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

Supersymmetry. Theory, Experiment, and Cosmology

.pdf
Скачиваний:
55
Добавлен:
01.05.2014
Размер:
12.6 Mб
Скачать

Dynamical supersymmetry breaking: an overview 199

As a first example consider a U (1) supersymmetric gauge theory with two supermultiplets of charge ±1. We denote their scalar components by φ±. In the case of a

vanishing superpotential, the scalar potential is simply given by the D-term

 

1

g2 φ+

φ+ − φφ

2

 

 

V =

 

.

 

(8.11)

2

where g is the gauge coupling. The D-flat direction is reached for φ± = ρe

 

± . By a

suitable gauge transformation, one may choose θ+ = θ≡ θ. Thus

 

 

φ+ = φ= a,

a = ρeC

 

(8.12)

characterizes the flat direction. This flat direction consists of all the degenerate vacua; borrowing the terminology to solitons, in particular monopoles (see Chapter 4), one often refers to this set of all degenerate vacua as the classical moduli space.

For any nonvanishing value of a, the U (1) gauge symmetry is spontaneously broken and the vector field becomes massive. In fact, because supersymmetry is not broken (the D-term vanishes), a whole vector supermultiplet becomes massive. As we have seen several times earlier (for example, in Section 5.1.2 of Chapter 5), such a supermultiplet has the same number of degrees of freedom as a massless vector supermultiplet plus a chiral supermultiplet (three vector, four spinor and one scalar): this may be interpreted as the supersymmetric version of the Higgs mechanism. Since we introduced in the theory one vector and two chiral supermultiplets, we are left with a single chiral superfield to describe the light degrees of freedom of the e ective theory much below the scale of gauge symmetry breaking. The light scalar degree of freedom is obviously X = φ+φ, since it must be gauge invariant. In the vacuum which is labelled by a, one has simply X = a2. The field X is called a modulus: its vacuum expectation value labels the classical moduli space. The dynamical field X(xµ) corresponds to a continuous variation through the space of vacua as one moves in spacetime.

There is obviously no classical potential for X and no strong interaction present to generate it dynamically: W (X) = 0.

The K¨ahler potential, which fixes the normalization of the kinetic term, is obtained from the K¨ahler potential in the original theory. Assuming normalized kinetic terms

for φ+ and φ, we have

 

K (φ+, φ) = φ+φ+ + φφ.

 

 

 

 

 

(8.13)

Since XX = φφ

 

φφ

 

 

, we have φ

 

 

= φ

+

, along the flat direction φ

+

= φ

φ

+

φ

=

+

 

 

 

 

 

 

 

+

 

 

 

and thus

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

XX

 

 

 

 

 

K(X) = 2

 

 

 

 

 

 

 

 

 

 

 

(8.14)

 

 

 

 

 

 

 

 

XX.

 

 

 

 

 

 

The corresponding K¨ahler metric gives the kinetic for X

 

 

 

 

 

 

 

 

 

 

 

 

 

2K

1

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

µXµX =

2

 

 

µXµX.

 

 

 

(8.15)

 

 

 

 

∂X∂X

 

 

 

 

 

 

 

 

XX

 

 

 

 

We see that the theory is singular as X → 0. This generally means that one is missing degrees of freedom in the e ective theory. Indeed, as a → 0, the symmetry remains unbroken and the low energy theory involves the two chiral supermultiplets.

The rest of this chapter lies in the “Theoretical Introduction” track.

200 Dynamical breaking. Duality

8.2Perturbative nonrenormalization theorems

8.2.1Wilson e ective action

Requiring holomorphy with respect to some parameters of a supersymmetric theory seems to be in contradiction with the renormalization program. Indeed, in the context of supersymmetry, all renormalization e ects appear through the wave function renormalization factors Z. Since these are intrinsically nonholomorphic (recall that scalar kinetic terms are given by a real function, the K¨ahler potential), it seems that the relation between renormalized and tree-level couplings introduces nonholomorphicity which endangers any of the arguments presented above. The solution to this puzzle has been explained by [337, 338].

