0% found this document useful (0 votes)
13 views61 pages

Introsusy

Uploaded by

luke091x
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
13 views61 pages

Introsusy

Uploaded by

luke091x
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 61

Introduction to supersymmetry

Borut Bajc
J. Stefan Institute, 1000 Ljubljana, Slovenia

Abstract
This is a pedagogical introduction for graduate students to the
(minimal) N = 1 supersymmetry in 4 dimensions. It ranges from the
supersymmetry algebra, superspace, explicit construction of a gen-
eral supersymmetric Lagrangian, to the minimal supersymmetric stan-
dard model. It discusses various issues like R-parity, the electroweak
symmetry breaking, renormalization and spontaneous supersymmetry
breaking.

Contents
1 Preface 2

2 Notation and conventions 3

3 A supersymmetric free action 5

4 The supersymmetry algebra 8

5 Superspace 10
5.1 Chiral superfields . . . . . . . . . . . . . . . . . . . . . 12
5.2 Vector superfields . . . . . . . . . . . . . . . . . . . . . 15
5.3 Supersymmetry invariants . . . . . . . . . . . . . . . . 16
5.4 The free Lagrangian again . . . . . . . . . . . . . . . . 17
5.5 Explicit formulae . . . . . . . . . . . . . . . . . . . . . 19

6 The Wess-Zumino model 21

7 Gauge theories 24
7.1 The abelian case . . . . . . . . . . . . . . . . . . . . . 24
7.2 The nonabelian case . . . . . . . . . . . . . . . . . . . 28

1
8 Renormalization 30

9 The minimal supersymmetric standard model (MSSM) 34


9.1 The supersymmetric Lagrangian . . . . . . . . . . . . 35
9.2 R-parity and other symmetries . . . . . . . . . . . . . 37
9.3 The supersymmetry breaking soft terms . . . . . . . . 39
9.4 Electroweak symmetry breaking . . . . . . . . . . . . . 40

10 The mass spectrum in MSSM 43


10.1 Higgs mass . . . . . . . . . . . . . . . . . . . . . . . . 43
10.2 Chargino mass . . . . . . . . . . . . . . . . . . . . . . 45
10.3 Neutralino mass . . . . . . . . . . . . . . . . . . . . . 45
10.4 Sfermion mass . . . . . . . . . . . . . . . . . . . . . . 46
10.5 Gluino mass . . . . . . . . . . . . . . . . . . . . . . . . 46

11 Possible experimental signatures


from supersymmetry 46

12 What is supersymmetry good for? 48


12.1 Gauge coupling unification . . . . . . . . . . . . . . . . 48
12.2 Dark matter . . . . . . . . . . . . . . . . . . . . . . . . 49
12.3 The hierarchy problem . . . . . . . . . . . . . . . . . . 50

13 Spontaneous supersymmetry breaking 53


13.1 MSSM is not enough . . . . . . . . . . . . . . . . . . . 53
13.2 Gravity mediation . . . . . . . . . . . . . . . . . . . . 55
13.3 Gauge mediation . . . . . . . . . . . . . . . . . . . . . 57

14 Exercises 59

1 Preface
There are many very good reviews and books on supersymmetry, for
example, among many others: [1], [2] and [3] are the classical refer-
ences, [4] is fastly becoming classical and it is continuously updated,
[5] is a very useful introduction (which strongly influenced the present
notes) with all computational details, [6] is a clear overview of the main
features, [7] and [8] are for those who like more formal approach, [9],
[10], [11] and [12] are reviews on susy breaking, [13] is part of the
Weinberg’s famous course on quantum field theory, [14] is for fans of
superspace.

2
These notes were written for a 10 lectures course of 45 minutes
each. At least the basics of field theory are a requisite, as it is the
usual course of particle physics, with the standard model.

2 Notation and conventions


We will use almost exclusively the 2 component spinor notation (a
very detailed review is found for example in [15]). Let us first see its
connection with the 4 component notation. The gamma matrices are

0 σµ
   
µ 5 1 0
0 1 2 3
γ = , γ ≡ iγ γ γ γ = , (1)
σ̄ µ 0 0 −1

with

σ µ = 1, −σ i , σ̄ µ = 1, σ i .
 
(2)
so that the usual anticommutation relations are satisfied

{γ µ , γ ν } = 2η µν (3)
with the metric tensor

η µν = diag (1, −1, −1, −1) (4)


As is well known, the Lorentz symmetry algebra SO(1,3) can be
equivalently represented by a product SU(2)× SU(2). The spinor in-
dices of the first SU(2) will be denoted by greek letters α, β, . . ., while
the dotted indices α̇, β̇, . . . will denote the second SU(2). The SU(2)
indices in (2) will be kept as

(σ µ )αα̇ , (σ̄ µ )α̇α (5)


while a general 4 component Dirac bispinor can be written as

   
ΨL ψα
ΨD = = (6)
ΨR χ̄α̇

The component ψα and χα̇ are representations (1/2, 0) and (0, 1/2)
respectively under the Lorentz SU(2)× SU(2). Bars on spinors de-
note usually complex conjugates (an exception is σ̄ µ which is not the

3
complex conjugate of σ µ , but defined explicitly by (5)). During the
operation of complex conjugation an undotted index becomes dotted
(and the opposite). SU(2) indices can be raised and lowered as

ψα = αβ ψ β , ψ α = αβ ψβ (7)


with the 2 dimensional antisymmetric Levi-Civita tensor defined by

12 = 1, 12 = −1 (8)


Similarly we have for the antispinors

ψ̄α̇ = α̇β̇ ψ̄ β̇ , ψ̄ α̇ = α̇β̇ ψ̄β̇ (9)


with the 2 dimensional antisymmetric Levi-Civita tensor defined by

1̇2̇ = 1, 1̇2̇ = −1 (10)


They enable to express the following useful relations:

(σ̄ µ )α̇α = α̇β̇ αβ (σ µ )β β̇ , (σ µ )αα̇ = αβ α̇β̇ (σ̄ µ )β̇β (11)
In a 4-component notation, when ψ = χ we have to do with a
Majorana bispinor:

 
ψα
ΨM = (12)
ψ̄ α̇

for which the usual reality condition applies:


T
ΨcM ≡ CΨM = ΨM , C = iγ 2 γ 0 (13)
Lorentz invariant quantities will alway have all SU(2) indices summed
(undotted indices with undotted indices, dotted with dotted, this is
why it is useful to keep dots), with equal number of upper and lower
components. The product convention is

ψχ ≡ ψ α χα (14)
(the first spinor has always its SU(2) index up), while opposite for
antispinors

ψ̄ χ̄ ≡ ψ̄α̇ χ̄α̇ (15)

4
(the first antispinor has always its SU(2) dotted index down).
The above order convention allows to change order at will in spite
of the anticommuting character of the (anti)spinors, i.e.

ψχ = χψ, ψ̄ χ̄ = χ̄ψ̄ (16)


The following relations, valid for any spinor ψ, ψ, will turn out to
be very useful

1
ψ α ψ β = − αβ ψψ (17)
2
α̇ β̇ 1
ψ ψ = + α̇β̇ ψψ (18)
2

3 A supersymmetric free action


Usual symmetries we know are Poincare or internal symmetries like
colour or electromagnetism. Generators of such symmetries transform
a bosonic field to a bosonic field, and a fermionic field to a fermionic
field. So it is natural to think of a diffferent possibility: a symmetry
that transforms a bosonic field to a fermionic field and viceversa:

BOSON (φ) <−> F ERM ION (ψ) (19)


Such symmetries are called supersymmetries. Elements of the
group of supersymmetric transformation interchange the bosonic and
fermionic components of the supersymmetric multiplet.
We expect that the number of degrees of freedom (d.o.f.) for bosons
will be the same as for fermions in a supersymmetric multiplet. This
is similar to the zero sum of charges in a multiplet (the sum of all
third component spin in a multiplet for example).
Since the Dirac field has 4 d.o.f. on-shell (i.e. after satisfying
the equation of motion - the Dirac equation), we need some invariant
projection, since in the bosonic sector with spin 0 fields we can have at
most a complex field (2 d.o.f.). There are two types of such projections,
i.e. the field can be a Weyl (massless) or a Majorana (neutral). We
will stick to the two component notation. In the special case below it
will be enough to consider the Majorana case, with the limit m → 0
corresponding to the Weyl case.

5
We first start with free massive fields. In this way we will learn all
we need about the supersymmetry algebra. Later on we will generalize
the situation for the interacting fields. Let us start then with

1 1
L = ∂ µ φ† ∂µ φ − |mB |2 φ† φ + ψiσ µ ∂µ ψ − mF ψψ − m†F ψψ (20)
2 2
We want the Lagrangian to be invariant under infinitesimal su-
persymmetry transformations. What could these transformations be?
The ansatz

δφ = α ψα ; δφ† = ψ α̇ α̇ (21)

seems reasonable, and tells us, that the mass dimension of the an-
ticommuting Grassman parameter  is −1/2. We then write down
for δψ just the most general expansion consistent with linearity in ,
Lorentz symmetry and dimensionality:

δψα = c (σ µ )αα̇ α̇ ∂µ φ − F α , δψ α̇ = c∗ α (σ µ )αα̇ ∂µ φ† − F † α̇ (22)

At this point we don’t yet know what are the complex number c,
and a mass dimension 2 object F . To see it, we just plug the above
transformations in

 
δL = ∂ µ δφ† ∂µ φ + ∂ µ φ† ∂µ δφ − |mB |2 δφ† φ + φ† δφ

+ δψiσ µ ∂µ ψ + ψiσ µ ∂µ δψ − mF ψδψ − m†F ψδψ (23)

A straightforward computation shows that the free Lagrangian is


invariant (up to total derivatives) to supersymmetry transformations
providing

c = −i, mF = mB (= m), F † = mφ (24)


In the derivation we used the relation

ψσ µ χ = −χσ µ ψ (25)
valid for any spinors ψ, χ.
At this point everything seems ok, except for the fact, that it is
a bit strange to have parameters of the Lagrangian (the mass m) in

