122
Introduzione alle teorie supersimmetriche Giovanni Ridolfi - INFN Genova Dottorato di Ricerca - Pavia, 6-8 Giugno 2007

Introduzione alle teorie supersimmetricheridolfi/pavia07.pdf · spinor ˘R under Lorentz transformations are the same as those of ˘ L, where is the antisymmetric matrix = 0 @ 0 1

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Introduzione alle teorie supersimmetriche

Giovanni Ridolfi - INFN Genova

Dottorato di Ricerca - Pavia, 6-8 Giugno 2007

Lay-out:

1. Why supersymmetry

2. General features of a supersymmetric theory

3. The minimal supersymmetric Standard Model

4. Breaking supersymmetry

5. Present experimental status

6. Conclusions and outlook

In many respects

we are happy with the Standard Model

• a unitary, renormalizable relativistic quantum field theory

• amazingly consistent with data.

The problem of spontaneous gauge symmetry breaking (short

range of weak interactions) is solved in a simple way by the Higgs

mechanism.

This also allows breaking of the flavour symmetry in a

phenomenologically consistent way (flavour breaking confined in

the charged-current interaction sector).

Achieving the same results without the Higgs mechanism is

extremely difficult.

At the same time,

we are unhappy with the Standard Model.

Many unsatisfactory aspects. Among others:

• What is the origin of flavour symmetry breaking?

• Is there a grand unification?

• Where is gravitation?

The last two points raise further problems:

• Hierarchy

• Naturalness

Two facts:

• The masses of all known particles (including the minimal

Standard Model Higgs boson) are not far from the weak scale,

∼ 200 GeV.

• Much larger energy scales become relevant at some point

A first question arises: why is the weak scale so much smaller than

the Planck scale, MP ∼ 1019 GeV, or the unification scale,

MGUT ∼ 1016 GeV?

This is usually referred to as the hierarchy problem.

Even worse: The Higgs mass (masses of scalar particles, in general)

is strongly sensitive to any large energy scale unless a fine tuning of

parameters is performed.

This is the so-called naturalness problem.

There is a very simple reason for this: scalar masses are not

naturally small, in the sense that no symmetry is recovered when

they are let go to zero.

Fermion and vector boson masses are naturally small: radiative

corrections are proportional to the masses themselves.

Naturalness and fine tuning in a simple example

Consider a theory of two real scalars fields:

L =1

2∂µφ∂µφ+

1

2∂µΦ ∂µΦ − V (φ,Φ)

with

V (φ,Φ) =m2

2φ2 +

M2

2Φ2 +

λ

4!φ4 +

σ

4!Φ4 +

δ

4φ2Φ2

Assume λ, σ, δ are all positive, small and comparable in

magnitude, and assume M 2 � m2 > 0.

Is the mass hierarchy m2 �M2 conserved at the quantum level?

Compute one-loop radiative corrections to m2 (e.g. by taking the

second derivatives of the effective potential at the minimum

φ = Φ = 0):

m2one loop = m2(µ2) +

λm2

32π2

(

logm2

µ2− 1

)

+δM2

32π2

(

logM2

µ2− 1

)

µ2 ∂m2

∂µ2=

1

32π2

(

λm2 + δM2)

Corrections proportional to M 2 appear at one loop. One can

choose µ2 ∼M2 in order to get rid of them, but they reappear

through the running of m2(µ2).

The mass hierarchy is preserved only if the parameters are such

that

λm2 ∼ δM2 → δ

λ∼ m2

M2

This is what we usually call a fine tuning of the parameters.

The same thing happens if m2 < 0, M2 �∣

∣m2∣

∣ > 0. In this case the tree-level

potential has a minimum at

Φ = 0, φ2 = −6m2/λ ≡ v2

and the symmetry φ → −φ is spontaneously broken. The degrees of freedom in

this case are Φ and φ′ ≡ φ − v, with

m2Φ = M2 m2

φ′ = −2m2 = λv2/3

At one loop, the minimization condition m2 + λv2/6 = 0 is replaced by

m2+

λv2

6= −

λ

32π2

(

m2+

λv2

2

)(

logm2 + λv2

2

µ2−1

)

δ

32π2

(

M2+

δv2

2

)(

logM2 + δv2

2

µ2−1

)

Following the same procedure as in the unbroken case one finds

m2φ′ =

λv2

3+

v2

32π2

[

λ2 logm2 + λ

2v2

µ2+ δ2 log

M2 + δ2v2

µ2

]

with v ∼ M without a suitable tuning of the parameters.

Naturalness: a closer look

The scalar potential in the Standard Model:

V (φ) = m2 |φ|2 + λ |φ|4

One-loop corrections to m2 due to fermionic (a) or bosonic (b)

degrees of freedom:

(a)

H

f

(b)

H

S

(

∆m2)

a=

|λf |216π2

[

−2Λ2 + 6m2f log

Λ

mf+ . . .

]

(

∆m2)

b=

λS

16π2

[

Λ2 − 2m2S log

Λ

mS+ . . .

]

Here Λ is an ultraviolet cut-off, to be identified with the energy

scale at which the SM is no longer reliable, and the dots stand for

terms that do not grow with Λ.

In dimensional regularization the Λ2 term would be absent, but

contributions proportional to m2f ,m

2S would still be there.

Even if the heavy degrees of freedom are not directly coupled to

the SM Higgs, it can be shown that similar contributions arise at

higher orders.

In the absence of very special cancellations, the Higgs boson

becomes as heavy as the heaviest degrees of freedom.

A symmetry that relates fermions to bosons would do the job, at

least at one loop. Suppose there are two scalars for each fermion:

(

∆m2)

a+b=λS − |λf |2

8π2Λ2 + . . .

For suitable values of the couplings the quadratic divergence

disappears.

No surprise: with bosons and fermions in the same multiplet,

scalar masses are protected by the same (chiral) symmetry that

protects fermion masses from large radiative corrections.

Clearly, more restrictions will be needed in order to guarantee that

the cancellation takes place at all orders.

Such a symmetry is called a supersymmetry:

Q|boson〉 = |fermion〉 Q|fermion〉 = |boson〉

The symmetry generator Q (and its hermitian conjugate Q†) carry

spin 1/2: it is a space-time symmetry.

The form of possible supersymmetry algebras is strongly

constrained on the basis of very general theorems in field theory.

For example, it is impossible with ordinary symmetry generators

(elements of a commutator algebra).

There is essentially one possibility:

{Q,Q†} = Pµ

{Q,Q} = {Q†, Q†} = 0

[Pµ, Q] = [Pµ, Q†] = 0

(more on this later). Further specifications:

• Q,Q† transform as spinors under the Lorentz group

• Q,Q† commute with gauge symmetry generators.

In principle, we may have more than one Q: Qi, i = 1, . . . , N

(extended supersymmetry).

Historical remark: symmetries that relate particles with different spin first

studied in the context of approximate symmetries of hadrons. Non relativistic

quark models have an approximate SU(6) symmetry (3 quark flavors with spin

1/2), which is observed hadron spectrum. It relates hadrons with the same

flavor content, but different spin (e.g. K ↔ K∗). Consequence of approximate

spin and flavor independence of quark-quark forces.

The Coleman-Mandula theorem tells us that this property cannot be

possessed by a relativistic theory: with some reasonable assumptions, the most

general Lie algebra of symmetry operators that commute with the S matrix

consists of Poincare generators Pµ and Jµν , plus ordinary internal symmetry

generators that act on one-particle states with matrices that are diagonal in,

and independent of, momentum and spin. Crucial point: the Poincare group is

non compact, it has no non-trivial unitary representations.