Let us illustrate this on the example of massive supersymmetric QED [338] (see Section C.3 of Appendix C or Exercise 5 of Chapter 3):

 

1

 

4

2

 

2

 

1

 

 

4

2

¯ ¯ 2

+

4

4

V

V

Φ

S =

 

 

d

xd

θW

 

+

 

 

 

d

xd

θW

d

xd

θ Φ+e

 

Φ+ + Φe

 

4g02

 

g02

 

 

 

+

d4xd2θ m0Φ+Φ+

 

d4xd2θ¯ m¯ 0Φ+Φ.

 

 

 

 

 

(8.16)

The low energy limit of this theory (i.e. the theory at a scale smaller to the mass scale m0) is a theory of free massless gauge boson and gaugino. The associated gauge coupling is simply given by the exact formula [337]

1

=

1

+

1

ln

ΛU V

,

(8.17)

g2

g02

 

4π2

m0

where ΛU V is the ultraviolet cut-o . Since the superpotential m0Φ+Φis not renormalized, one obtains for the low energy physical mass m = Zm0, where Z is the common field renormalization factor. One can thus express the low energy gauge coupling in terms of the physical mass as3:

1

=

1

+

1

ln

ΛU V

+

1

ln Z.

(8.19)

g2

g02

4π2

m

 

4π2

The factor Z is by essence nonholomorphic: it can be computed through the D-term renormalization and is a real function of ln |ΛU V /m|. Its presence accounts for the multiloop contribution.

It turns out that this nonholomorphic contribution is due to the presence of massless fields. The approach to e ective theories that we have implicitly followed is based on the e ective action formalism described in Section A.5.3 of Appendix Appendix A. The e ective action Γ, which generates proper Green’s functions, is expressed as a series in , which corresponds to an expansion in the number of loops. Because loop

3For a scale µ m, this becomes

 

 

 

 

ΛU V

 

 

 

 

 

1

=

1

+

1

ln

+

1

ln Z,

(8.18)

g2

g2

4π2

 

4π2

 

 

µ

 

 

 

 

 

 

0

 

 

 

 

 

 

 

 

which gives the standard super-QED relation β(α) = α2 (1 − γ(α)) with β(α) = ∂α/∂ ln µ and γ(α) = ln Z/∂ ln µ (see, for a similar relation in the Wess–Zumino model, (8.39)).

Perturbative nonrenormalization theorems 201

integration is performed down to zero momentum, infrared e ects due to the presence of massless fields are naturally included into the e ective action.

An alternative approach has been pursued by [367] and [310]. If one is interested in studying the e ects of a theory at a scale µ, one decomposes the quantum fields into a high frequency part (E > µ) and a low frequency part (E < µ). The Wilson e ective action is then obtained by integrating over the high frequency modes. At a scale µ m, it simply reads

 

 

S

W =

1

 

d4xd2θ W 2 + h.c.

(8.20)

 

 

4gW2

with

 

 

 

 

 

 

 

 

 

 

 

 

 

1

=

 

1

+

 

1

ln

ΛU V

, g02

= gW2

U V ),

(8.21)

 

g2

 

g2

4π2

µ

 

 

 

 

 

 

 

 

 

 

W

 

0

 

 

 

 

 

 

 

 

 

which does not contain higher powers of g i.e. it is renormalized only at one loop. Thus, the relation between the gauge coupling (8.19) appearing in the e ective action Γ and the one appearing in the Wilson e ective action SW is

1

=

1

+

1

ln Z.

(8.22)

g2

g2

 

4π2

Γ

 

W

 

 

 

 

 

The last term may be interpreted in terms of the [259]. The parameters of the Wilson e ective action are free from infrared contributions, and one-loop contributions exhaust all renormalization e ects (see (8.21)), which saves holomorphicity. This is not so for the e ective action Γ parameters. Infrared contributions lead to ln p2 terms at the loop level, which results in nonholomorphicity of the nonlocal action.

