscopic objects (a ball at the top of a Mexican hat falls into the bottom of the hat, a vertical force down a column will shatter it into pieces). In ordinary quantum mechanics, one can always

374 A review of the Standard Model and of various notions of quantum field theory

V (φ)

φ 0

Re φ

 

Im φ

Fig. A.2

form linear combinations of ground states which are invariant under the symmetry. This is not possible in the case of quantum field theory because of the infinite number of degrees of freedom (infinite volume): the mixed Hamiltonian matrix elements between two di erent ground states vanish.

Now, following our remark above, let us return to our example and express the scalar field around its ground state value, say e0 "m2. We write

 

 

 

 

 

 

 

 

 

 

+ 2

!

 

 

 

 

 

m2

ϕ1

 

 

φ = e0

 

 

 

 

 

 

 

 

 

+

 

 

 

.

(A.67)

 

 

 

 

λ

 

 

 

 

 

 

2

Then

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

V (φφ) =

m4

+ m2

ϕ2

+ cubic and quartic terms.

(A.68)

2λ

 

 

1

 

 

 

 

 

 

 

 

 

We conclude that the field ϕ1 has mass squared 2m2 (beware of the normalization of the kinetic term 12 µϕ1µϕ1) whereas ϕ2 is massless. Thus the mass spectrum does not reflect the symmetry: even though ϕ1 and ϕ2 are interchanged in the symmetry transformation, they are not degenerate in mass. This is a consequence of the spontaneous breaking of the symmetry. Moreover, the fact that ϕ2 is massless is not a coincidence. Indeed, it is a general theorem due to [198] that to every continuous global symmetry spontaneously broken, there corresponds a massless particle. In the case of internal symmetries such as considered here, it is a boson called the Goldstone boson.

Spontaneous breaking of symmetry 375

Before proceeding to prove and discuss Goldstone’s theorem, let us stress one important property of a Goldstone boson. If we write the infinitesimal transformation δφ = −iθφ in the parametrization (A.67), we obtain

δϕ1 = θϕ2, δϕ2 = −θ

2

m2

(A.69)

λ θϕ1.

The Goldstone transformation is nonhomogeneous and is characterized by a constant term proportional to the field vacuum expectation value.

This is even better seen by using the following more appropriate parametrization of the scalar field φ:

φ(x) = ei(θ0+γ(x))

m2

 

λ + ρ(x) .

(A.70)

"

Since V (φφ) = V ([ρ + m2]2), the field γ(x) does not appear in the potential and is thus massless: it is the Goldstone boson, i.e. the massless excitation along the continuum of vacua. Under a gauge transformation we simply have γ(x) → γ(x) − θ. We will see that, when the symmetry is local (θ(x)), this allows us to gauge away the corresponding degree of freedom.

A.2.2 Goldstone theorem

We have defined two distinct notions of a symmetry:

A symmetry of the Hamiltonian: the corresponding charge commutes with the Hamiltonian

[Q, H] = 0

(A.71)

which is a statement of the conservation of the associated current

µjµ = 0.

It also implies that the finite transformation U ≡ e−iαQ invariant:

U HU = H = U HU.

(A.72)

leaves the Hamiltonian

(A.73)

A symmetry of the vacuum, i.e. a symmetry which leaves the ground state |0 invariant:

U |0 = |0

(A.74)

or

 

Q |0 = 0.

(A.75)

1In terms of the current jµ, (A.75) reads

376A review of the Standard Model and of various notions of quantum field theory

[83]has shown that a symmetry of the vacuum is a symmetry of the Hamiltonian1. But the converse is not true. If [Q, H] = 0 or µjµ = 0, then one may distinguish two cases:

(a) Wigner realization Q|0 = 0

This has well-known properties. Two states which transform into one another under the symmetry are degenerate in energy. Indeed, if φ = U φU , then

0Hφ |0 = 0|U φU HU φ U |0 = 0Hφ|0 ,

using both (A.73) and (A.74).