Rules out SU(6), but also Lie algebras of supersymmetry generators.

A few basic properties of a supersymmetric theory can already be

recognized:

• particles in the same supersymmetric multiplet (which we will

call a supermultiplet) have equal masses and equal gauge

transformation properties (electric charge, weak isospin and

color)

• within the same supermultiplet, there is an equal number of

bosonic and fermionic degrees of freedom (a proof on the next

slide)

A proof (taken from S. Martin):

(−1)2s|boson〉 = +|boson〉

(−1)2s|fermion〉 = −|fermion〉

}

⇒{

(−1)2s, Q}

={

(−1)2s, Q†}

= 0

Consider the subspace of states |i〉 within a supermultiplet with the same

eigenvalue pµ of the four-momentum operator P µ.∑

i|i〉〈i| = 1 within this

subspace of states.

i

〈i|(−1)2sPµ|i〉 =∑

i

〈i|(−1)2sQQ†|i〉 +∑

i

〈i|(−1)2sQ†Q|i〉

=∑

i

〈i|(−1)2sQQ†|i〉 +∑

i

j

〈i|(−1)2sQ†|j〉〈j|Q|i〉

=∑

i

〈i|(−1)2sQQ†|i〉 +∑

j

〈j|Q(−1)2sQ†|j〉

=∑

i

〈i|(−1)2sQQ†|i〉 −∑

j

〈j|(−1)2sQQ†|j〉

= 0.

Fermionic fields: a reminder

We use the Weyl representation of the Dirac matrices:

γµ =

0 σµ

σµ 0

σµ = (σ0, ~σ) σµ = (σ0,−~σ) γ5 =

−1 0

0 1

where ~σ are the three Pauli matrices

σ1 =

0 1

1 0

σ2 =

0 −ii 0

σ3 =

1 0

0 −1

and

σ0 =

1 0

0 1

The familiar lagrangian for a free, massive Dirac spinor is

L = ψ (i∂/−m)ψ → (i∂/−m)ψ = 0

Four-component Dirac spinors realize a reducible representation of

the Lorentz group:

ψ =

ξL

ξR

ψL =

ξL

0

=1

2(1 − γ5)ψ ψR =

0

ξR

=1

2(1 + γ5)ψ

the two-component spinors ξL, ξR transform independently under

Lorentz transformations.

In terms of ξL, ξR we have

L = iξ†L σµ∂µ ξL + iξ†R σ

µ ∂µξR −m(ξ†LξR + ξ†RξL)

It can be shown that the transformation law of the two-component

spinor ξR under Lorentz transformations are the same as those of

εξ∗L, where ε is the antisymmetric matrix

ε =

0 1

−1 0

.

It follows that any Dirac spinor may always be written in terms of

two left-handed Weyl spinors ξ, χ as

ψ =

ξ

−εχ∗

(the minus sign on the second Weyl spinor is conventional).

The free lagrangian for a massive Dirac spinor is

L = ψ (i∂/−m)ψ = iχT εTσµε ∂µχ∗ + iξ† σµ∂µξ −m(χT εξ − ξ†εχ∗)

= iχ† σµ ∂µχ+ iξ† σµ∂µξ −m(χT εξ − ξ†εχ∗)

Note that

χT εξ = ξT εχ

because of the anticommutation properties of fermion fields, and

that

(χT εξ)† = ξ†εTχ∗ = −ξ†εχ∗

The shorthand notation

χξ = ξχ = χT εξ ξ†χ† = χ†ξ† = −ξ†εχ∗

is often used.

Any theory involving spin-1/2 particles may be written as a

collection of left-handed Weyl spinors ψi.

Left-handed fermions in the Standard Model:

ui, di, ei, νi

ui, di, ei

where i = 1, 2, 3 is a generation index. For example, the Dirac

spinor for the up quark is written in terms of two left-handed Weyl

spinors u and u as

ψu ≡

uL

uR

=

u

−εu∗

(NB: the bar on u here has no special meaning, it is just part of

the name)

The Weyl notation is convenient, since left- and right-handed

components of Dirac fermions, which behave differently in weak

interactions, are treated separately.

The simplest (irreducible) representations of supersymmetry

contain Weyl fermions.

Massless supermultiplets of N = 1 supersymmetry are formed with

a state with helicity λmax and a state with helicity λmax − 1/2.

The simplest supermultiplet candidate: λmax = 1/2, λmin = 0

a Weyl fermion ψ and two real (or one complex) scalars φ1, φ2.

This is called a chiral or matter supermultiplet.

Next-to-simplest possibility: λmax = 1, λmin = 1/2

a massless gauge vector boson Aaµ and a Weyl fermion, called a

gaugino, λa

This is called a vector supermultiplet.

Gauge bosons belong to the adjoint representation of the gauge

group, so the same is true for their supersymmetric partners.

Thus,

gauginos are not chiral fermions

because the adjoint representation is equivalent to its conjugate.

The pair Aaµ, λ

a is called a gauge supermultiplet.

It turns out that any other representation of the supersymmetry

algebra can be reduced to a combination of chiral and gauge

supermultiplets, at least if there is only one supersymmetry

generator Q (N = 1 supersymmetry).

We will not consider extended (N > 1) supersymmetries, because

they cannot accommodate chiral fermions.

A word on extended supersymmetries

If λmax is the maximum helicity in a supermultiplet, then λmin = λmax − N/2.

Hence, in order to exclude states with |λ| > 2, we cannot allow N > 8.

N = 8 and N = 7 supergravity theories are entirely equivalent (same particle

content); N ≤ 6 are all different from each other.

Global supersymmetry (no spin 2 and 3/2 particles) is possible with N ≤ 4.

N = 4 and N = 3 global supersymmetries equivalent.

Fermions in extended supersymmetries are either partners of gauge bosons, or

have both helicity components in the same supermultiplet; impossible to

accommodate chiral fermions.

The particle content of a supersymmetric SM

• No supermultiplet can be formed out of standard particles

(e.g. γ, ν).

• Matter fermions must belong to chiral supermultiplets, because

left and right fermions transform differently under the weak

gauge group.

• Gauge bosons must go into gauge supermultiplets.

• At least two Higgs doublets, with their fermionic partners:

cancellation of the axial anomaly.

spin 0 spin 1/2 spin 1 SU(3)C SU(2)L U(1)Y

uL, dL uL, dL 3 2 + 13

uR uR 3 1 + 43

dR dR 3 1 − 23

ν, eL ν, eL 1 2 −1

eR eR 1 1 −2

H+u , H

0u h+

u , h0u 1 2 +1

H0d , H

−d h0

d, h−d 1 2 −1

g g 8 1 0

w±, w0 W±,W 0 1 3 0

b0 B0 1 1 0

Supersymmetric partners of standard particles (e.g. a scalar

electron) with the same masses would have been detected in

experiments. Since none of them has been observed so far, we

must conclude that

supersymmetry must be broken

in a realistic theory. Recall our formula for scalar mass corrections:

(

∆m2)

a+b=λS − |λf |2

8π2Λ2 + . . .

Supersymmetry forces λS = |λf |2 to all orders in perturbation

theory, so that quadratic divergences are systematically cancelled.

This feature must be preserved in the broken theory:

L = Lsupersymmetric + Lsoft

where Lsoft only contains mass terms and couplings with positive

mass dimension.

Supersymmetric partners of ordinary particles are so heavy that

they have escaped detection so far.

Is there a reason for that?