One may now turn to nonabelian gauge theories. In the case of a supersymmetric SU (N ) gauge theory, direct instanton calculations give the following result for the

gluino condensate

exp

 

 

 

1

8π2

 

λλ = CΛU3 V

 

 

,

(8.23)

g02

N g02

where C is a constant and the prefactor 1/g02 accounts for zero modes (infrared e ects).

Since we are discussing holomorphy, one may allow a θFµν ˜µν term which has a

F

physical e ect in an instanton background. Introducing thus the complex parameter

 

θ

4π

 

 

τ ≡

 

+ i

 

,

(8.24)

2π

g2

the gauge action simply reads

S =

1

τ

 

d4xd2θ WαWα .

 

16π Im

(8.25)

However, if we replace all factors 1/g02 in (8.23) by −iτ /4π, we find a dependence which is not periodic in the vacuum angle θ. Hence, the expression (8.23) does not seem to allow for a complex gauge coupling, i.e. to be consistent with holomorphy.

202 Dynamical breaking. Duality

Again, the way out of this dilemma is to go to the Wilson e ective action, whose coupling is given in terms of the coupling in the e ective action as:

1

 

1

 

 

N

1

 

 

 

 

 

 

=

 

 

ln

 

 

,

 

 

(8.26)

gW2

g02

8π2

g02

 

 

when θ = 0. Then, (8.23) simply reads

 

λλ

 

 

3

 

 

 

2

2

 

 

= CΛU V exp 8π /(N gW ) , which, once

the θ term is restored, reads

 

 

λλ = CΛU3 V exp (2iπτ /N ) = CΛU3 V exp

 

8π2

exp (iθ/N ) .

(8.27)

N gW2

We see that, as θ changes continuously from 0 to 2π, the N distinct vacua (with a phase dependence e2iπk/N , k = 0, . . . , N −1) are covered. Finally, just as in the abelian case, the evolution of the Wilson e ective action gauge coupling is purely one-loop:

1

=

1

3N

ln

ΛU V

,

(8.28)

 

gW2 (µ)

 

gW2

8π2

µ

which allows us to write the θ-independent part of the condensate (8.27) as 3 exp

 

8π2/N gW2 (µ) .

One may

note that, in the Wilson approach, cut-o s play a central rˆole. Just as

 

parameters can be viewed as the vevs of chiral fields, one may consider the cut-o s as field-dependent. In order to have a holomorphic description in the e ective theory, one may need to introduce di erent field-dependent regulators for the di erent sectors of the theory.

8.2.2Flat directions

To discuss the rˆole played by flat directions of the scalar potential, we have relied mostly on the tree level potential. A key property is that flat directions remain flat to all orders of perturbation theory [371]. We will prove here this statement in the context of global supersymmetry since this is the relevant framework when one discusses flat directions (see, for example, Section 6.12 of Chapter 6).

Flat directions are associated with the vanishing of auxiliary fields and are thus referred to as F -flat or D-flat directions. They are lifted if one of the auxiliary fields acquires a vacuum expectation value. The tree level potential is quadratic in these fields. In the context of renormalizable theories, a nontrivial vacuum appears if a

linear term is generated at higher orders.

 

d4xd2θR(x, θ) where R is a

A term linear in F would necessarily be of the form

superfield which depends only on x and θα. The

nonrenormalization theorems [210,235]

 

 

discussed in Chapter 1 precisely forbid the generation of such terms at higher orders. If a nonvanishing D-term is to lead to spontaneous supersymmetry breaking, it must be associated with a massless fermion (the Goldstino). This is only possible if the corresponding gauge symmetry is not broken. If the gauge symmetry group is semisimple (no U (1) factor), a term linear in D always appears multiplied with fields which transform nontrivially under the gauge symmetry. Since the symmetry must remain unbroken, these fields have a vanishing vacuum expectation value, which

Perturbative nonrenormalization theorems 203

forbids a nontrivial value for D. If the gauge symmetry is abelian, then a Fayet– Iliopoulos term is allowed and will be generated at higher orders, unless forbidden by symmetries (D is pseudoscalar). It has to be included from the start when discussing D-flat directions (see next section).