6
the transformation properties. This is connected to the fact that the
d.o.f. for a Majorana field and a complex scalar field are the same
only on-shell, while off-shell the Majorana field have 4 real d.o.f. It is
thus useful to promote the above quantity F to an auxiliary field (2
bosonic d.o.f. off-shell), which equation of motion fixes to (24) (and
thus counts zero d.o.f. on-shell). The new Lagrangian can be easily
written as

L = ∂ µ φ† ∂µ φ + ψiσ µ ∂µ ψ + F † F
1 1
− F mφ − F † m† φ† − mψψ − m† ψψ (26)
2 2
Considering now everything off-shell, we need on top of (21) and
(22) also the transformation of F :

α̇
δF = −i∂µ ψ α (σ µ )αα̇ α̇ ; δF † = iα (σ µ )αα̇ ∂µ ψ (27)

The Lagrangian (26) is off-shell (i.e. without the use of the equa-
tions of motion) invariant (up to total derivatives) under the infinitesi-
mal supersymmetric transformations (21), (22) and (27) collected here
below:

δφ = ψ ; δφ† = ψ (28)
µ µ † †
δψ = −iσ ∂µ φ − F  ; δψ = iσ ∂µ φ − F  (29)
µ † µ
δF = −i∂µ ψσ  ; δF = iσ ∂µ ψ (30)

The equation of motion for the auxilary fields F and F †

∂L ∂L
= 0, =0 (31)
∂F ∂F †
can be solved explicitly:

F † = mφ, F = m † φ† (32)
giving back the free Lagrangian for a complex boson field and a Ma-
jorana fermion field with equal masses:

1 1
L = ∂ µ φ† ∂µ φ − |m|2 |φ|2 + ψiσ µ ∂µ ψ − mψψ − m† ψψ (33)
2 2

7
4 The supersymmetry algebra
The next step is to obtain explicitly the algebra of the generators. To
find it out, we can act twice with a general supersymmetry transfor-
mation on our fields. Let us show it for the scalar field:

δ1 δ2 φ = δ1 (α2 ψα ) = −iα2 (σ µ )αα̇ 1 α̇ ∂µ φ − F α2 1α (34)


which gives

[δ1 , δ2 ] φ = −i (2 σ µ 1 − 1 σ µ 2 ) ∂µ φ (35)


The same exercise can be repeated for the other fields, giving in
general

[δ1 , δ2 ] = −i (2 σ µ 1 − 1 σ µ 2 ) ∂µ (36)


Since this is valid always, i.e. when applied on any field, one can
replace

δ → i Q + Q (37)
where we introduced a generator for each component of the infinites-
imal parameter α (α̇ ).
Comparing the commutator

  
i 1 Q + Q1 , i 2 Q + Q2 = (38)
n o n o
β β
α1 2 {Qα , Qβ } + 2 1 α̇ Qα̇ , Qβ − α1 2 β̇ Qα , Qβ̇ + 1 α̇ 2 β̇ Qα̇ , Qβ̇


with (36) one can easily get

n o
{Qα , Qβ } = Qα̇ , Qβ̇ = 0 (39)
= −i (σ µ )αα̇ ∂µ

Qα , Qα̇ (40)

As usual, we assume Poincaré invariance: transformations are gen-


erated by the translations

Pµ = −i∂µ (41)
and Lorentz transformations, which are the sum of the ”angular mo-
mentum” and ”spin” part

8
Mµν = Lµν + Sµν (42)
The ”angular momentum” part is

Lµν = i (xµ ∂ν − xν ∂µ ) (43)


Their algebra is as usual

[Pµ , Pν ] = 0 (44)
[Pµ , Mρσ ] = i (ηµρ Pσ − ηµσ Pρ ) (45)
[Mµν , Mρσ ] = −i (ηµρ Mνσ + ηνσ Mµρ − ηµσ Mνρ − ηνρ Mµσ ) (46)

How do the supersymetry generators Qα (Qα̇ ) commute with the


generators of the Poincaré algebra Pµ and Mµν ?
On the one hand it is difficult to think that a translation can change
the effect of a supersymmetric transformation, which reduces to
 
[Pµ , Qα ] = Pµ , Qα̇ = 0 (47)
On the other hand similar commutation relations cannot be true
for Lorentz transformations: changing spin with Q or Q automatically
changes also the Lorentz character. The most general ansatz we can
think of is

[Mµν , Qα ] = c(σµ σ ν − σν σ µ )α β Qβ , (48)


Mµν , Qα̇ = c̄Qβ̇ (σ µ σν − σ ν σµ )β̇ α̇
 
(49)

Using (41) we commute equation (40) with Mµν . The righthand-


side is

[(σ ρ )αα̇ Pρ , Mµν ] = i (σ ρ )αα̇ (ηρµ Pν − ηρν Pµ ) (50)


while the lefthandside becomes

     
Qα , Qα̇ , Mµν = Qα , Qα̇ , Mµν + Qα̇ , [Qα , Mµν ] (51)

Using (µνρχ is a completely antisymmetric Levi-Civita tensor, 0123 =


+1)

9
σ µ σ ν σ ρ = η µν σ ρ − η µρ σ ν + η νρ σ µ − iµνρχ σχ (52)
it is straightforward to obtain c = −c̄ = −i/4 after comparison with
(50). Notice that this has been dictated simply by Lorentz invariance.
The complete Poincaré and supersymmetry algebra is thus

[Pµ , Pν ] = 0 (53)
[Pµ , Mρσ ] = i (ηµρ Pσ − ηµσ Pρ ) (54)
[Mµν , Mρσ ] = −i (ηµρ Mνσ + ηνσ Mµρ − ηµσ Mνρ − ηνρ Mµσ ) (55)
n o
{Qα , Qβ } = Qα̇ , Qβ̇ = 0 (56)
= (σ µ )αα̇ Pµ

Qα , Qα̇ (57)
 
[Pµ , Qα ] = Pµ , Qα̇ = 0 (58)
i
[Mµν , Qα ] = − (σµ σ ν − σν σ µ )α β Qβ , (59)
4
i
Q (σ µ σν − σ ν σµ )β̇ α̇
 
Mµν , Qα̇ = (60)
4 β̇
The system is clearly closed. Additional internal (gauge or global)
symmetry generators typically commute with the above generators,
although there could be exceptions.

5 Superspace
It is suggestive that an arbitrary function of translated or Lorentz
transformed coordinates can be represented by

µ
f (x + a) = exp (iaµ Pµ )f (x) = ea ∂µ f (x) (61)
 
i 1 µν
f (Λx) = exp − θµν Lµν f (x) = e 2 θ (xµ ∂ν −xν ∂µ ) f (x) (62)
2
Can something analogue be possible for supersymmetry transfor-
mations? In other words, can we generalize space adding new coordi-
nates, such that a supersymmetry transformation will be nothing else
than a translation in these new coordinates? The answer is, thanks
to Salam and Strathdee, yes, and simplifies a lot the construction of
supersymmetry invariant field theories. Since the supersymmetry gen-
erators Qα and Qα̇ have fermionic anticommuting (called Grassmann)

10
character, the same will be necessarily true for these new coordinates,
α̇
θα and θ .
In doing that, let us rewrite the generators of supersymmetry
transformations with these coordinates and their derivatives. Again,
using Lorentz invariance, we can expand


+ b σ µ θ α ∂µ

Qα = a α
(63)
∂θ

Qα̇ = a α̇ + b (θσ µ )α̇ ∂µ (64)
∂θ
with a, b, a, b complex numbers (a and b are in general not the complex
conjugate of a and b). We will try to fix them by requiring the same
commutation relations as in (28)-(30).
α̇
A generic field of the coordinates xµ , θα and θ is called a super-
1̇ 2̇
field. Due to the limited number (θ1 , θ2 , θ , θ ) and nice properties
(a square is always zero) of the Grassmann variables, a superfield can
be expanded in a finite series of ordinary space coordinate functions:


F x, θ, θ = f1 (x) + θψ1 (x) + θψ2 (x) + θθf2 (x) + θθf3 (x) (65)
+ θσ µ θvµ (x) + θθθψ3 (x) + θθθψ4 (x) + θθθθf4 (x)

This is the most general expansion of a function of Grassmann


variables, a finite analogue of the Laurent expansion.
The supersymmetry transformation of a superfield is given by

δF (x, θ, θ̄) = i(Q + Q)F (x, θ, θ) (66)


Such general superfield has however much more components than
necessary: off-shell we have 8 complex functions and 4 Majorana
spinors, together 16 bosonic and 16 fermionic d.o.f. Fortunately this
representation is reducible. In order to do that, we need a constraint
that (anti)commutes with the generators (63) and (64), so that it is
preserved during the supersymmetry transformations (66). Formally
we need such operators Ôi that reduce the number of d.o.f. (projects
any function in superspace into a subsuperspace) through

Ôi F (x, θ, θ) = 0 (67)


and satisfy

11
n o n o
Ôi , Qα = Ôi , Qα̇ = 0 (68)
if they have a fermionic character, or
h i h i
Ôi , Qα = Ôi , Qα̇ = 0 (69)
if they have a bosonic character.