All ordinary particles, including the W,Z bosons and the top

quark, would be massless in the absence of spontaneous breaking

of the electroweak gauge symmetry.

The contrary is true for their partners: scalar masses are always

allowed by gauge symmetries, and gaugino can be massive because

they belong to a real representation of the gauge group.

The only exception is the Higgs boson.

A rough estimate of superparticle masses

Call msoft the largest mass scale present in Lsoft. Corrections to m2

arising from Lsoft must vanish as msoft → 0, so they cannot grow as

Λ2 (corrections proportional to msoftΛ are also forbidden: UV

divergences are either quadratic or logarithmic). Therefore

∆m2 ∼ λm2soft log

Λ

msoft

(λ a generic coupling in Lsoft). It follows (λ ∼ 1,Λ ∼MP ) that

msoft<∼ 1 TeV

in order to get the correct value of the Higgs v.e.v. Furthermore,

m2F −m2

B ∼ m2soft

within a supermultiplet: superpartners cannot be too heavy.

Building a supersymmetric theory

A simple example: one left-handed Weyl fermion

ψ ≡ PL ψ PL,R =1

2(1 ∓ γ5)

and one complex scalar φ, no masses, no interactions:

Lchiral = Lf + Ls

Lf = ψ i∂/ ψ

Ls = ∂µφ∗∂µφ

Is it possible to define a set of transformation rules on the fields,

such that scalars are transformed into fermions, under which Lchiral

remains unchanged?

Let us consider the scalar field first, and tentatively define

δφ = η ψ δφ∗ = ψ η

where η is a constant, infinitesimal, anticommuting spinor. Then

δLs = (∂µψ) (∂µφ) η + η (∂µφ∗) (∂µψ)

How does ψ tranform? δψ must be left-handed, linear in η, and

must contain one derivative. Only one possibility (up to a

proportionality factor):

δψ = −i(∂µφ) γµη δψ = iη γµ∂µφ∗

which tells us that η is right-handed, PRη = η, and gives

δLf = −η (∂µφ∗) γµγν∂ν ψ + ψ γνγµ∂ν (∂µφ) η

Using {γµ, γν} = 2gµν and ∂µ∂ν = ∂ν∂µ we get

δLf = −η (∂µφ∗) (∂µψ) − (∂µψ) (∂µφ) η

−η ∂ν [(∂µφ∗) γµγν ψ] + η ∂µ[(∂µφ∗)ψ] + ∂µ(ψ ∂µφ) η

The first row cancels against δLs, and the rest is a total derivative.

So,

δ

d4xLchiral = 0

Not yet enough to declare that Lchiral is supersymmetric: we must

check that the commutator of two transformations is a symmetry

of the theory. For the scalar fields we find

(δη1δη2

− δη2δη1

)φ = −(η2 γµ η1 − η1 γ

µ η2) i∂µφ

Good news: i∂µ is nothing but the four-momentum operator Pµ,

the generator of space-time translations, a symmetry of

space-time.

Now consider the fermion fields. After some tedious algebra,∗ we

get

(δη1δη2

− δη2δη1

)ψ = −(η2 γµ η1 − η1 γ

µ η2) i∂µψ

+1

2(η2 γ

µ η1 − η1 γµ η2) iγµ∂/ψ

The first term is similar to what we got for φ, while the second

one vanishes on the mass shell,

∂/ψ = 0.

OK, but it might be useful to close the algebra even off shell.

∗ For those who wish to try: you need the Fierz identities (and some patience).

This can be done. Let us introduce an auxiliary scalar field F , with

the lagrangian

La = F ∗F

(note the unusual dimension: m2 instead of m). It is harmless: the

equation of motion is F = 0. Now assume

δF = −i η γµ ∂µψ δF ∗ = i ∂µψ γµ η

and modify the fermion transformation rules as follows:

δψ = PL [−i∂µφ γµη + Fη] δψ = [iη γµ∂µφ

∗ + ηF ∗]PR

(which amounts to nothing, because the auxiliary field vanishes on

shell).

Things start getting complicated:

1. We can no longer assume that PRη = η, we must promote it to a

full four-component spinor. In order not to increase the

number of independent parameters, we assume it is a Majorana

spinor:

η =

ηL

ηR

ηL = ε η∗R

2. We now have to write explicitly the projector PL in front of δψ,

in order to keep ψ left-handed.

Nevertheless, let us press on ...

... It is immediate to check that

Lchiral = Ls + Lf + La

is our first supersymmetric lagrangian (not very exciting: it’s a

non-interacting theory), invariant under the transformation

δφ = η ψ δφ∗ = ψ η

δψ = PL [−i∂µφ γµη + Fη] δψ = [iη γµ∂µφ

∗ + ηF ∗]PR

δF = −i η γµ ∂µψ δF ∗ = i ∂µψ γµ η

with

(δη1δη2

− δη2δη1

)X = −(η2 γµPR η1 − η1 γ

µPR η2) i∂µX

for X = φ, ψ, F , both on and off the mass shell. This structure can

be replicated for an arbitrary number of chiral supermultiplets.

The same results take a much simpler form when recast in Weyl

language. Define Weyl spinors through

ψ →

ψ

0

η →

η

−εη∗

Recalling the definitions of γ matrices, we get

δφ = η ψ

δψ = i(σµεη∗) ∂µφ+ Fη

δF = −i η† σµ ∂µψ

and

(δη1δη2

− δη2δη1

)X = −(η†1 σµ η2 − η†2 σ

µ η1) i∂µX

Why do we need auxiliary fields? Count fermionic and bosonic

degrees of freedom:

• on shell: ψ has two helicity states (it is a Weyl fermion): nf = 2

φ is a complex scalar: nb = 2

• off shell: ψ is a complex two-component spinor: nf = 4

φ is a complex scalar: nb = 2

F is a complex scalar: nb = 2

Without the auxiliary fields, the counting rule nf = nb would be

violated off the mass shell.

Ordinary symmetries: a reminder

Consider a lagrangian density

L(φ, ∂φ)

functions of a set of fields φi, i = 1, . . . , n. The transformation

δφi = iηATAijφj ;

(

TA)†

= TA;[

TA, TB]

= ifABCTC

is a symmetry transformation if

δL = ηA∂µKAµ ⇒ δS[φ] = δ

d4xL = 0

As a consequence, a set of conserved currents exists:

∂µJAµ = 0; JA

µ = TAijφj

∂L∂∂µφi

−KAµ

QA(t) =

d3x JA0 (t,x);

d

dtQA(t) = 0;

[

QA, QB]

= ifABCQC

The same procedure can be applied to our supersymmetry: a

conserved (spinor) current Jµ and a conserved charge

Q =√

2

d3x J0(x)

can be built by tedious but straightforward calculations

The supersymmetry charge Q (a Weyl spinor!) obey the

anticommutator algebra

{

Q,Q†}

= 2σµPµ; {Q,Q} ={

Q†, Q†}

= 0

as anticipated.

Summary:

Lchiral = ψ† iσµ∂µ ψ + ∂µφ∗∂µφ+ F ∗F

(ψ left-handed fermion, φ and F complex scalars) is invariant under

the supersymmetric transformation

δφ = η ψ

δψ = i(σµεη∗) ∂µφ+ Fη

δF = −i η† σµ ∂µψ

The field F is auxiliary: F = 0 on shell. For all fields X = ψ, φ, F

(δη1δη2

− δη2δη1

)X = −(η†1 σµ η2 − η†2 σ

µ η1) i∂µX

Masses and non-gauge interactions

The most general renormalizable supersymmetric term has the

form (in Weyl notation)

Lng = −1

2Wij ψiψj +Wi Fi + h.c.

where

Wi =∂W

∂φiWij =

∂2W

∂φi∂φj

and W is the superpotential:

W =1

2Mij φiφj +

1

6yijk φiφjφk,

The constants Mij and yijk must be totally symmetric in their

indices.