Thus, global supersymmetry remains unbroken to any finite order if it is unbroken at tree level. In other words, vanishing vacuum energy at tree level ensures vanishing vacuum energy at all finite orders: flat directions remain flat to all orders.

8.2.3More on holomorphy

We emphasized the rˆole of holomorphy in discussing F -flat directions. Surprisingly as it may seem, holomorphy also helps in classifying D-flat directions [1, 2, 55].

For example, when we considered in Section 8.1.3 the example of super-QED with two chiral superfields φ± of charge ±1, we concluded that the classical modulus space is parametrized by the holomorphic gauge invariant monomial X ≡ φ+φ.

A second example may be borrowed to the lepton sector of the MSSM. The D-term restricted to the lepton superfields reads:

 

 

1

 

 

 

 

1

g

2

 

 

 

 

 

 

 

 

i

 

 

 

 

1

 

 

 

 

 

 

 

 

 

VD =

2

g2 D12 + D22 + D32

+

2

 

2

 

 

 

DY2 ,

D1

=

 

i

(Ei Ni + Ni Ei) , D2 =

 

i

(Ei Ni − Ni Ei)

2

2

D3

=

1

 

|Ni|2 − |Ei|2 , DY =

 

2 |Eic|2 − |Ei|2 − |Ni|2 .

 

 

 

 

 

 

 

 

 

 

 

2

i

 

 

 

i

For any given set i0, j0, k0 (i0 = j0) in (1, 2, 3), the class of solutions:

Ni0

=

a

0

c

 

Li0 Ei0

0

, Lj0 = a

, Ek0

= a

(8.29)

(8.30)

corresponds to a flat direction of the potential VD. It may be parametrized by the gauge invariant monomial Li0 · Lj0 Ekc0 = a3, which is a holomorphic function of the superfields in the MSSM. This can obviously be generalized: every class of D-flat directions (see Exercise 1 for the MSSM) may be associated with a holomorphic gauge invariant monomial. Since this combination is allowed by gauge symmetry, it should be present in the superpotential and play a rˆole in the discussion of F -flat directions.

We see that the moduli space of D-flat directions is parametrized by a finite set of gauge-invariant polynomials holomorphic in the chiral superfields of the theory4. More precisely, the classical set of vacua can be parametrized in terms of holomorphic

4In a supersymmetric context, the gauge transformation Φ → e2igqΛΦ promotes the real gauge parameter to a complex parameter and thus the gauge symmetry group G to its complex extension Gc. In this context, the D-flatness condition may be seen as a gauge-fixing condition which breaks Gc down to G. Gauge-invariant holomorphic polynomials distinguish any two distinct Gc extended orbit [278].

204 Dynamical breaking. Duality

monomials subject to polynomial constraints. In the case of a nonvanishing superpotential, F -flatness conditions induce additional constraints, which are also obviously holomorphic in the fields.

The presence of a Fayet–Iliopoulos term tends to lift some of the flat directions (some fields acquire a vacuum expectation value of order the Fayet–Iliopoulos coupling; see for example Exercise 2) but remains usually insu cient to lift completely the degeneracy. In any case, the corresponding vacua are still associated with holomorphic polynomials.

As an important example for the rest of this chapter, we illustrate the above considerations by considering in some details a SU (3) gauge theory with two quark/antiquark

¯αi

¯

flavors [2]. We denote respectively by Qαi 3 and Q

3 the quark and antiquark

superfields as well as their scalar field components: α = 1, 2, 3 is a color index and i = 1, 2 refers to flavor. The D-term reads5

 

 

 

8

 

 

 

 

 

 

2

 

1

 

 

λa

β

 

 

λa β

 

 

VD =

 

g2

a=1 Q†αi

α

 

Qβi − Q¯

βi

 

α

Q¯αi

2

2

 

2

 

 

 

 

8

 

 

 

 

 

 

 

=

1

g2

 

 

2

 

 

 

(8.31)

 

λaαβ Dβ α

 

 

 

 

 

8

 

a=1

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

with

 

 

Dβ α = Qβi Q†αi − Q¯βi

Q¯αi.