5.1 Chiral superfields


Operators of the form (63) and (64) with properly chosen constants
a, b, a, b are certainly natural candidates for the operators Ôi above.
It is an easy exercise to find out that operators

 
∂ b µ

Dα = − σ θ ∂
α µ
(70)
∂θα a
 
∂ b
Dα̇ = α̇
− (θσ µ )α̇ ∂µ (71)
∂θ a
do exactly the job we need.
We will not apply both types of operators on the same field, but
instead define the chiral superfields those that satisfy

Dα̇ Φ = 0 (72)
and antichiral superfields those that satisfy

Dα Φ = 0 (73)
Let us concentrate on the chiral superfield defined by (72). While
α̇
a general superfield is a function on xµ , θα and θ , a chiral superfield
is a function of only θα and the combination
 
µ µ b
y =x − θσ µ θ (74)
a
Due to that, when acting on a chiral superfield, the supersymmetry
generators get simplified:


Qα → a (75)
∂θα
ab + ba
Qα̇ → (θσ µ )α̇ ∂µy (76)
a

12
where the spacetime derivatives ∂µy act now on y coordinates.
As a check one can see, that any Φ = Φ(y, θ) automatically satisfies
(72) and is thus a superfield. It can be expanded as

Φ(y, θ) = Aφ(y) + Bθψ(y) + CθθF (y) (77)


A supersymmetry transformation (66) is then

Aδφ(y) + Bθδψ(y) + CθθδF (y) = (78)


 
∂ ab + ba
i aα α + (θσ µ )∂µy (Aφ(y) + Bθψ(y) + CθθF (y))
∂θ a

To obtain (28)-(30) the following relations must be satisfied:

A A
B= , C = 2, ab + ba = i (79)
ia 2a
In doing that we used the relation
1
θα θβ = − αβ θθ (80)
2
Similarly an antichiral superfield defined by (73) is a function of
α̇
only θ and the combination
 
b
y µ = xµ + θσ µ θ (81)
a
so that it can be expanded as

Φ(y, θ) = Aφ(y) + Bθψ + CθθF (y) (82)


When acting on an antichiral superfield, the supersymmetry gen-
erators can be written as

ab + ba µ  y
Qα → σ θ α ∂µ (83)
a

Qα̇ → a α̇ (84)
∂θ
A supersymmetry transformation (66) on a chiral superfield is

13
Aδφ(y) + Bθδψ(y) + CθθδF (y) = (85)
 
∂ ab + bb µ y
i a α̇ α̇ +

σ θ∂µ Aφ(y) + Bθψ(y) + CθθF (y)
∂θ a
and the correct tranasformation properties are obtained only when

A A
B=− , C = 2 , ab + ba = i (86)
ia 2a
which is consistent with (79). Here we used

α̇ β̇ 1
θ θ = + α̇β̇ θθ (87)
2
Demanding that the hermitian conjugate of the chiral superfield is
an antichiral superfield we immediately get also

A = A∗ , a = a∗ (88)
where in this special case the bars on the fields φ, ψ, F denote the
hermitian conjugation.
The expansion of the chiral and antichiral superfields is then

1 1 1
Φ(y, θ) = φ(y) + θψ(y) + 2 θθF (y) (89)
A ia 2a
1 1 1
Φ(y, θ) = φ(y) − θψ(y) + ∗2 θθF (θ) (90)
A∗ ia ∗ 2a
while the transformations of the fields can be simplified by


δΦ(y, θ) = i Q + Q Φ(y, θ)
 
α ∂ i µ y
= i a + θσ ∂µ Φ(y, θ) (91)
∂θα a

δΦ(y, θ) = i Q + Q Φ(y, θ)
 
∗ ∂ α̇ i µ y
= i a α̇
 + ∗ σ θ∂µ Φ(y, θ) (92)
∂θ a
We will determine the last constants later on from the canonical
normalization of fields.
A crucial property of chiral multiplets is that a product of chiral
multiplets is again a chiral multiplet. This follows simply from the

14
representation of the covariant derivatives (120), i.e. they are deriva-
tives, and so for any superfields Φ1,2
 
Dα̇ (Φ1 Φ2 ) = Dα̇ Φ1 Φ2 + Φ1 Dα̇ Φ2 (93)
In the special case that Φ1,2 are chiral superfields

Dα̇ Φ1,2 = 0 (94)


the product is also a chiral superfield since due to (93)

Dα̇ (Φ1 Φ2 ) = 0 (95)


The same is obviously true for antichiral fields.

5.2 Vector superfields


Another possible submultiplet of the general superfield is the vector
superfield, which is nothing else than a real superfield:
 †
V (x, θ, θ) = V (x, θ, θ) (96)
In general it can be expanded as

V (x, θ, θ) = C(x) + iθχ(x) − iθχ(x) + θσ µ θvµ (x)


i i
+ θθ [M (x) + iN (x)] − θθ [M (x) − iN (x)]
2  2  
i µ i µ
+ iθθθ λ(x) + ∂µ χ(x)σ − iθθθ λ(x) − σ ∂µ χ(x)
2 2
 
1 1
+ θθθθ D(x) − ∂ µ ∂µ C(x) (97)
2 2

We still seem to have a lot of degrees of freedom, 8 bosonic (the real


C, M , N , D and vµ ) and 8 fermionic (the two-component complex χ
and λ). However, if V is a vector multiplet (i.e. real), so is V + Φ + Φ†
with Φ a chiral superfield. It is thus possible to show that the gauge
transformation

V → VW Z = V + Φ + Φ† (98)
with properly chosen chiral multiplet Φ can bring the vector multiplet
into a simple form

15
1
VW Z (x, θ, θ) = θσ µ θvµ (x) + iθθθλ(x) − iθθθλ(x) + θθθθD(x) (99)
2
This Wess-Zumino gauge is useful because it explicitly reduces the
number of degrees of freedom, but one should keep in mind, that a
supersymmetry transformation does not preserve it, i.e. δVW Z cannot
be written anymore in the form (99).

5.3 Supersymmetry invariants


We want now see how to obtain supersymmetric invariants in the La-
grangian. The idea is very simple, and can be seen from (30). The
highest component (F ) of a chiral superfield (89) transforms under su-
persymmetry translation as a total derivative, so its spacetime integral
is a supersymmetry invariant. This property is due to the dimension-
ality of the auxiliary field F : since from (26) the mass dimensions of
the fields are

[φ] = 1, [ψ] = 3/2, [F ] = 2 (100)


and that of the transformation parameters

[] = [] = −1/2 (101)


the field with the largest dimension (F in this case) can transform
only through a spacetime derivative.
So the θθ component of the chiral superfield is a supersymmetry
invariant (up to total derivatives) and thus a possible invariant term
in the Lagrangian:

[Φ]θθ (102)
Due to the known properties of chiral superfields, any combination
" n #
Y n
Φi i (103)
i=1 θθ
is also a candidate for a supersymmetric invariant term in the La-
grangian.
Similarly the θθ component of any antichiral superfield is an in-
variant

16
 
Φ θθ (104)
which means that any combination
" n #
Y ni
Φi (105)
i=1 θθ
is also a possible supersymmetry invariant combination in the La-
grangian. Notice that above only pure holomorphic (103) or antiholo-
morphic (105) products of chiral or antichiral fields are allowed.
For mixed products one should resort to the vector multiplets.
As for chiral superfields, also the highest component of the vector
multiplet (the auxiliary field D) transforms as a total derivative. So
the term

[V ]θθθθ (106)
is again a supersymmetric invariant (up to total derivatives) and thus
a good candidate for a supersymmetry invariant Lagrangian. Notice
that it does not matter in which gauge the vector multiplet is writ-
ten down: the difference is again proportional to a total spacetime
derivative, and so irrelevant for the dynamics.
A general supersymmetric invariant Lagrangian can be thus writ-
ten as
h i h i
L = K(Φ, Φ† ) + [W (Φ)]θθ + W † (Φ† ) (107)
θθθθ θθ

where the real function K(Φ, Φ† ) (which is thus a vector superfield) is


called the Kähler potential, while the holomorphic function W (Φ) is
the superpotential. Notice that we have chosen W = W † in order to
satisfy the hermiticity (reality) condition for the Lagrangian.

5.4 The free Lagrangian again


At this point we have all the ingredients to derive explicitly the free
Lagrangian of section 3. We choose for the Kähler potential and su-
perpotential
m 2
K(Φ, Φ† ) = Φ† Φ, W (Φ) = Φ (108)
2

17
Then we expand the chiral superfield using (89). Of course the
spacetime coordinates are xµ , not y µ (74), so we have to expand fur-
ther the single functions as

1 b 2
   
b µ
φ(y) = φ(x) − θσ θ∂µ φ(x) + θθθθ∂ µ ∂µ φ(x)
a 4 a
 
1 b
θθ ∂µ ψ(x)σ µ θ

θψ(y) = θψ(x) + (109)
2 a
θθF (y) = θθF (x)

where we used

(σ µ )αα̇ (σ ν )α̇α = 2η µν (110)


Similarly we expand the antichiral superfield Φ† as in (90) and
then using (81)

   2
∗ ∗b ∗ 1 b
φ (y) = φ (x) + µ
θσ θ∂µ φ (x) + θθθθ∂ µ ∂µ φ∗ (x)
a 4 a
 
1 b
θθ θσ µ ∂µ ψ(x)

θψ(y) = θψ(x) − (111)
2 a
θθF ∗ (y) = θθF ∗ (x)

The comparison between (109) and (111) gives


 ∗
b b
=− (112)
a a
and due to (79) (or (86)) and (88) we find

b b i
+ = 2 (113)
a a |a|
Up to total derivatives we get

2
h i A
Φ† Φ ∂ µ φ∗ ∂µ φ + ψiσ µ ∂µ ψ + F ∗ F

= (114)
θθθθ 2a2

At this point we can choose

18
i 1
A = A = 1, a = −a = − √ , b = −b = √ (115)
2 2
In a similar way we can calculate the second part, i.e.

hm i m
Φ2 = −mφF − ψψ (116)
2 θθ 2
m∗
 
m 2  †
Φ = −m∗ φ∗ F ∗ − ψψ (117)
2 θθ 2

so that the total Lagrangian (107) with the choice (108) coincides with
the free (supersymmetry invariant) Lagrangian (26).