A proof

Let us write in general

Lng = −1

2Wijψiψj +WiFi + h.c.

where Wij and Wi are some functions of the bosonic fields with

dimensions of (mass) and (mass)2 respectively. Wij is symmetric

under i↔ j.

No term which is a function of the scalar fields φi, φ∗i only can be

included in Lng: under a supersymmetry transformation it would

go into another function of the scalar fields only, multiplied by ηψi

or η†ψ†i, and with no spacetime derivatives or Fi, F∗i fields: nothing

of this form can possibly be cancelled by the supersymmetry

transformation of any other term in the lagrangian.

Renormalizability: each term has field content with mass

dimension ≤ 4. Hence

• Wij and Wi cannot be functions of the fermionic or auxiliary

fields

• Wi is a quadratic polynomial, and Wij is linear, in the fields φi

and φ∗i .

Now require supersymmetry invariance of Lng. We do it in steps:

1) Part which contains four spinors:

[δLng]4−spinor = −1

2

∂Wij

∂φk(ηψk)(ψiψj) −

1

2

∂Wij

∂φ∗k(η†ψ†

k)(ψiψj) + h.c.

The first term cannot cancel against any other. It vanishes if and

only if ∂Wij/∂φk totally symmetric in i, j, k, because of the Fierz

identity:

(ηψi)(ψjψk) + (ηψj)(ψkψi) + (ηψk)(ψiψj) = 0.

No such identity for the second term. Hence, supersymmetry

requires∂Wij

∂φ∗k= 0

An important feature: the superpotential W must be an analytic

function of the scalar fields. It is a function of φ, not of φ and φ∗.

A complex function f(z, z∗) = u(x, y) + iv(x, y) of a complex variable z = x + iy is

analytic if∂u

∂x=

∂v

∂y

∂u

∂y= −

∂v

∂x

This implies

∂f

∂z∗=

∂(u + iv)

∂x

∂x

∂z∗+

∂(u + iv)

∂y

∂y

∂z∗

=∂(u + iv)

∂x+ i

∂(u + iv)

∂y

=

(

∂u

∂x−

∂v

∂y

)

+ i

(

∂v

∂x+

∂u

∂y

)

= 0

It follows that

Wij = Mij + yijkφk

with Mij and yijk totally symmetric. We define the superpotential

W =1

2Mijφiφj +

1

6yijkφiφjφk

so that

Wij =∂

∂φi

∂φjW

Mij fermion mass matrix, yijk Yukawa couplings.

2) Parts of δLng which contain a spacetime derivative:

[δLng]∂ = −iWij ψi∂µφjσµ ε η∗ − iWi η

† σµ ∂µψi + h.c.

= −i (Wij ψi∂µφj +Wi ∂µψi) σµ ε η∗ + h.c.

Using

Wij∂µφj = ∂µ∂W

∂φi

this becomes

[δLng]∂ = −i(

ψi ∂µδW

δφi+Wi ∂µψi

)

σµ ε η∗ + h.c.

which is a total derivative provided

Wi =∂W

∂φi= Mijφj +

1

2yijkφjφk

The remaining terms in δLng cancel, provided Wi and Wij are the

first and second derivatives of the same function W .

The field equations for the auxiliary fields Fi are now

F ∗i = −Wi = −Mij φj −

1

2yijk φjφk

and therefore

Lng + La = −1

2Wij ψiψj −

1

2W ∗

ij ψ†iψ

†j +WiFi +W ∗

i F∗i + Fi F

∗i

= −1

2Wij ψiψj −

1

2W ∗

ij ψ†iψ

†j −WiW

∗i

The scalar potential is entirely determined by the superpotential:

V (φ) = WiW∗i = Fi F

∗i

= MijM∗il φjφ

∗l scalar masses

+1

2Mijy

∗ilm φjφ

∗l φ

∗m +

1

2M∗

ilyijk φjφkφ∗l trilinear couplings

+1

4yijky

∗ilm φjφkφ

∗l φ

∗m quartic couplings

The remaining terms contain mass terms for the fermions and

Yukawa couplings:

−1

2Wij ψiψj −

1

2W ∗

ij ψ†iψ

†j = −1

2Mij ψiψj −

1

2M∗

ij ψ†iψ

†j

−1

2yijk φiψjψk − 1

2y∗ijk φ

∗iψ

†jψ

†k

Many interaction terms, but couplings related by supersymmetry.

A simple example: just one supermultiplet φ, ψ.

Lng + La = |M |2 |φ|2 +1

2M y∗ φφ∗φ∗ +

1

2M∗ y φφφ∗ +

1

4|y|2 |φ|4

−1

2M ψψ − 1

2M∗ ψ†ψ†

−1

2y φψψ − 1

2y∗ φ∗ ψ†ψ†

Some comments:

• The scalar potential is positive semi-definite: (almost) no

problems with the stability of the ground state.

• The quartic scalar self-coupling and the fermion-fermion-scalar

Yukawa coupling are related: λS = |λf |2.

• It is easy to check that scalars and fermions have the same

squared mass matrix

(M2)ij = MikM∗kj

which is symmetric and can be diagonalized.

• The analiticity of W requires the introduction of two Higgs

doublets in the supersymmetric version of the Standard Model.

In summary, we have the following

Recipe

for non-gauge interaction terms of a set of chiral supermultiplets

φi, ψi, Fi:

1. Give a superpotential

W =1

2Mij φiφj +

1

6yijk φiφjφk,

with Mij and yijk completely symmetric.

2. Build the interaction lagrangian as

Lng = −1

2Wij ψiψj +Wi Fi + h.c.

where

Wi =∂W

∂φiWij =

∂2W

∂φi∂φj

Gauge interactions

A gauge supermultiplet contains a vector field Aaµ and a Weyl

fermion λa. An infinitesimal gauge symmetry transformation with

parameter αa acts on Aaµ and λa as

δgaugeAaµ = g fabcAb

µαc − ∂µα

a

δgaugeλa = g fabcλbαc

color SU(3) a = 1, . . . , 8 fabc → fabcSU(3) g → gs

weak isospin SU(2) a = 1, . . . , 3 fabc → εabc g → g

weak hypercharge U(1) a = 1 fabc → 0 g → g′

Counting fermionic and bosonic degrees of freedom, we see that we

should add one real scalar auxiliary field, Da (off shell, nb = 3 from

Aµ and nf = 4 from λ). Then

Lgauge = −1

4F a

µνFaµν + iλa† σµDµ λ

a +1

2DaDa

where

Dµλa = ∂µλ

a − gfabcAbµλ

c

The corresponding supersymmetry transformation laws are easily

worked out, requiring that δAaµ be real, and δDa proportional to

the equation of motion of λa:

δAaµ = − 1√

2

(

η† σµ λa + λa† σµ η

)

δλa = − i

2√

2σµσνη F a

µν +1√2Daη

δDa = − i√2

[

η† σµDµ λa − (Dµλ

a)† σµ η]

With these definitions,

(δη1δη2

− δη2δη1

)X = −(η†1 σµ η2 − η†2 σ

µ η1) iDµX

on X = F aµν , λ

a, Da.