 

 

 

 

 

 

(8.32)

We have used the fact that the Gell-Mann matrices are hermitian: λa = λa† = (λa )T . Then it is easy to show that the flat direction corresponds to

VD = 0 ↔ Dβ

α = ρ0 δβα, ρ0 R.

(8.33)

b#

8

 

8

 

 

To prove this, write D = ρbλb with λ0 1l. Then the flat direction VD = 0

=0

corresponds to λaα β Dβ α = 0 for all 1 ≤ a ≤ 8. Hence

Tr(λaλb)ρb = 2ρa = 0 and

 

=0

D = ρ01l.

b#

We next try to find a solution to (8.33). If we define Rαβ ≡ Q†αiQβi, then

det(Rαβ ) = Q1i Q2j Q3k εαβγ Qαi Qβj Qγk = 0,

since the latter part of the expression is antisymmetric under the exchange of i, j, k {1, 2}. Hence Rαβ is a positive semidefinite hermitian matrix of rank 2. Diagonalizing by a SU (3) rotation yields:

 

v12

 

 

0

 

Q†αi Qβi =

v22

.

(8.34)

5Note that, if φ → eaλa φ, then φ → e−iαaλa φ .

 

 

Key issues in dynamical breaking 205

¯

αi ¯

¯

αi ¯

is also given by

A similar reasoning for Q

Qβi

shows that ρ0 = 0 and thus Q

Qβi

(8.34). Using the symmetries to redefine the fields, one concludes that the vacuum solution is given by

Qβi = Q¯βi=

 

v δ

= 1, 2

 

i0βi

ββ = 3

 

 

 

 

 

 

 

 

 

 

 

which we may write, using a matrix notation,

 

 

 

 

v

0 0

 

Q = Q¯

= 01

v2 0 .

(8.35)

The fields of the low energy e ective theory should be SU (3) singlets. Thus the classical moduli space is labelled by the vacuum expectation values of the meson fields described

 

j

 

¯

αj

:

 

by the gauge invariant monomial Mi

 

≡ QαiQ

 

 

 

M

 

 

v12

0

.

(8.36)

0 v22

 

 

=

 

 

This is easily generalized to a gauge symmetry SU (Nc) with Nf flavors of quarks

¯αi

(Qαi, α = 1, . . . , Nc, i = 1, . . . , Nf ) and antiquarks (Q , α = 1, . . . , Nc, i = 1, . . . , Nf ). The classical moduli space is then given by:

 

 

 

 

 

 

 

Nc

 

 

 

 

 

 

 

 

 

 

 

 

 

if Nf < Nc,

Q = Q¯ = v1 . . .

 

 

 

Nf

 

 

 

 

 

 

 

 

 

 

0

 

 

 

 

 

 

 

 

 

 

 

 

 

 

v

Nf

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

j

 

¯

αj

 

 

 

 

The e ective theory consists of meson fields M

i

= QαiQ .

 

 

 

 

 

 

 

 

 

 

Nc

 

 

 

 

 

 

 

 

 

Nc

 

 

 

 

 

v1 . . .

 

 

 

 

 

 

. . .

 

 

 

 

 

 

 

 

 

v1

 

 

 

 

0

 

 

 

 

 

 

0

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Q¯ =

 

 

 

 

 

if Nf > Nc,

 

 

 

vNc

 

Nf

 

 

 

 

vNc

 

Nf

 

Q =

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

with |vf |2 − |v¯f |2 independent of f .

 

 

 

 

 

 

 

 

 

 

 

 

 

 

The e ective theory consists of:

 

 

 

 

 

 

 

 

 

 

 

 

 

 

meson fields

j

¯

αj

,

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Mi

= QαiQ

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

baryon fields

Bi1···iNc = εα1···αNc

Qα1i1 · · · QαNc iNc ,

 

 

 

 

antibaryon fields

B¯i1···iNc = εα1···αNc

Q¯α1i1 · · · Q¯αNc iNc .