Also, Green’s functions (i.e. the vacuum values of chronological products of fields0|T φ(x1) · · · φ(xn)|0 ) are invariant under the symmetry: this leads to Ward identities which translate the symmetry at the level of Green’s functions. Indeed, under an infinitesimal transformation, using (A.17) (which is derived from (A.13) or (A.71)), we have, for x01 > x02 > · · · > x0n,

δ 0|T φ(x1) · · · φ(xn)|0 = 0|δφ(x1)φ(x2) · · · φ(xn)|0

+ 0(x1)δφ(x2) · · · φ(xn)|0 + · · ·

=iδα 0|((x1) − φ(x1)Q)φ(x2) · · · φ(xn)|0

+iδα 0(x1)((x2) − φ(x2)Q) · · · φ(xn)|0 + · · ·

=0.

(b)Goldstone realization Q|0 = 0

Two states which are exchanged under the symmetry are not necessarily degenerate. Hence the symmetry is not manifest in the energy (mass) spectrum. We have seen an explicit example of this in the previous Section. The symmetry is said2 to be spontaneously broken.

According to case (a), it su ces to prove that at least one Green’s function is not invariant: δ 0|T φ(x1) · · · φ(xn)|0 = 0. In most cases, it is the one-field Green’s function 0(x)|0 which is not invariant: in other words, the vacuum expectation value of a field is not invariant. This was precisely the case of the explicit example presented in the previous Section. We now prove Goldstone’s theorem in this case.

Defining

 

 

η ≡ 0|[Q(t), φ(0)]|0 = d3x 0| [J0(t, x), φ(0)] |0

(A.76)

we see that η = 0 because the symmetry is not an invariance of the vacuum (expressed as 0 = 0|δφ(x)|0 = i 0[Q, φ(x)]|0 using (A.17)) whereas dη/dt = 0 because the symmetry is an invariance of the Hamiltonian (i.e. the charge is a constant of motion: dQ/dt = i[H, Q] = 0).

d3xj0(t, x)|0 = 0. If |n is a state of vanishing 3-momentum, one may apply the bra n| to the previous relation. Thus, n| d3xe−ik·xj0(t, x)|0 = 0 and n|j0(x)|0 = 0 from which follows n|∂µjµ(x)|0 = 0. Lorentz invariance implies that this is true for any state |n and thus µjµ(x)|0 = 0 from which (A.72) follows.

2Somewhat improperly since the symmetry remains an invariance of the Hamiltonian.

Spontaneous breaking of symmetry 377

We introduce a complete basis of eigenstates |n of energy (H|n = En|n ) and

3-momentum (P|n = pn|n ). The zero of energy is taken at the ground state: H|0 =

#

E0|0 = 0, P|0 = p0|0 = 0. Then using the closure relation n |n n| = 1l, and translation operators:

J0(t, x) = e−ip·xeiHtJ0(0)e−iHteip·x,

we may write η as:

η =

d3x [ 0|J0(t, x)|n n|φ(0)|0 0(0)|n n|J0(t, x)|0 ]

n &

=d3x e−iEnteipn·x 0|J0|n n|φ(0)|0

n

'

 

−eiEnte−ipn·x 0(0)|n n|J0)|0

&

=(2π)3δ3(pn) e−iEnt 0|J0|n n|φ(0)|0

 

n

 

 

−eiEnt 0(0)|n n|J0(0)|0 '.

 

Obviously,

 

 

(A.77)

 

 

 

 

 

= −i

(2π)3Enδ3(pn)&e−iEnt 0|J0(0)|n n|φ(0)|0

 

 

 

 

 

 

 

 

 

dt

n

 

 

 

+ eiEnt 0(0)|n n|J0(0)|n '.