Back to the chiral supermultiplet X = φ, ψ, F . All of them must be

in the same representation of the gauge group:

δgaugeXi = i g αa (T a)ij Xj

where T a are the generators of the gauge group in the

representation of X. Ordinary derivatives must be replaced by

covariant derivatives everywhere:

∂µ → Dµ = ∂µ + i g Aaµ T

a

Is this enough to turn our simple model into a gauge-invariant

supersymmetric theory?

Not quite: there are other possible gauge invariant and

renormalizable terms that can be formed out of the fields in the

gauge and chiral supermultiplets:

φ∗ T a ψ λa (φ∗ T a φ)Da

They are not taken into account by the covariant derivatives,

because they involve the superpartners of the gauge fields.

They can be added to our supersymmetric lagrangian, with

appropriate coefficients, provided the transformation law of the

auxiliary field F is slightly modified:

δφ = η ψ

δψ = i(σµεη∗)Dµφ+ Fη

δF = −i η† σµ Dµψ + g√

2(T aφ)λa†η†

Then

L = Lgauge + Lchiral + g√

2(

φ∗ T a ψ λa + φT a λa† ψ†)

+ g(φ∗ T a φ)Da

is supersymmetric, provided the superpotential is gauge-invariant:

δgaugeW = Wi (T a)ijφj = 0

The equations of motion for Da are now

Da = −g(φ∗ T a φ)

and the scalar potential becomes

V (φ) = F ∗i Fi +

1

2DaDa = W ∗

i Wi +1

2g2(φ∗ T a φ)(φ∗ T a φ)

completely determined by the auxiliary fields. This fact is relevant

for the discussion of spontaneous supersymmetry breaking.

• The scalar potential is positive semidefinite: a sum of squares.

• A unique feature of supersymmetric theories: the scalar

potential is entirely determined by other interactions. F - terms

are fixed by Yukawa couplings, D- terms by gauge couplings

In particular, the quartic Higgs couplings are no longer free

parameters.

Superpotential couplings

i

k

j

(a) (b)

j

i

l

k

yijk yijmy∗klm

i

k

j

(a) (b)

i j

(c)

i j

M∗imyjkm Mij M∗

ikMkj

Gauge couplings

(a) (b) (c) (d)

(e) (f) (g) (h)

Comments

• Mij and yijk in general restricted by gauge invariance

• Pattern of Yukawa couplings required for the cancellation of

quadratic divergences

• Fermion mass terms flip chiality, scalar masses do not

Soft supersymmetry breaking

The most general soft supersymmetry breaking lagrangian is given

by

Lsoft = −m2ijφiφ

∗j −

(

1

2mλ λ

aλa +1

2bijφiφj +

1

6aijkφiφjφk + h.c.

)

−1

2cijkφ

∗iφjφk + h.c.

It can be shown that a theory with exact supersymmetry, plus

Lsoft, is free of quadratic divergences.

The last term is usually negligible in most supersymmetry

breaking scenarios.

A word on gaugino masses. A mass term for a left-handed Weyl spinor

ξT ε ξ + h.c.

is called a Majorana mass term. It is allowed by Lorentz invariance, but

generally forbidden if ξ belongs to a complex representation of a gauge group.

A simple example: a phase transformation δξ = iαξ. Matter fermions in the

Standard Model (with the possible exception of a right-handed neutrino)

cannot have a Majorana mass term.

The case of gauginos λa is different: they transform according to the adjoint

representation,

δgaugeλa = g fabcλbαc

which is self-conjugate. This gives

δgauge(λTa ελa) = g fabcαc(λ

Tb ελa + λT

a ελb) = 0

because fabc is completely antisymmetric.

The minimal supersymmetric Standard Model

With some specifications, there is a unique superpotential which is

allowed by the Standard Model gauge invariance and by

renormalizability:

WMSSM = u∗R yu qLHu − d∗R yd qLHd − e∗R yl˜LHd + µHuHd

(call it W from now on). The first term in detail:

u∗R yu qLHu = (u∗R)jf (yu)fg(qL)gj

α (Hu)βεαβ

j is an SU(3)color index, j = 1, 2, 3

α, β are SU(2)L indices, α = 1, 2, β = 1, 2

f, g are generation indices, f = 1, 2, 3, g = 1, 2, 3

If φ1 and φ2 are SU(2) doublets,

δφ1 = iαaσa φ1 δφ2 = iαaσa φ2

then

φT1 ε φ2, ε =

(

0 1

−1 0

)

≡ iσ2

is an SU(2) scalar:

δ(φT1 ε φ2) = iαaφT

1 (σTa ε + εσa) φ2 = 0

because σ1, σ3 are symmetric, σ2 antisymmetric, and {σa, σb} = 2δab. The

quantities

φ†1φ1 φ†

1φ2 φ†

2φ2

are also, obviously, SU(2) invariants, by they cannot enter the superpotential,

which is an analytic function of all scalar fields.

Some comments:

• It is now clear why we need two Higgs doublets: in the

Standard Model, H gives mass to up-type quarks, while

Hc = εH∗ gives mass to down quarks and lepons. Here, H and

H∗ cannot appear simultaneously in the superpotential.

• yu, yd, yl are 3 × 3 matrices in family space of dimensionless

Yukawa couplings. They contain fermion masses and

generation mixing. A good approximation:

y33u = yt, y

33b = yb, y

33l = yτ , all other entries equal to 0.

• No bilinear term other than µHuHd is allowed by gauge

symmetry. µ is the only dimensionful parameter in the

superpotential; in particular, mass terms for fermions are

forbidden.

The superpotential originates a large variety of interaction

vertices, with coupling constants related by supersymmetry.

Recall our recipe,

Lng = −WiW∗i − 1

2Wij ψiψj −

1

2W ∗

ij ψ†iψ

†j

where

Wi =∂W

∂φi,Wij =

∂2W

∂φi∂φj

We find e.g.

• Yukawa couplings, such as

∂W

∂u∗R∂uLuu = u yu uH

0u

as in the Standard Model.

• quark-squark-higgsino couplings,

∂W

∂u∗R∂H0u

uh0u = u yu qL h

0u

• Quartic scalar couplings:

∂W

∂u∗R

2

= |yu qLHu|2

The term µHuHd provides a higgsino mass term

−µ (h+u h

−d − h0

uh0d + h.c.)

and a scalar Higgs mass term

− |µ|2(

∣H+u

2+∣

∣H0u

2+∣

∣H0d

2+∣

∣H−d

2)

This term cannot induce spontaneous breaking of the electroweak

gauge symmetry. We must rely upon soft breaking terms.

The value of µ is a problem: we expect it to be of the same order

(∼ 100 GeV) as the soft breaking terms, in order to provide the

correct value of the W and Z masses, but it has a completely

different origin.

Accidental symmetries and R-parity

Baryon and lepton number conservation in the Standard Model

follow from accidental symmetries: a consequence of

renormalizability.

This is no longer true in supersymmetry:

W∆L=1 =1

2λijk

˜iL

˜jL(ek

R)∗ +1

2λ′ijk

˜iLq

jL(dk

R)∗ + µ′i˜iLHu

W∆B=1 =1

2λ′′ijk (ui

R)∗(djR)∗(dk

R)∗

are allowed by renormalizability and gauge symmetry.

In order to implement B and L conservation we must impose some

extra symmetry (not B and L themselves: they are violated by

nonperturbative effects!).

Consider the so-called matter parity:

PM = (−1)3(B−L)

where B = 1/3 for quarks and squarks and 0 otherwise, L = 1 for

leptons and sleptons, 0 otherwise.