 

 

 

 

8.3Key issues in dynamical breaking

We continue this presentation of the main aspects of dynamical symmetry breaking by discussing the variety of phases in gauge theories. We then describe some of the key tools necessary for a detailed analysis: renormalisation group, anomalies, decoupling.

206 Dynamical breaking. Duality

8.3.1Phases of gauge theories

The simplicity of gauge theories is somewhat misleading. They can represent very diverse regimes with strikingly di erent properties (mass gap, confinement, etc.) as one probes deeper into the infrared, i.e. larger distances. Let us consider two test charges separated by a large distance r.

If we are in a regime where there are unconfined massless gauge fields, the interaction potential V (r) is typically of the form g2(r)/r, where the gauge coupling is evaluated at the distance r. The gauge theory in this regime is not asymptotically free and the gauge coupling diverges at some ultraviolet scale Λ: g2(r) 1/ log (rΛ). Thus V (r) 1/ (r log(rΛ)). This is referred to as the free electric phase. But, it is possible that g2 is constant (free abelian gauge theory or renormalization group fixed point of a nonabelian gauge theory): this is the Coulomb phase where V (r) 1/r. The third possibility is the presence of massless magnetic monopoles. It is now 4π/g2 which is renormalized and V (r) log(rΛ)/r. This is known as the free magnetic phase.

There is, however, the further possibility that the gauge symmetry be spontaneously broken, in which case the gauge bosons become massive through the Higgs mechanism. The vacuum condensate provides a constant contribution to the potential, whereas the Higgs field exchange yields a Yukawa interaction, which is exponentially decreasing at large distances. Thus V (r) is constant in the Higgs phase.

Finally, in the case of asymptotically free nonabelian gauge theories, the interaction potential is confining at large distance: V (r) r.

8.3.2Renormalization group fixed points

In the renormalization group evolution, fixed points correspond to values of the coupling for which the beta function vanishes. The presence of nontrivial fixed points may enhance the symmetries of the theory. Indeed, at a fixed point, a nonsupersymmetric theory has conformal invariance. One may use this to prove general results [279]. For example, the scaling dimension d of a scalar field satisfies d ≥ 1, with equality only for a free field.

Supersymmetry may provide useful constraints on possible fixed points. As a first example, we consider the Wess–Zumino model and show that the nonrenormalization theorems discussed above forbid the existence of nontrivial fixed points [158]. We have seen in Section 8.1.2 that the superpotential (8.3) is not renormalized. This leaves us only with wave function renormalization. Adding the appropriate counterterm to the scalar field kinetic term in the renormalized Lagrangian yields

Lkin = 21 Z ∂µφ∂µφ,

(8.37)

where Z is the wave function renormalization constant: the bare field is φ0 = Z1/2φ. Then, expressing the superpotential (8.3) in terms of φ0,

W (φ) = 21 mZ1 φ02 + 31 λZ3/2 φ03

(8.38)

gives the renormalized coupling λ in terms of the bare coupling λ0: λ = Z3/2λ0. We deduce a relation between the beta function β(λ) and the anomalous dimension γ(λ) of the field φ:

β(λ) = µ

= 3

λµ

d ln Z

= 3

λγ(λ).

(8.39)

 

 

 

2

 

2

 

 

 

 

 

 

 

 

Key issues in dynamical breaking 207

This relation is obviously a mere consequence of the nonrenormalization theorems. It forbids any nontrivial fixed point. Indeed if λ = 0 was such a point (β(λ ) = 0), it would follow that γ(λ ) = 0. Since its anomalous dimension vanishes, the scalar field has its scaling dimension d coinciding with its canonical dimension: d = 1. Conformal symmetry implies that the field is free, in contradiction with our starting assumption that we are in a nontrivial fixed point regime.

For our next example, let us turn to nonabelian gauge theories. We suppose that the beta function reads

β(g) =

1

−b1g3 + b2g5 + O(g7)

(8.40)

16π2

with b > 0, b > 0 and b b . There is then a nontrivial infrared fixed point

"1 2 1 2

g b1/b2 in the perturbative regime (see Fig. 8.1).