5.5 Explicit formulae


At this point, after having fixed all the different constants, it is maybe
useful to rewrite some of the above formulae.
The supersymmetry generators are

 
i ∂ µ

Qα = −√ + i σ θ α ∂µ (118)
2 ∂θα
 
i ∂ µ
Qα̇ = √ + i (θσ )α̇ ∂µ (119)
2 ∂θα̇
and the covariant derivatives


− i σ µ θ α ∂µ

Dα = α
(120)
∂θ

Dα̇ = α̇
− i (θσ µ )α̇ ∂µ (121)
∂θ
A chiral superfield is a function of θα and

y µ = xµ − iθσ µ θ (122)
and can be expaded as

Φ(y, θ) = φ(y) + 2θψ(y) − θθF (y) (123)
When acting on a chiral superfield the supersymmetry generators
becomes

19
i ∂
Qα → − √ (124)
2 ∂θα

Qα̇ → − 2(θσ µ )α̇ ∂µy (125)

so that a general supersymmetry transformation is

 
 1 ∂ α µ y
δΦ(y, θ) = i Q + Q Φ(y, θ) = √  − 2iθσ ∂µ Φ(y, θ)
2 ∂θα
(126)
The single components of the chiral superfield are

1
φ(y) = φ(x) − iθσ µ θ∂µ φ(x) − θθθθ∂ µ ∂µ φ(x) (127)
4
i µ

θψ(y) = θψ(x) + θθ ∂µ ψ(x)σ θ (128)
2
θθF (y) = θθF (x) (129)

Putting all together, a chiral superfield gets expanded in general


as


Φ(x, θ, θ) = φ(x) + 2θψ(x) − θθF (x) − i θσ µ θ ∂µ φ(x)


i  1
+ √ θθ ∂µ ψ(x)σ µ θ − θθθθ∂ µ ∂µ φ(x) (130)
2 4
α̇
In a similar way the antichiral superfield is a function of θ and

y µ = xµ + iθσ µ θ (131)
and can be expanded as

Φ(y, θ) = φ(y) + 2θψ(y) − θθF (y) (132)
When acting on an antichiral superfield the supersymmetry gen-
erators becomes


2 σµθ ∂µy

Qα → α
(133)
i ∂
Qα̇ → √ (134)
2 ∂θα̇

20
so that a general supersymmetry transformation is

 
1 ∂ α̇ µ
θ∂µy

δΦ(y, θ) = i Q + Q Φ(y, θ) = − √ α̇
 − 2iσ Φ(y, θ)
2 ∂θ
(135)
The single components of the chiral superfield are

1
φ(y) = φ(x) + iθσ µ θ∂µ φ(x) − θθθθ∂ µ ∂µ φ(x) (136)
4
i µ

θψ(y) = θψ(x) − θθ θσ ∂µ ψ(x) (137)
2
θθF (y) = θθF (x) (138)

Putting all together, an antichiral superfield gets expanded in gen-


eral as


Φ(x, θ, θ) = φ(x) + 2θψ(x) − θθF (x) + i θσ µ θ ∂µ φ(x)


i  1
− √ θθ θσ µ ∂µ ψ(x) − θθθθ∂ µ ∂µ φ(x) (139)
2 4

6 The Wess-Zumino model


We can now generalize the free Lagrangian for an arbitrary interacting
case. In this section we will consider only theories without gauge
interactions, called Wess-Zumino models.
It is not difficult to derive the θθ component for a general super-
potential of a single chiral superfield W (Φ). All we need is just to
expand the superpotential around the bosonic component φ:

∂W
W (Φ) = W (φ) + (φ) (Φ − φ) + . . . (140)
∂φ
which gives

∂W 1 ∂2W
[W (Φ)]θθ = − (φ)F − (φ)ψψ (141)
∂φ 2 ∂φ2
For one chiral superfield, the total Lagrangian for the canonical
Kähler

K(Φ, Φ† ) = Φ† Φ (142)

21
and a general superpotential W (Φ) gives

L = ∂ µ φ∗ ∂µ φ + ψiσ µ ∂µ ψ + F ∗ F
1 ∂2W
 
∂W
− (φ)F + (φ)ψψ + h.c. (143)
∂φ 2 ∂φ2
The equation of motion for the auxiliary field is enough to deter-
mine it:

∂W
F∗ = (144)
∂φ
and the single field Wess-Zumino Lagrangian with the canonical Kähler
potential looks like

2
∂W 1 ∂2W 1 ∂2W ∗
L = ∂ µ φ∗ ∂µ φ − + ψiσ µ ∂µ ψ − ψψ − ψψ (145)
∂φ 2 ∂φ2 2 ∂φ∗2
Needless to say, once we get rid of the auxiliary fields, we are left
with just a field theory, although a bit peculiar. In principle we don’t
need even to know that the theory is supersymmetric. We could in
principle use just the usual field theory method, although using the
power of supersymmetry makes any analysis simpler and it is thus
worth to make it explicit.
Notice that the potential is
2
∂W
V = |F |2 = (146)
∂φ
and the energy always positive.
From the mass dimensions

[K] = 2, [W ] = 3 (147)
it can easily be seen that any noncanonical Kähler potential with
cubic or higher powers of the chiral and antichiral superfields and any
superpotential with quartic or higher powers of chiral superfields is
nonrenormalizable.
So the most general renormalizable single field Wess-Zumino model
is for

m 2 λ 3
W (Φ) = aΦ + Φ + Φ (148)
2 3

22
giving for the auxiliary field

F ∗ = a + mφ + λφ2 (149)
The explicit form of the Lagrangian is

2
L = ∂ µ φ∗ ∂µ φ − a + mφ + λφ2 (150)
1 1
+ ψiσ µ ∂µ ψ − (m + 2λφ) ψψ − (m∗ + 2λ∗ φ∗ ) ψψ
2 2
At first glance the Lagrangian does not look supersymmetric. For
example, it is not clear whether the bosonic and fermionic masses are
equal. But the mass is a concept defined only in the minimum of the
potential. This is seen explicitly if we expand as usual the bosonic
field

φ=v+ϕ (151)
with

a + mv + λv 2 = 0, hϕi = 0 (152)
The above Lagrangian (150) becomes

2
L = ∂ µ ϕ∗ ∂µ ϕ − µϕ + λϕ2 (153)
1 1
+ ψiσ µ ∂µ ψ − (µ + 2λϕ) ψψ − (µ∗ + 2λ∗ ϕ∗ ) ψψ
2 2
with

µ = m + 2λv (154)
the common bosonic and fermionic mass. The Lagrangian above has
most of the terms a renormalizable (also nonsupersymmetric) field
theory would have. What is special in the supersymmetric Lagrangian
(153) is that there are relations between different parameters. For
example, only two parameters µ and λ describe the mass, the trilinear
and quadrilinear terms of the boson field, as well as the fermion mass
and Yukawa couplings. This is special with supersymmetry. These
relations get maintained by radiative corrections, which would not be
true in nonsupersymmetric models, even if we imposed them at tree
order.
There are different ways in which the above results can be gener-
alized. We leave it for the exercise.

23
7 Gauge theories
We have already introduced the vector multiplet. It contains a vector
field vµ which seems a good candidate for a gauge boson. The problem
is that the θθθθ of just powers of the vector superfield will not give
enough structure. For example, we have in the Wess-Zumino gauge

1  2  1
[VW Z ]θθθθ = D, VW Z θθθθ = v µ vµ , n
[VW Z ]θθθθ = 0 for n ≥ 3
2 2
(155)
This is due to the fact that the fields in VW Z are multiplied by
already some powers of θ and/or θ. In practice, what we need, is to
reduce their number. We can obtain that by applying the covariant
derivative on the vector superfield. This will also introduce automat-
ically the spacetime derivatives, needed for the kinetic terms. Let us
concentrate first on the abelian case, i.e. a U(1) gauge theory.

7.1 The abelian case


A natural first possibility would be to consider fields like Dα V and
Dα̇ V . This makes things just complicated, because these fields are no
more real on one side, i.e. (Dα V )† 6= Dα V , and not gauge invariant
on the other, i.e. Dα (Φ + Φ† ) 6= 0 for any chiral supermultiplet Φ. We
can solve both problems by introducing a chiral superfield
1 
Wα = − DD Dα V (156)
4
3
The field is chiral, because D = 0 as usual (the operator D is a
spinor operator) and so

Dα̇ Wα = 0 (157)
and it is gauge invariant because

 
DD Dα Φ + Φ† = α̇β̇ Dα̇ Dβ̇ Dα Φ = −2iα̇β̇ Dα̇ (σ µ )αβ̇ ∂µ Φ = 0


(158)
where we used the explicit representations (120) and (121).
From the general definition (97) we can now evaluate the chiral
superfield

24
i
(θσ µ σ ν )α Fµν (y) − θθ σ µ ∂µ λ(y) α

Wα (y, θ) = −iλα (y) + θα D(y) +
2
(159)
with the usual gauge field strength

Fµν = ∂µ vν − ∂ν vµ (160)
The kinetic term is obtained from the Lorentz invariant product
of two such chiral superfields. Up to a total derivative we get

1 i
[W α Wα ]θθ = 2λiσ µ ∂µ λ + D2 − F µν Fµν − µνρσ Fµν Fρσ (161)
2 4
In a completely analogous way we expand

i µ ν 
W α̇ (y, θ) = iλα̇ (y) + θα̇ D(y) − σ σ θ α̇ Fµν (y) − θθ (∂µ λ(y)σ µ )α̇
2
(162)
and get

h
α̇
i 1 i
W α̇ W = 2λiσ µ ∂µ λ + D2 − F µν Fµν + µνρσ Fµν Fρσ (163)
θθ 2 4
The correct supersymmetric kinetic term for a U(1) gauge field vµ
and its partner gaugino λα is thus

1 1h α̇
i 1 1
L= [W α Wα ]θθ + W α̇ W = − Fµν F µν + λiσ µ ∂µ λ + D2
4 4 θθ 4 2
(164)
The theory has four bosonic d.o.f. off-shell (3 from vµ and one real
D) and 4 fermionic (λα ), which reduce to 2 d.o.f. each when equations
of motion are applied.
How do we couple the U(1) vector superfield to a charged chiral
superfield? One is tempted to assume a gauge transformation for the
whole superfield, since all the single components presumably carry the
same U(1) charge. But the transformed superfield

Φ0 = eiΛ(x) Φ(y, θ) (165)