(A.78)

For this to be identically vanishing, whereas η = 0, there must be a state |n0 such that En0 δ3(pn0 ) = 0 (i.e. the state |n0 is a massless excitation of the system) and

0|J0(0)|n0 = 0 and n0(0)|0 = 0.

(A.79)

This massless excitation is the Goldstone boson.

A.2.3 General discussion in the case of global symmetries

We now consider the general case of a group G of continuous global transformations. The scalar fields are taken to be real and transform under G as:

φ (x) = e−iαata φ(x)

(A.80)

where the ta are the generators of the group, in a n-dimensional representation3. Infinitesimally

φ

(x) = φ

m

(x)

ata φ

n

(x)

(A.81)

m

 

 

mn

 

 

where m = 1, . . . , n.

3The ta are taken to be hermitian, ta= ta. Since φ is real, (ita) = ita. Hence ta = −ta and taT = −ta: the generators are antisymmetric.

378 A review of the Standard Model and of various notions of quantum field theory

 

Since the potential V (φ) is invariant, we have

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

∂V

 

αa tmna

φn = 0

 

 

 

(A.82)

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

∂φm

 

 

 

 

 

 

 

 

 

 

 

 

from which we obtain, by di erentiating with respect to φ ,

 

 

 

 

 

 

2V

 

 

 

φn +

∂V

 

 

αa tma = 0.

(A.83)

 

 

 

 

 

 

αa tmna

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

∂φm∂φ

 

 

 

 

 

∂φm

 

 

 

 

 

At the minimum φ0 of the potential, we thus have

 

 

 

 

 

 

 

 

 

 

 

M 2m αa tmna

φ0n = 0

 

 

 

(A.84)

where

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

2V

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

M 2m =

∂φ ∂φm

φ0

 

 

 

 

(A.85)

is the scalar squared mass matrix. In the case of spontaneous breaking, the vacuum

φ is not invariant and thus αata

φ

0n

is a nonzero eigenvector of the squared mass

0

 

 

 

 

 

mn

 

 

 

 

 

 

 

 

 

 

 

 

 

matrix with vanishing eigenvalue: it corresponds to a Goldstone boson.

 

It is possible to separate the generators ta (a = 1 · · · dim G) into:

Generators ti

which leave the vacuum invariant: tiφ0 = 0. These generators cor-

 

respond to the residual symmetry of gauge group H and i = 1, . . . , dim H.

Generators θα

which do not leave the vacuum

invariant: θαφ

= 0 (α = 1, . . . ,

 

 

 

 

α

 

 

0

 

dim G − dim H). Let us show that the vectors θ

 

φ0

are independent and thus

 

generate a

vector space of dimension dim G

dim H. We introduce a scalar prod-

 

 

 

n

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

uct (x, y) =

 

 

x yi and consider the matrix Xαβ = (θαφ0, θβ φ0). Since X is

 

 

 

 

i=1 i

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

can be diagonalized: D = OXOT

with

 

 

 

symmetric, it#

 

 

 

 

 

 

 

 

 

 

 

 

 

 

.

 

 

 

 

 

 

Dαα = Oαγ θγ φ0, Oαδ θδ φ0

(A.86)

But Oαγ θγ φ0 = 0. Hence all diagonal elements are positive definite and the vector space generated by the Oαγ θγ or the θγ is (dim G − dim H) dimensional.

[We may parametrize nonlinearly the scalar degrees of freedom as in (A.70) to make explicit the rˆole of the Goldstone bosons. Indeed, let us consider the function

F (g, x) = (φ(x), gφ0),

(A.87)

where g = eata is an element of G. Since, for every element h of the residual symmetry group H (0 = φ0), we have F (g, x) = F (gh, x), F is in fact a function on right cosets of G with respect to H, i.e. on G/H.4 Since G is compact, the continuous function

4One defines the right coset with respect to H of any element in G as the equivalence class g˙ = {γ G; h H, γ = gh} = gH. If H is an invariant subgroup of G, the set of these right cosets has a group structure, the quotient group G/H.