If we take the example of nonsupersymmetric SU (Nc) with Nf flavors, then b1 = 113 Nc 23 Nf whereas b2 is of order Nc2 in the limit of large Nc. Thus if we choose enough

flavors to almost compensate the one-loop beta function (Nf 11Nc/2), then b1 is of order 1 and g is or order 1/Nc at large Nc, hence in the perturbative regime [21].

In the supersymmetric case, we have [298]

 

 

β(g) =

g3

 

3Nc − Nf + Nf γ(g2)

,

(8.41)

16π2

1 − Ncg2/(8π2)

 

 

 

 

γ(g) =

g2

 

Nc2 1

+ O(g4),

 

 

 

 

 

 

 

8π2

 

Nc

 

 

where γ(g) is the anomalous mass dimension. If again we choose a regime of large Nc where the one-loop beta coe cient almost cancels, i.e. Nf /Nc = 3with 1, then we have a zero of the beta function for γ = 1 3Nc/Nf − /3: we find a nontrivial infrared fixed point at

 

8π2

 

g 2Nc =

 

+ O( 2).

(8.42)

3

 

 

 

The theory is then in a Coulomb phase, which is very di erent from the confining phase observed with a smaller number of flavors, e.g. QCD with three flavors.

β (g)

g

g

Fig. 8.1 The beta function of equation (8.40): the arrows denote how g(µ) varies as one goes into the infrared, i.e. for decreasing µ.

208 Dynamical breaking. Duality

In the case of supersymmetric theories, the fixed point regime has superconformal invariance, which has some nontrivial consequences. For example, if we consider the supermultiplet of currents (Tµν , J, JµR) which consists of the energy–momentum tensor, the supersymmetric current and the R-symmetry current (see Section 4.2 of Chapter 4), superconformal invariance implies:

Θµµ = 0, γrsµ J= 0, ∂µJµR = 0.

(8.43)

Hence, at the fixed point, the R-symmetry is conserved. Moreover scaling dimensions of superfields satisfy

d ≥ 23 |R|,

(8.44)

where R is the R-charge, with equality for chiral or antichiral superfields. We will make use of these results below.

8.3.3Symmetries and anomalies. The ’t Hooft consistency condition

In a confining regime (such as in QCD), it is important to identify correctly the light (massless) degrees of freedom. This is not a problem for QCD since we observe them: they are for example the nucleons. When we consider other theories, identifying them might be a challenge.

’t Hooft [349] has devised a general consistency condition which has proved to be a very powerful tool to identify massless fermionic bound states. It rests on a clever use of the cancellation of anomalies.

Following ’t Hooft, we consider a Yang–Mills theory with gauge symmetry group G, coupled to chiral fermions in various representations of G. These massless fermions form multiplets of a global symmetry group GF . If we now consider the low energy theory, one encounters massless composite bound states. If we now consider three currents associated with the global symmetry GF , the corresponding triangle anomalies may be computed using the elementary fermions of the short-distance description or using the composite massless bound states. Consistency requires that the two computations coincide.

The proof is rather straightforward. Let us gauge GF by introducing weaklycoupled “spectator” gauge bosons. In order to cancel the anomalies of the short distance theory, one must introduce massless “spectator” fermions. At low energy (large distance), one finds the gauge symmetry GF with the massless bound states and the “spectator” fermions: anomalies must cancel i.e. the contribution of spectators must cancel that of bound states. Hence the latter coincides with the anomalies computed at short distance with the elementary (nonspectator) fermions.

8.3.4E ective theories and decoupling

Another way of extracting information is based on the decoupling of heavy flavors [8]. If we make one of the fundamental fermions massive, then, in the limit of large mass, this flavor decouples: in the low energy theory, all bound states containing this fermion disappear. In the context of supersymmetric theories, this means that if we start with a theory with Nf + 1 massless flavors with superpotential W 1, . . . , Φn+1) and make the (n+1)th flavor heavy by adding a mass term mΦ2n+1, one may solve the F -flatness condition for Φn+1 and thus eliminate this field from the low energy action, recovering an e ective theory with Nf fundamental flavors.