25
is not a chiral superfield1 anymore since Dα̇ Λ(x) 6= 0 . This difficulty is
immediately solved if Λ(x) gets promoted to a chiral superfield Λ(y, θ),
so that the transformed superfield remains chiral:

Φ0 (y, θ) = eiΛ(y,θ) Φ(y, θ) (166)


But in this case Λ† 6= Λ and so the usual canonical Φ† Φ in the
Kähler potential is not gauge invariant anymore. This is expected,
since we need the equivalent of gauge covariant derivatives. We know
already the invariance of the gauge kinetic terms under the transfor-
mation (98). We called it a gauge transformation and we can now
understand the reason. In fact, assuming
 
V → V − i Λ − Λ† (167)
as equivalent to the gauge tansformation, we find immediately an in-
variant Kähler:
 0 † †
Φ† eV Φ = Φ† e−iΛ eV −i(Λ−Λ ) eiΛ Φ = Φ† eV Φ (168)

since Λ, Λ† and V all commute. Let us see explicitly what this term
is. In doing that we expand the exponential and find, using the Wess-
Zumino gauge (a different gauge would give the same result, since the
Kähler is gauge invariant)

h i
Φ† Φ = ∂ µ φ∗ ∂µ φ + ψiσ µ ∂µ ψ + F ∗ F (169)
θθθθ
h i i 1
Φ† V Φ = (∂µ φ∗ v µ φ − φ∗ v µ ∂µ φ) − ψσ µ vµ ψ
θθθθ 2 2
D 2 i
|φ| + √ φ∗ λψ − ψλφ

+ (170)
2 2
1h † 2 i 1 µ
ΦV Φ = v vµ |φ|2 (171)
2 θθθθ 4
To get the usual normalization, we redefine V → 2gV in the Kähler
potential (and in the transformation rule), with g the gauge coupling:

h i
Φ† e2gV Φ = (Dµ φ)∗ Dµ φ + ψiσ µ Dµ ψ + F ∗ F
θθθθ
√ √
+ gD|φ|2 + ig 2φ∗ λψ − ig 2ψλφ (172)
1
Notice that this problem does not arise for global symmetries, when Λ is a constant
and thus automatically a chiral superfield.

26
where the covariant derivatives (not to be confused with the super-
symmetry covariant derivatives Dα , Dα̇ ) are defined as usual

Dµ φ = (∂µ + igvµ ) φ, Dµ ψ = (∂µ + igvµ ) ψ (173)


We would have guessed the first three terms, i.e. they are obtained
with the usual replacement ∂µ → Dµ , but the last three terms are of
purely supersymmetric origin.
Another possible supersymmetry and gauge symmetry invariant
term is the so-called Fayet-Iliopoulos term, which is nonzero only in
the U(1) case:

[ξ2gV ]θθθθ = gξD (174)


where ξ is an arbitrary real number.
For more chiral superfields, the Lagrangian, invariant under a U(1)
gauge symmetry and supersymmetry, can be written as

" #   
1 α
Φ†i e2gqi V Φi
X
L= + ξ2gV + W (Φi ) + W Wα + h.c.
4 θθ
i θθθθ
(175)
The U(1) gauge invariance can be explicitly seen (up to total
derivatives or field independent terms) from
i  
V →V − Λ − Λ† , Φi → eiqi Λ Φi (176)
2g
where qi is the U(1) charge of Φi . Notice that the superpotential
W (Φi ) must also be invariant under the U(1) guage transformation
(176).
Let us now rewrite (175) with component fields. Using (172), (174),
(141) and (164) we get

L = (Dµ φi )∗ Dµ φi + ψ i iσ µ Dµ ψi + |Fi |2
1 1
− Fµν F µν + λiσ µ ∂µ λ + D2
4 2
1 ∂2W
 
∂W
− Fi + (φ)ψi ψj + h.c.
∂φi 2 ∂φi ∂φj
√ √
+ gDqi |φi |2 + ig 2qi φ∗i λψi − ig 2qi ψ i λφi + gξD (177)

with the covariant derivatives

27
Dµ (φ, ψ)i = (∂µ + iqi gvµ ) (φ, ψ)i (178)
One can easily integrate out the auxiliary fields Fi and D:

∂W
Fi∗ = D = −g qi |φi |2 + ξ

, (179)
∂φi
and finally obtain

1
L = (Dµ φi )∗ Dµ φi + ψ i iσ µ Dµ ψi − Fµν F µν + λiσ µ ∂µ λ
4
1 ∂2W √
 

− (φ)ψi ψj − ig 2qi φi λψi + h.c. − V (180)
2 ∂φi ∂φj

where the potential V = V (φi , φ∗i ) is the sum of the F-term

X X ∂W 2
2
VF = |Fi | = (181)
∂φi
i i

and D-term
!2
1 g2 X
VD = D2 = qi |φi |2 + ξ (182)
2 2
i
i.e.
!2
2
X ∂W g2 X
V (φi , φ∗i ) = + qi |φi |2 + ξ (183)
∂φi 2
i i

7.2 The nonabelian case


Finally, the whole procedure gets a bit complicated if non-abelian
theories are considered. This amounts in changing the formulae above
as

V → V aT a, Λ → Λa T a (184)
where the generators of gauge transformation satisfy as usual
h i
T a , T b = ifabc T c (185)

28
The problem is, that now Λ, Λ† and V do not commute, and the
combination Φ† eV Φ is not invariant anymore, see eq. (168). We thus
generalize the gauge transformation of the vector superfield as
0 †
Φ0 = eiΛ Φ, e2gV = eiΛ e2gV e−iΛ (186)
which reduces to (176) in the U(1) case. Such a change would make
Wα from (156) gauge noninvariant, so we have to generalize it by
1
DD e−2gV Dα e2gV

Wα = − (187)
4
The function of the new terms with respect to (156) is simply to
introduce the covariant derivative on a gaugino λ and generalize the
expression for the gauge field strength:
 
Wα = T a Wαa , T r T a T b = Cδ ab (188)

Due to that we have to properly normalize the gauge kinetic term


1
[T r (W α Wα )]θθ + h.c. (189)
16g 2 C
A second difference with respect to the abelian case, is that there
is no Fayet-Iliopoulos term (174) now. The reason is simple: it would
be either gauge nonivariant (if left as it is) or zero (if taken with a
trace).
The Lagrangian for a supersymmetric nonabelian gauge theory can
thus be written first as

  
h i 1
Φ†i 2gV
W aα Wαa

L= e ij
Φj + W (Φi ) + + h.c.
θθθθ 16g 2 θθ
(190)
Then (177) generalizes to

L = (Dµ φi )∗ Dµ φi + ψ i iσ µ Dµ ψi + |Fi |2
1 a aµν a 1
− Fµν F + λ iσ µ Dµ λa + Da Da (191)
4 2
1 ∂2W
 
∂W
− Fi + (φ)ψi ψj + h.c.
∂φi 2 ∂φi ∂φj
√ √ a
+ gDa φ∗i (T a )ij φj + ig 2φ∗i λa (T a )ij ψj − ig 2ψ i λ (T a )ij φj

29
with

Dµ φi = ∂µ φi + igvµa (T a )ij φj (192)


Dµ ψi = ∂µ ψi + igvµa (T a )ij ψj (193)
Dµ λa = ∂µ λa − gfabc vµb λc (194)
a
Fµν = ∂µ vνa − ∂ν vµa − gfabc vµb vνc (195)

Integrating the auxiliary fields

∂W
Fi∗ = , Da = −gφ∗ (T a )ij φj (196)
∂φi
leads to the final result

1 a aµν a
L = (Dµ φi )∗ Dµ φi + ψ i iσ µ Dµ ψi − Fµν F + λ iσ µ Dµ λa
4
1 ∂2W √ ∗ a a
 
− ψi ψj − ig 2φi λ (T )ij ψj + h.c. − V (197)
2 ∂φi ∂φj
where the potential V = V (φi , φ∗i ) is again a sum of the F-terms

X X ∂W 2
VF = |Fi |2 = (198)
∂φi
i i
and D-terms
X1 g2  ∗ a 2
VD = Da Da = φi (T )ij φj (199)
a
2 2
giving
2
X ∂W g2  ∗ a 2
V (φi , φ∗i ) = + φi (T )ij φj (200)
∂φi 2
i

8 Renormalization
Although in principle a supersymmetric theory is just a field theory
and so one can use for renormalization the usual rules, there are some
particularities that is good to keep in mind.
First, to keep supersymmetry explicit, one needs the same num-
ber of degrees of freedom for bosons and fermions. This means that

30
in using dimensional regularization, one should always run the space-
time index µ of gauge bosons from 0 to 3. Such a scheme is called
dimensional reduction (DRED) instead of dimensional regularization
(DREG). Since at one loop order they both lead to same results, we
will not make the above statements more explicit.
Second, as known, in many processes radiative contributions of
fermion loops have a minus sign with respect to boson contributions
in the loops. Since in supersymmetry couplings are related, sometimes
these cancellations are exact, and true at any order of perturbation
theory. This is the case of the superpotential, for which it has been
shown, that, due to its holomorphicity, it does not get renormalized
at any order in perturbation theory (the famous non-renormalization
theorem). Since we are considering only renormalizable models, only
the renormalization of the (canonical) Kähler is possible: there is
nothing more than wave-function renormalization:
1/2
Φ = ZΦ ΦB (201)
where Φ is the renormalized field and ΦB the bare one. Of course the
other parameters in the superpotential gets renormalized by induction.
Take for example the simplest Wess-Zumino model

m 2 λ 3
W = Φ + Φ (202)
2 3
The nonrenormalization theorem tells us that the superpotential
looks the same written in original (ΦB , mB , λB ) or renormalized (Φ,
m, λ) quantities, i.e.