PM commutes with supersymmetry: all particles in the same

supermultiplet have the same PM (−1 for quarks and leptons, +1

for Higgs and gauge).

Requiring PM = +1 for all terms in the lagrangian (or in the

superpotential) eliminates B and L violating terms.

A different, but equivalent language: define R−parity as

PR = (−1)3(B−L)+2s

Conserved if PM is conserved, but does not commute with

supersymmetry:

PR = +1 for standard particles

PR = −1 for superpartners

R-parity conserved ⇔ matter parity conserved.

Useful for phenomenology: each vertex must contain an even

number of fields with PR = −1.

Consequences of R-parity conservation

• The lightest particle with PR = −1 is stable. It is called the

LSP. A candidate for dark matter.

• Supersymmetric particles decay into states with an odd

number of PR = −1 particles.

• Supersymmetric particles are always produced in even numbers

(usually 2).

Soft supersymmetry breaking in the MSSM

The most general soft supersymmetry breaking term for the

MSSM is

Lsoft = −1

2

(

M3gg +M2ww +M1bb+ h.c.)

−(

u∗R Au qL Hu − d∗RAd qL Hd − e∗RAe˜L Hd

)

+ h.c.

−q†Lm2Q qL − ˜†

Lm2L

˜L − u†R m

2U uR − d†Rm

2D dR − e†R m

2E eR

−m2Hu

H†uHu −m2

HdH†

d Hd − (BHuHd + h.c.)

An enormous number of new independent parameters (105).

All of them are expected to be of order msoft (to the appropriate

power), something between 100 GeV and 1 TeV.

Do we have any guiding principle that can help us reducing the

number of arbitrary parameters?

m2Q,m

2U ,m

2D,m

2E are arbitrary 3 × 3 matrices in flavour space; they

are subject from strong constraints from

• flavour changing neutral current phenomena (e.g. K0K0 mixing)

• individual leptonic number conservations (e.g. µ→ eγ)

• CP violation

The (unknown) mechanism that generates soft breaking terms

must respect these constraints.

All these constraints are satisfied with the simple assumption that

flavour mixing and CP violation are only originated by Yukawa

couplings. This is achieved by assuming that

• (m2Q)fg = m2

Qδfg (m2U )fg = m2

Uδfg (m2D)fg = m2

Dδfg

(m2E)fg = m2

Eδfg

• (Au)fg = Au(yu)fg (Ad)fg = Ad(yd)fg (Ae)fg = Ae(ye)

fg

• all soft breaking parameters are real.

This corresponds to assuming that the sector that generates

soft-breaking terms is flavor-blind.

This (or any other) structure of soft breaking terms results from

some underlying mechanism that breaks supersymmetry at a scale

Q0, presumably very large (e.g. the grand-unification scale).

Predictions at energy scales relevant for our experiments, mexp,

typically of order 100 to 1000 GeV, will therefore contain large

logarithms

logQ0

mexp

induced by radiative corrections. These must be resummed by

standard renormalization group techniques.

An encouraging fact: coupling constant unification works much

better in the MSSM than in the Standar Model. Define

g3 = gs

g2 = g =e

sin θW

g1 =

5

3g′ =

5

3

e

cos θW

(in order to have the same normalization for symmetry

generators). The corresponding RGEs at one loop are

µ2 dgi

dµ2=

bi8πg3

i ⇒ µ2 d

dµ2

1

αi= −bi

with αi = g2i /(4π).

0

10

20

30

40

50

60

10 5� 10 10 10 151µ (GeV)

α i-1(µ

)

SMWorld Average�α1

α2�

α3 αS(MZ)=0.117±0.005�

sin2ΘMS__=0.2317±0.0004�

0

10

20

30

40

50

60

10 5� 10 10 10 151µ (GeV)

α i-1(µ

)

MSSMWorld Average�68%�

CL�

M S=103.7±0.8 ±0.4� GeV M� U =1015.9±0.2� ±0.1�U.A.�W.d.B�H.F.�

α1

α2

α3

2 4 6 8 10 12 14 16 18Log10(Q/1 GeV)

0

10

20

30

40

50

60

α−1

α1−1

α2−1

α3−1

The mass spectrum

1. The Higgs sector

The Higgs potential of the MSSM is

VHiggs = (|µ|2 +m2Hd

)(|H0d |2 + |H−

d |2) + (|µ|2 +m2Hu

)(|H0u|2 + |H+

u |2)+B

(

H+u H

−d −H0

uH0d

)

+ h.c.

+1

8(g2 + g′

2)(|H0

u|2 + |H+u |2 − |H0

d |2 − |H−d |2)2

+1

2g2|H+

u H0∗d +H0

uH−∗d |2

No arbitrary quartic couplings: they are fixed by the D terms.

Minimization: We can take H+u = H−

d = 0 (good news: EM

unbroken) and H0u, H

0d real and positive at the minimum without

loss of generality. Then we restrict to the neutral sector:

VHiggs = (|µ|2 +m2Hd

)|H0d |2 + (|µ|2 +m2

Hu)|H0

u|2 − (BH0uH

0d + h.c.)

+1

8(g2 + g′

2)(|H0

u|2 − |H0d |2)2

B may also be taken to be real and positive: no CP violating

phases in the Higgs sector.

Perform an SU(2)L gauge rotation that makes 〈H+u 〉 = 0; then, the condition

∂V/∂H+u = 0 is only fulfilled if also 〈H−

d〉 = 0.

The only place where the phases of H0u, H0

dappear is the term

−BH0uH0

d + h.c.

We can make B real and positive by a suitable phase choice for H0u, H0

d.

Clearly, the potential has a minimum if also H0uH0

dis real and positive, so 〈H0

u〉

and 〈H0d〉 must have opposite phases, which can be rotated away by a U(1)Y

gauge transformation.

Conditions for electroweak spontaneous symmetry breaking:

1. The origin H0u = H0

d = 0 is not a minimum of the potential

2. The potential is bounded from below

Condition 1. leads to the constraint

(|µ|2 +m2Hd

)(|µ|2 +m2Hu

) < B2

In some cases, guaranteed by large negative radiative corrections

to m2Hu

.

Condition 2. is almost automatic: the quartic couplings are

positive definite. There are however D-flat directions∣

∣H0u

∣ =∣

∣H0d

∣.

We must require that the potential in these directions is stabilized

by quadratic terms:

2B < 2 |µ|2 +m2Hd

+m2Hu

Define vu = 〈H0u〉 and vd = 〈H0

d〉. We find

m2Z =

1

2(g2 + g′

2)(v2

u + v2d) ⇒ v2 ≡ v2

u + v2d ' (174 GeV)2

Following tradition, we also define an angle β through

tanβ =vu

vd

Clearly 0 ≤ β ≤ π/2. The minimization conditions are

|µ|2 +m2Hd

= B tanβ − m2Z

2cos 2β

|µ|2 +m2Hu

= B cotβ +m2

Z

2cos 2β

The µ problem is now evident.

Mass eigenstates:

1. Two CP-odd neutral scalars

G0

A0

=√

2

sinβ − cosβ

cosβ sinβ

ImH0u

ImH0d

2. Two charged scalars

G+

H+

=

sinβ − cosβ

cosβ sinβ

H+u

H+d

G0, G± are pseudo-Goldstone bosons: they can be removed from

the spectrum by an appropriate gauge transformation.

A0, H± are physical degrees of freedom.