mB 2 λB 3 m λ
ΦB + ΦB = Φ2 + Φ3 (203)
2 3 2 3
This means that
−3/2
m = ZΦ−1 mB , λ = ZΦ λB (204)
In other words, once we know the wave-function ZΦ , we know
everything needed about renormalization.
So for example when we are looking for the renormalization group
equation for the coupling constant, we get for t = log µ (remember
that the bare parameters are independent of µ)

dλ d −3/2 3 d
= λ B ZΦ = − λ log ZΦ (205)
dt dt 2 dt

31
In a similar way the mass is also multiplicatively renormalizable:

dm d
= −m log ZΦ (206)
dt dt
On top of that, in gauge theories, the gauge couplings also undergo
the process of renormalization. In an ordinary theory, the one loop
result for the running of the gauge coupling constant is given by

dg b
=− g3 (207)
dt (4π)2
where the 1-loop β function coefficient is
11 2 1
b= C(G) − TF − TB . (208)
3 3 3
The Dynkin index

T (R)δab = T r (Ta (R)Tb (R)) (209)


and the second Casimir
X
C(R)δ ij = (Ta (R)Ta (R))ij (210)
a
depend on the choice of the gauge group and on the representation
involved. The indices a, b run over the generators of the group (N 2 −1
in SU(N)), while i, j run from 1 to the dimension of the representation.
The normalization usually chosen is T = 1/2 for the fundamental
representation (quarks, leptons). Then one has in the SU(N) group
for the fundamental representation C = (N 2 − 1)/(2N ), and for the
adjoint T = C = N . The dimension of the representation is N for
fundamentals and N 2 − 1 for adjoint. To remember also that in SU(2)
the generators in the fundamental are the Pauli matrices Taij = τaij /2,
while in the adjoint representations are the Levi-Civita antisymmetric
tensor Taij = −iaij .
For supersymmetric theories we know that for each fermion (boson)
there is a boson (fermion) in the same group representation, so (208)
can be written more compactly as

b = 3C(G) − T . (211)
Now let us consider a specific example from [16]

W (Φ) = λΦ1 Φ2 Φ3 (212)

32
The RG equation for the wave-functions are
 
d 1 Di
log Zi = Ci α − αλ (213)
dt π 2
where as usual α = g 2 /4π, αλ = λ2 /4π. Ci is the quadratic Casimir
of Φi and Di is the number of internal fields involved in the Yukawa
loop of the Φi propagator. The RGE for the coupling
P constant
P can be
derived in a similar way as (205) giving (C = i Ci , D = i Di and
the sum goes over all the fields in (212))
 
d αλ D
αλ = αλ − Cα (214)
dt π 2
which together with (207)

d −1 b
α = (215)
dt 2π
form a closed system of differential equations.
All this simplifies the calculation with respect to ordinary, non
supersymmetric, theories. On top of that, all parameters (massive
or dimensionless) are multiplicatively renormalized. This means that
its radiative correction is proportional to the tree order value. We
have always been used to that for example for fermion and gauge
boson masses, because the zero mass limit represented a new symme-
try (chiral or gauge). But in ordinary theories this was not true for
the mass of a spin 0 field: a tree order massless spin 0 boson could
get through radiative corrections in principle arbitrary contributions
to its mass. Supersymmetry prevents that, since it links the spin 0
boson mass to the spin 1/2 fermion mass, which is protected by chi-
ral symmetry. In the technical language this is a consequence of the
non-renormalization theorem of the superpotential.
In practice the above results are connected with the absence of
quadratic divergences in supersymmetric field theories. What happens
is that loops with internal fermions cancel the quadratic divergent
piece of loops with internal bosons, leaving in the final result at most
logarithmic divergences.

33
9 The minimal supersymmetric stan-
dard model (MSSM)
Now we have enough knowledge to supersymmetrize a realistic model,
i.e. the standard model. Is it enough to promote each field of the
standard model to a (chiral or vector) superfield? It turns out that
two are the things we have to worry about.
First, as is clear from the above, the superpotential W is a function
of chiral superfields. Superfields contain the spinor ψ, which is nothing
else than the left-handed field in the Dirac notation. On the contrary
its hermitian conjugate W ∗ can contain only right-handed spinors ψ.
But the standard model needs both left-handed fields QL , LL , and
right-handed fields uR , dR , eR . This apparent problem is fortunately
only a problem in notation: defining the conjugated field as
α̇
ψαc ≡ i σ 2

αα̇
ψ (216)
we see that ψ c is a left-handed field (has Lorentz index α), although
the original field ψ was a right-handed field (dotted index α̇). Of
course this is nothing else than the 2-component analogue of the 4-
component
T
Ψc = CΨ (217)
Since C = iγ 2 γ 0 it automatically follows that

1 − γ5 1 + γ5
   
Ψ=Ψ → Ψc = Ψc (218)
2 2
i.e. if Ψ is right-handed, Ψc is left-handed.
So the fermion fields of the standard model (and its bosonic part-
ners) will be described by the following chiral superfields:

Q ∼ (3, 2, 1/6)
L ∼ (1, 2, −1/2)
uc ∼ (3̄, 1, −2/3) (219)
c
d ∼ (3̄, 1, 1/3)
c
e ∼ (1, 1, 1)

34
with the SU(3)C ×SU(2)L ×U(1)Y /2 numbers explicitly specified. No-
tice the reversed representations or opposite quantum numbers of the
conjugated fields.
The second complication comes from the fact that transforming
the Higgs boson into a chiral superfield introduces a new fermionic
field (the spin 1/2 Higgsino), so that the anomaly constraints of the
standard model gets automatically changed. This would spoil the
anomaly cancellation conditions and thus the renormalizability. It is
thus mandatory to add another Higgs superfield with the opposite
hypercharge. Together we need two chiral superfields (the notation
will become clear in the next subsection):

Hu ∼ (1, 2, 1/2)
Hd ∼ (1, 2, −1/2) (220)

Finally we supersymmetrize the gauge bosons, introducing the vec-


tor superfields

VB ∼ (1, 1, 0)
VW ∼ (1, 3, 0) (221)
VG ∼ (8, 1, 0)

9.1 The supersymmetric Lagrangian


Let us concentrate first on the superpotential. To mimic as much
as possible the usual Yukawa terms, see (197) and assuming a non-
zero (µ) mass for the Higgs superfields we write the following gauge
invariant terms:

W = µHu Hd + YUij Hu Qi ucj + YDij Hd Qi dcj + YEij Hd Li ecj (222)

where i, j are generation indices. The above compact notation should


read for example

YUij Hu Qi ucj → YUij (Hu )m (iτ 2 )mn (Qi )αna (ucj )αa (223)
where α = 1, 2 is the Lorentz (left) SU(2) index, m, n = 1, 2 the gauge
SU(2) indices, and a = 1, 2, 3 the gauge SU(3) index .

35
It is now clear that the notation Hu,d tells us, to which right-
handed quark each Higgs is coupled in the superpotential. Also, we
see that there is another reason why it is not enough to introduce
one single Higgs in supersymmetry: either the up quarks or the down
quarks would have to be massless, since terms with Hu,d∗ cannot be

introduced due to the holomorphicity of the superpotential.


The Kähler potential is the canonical one:

= Q† exp (2gs VG ) exp (2gVW ) exp g 0 VB /3 Q



KM SSM
+ (uc )† exp (−2gs VG ) exp −4g 0 VB /3 uc


+ (dc )† exp (−2gs VG ) exp 2g 0 VB /3 dc




+ L† exp (2gVW ) exp −g 0 VB L



(224)

+ (ec ) exp 2g 0 VB ec


+ (Hu )† exp (2gVW ) exp g 0 VB Hu




+ (Hd )† exp (2gVW ) exp −g 0 VB Hd




Finally the gauge kinetic terms are found from

8 3
" #
X W aα (VG )W a (VG )
α
X W aα (VW )W a (VW )
α W α (VB )Wα (VB )
+ + +h.c.
16gs2 16g 2 16g 02
a=1 a=1 θθ
(225)
where

8
X λa 1
Wαa (VG ) DD e−2gs VG Dα e2gs VG

= −
2 4
a=1
3
X τa 1
Wαa (VW ) DD e−2gVW Dα e2gVW

= − (226)
2 4
a=1
1
Wα (VB ) = − DD Dα 2g 0 VB
 
4
where λa and τ a are the Gell-Mann and Pauli matrices respectively.
The particles of the MSSM are the one from the SM, quarks (Q,
uc , dc ), leptons (L, ec ), Higgses (Hu , Hd ), gauge bosons (g, W , B),
and their supersymmetric partners, squarks (Q̃, ũc , d˜c ), sleptons (L̃c ,
ẽc ), Higgsinos (H̃u , H̃d ), gauginos (g̃, W̃ , B̃). There are twice as much
particles than in the standard model, plus one extra Higgs supermul-
tiplet.

36
9.2 R-parity and other symmetries
The above superpotential (222) has 7 type of fields: Q, L, uc , dc , ec ,
Hu , Hd . Assuming only generation independent symmetries, we have
4 terms, so there can be 7 − 4 = 3 independent U(1) symmetries.
These are the gauge hypercharge, the baryon and the lepton numbers.
In the limit µ → 0 there is another symmetry, under which the
matter fields (Q, uc , dc , L, ec ) have charge +1, while the Higgses
(Hu , Hd ) have charge −2. This is called the Peccei-Quinn symmetry,
since it transforms in the opposite way left-handed and right-handed
fermions.
This would be all if we were not in supersymmetry. In fact there are
other fermions in the theory, gauginos, that could in principle rotate
by a (axial) phase. One could think that any gaugino could have its
own independent rotation, but a quick glance to (99) is enough to
convince us of the contrary. Since any gauge boson is a real field,
it cannot have any phase redefinition, and so the U(1) rotation of
the gaugino λ must be neutralized by an equal rotation of the spinor
coordinate θ. The same spinor θ appears in all vector and chiral
superfields, so all gauginos have the same rotation, which is also equal
to the difference between the rotation of the boson φ and fermion ψ
in any chiral multiplet.
This is confirmed by an explicit check in (197). We can thus sum-
marize the charges of this so called U(1) R-symmetry:

R(λ) = R(φi ) − R(ψi ) = R(θ) = 1, (227)


α
R(W ) = R(W ) = R(φ) − R(Fi ) = 2

For example, MSSM is invariant under the following R-symmetry:

R(Hu ) = 2, R(Hd ) = 0, R(uc ) = −1 (228)


c c
R(Q) = R(d ) = R(L) = R(e ) = R(λ) = R(θ) = 1

The mentioned symmetries are not necessary anomaly free, al-


though some linear combinations are (for example the hypercharge,
or the baryon minus lepton number).
At this point we must however realize, that the situation is different
from the standard model. In fact, in the non-supersymmetric stan-
dard model gauge symmetry was enough to forbid any renormalizable

37
baryon or lepton number violating term. In some sense, the (approx-
imate) baryon and lepton conservation is a success of the standard
model. Any violation of them must automatically include a new scale
and it is suppressed by it. This is not true in the supersymmetric
version of the standard model. In fact, although (222) is baryon and
lepton conserving, not all gauge invariant renormalizable operators
have been included. One could as well add

δW = i Hu Li + λijk Li Lj eck + λ0ijk Li Qj dck + λ00ijk uci dcj dck (229)

The total superpotential with these terms included would in gen-


eral be invariant only under the gauge symmetries. Such terms would
violate baryon and lepton numbers, and thus mediate proton decay.
Due to strong constraints the new baryon and/lepton number violat-
ing parameters in (229) should typically be small. It is thus tempting
to regulate this smallness with a symmetry. An ideal candidate is a
Z2 subgroup of the U(1) R-symmetry (228). If we write the above
symmetry transformation as

Φ0i = eiR(Φi )α Φi (230)


then it is enough to keep the element denoted by α = π. We call this
discrete symmetry R-parity. All the fields of the SM have a positive
parity, while all the superpartners have negative R-parity. It can be
written as

R = (−1)3(B−L)+2S (231)
where B, L and S are baryon number, lepton number, and spin.
Clearly, this parity, as all R-symmetries, does not commute with su-
persymmetry (bosons and fermions in the same multiplet have differ-
ent R-charges).
Most works in the MSSM assume this parity. It must be kept in
mind however, that there is really no good reason to believe it exact.
In fact, all we know is that constraints from nucleon lifetime (and
others) must be satisfied. But this still allows at least some of the
baryon and/or lepton number violating parameters nonzero, and in
some cases also large. Not only, with for example just lepton number
violating terms (and without introducing any other degree of freedom
like for example gauge singlet right-handed neutrinos) we can describe
the nonzero neutrino masses, without entering into conflict with any

38
experimental result. So, although we will assume in the following for
simplicity that this R-symmetry is exact, we must remember that this
is just an assumption, which will have to be experimentally verified
sooner or later.

9.3 The supersymmetry breaking soft terms


As we repeatedly said, supersymmetry implies exact degeneracy among
members of the same supermultiplet. This means that for example
a spin 0 particle with the same mass and quantum numbers of the
electron must exist. Since this is experimentally ruled out, super-
symmetry, if it exists, must be broken. In this section we will not
be interested in how supersymmetry is broken, i.e. in the dynamics,
which study we postpone. All we need now is the parametrization of
this breaking, applied to the (R-parity conserving) MSSM.
A strongly broken symmetry is not a symmetry even as an approx-
imation, so we will limit ourselves to soft breaking. Technically this
means that only operators with mass dimension less or equal 3 are
allowed, which in turn implies in theories without gauge singlets (like
the MSSM) that no quadratic divergences will be generated.
Let’s now classify them (we limit ourselves to the R-parity con-
serving ones):
• Gaugino masses: mλ λλ + h.c.

mB̃ B̃ B̃ + mW̃ W̃ W̃ + mG̃ G̃G̃ + h.c. (232)


 
• Boson masses: m2φ̃ φ̃i φ̃∗j
ij


   

m2Q̃ Q̃i Q̃∗j + m2ũc ij ũci u˜c j + m2d˜c d˜ci d˜c j

(233)
ij ij
 

m2L̃ L̃i L̃∗j + m2ẽc ij ẽci e˜c j + m2Hu |Hu |2 + m2Hd |Hd |2

+
ij

• bilinear terms (B terms): Bij φi φj + h.c.

BHu Hd + h.c. (234)


• trilinear terms (A terms): Aijk φi φj φk + h.c.

Aij c ij ˜c ij c
U Hu Q̃i ũj + AD Hd Q̃i dj + AE Hd L̃i ẽj + h.c, (235)

39
• non-holomorphic trilinear terms (A0 terms): A0ijk φ∗i φj φk + h.c.

A0ij † c 0ij † ˜c 0ij † c


U Hd Q̃i ũj + AD Hu Q̃i dj + AE Hu L̃i ẽj + h.c. (236)

Notice that there is no need to add the fermionic mass terms, since
with proper redefinition of the parameters in the superpotential and
the soft terms of the form above one can always get rid of them.
What are the experimental constraints on these parameters? First
of all, let’s remind the reader, that the parameters in the superpo-
tential are (with the exception of µ) the same as in the standard
model. The soft terms however describe interactions of yet unfound
particles, like Higgses, sfermions or gauginos. So the first constraints
come already from direct searches at LEP, which typically give for the
lower limit for masses something around 100 GeV or so. Another way
of constraining them are through rare processes, for example flavour
changing neutral currents (FCNC) mediated K − K mixing, b → sγ,
µ → eγ, µ → 3e, µ − e conversion, etc. Light sfermions could in prin-
ciple contribute hugely to these processes. This limits the above soft
parameters, although due to the large number of them, not uniquely.
For example, the easiest way to get rid of such unwanted contribu-
tion is to substantially increase the masses of the spartners, since all
the above rare processes are inversely proportional to these masses.
Masses of order 10 to 100 TeV would suffice, but this would be sad
for our perspectives at LHC. Other possibilities in the huge MSSM
parameter space are however good as well. For example, if the off-
diagonal m̃2 , A and A0 terms are very small (or the A and A0 have
a similar flavour structure as their corresponding Yukawa matrices),
then the FCNC contributions will be similarly small as in the SM.
A detailed analysis of these processes and constraints is beyond the
scope of this introduction.

9.4 Electroweak symmetry breaking


There is another experimental constraint on the MSSM parameters:
this is the spontaneous electroweak symmetry breaking. It is easy to
see, that without soft terms and nonzero µ the Higgses cannot get
nonzero vacuum expectation values. So here the soft contributions
are crucial. Let us study this in more detail. The Higgs potential is

40
|µ|2 + m2Hu |Hu |2 + |µ|2 + m2Hd |Hd |2 + (BHu Hd + h.c.)
 
V (Hu , Hd ) =
g2  † a 2 g 02  2
+ Hu τ Hu + Hd† τ a Hd + Hu† Hu − Hd† Hd (237)
8 8
The terms proportional to |µ|2 come from the superpotential, the
last two are the D-terms for SU(2) (∝ g 2 ) and U(1) (∝ g 02 ), while the
other are supersymmetry breaking soft terms.
In terms of the U(1)EM and CP (i.e. vu,d real) preserving vevs

   
0 vd
hHu i = , hHd i = (238)
vu 0

the potential becomes (with a rotation of Hu,d we can always make B


real and positive, so vu,d will also be positive)

g 2 + g 02 2 2
V (vu , vd ) = m2u vu2 + m2d vd2 − 2Bvu vd + vu − vd2 (239)
8
where we redefined

m2u,d = |µ|2 + m2Hu,d (240)


By inspecting the potential in the vu = vd direction we can imme-
diately see that it is bounded from below only if

m2u + m2d − 2B ≥ 0 (241)


The equations of motion are

∂V g 2 + g 02
= 2m2u vu − 2Bvd + vu vu2 − vd2 = 0

(242)
∂vu 2
∂V g + g 02
2
= 2m2d vd − 2Bvu − vd vu2 − vd2 = 0

(243)
∂vd 2
From
∂V ∂V g 2 + g 02 2 2
0 = vu + vd = 2m2u vu2 + 2m2d vd2 − 4Bvu vd + vu − vd2
∂vu ∂vd 2
(244)
we immediately find out that in the minimum

41
g 2 + g 02 2 2
V =− vu − vd2 (245)
8
i.e. any minimum will be lower than the one at vu = vd .
Now we write (experimentally v = 174 GeV)

vu2 + vd2 = v 2 , g 2 + g 02 v 2 = 2MZ2



(246)
and introduce
vu
tan β = (247)
vd
On the one side, from

∂V ∂V
= 2 m2u + m2d vu vd − 2B vu2 + vd2
 
0 = vd + vu (248)
∂vu ∂vd
we get

B
sin β cos β = (249)
m2u + m2d
On the other side, from

∂V ∂V
= 2 m2u + MZ2 /2 vu2 − 2 m2d + MZ2 /2 vd2 (250)
 
0 = vu − vd
∂vu ∂vd
it follows

m2d + MZ2 /2
tan2 β = (251)
m2u + MZ2 /2
Comparing now (249) and (251) we obtain a nontrivial relation
between the parameters

2
M2

m2u MZ2 /2 m2d MZ2 /2 2
1+ 2 Z 2
 
+ + =B (252)
mu + md

Since the sum of the two factors on the left-handside is positive


(see (241) and use the positivity of B) each factor separately must
also be positive.
A little algebra gives

42
2 2
m2u + m2d + MZ2 B − m2u m2d

2
m2u − m2d m2u + m2d + MZ2 /2 MZ2 /2

= (253)

from which it follows that

B 2 ≥ m2u m2d (254)


This inequality means that there is a tachyonic state at vu = vd =
0, and so not even a local minumum is possible at the origin.
What would change if we added the Fayet-Iliopoulos term to the
MSSM? Effectively this would amount in an extra factor (183)

g 02
ξ |Hu |2 − |Hd |2

δF I V (Hu , Hd ) = (255)
2
The same results as above would still be valid, providing now

m2u = |µ|2 + m2Hu + g 02 ξ/2, m2d = |µ|2 + m2Hd − g 02 ξ/2 (256)

instead of (240).
To summarize, the constraints (252) and (241) must be satisfied
by the MSSM parameters.