3. Two CP-even neutral scalars

h0

H0

=√

2

cosα − sinα

sinα cosα

ReH0u − vu

ReH0d − vd

Finding the mass eigenvalues is a matter of algebra:

m2A =

2B

sin 2β

m2H± = m2

A +m2W

m2h,H =

1

2

(

m2A +m2

Z ∓√

(m2A +m2

Z)2 − 4m2Zm

2A cos2 2β

)

sin 2α

sin 2β= −m

2A +m2

Z

m2H −m2

h

cos 2α

cos 2β= −m

2A −m2

Z

m2H −m2

h

It is easy to show that

m2H± ≥ m2

A

mh ≤ mA ≤ mH

mh < |cos 2β|mZ

Radiative corrections weaken this bound. At one loop, and

neglectic squark mixing, one finds

∆m2h ' 3g2

8π2

m4t

m2W

logmt1

mt2

m2t

An extra term appears if mixing is taken into account:

(

∆m2h

)

mix' 3g2

8π2

m4t

m2W

(At − µ cotβ)2

m2t1−m2

t2

[

logm2

t1

m2t2

+ (At − µ cotβ)2 f(m2t1,m2

t2)

]

f(a, b) =1

a− b

[

1 − 1

2

a+ b

a− blog

a

b

]

Including all corrections, one obtains approximately

mh<∼ 130 GeV

still very interesting in view of LHC.

Yukawa couplings are related to fermion masses in the third

generation by

yt =g mt√

2mW sinβyb =

g mb√2mW cosβ

yτ =g mτ√

2mW cosβ

A difference with respect to the Standard Model: yb and yτ may be

as large as yt, if tanβ is large enough (tanβ < 1 disfavoured: yt

becomes nonperturbative for tanβ > 1.2).

The couplings of h,H,A to standard particles are the same as in

the Standard Model, rescaled by α− and β−dependent factors:

dd, ss, bb uu, cc, tt W+W−, ZZ

e+e−, µ+µ−, τ+τ−

h − sinα/ cosβ cosα/ sinβ sin(β − α)

H cosα/ cosβ sinα/ sinβ cos(β − α)

A −iγ5 tanβ −iγ5 cotβ 0

At tree level, the Higgs sector is described by two parameters (a

common choice: m2A and tanβ).

In addition, there are h(H)AA and h(H)AZ couplings.

2. Neutralinos and charginos

The supersymmetric partner of neutral gauge bosons, w0 and b,

and of Higgs scalars, h0u, h

0d, are not mass eigenstates. We find

Ln = −1

2ψT

0 Mn ψ0 + h.c.

where the matrix Mn in the basis b, w0, h0d, h

0u is given by

M1 0 −mZ cosβ sin θW mZ sinβ sin θW

0 M2 mZ cosβ cos θW −mZ sinβ cos θW

−mZ cosβ sin θW mZ cosβ cos θW 0 −µmZ sinβ sin θW −mZ sinβ cos θW −µ 0

The mass matrix Mn is symmetric, and can be diagonalized by a

unitary transformation N . The four mass eigenstates

χ0i = Nij ψ

0j ; Mn = NT Mn N

are called neutralinos. Their coupling are entirely determined by

the matrix N , and thus by the four parameters

M1,M2, µ, tanβ

The lightest neutralino is in most models (= for most of the

allowed parameter choices) the LSP; neutralinos play a crucial role

in phenomenology.

Charged gauginos and higgsinos

w+, h+u , w

−, h−d

mix in a similar way:

Lc = −1

2ψT

c Mc ψc + h.c.

Mc =

0 XT

X 0

X =

M2

√2mW sinβ

√2mW cosβ µ

The mass eigenstates (called charginos) are found by diagonalizing

X with a bi-unitary transformation:

χ+1

χ+2

= V

w+

h+u

χ−1

χ−2

= U

w−

h−d

X = UT

mχ±

1

0

0 mχ±

2

V

m2χ±

1,χ±

2

=1

2

[ (

|M2|2 + |µ|2 + 2m2W

)

∓√

(

|M2|2 + |µ|2 + 2m2W

)2

− 4 |µM2 −m2W sin 2β|2

]

As in the case of neutralinos, the matrices U, V contain all the

relevant information about chargino couplings.

It is quite natural to assume that gaugino masses have a common

value m1/2 at the reference scale Q0, where the (properly

normalized) gauge couplings take approximately the same value

αU ∼ 0.04.

It turns out that gaugino running masses obey the same

renormalization group equations. Therefore, with the above

assumption,

Mi(Q) =αi(Q)

αUm1/2

This assumption reduces (from 4 to 3) the number of arbitrary

parameters in the gaugino sector.

3. Squarks and sleptons

The scalar partners of matter fermions (squarks and sleptons) have

mass matrices with contributions from various terms:

• quartic Higgs-Higgs-squark-squark vertices

• trilinear Higgs-squark-squark couplings

• soft-breaking masses

• D-terms

In general, there is mixing among different families, and mixing

betweel scalar parners of left- and right-handed fermions. This can

be sizeable in the third generation.

Mass matrices for squarks and sleptons in the third generation, in

the simplified situation outlined above:

s− top : (t∗L t∗R)

m2Q +m2

t + ∆Q mt(At − µ cotβ)

mt(At − µ cotβ) m2U +m2

t + ∆U

tL

tR

s− bottom : (b∗L b∗R)

m2Q +m2

b + ∆Q mb(Ab − µ tanβ)

mb(Ab − µ tanβ) m2D +m2

b + ∆D

bL

bR

s − tau : (τ∗L τ∗R)

m2L +m2

τ + ∆L mτ (Aτ − µ tanβ)

mτ (Aτ − µ tanβ) m2E +m2

τ + ∆E

τL

τR

where

∆ = m2Z(T3 −Q sin2 θW ) cos 2β

Spontaneous supersymmetry breaking

How are soft supersymmetry-breaking terms generated? We need

a mechanism of spontaneous supersymmetry breaking (explicit

breaking not acceptable).

A symmetry is said to be spontaneously broken when the vacuum

state is not invariant under that symmetry:

Q|0〉 6= 0

Supersymmetry is no exception. This can happen only if some of

the auxiliary fields Fi or Da acquires a non-zero vacuum

expectation value.

Proof: the supersymmetry algebra implies

H = P 0 =1

4

(

Q1Q†1 +Q†

1Q1 +Q2Q†2 +Q†

2Q2

)

which implies

Qi|0〉 6= 0 ⇔ 〈0|H|0〉 > 0 ⇔ 〈0|V |0〉 > 0

for a translation-invariant vacuum state. But

V = FiF∗i +

1

2DaDa

so supersymmetry is spontaneously broken if and only if the

system

Fi = 0 Da = 0

has no solution.

Spontaneous breaking based on 〈Da〉 6= 0 for some Da is called a la

Fayet-Iliopoulos. If there is a U(1) factor in the gauge group, a

term

LD−breaking = kD

can be added to the lagrangian. Then

V =1

2D2 − kD + gD

i

qiφiφ∗i

D = k − g∑

i

qiφiφ∗i

D is different from zero if for some reason all scalar fields are zero

at the minimum. This is not the case in the MSSM.

Not very promising, although not completely ruled out.

Models in which supersymmetry is spontaneously broken by

〈Fi〉 6= 0 for some Fi are called O’Raifeartaigh models. An explicit

example: three chiral superfields, and

W = −kφ1 +mφ2φ3 +y

2φ1φ

23

where φ1 must be a gauge singlet. Then

V = |F1|2 + |F2|2 + |F3|2

F1 = k − y

2φ∗3

2 F2 = −mφ∗3 F3 = −mφ∗2 − yφ∗1φ∗3

F1 = F2 = F3 = 0 has no solution. For m2 > yk, the minimum of V is

for φ2 = φ3 = 0, φ1 undetermined (flat direction).

The solution is F1 = k, V = k2 at the minimum. The flat direction is

no longer flat after radiative corrections, and the minimum is at

φ1 = 0. The spectrum: six scalars with masses

0 0 m2 m2 m2 − yk m2 + yk

and three fermions with masses

0 m m

The zero-mass fermion is the analog of a Goldstone boson in

broken ordinary symmetry: we call it a goldstino. It is the partner

of the auxiliary field F1, responsible for supersymmetry breaking.

A general conclusion: spontaneous breaking of supersymmetry

cannot take place within the MSSM. There is no gauge-singlet

supermultiplet whose auxiliary field can take a non-vanishing

vacuum expectation value.

Hard to do by simply extending the MSSM with new fields and

new renormalizable interactions:

• gaugino masses cannot arise from renormalizable couplings

• the spectrum must include light scalars:

TrM2real scalars = 2 TrM2

chiral fermions

even after supersymmetry breaking.

Mass terms are generated after spontaneous symmetry breaking

as, for example, quark masses in the standard model:

y φψψ → y 〈φ〉ψψ = mψψ

In supersymmetry, the same mechanism is at work. A gaugino

mass could be generated either by a coupling

φλλ→ 〈φ〉λλ

but such a term is absent in a supersymmetric theory, or by

Fλλ→ 〈F 〉λλ

which is not renormalizable.

Way out: assume that supersymmetry is broken in a hidden sector

of non-observable degrees of freedom, that communicate somehow

with the observable sector. Historically, two proposals:

supersymmetry breaking communicated to the observable sector by

1. gravitational interactions

2. gauge interactions

The differences between the two cases are due to a different size of

the supersymmetry breaking scale

〈0|F |0〉

In gravity mediated supersymmetry breaking, soft breaking terms

arise from (in general nonrenormalizable) interaction terms (of

gravitational strength) of observable fields with combinations of

fields in the hidden sector that acquire a non-zero vev. We expect

therefore

msoft ∼〈F 〉MP

on dimensional grounds: msoft must vanish as 〈F 〉 goes to zero (no

spontaneous breaking) and as MP → ∞ (no interaction between

hidden and observable sectors.

For msoft ∼ 1 TeV we get

〈F 〉 ∼ 1011 GeV

In the gauge mediated scenario, superymmetry breaking is

communicated to the observable sector by the ordinary gauge

interactions via loop effects due to particles called messengers.

A rough estimate of the relevant supersymmetry breaking scale

can be obtained. We have in this case

msoft ∼α

〈F 〉M

where M is the typical messenger mass. This corresponds to a

much lower value for the supersymmetry order parameter:

〈F 〉 ∼ 104 − 105 GeV

if 〈F 〉 and M2 are assumed to be roughly of the same size.

A closer look at the goldstino

The goldstino field can be easily identified. The full fermion mass

matrix of a generic supersymmetric theory is

Mf =

0√

2ga(T a〈φ〉)i√2ga(〈φ∗〉T a)j 〈Wij〉

There is always one combination of fermion fields that corresponds

to zero mass:

G =

〈Da〉/√

2

〈Fi〉

is an eigenvector of Mf with zero eigenvalue. This is the goldstino.

The first component of Mf G is

√2〈ga(T aφ)iFi〉 = −

√2〈ga(T aφ)iW

∗i 〉 = −

√2〈∂W

∂φ∗iδgaugeφ

∗i 〉 = 0

by gauge-invariance of the superpotential.

The second component is given by

〈ga(φ∗T a)jDa +WijFi〉

The scalar potential is

V = FiF∗i +

1

2DaDa = WiW

∗i +

1

2g2

a(φ∗T aφ)2

It follows that

〈 ∂V∂φk

〉 = 〈W ∗i Wik + g2

a(φ∗T aφ)(φ∗T a)k〉

= −〈FiWik + gaDa(φ∗T a)k〉 = 0

Let us consider the case of local supersymmetry: η = η(x). Then

(δη1δη2

− δη2δη1

)X = −(η†1 σµ η2 − η†2 σ

µ η1) i∂µX

is a local translation, i.e. a general coordinate transformation.

Gravitation is included in a natural way.

Local supersymmetry is called supergravity.

In this case, the so-called super-Higgs effect takes place: the

spin-3/2 partner of the graviton, the gravitino, acquires a mass.

The missing spin degrees of freedom are provided by the goldstino.

By dimensional analysis,

m3/2 ∼ 〈F 〉MP

(

m3/2 =〈F 〉√3MP

)

Clearly, the gravitino mass has very different values in

gravity-mediated and gauge mediated scenarios:

• gravity-mediated: the gravitino mass

m3/2 ∼ 〈F 〉MP

∼ msoft

and its interactions are gravitational. Almost no interest for

phenomenology at colliders.

• gauge-mediated:

m3/2 ∼ 〈F 〉MP

∼ 105

1019GeV

if M ∼ 〈F 〉 �MP . Its goldstino components may have

non-negligible couplings, and may therefore play a role in

collider physics (X → XG).

Gravity-mediated models: some details

The supergravity lagrangian is non-renormalizable (as expected).

Assume there is a hidden chiral superfield X whose auxiliary

component FX gets a non-zero vev. The lagrangian includes

L = − 1

MPFX

a

1

2faλ

aλa + h.c.

+1

M2P

FXF∗Xkijφ

∗iφj

− 1

MPFX

(

1

6y′ijkφiφjφk +

1

2µ′

ijφiφj + h.c.

)

When FX → 〈FX〉, this gives precisely the soft terms we need.

In a minimal version of supergravity, there are considerable

simplifications:

fa = f kij = kδij y′ = αy µ′ = βµ

This leads to

M1 = M2 = M3 = m1/2

m2Q = m2

U = m2D = m2

L = m2E = m2

Hu= m2

Hd= m2

0

Au = A0yu Ad = A0yd Ae = A0ye

B = B0µ

Neutral flavour changing effects automatically suppressed.

Gauge-mediated models: some details

In the simplest version, the model contains four messenger

supermultiplets:

q ∼(

3,1,−2

3

)

q ∼(

3,1,+2

3

)

` ∼ (1,2,+1) ¯∼ (1,2,−1)

and a chiral multiplet S, whose scalar and auxiliary components

take a vev (e.g. by the O’Raifeartaigh mechanism). They interact

through a superpotential

Wmessengers = y2S`¯+ y3Sqq

which becomes effectively

Wmessengers = y2〈S〉`¯+ y3〈S〉qq

Gauginos get masses

Ma =αa

〈FS〉〈S〉

through one-loop diagrams:

⟨ S ⟩

⟨ FS ⟩

B, W, g

Soft scalar masses and trilinear couplings arise at two-loops. One

gets

Au = Ad = Ae = 0

at the messenger scale.

General features of soft masses in gauge-mediated models:

• larger for strongly-interacting particles

• determined by gauge properties only

The second properties leads to a degeneracy in squark and slepton

masses, that suppresses FCNC effects.

Summary

• Supersymmetry provides a solution to the naturalness problem

without spoiling the good features of the Standard Model.

• It has a number of welcome side-effects:

– a better context for grand unification

– a natural candidate for cold dark matter.

• A large portion of the parameter space for supersymmetric

theories has already been explored.

• Supersymmetry breaking in realistic models is still an open

problem.