10 The mass spectrum in MSSM


As we already saw, there are not many hints on where should the
spartners lie. I spite of this we will see that there are some relations
among different masses that can represent a test of the MSSM. Let us
discuss the different options.

10.1 Higgs mass


This can be easily found out from

V (Hu , Hd ) = m2u |Hu |2 + m2d |Hd |2 + 2B (Hu Hd + h.c.) (257)


2
g +g 02 2 g 2    
+ |Hu |2 − |Hd |2 + Hu† Hd Hd† Hu
8 2
Then we first explicitly expand in terms of

43
φ+ vd + φ0∗
   
u d
Hu = , Hd = (258)
vu + φ0u φ−
d

Using the obvious notation


1
φ−∗ +
φ0u,d = √ (Ru,d + iIu,d )
u,d = φu,d , (259)
2
we can rewrite the Higgs quadratic terms in the form

1 + +  2 φ−
     
1  2 Ru 1  2 Iu u
Ru , Rd MR + Iu , Id MI + φ , φ M± (260)
2 Rd 2 Id 2 u d φ−
d

The matrices MI2 and M±2 have one zero eigenvalue each, i.e. due
the would-be Goldstones eaten by the Z and W ± . The eigenvectors of
MR2 are the two CP even neutral scalars h0 (the lighter) and H 0 (the
heavier), while the physical remaining eigenvalues of MI2 and M±2 are
the CP odd neutral scalar A0 and the charged H ± .
The masses of h0 , H 0 , A0 and H ± cannot be arbitrary. There are
experimental lower limit constraints as well as theoretical relations
among. The most interesting is the theoretical upper bound for the
mass of h0 . In fact, one finds at tree level

mh0 < MZ |cos (2β)| (261)


This upper limit is unexpected (nothing of this kind happens in
the standard model for example), but it gets corrected at 1-loop. The
full result is quite complicated but in the large stop mass mt̃  mt , At
(from the soft term At Hu t̃t̃c ) and large tan β limit it reduces to [17]
!
2 m4 m 2
3g t
m2h0 < 2
∼ MZ + 8π 2 M 2 log m2

(262)
W t

The bound now depends on the value of the unknown stop mass,
although only logarithmically. It can be further relaxed by a larger At
value. For not too large stop mass mt̃ < ∼ 100 TeV), the lightest Higgs
boson is still quite light, less that 200 GeV or so. For low tan β the
experimental bound on the Higgs mass can be reached and constitutes
a constraint on tan β.
It is often claimed that MSSM predicts for the lightest Higgs to
be lighter than 130 GeV or so. Strictly speaking, since the other

44
MSSM parameters (mt̃ , At , etc) are not known, such a statement is not
correct, and these type of constraints should actually be interpreted
as relations among the (unknown) supersymmetric parameters.

10.2 Chargino mass


We have originally four charged massless Weyl spinors, W̃ + , W̃ − , H̃u+ ,
H̃d− , in the supersymmetric Lagrangian with µ = 0 and unbroken
electroweak symmetry. Notice that their charged conjugated spinors
are W̃ + , W̃ − , H̃u+ , H̃d− , which are not W̃ − , W̃ + , H̃u− , H̃d+ . In fact W̃ +
and W̃ − are two independent spinors, not connected by conjugation,
while H̃u− and H̃d+ do not even exist. The charges on a spinor denote
a different state, not only a charge. It is more like writing ψ̃ ± = a±ib,
with a and b complex.
Once µ, supersymmetry and electroweak symmetry breaking are
introduced, the four spinors have a 4 × 4 mass matrix


W̃ +
1 +  −
− +

− M  W̃ 
W̃ , W̃ , H̃u , H̃d C̃   + h.c. (263)
2  H̃u+ 
H̃d−

There are only two different eigenvalues, as expected: a charged


spinor can be either massless (not our case) or of a Dirac type (i.e.
two spinors have the same mass and thus form a massive Dirac field).
Only neutral spinors can have Majorana masses.
The eigenvectors are called chargini and are conventionally denoted
by C̃i± , with i = 1 for the lighter, i = 2 for the heavier.

10.3 Neutralino mass


The four neutral fermionic spartners B̃, W̃ 0 , H̃u0 , H̃d0 , have a 4 × 4
Majorana mass matrix

 

1  W̃ 0 
B̃ , W̃ 0 , H̃u0 , H̃d0 MÑ 
 H̃ 0  + h.c.
 (264)
2 u
H̃d0

45
The four linear combinations (mass eigenstates) of the original
fields are called neutralini. They are denoted by Ñi , i =1 for the
lightest, i = 4 for the heaviest. In a generic point of the MSSM pa-
rameter space they are massive Majorana particles, although neither
a Weyl (one massless eigenvalue) or a Dirac (two equal eigenvalues)
possibility is excluded.

10.4 Sfermion mass


Finally we have the sfermions, i.e. the spin 0 partners of the SM
fermions. They are three generations of q̃ = ũ or d, ˜ charged leptons ˜l
and neutral leptons (sneutrinos) ν̃. Every charged sfermion has both
the left (f˜) and right component (f˜c ) and so its mass is a 6×6 matrix.
Since there is no right-handed neutrino superfield (ν̃ c ) in the MSSM
the mass matrix for the neutral ν̃ is only 3 × 3.

 2 d˜
   
ũ ˜ ˜
ũ∗ , ũc Mũ2 ∗
 c
+ d ,d M d˜ d˜c∗
ũc∗
 

ẽ∗ , ẽc Mẽ2 c∗ + ν̃ ∗ Mν̃2 ν̃

+ (265)

Notice that the above masses are already Hermitian, so that the
eigenvectors are complex fields, as they must, being charged massive
bosons.

10.5 Gluino mass


There is not much to say here. Gluino cannot mix with any other field
due to the unbroken colour SU(3). The Majorana mass term is simply
1
g̃Mg̃ g̃ (266)
2
with Mg̃ a complex number (not a matrix).

11 Possible experimental signatures


from supersymmetry
As we have seen the MSSM has more than 100 parameters. Although
some combinations of them are actually excluded already by current

46
data, almost all of the parameter space is still available. So predic-
tions are very model dependent. Essentially there are two ways of
constraining the parameter space: one is through rare processes, the
other through direct collider searches. This has led so far to various
lower limits for the sparticle masses (from LEP II and Tevatron they
must be > ∼ 100 GeV or so, although exceptions are possible) and small
mixings in the sfermion mass matrices. We will assume in the fol-
lowing that the spectrum is low enough to be detectable at the LHC.
However, nothing is known about which sparticle is lighter which is
heavier (with few exceptions, like the CP even Higgs boson mentioned
before), so that a precise prediciton of the processes that will dominate
and the decays involved is very difficult. In other words, the super-
symmetric spectrum is practically arbitrary. Most of the analysis are
done having some particular model in mind, and so the expectations
should be taken with care.
Anyway, generically sparticles are generated in colliders mostly
through the following interactions

Ṽ Ṽ V, Ṽ f˜f¯, f˜f˜∗ V, f˜f˜∗ V V (267)


where V is a gauge boson and f a fermion.
Things simplify if R-parity is assumed. Then sparticles are always
produced in pairs. The lightest supersymmetric partner (LSP) is sta-
ble and escapes the detector (it behaves like a massive neutrino). So
large missing energy is a common prediction of R-parity conserving
low energy supersymmetry models.
If the squarks and gluinos are not too heavy (< ∼ 1 TeV), the Large
Hadron Collider (pp collider at 14 (10) TeV), will produce them in
pairs mainly through strong interactions with gg → g̃g̃, q̃ q̃ ∗ , presum-
ably the dominant modes through s-channel gluon and t-channel g̃, q̃
exchange, or gq → g̃ q̃ through t-channel g̃ exchange. Other possibili-
ties are productions of chargini and/or neutralini through electroweak
processes. Of course all these final state sparticles decay into SM
particles plus lighter sparticles again, until only jets, leptons and two
LSPs are left.
In most cases such processes have large SM backgrounds that can
hide the signal. This makes some processes that have small SM back-
ground particularly interesting. For example, in the SM missing en-
ergy comes from neutrini. Since they are mostly produced in combina-
tion with charged leptons through W , missing energy without charged
leptons is predicted relatively small in the SM, and it is considered one

47
of the golden plates of supersymmetry. On top of the choice of the
channel, particular care must be taken to put kinematical cuts, as
usual.
Notice that due to the non-detectability of the lightest neutralino,
finding the masses of the sparticles produced is not so easy. It turns
out that careful analysis (through kinematical edges) could give some
information on combinations of masses, mainly differences. In any
way, even if LHC finds some spartilcles, this will probably be just the
beginning of a long search, and other, more precise colliders will have
to be used.

12 What is supersymmetry good for?


We have seen that supersymmetry is an interesting theoretical con-
struction that could give us a hint of what is the physics beyond the
standard model. If MSSM is really realized in nature, then there is a
very rich phenomenology waiting us in future. Unfortunately there is
no hint of where supersymmetry should be broken, i.e. what are the
values of the soft terms. This seems to make our possibility to find
supersymmetry in near future colliders wishful thinking, since in prin-
ciple any scale between MZ and MP lanck looks equally good. We will
see in this section that there are actually various arguments of why
at least some of the spartners should lie in the TeV region, possibly
accessible at the LHC.

12.1 Gauge coupling unification


The first hint of why this should happen comes form the possibility of
gauge coupling unification. This idea postulates that at some large en-
ergy the three gauge interactions of the standard model are described
by just one type of gauge interaction, that embodies the known three
SM ones. The SM is thus just the low energy approximations of such
a grand unified theory (GUT), that unifies not only the gauge interac-
tions, but also the matter fields (leptons and quarks typically behave
similarly, although details depend on the choice of the grand unified
group and representations).
As we know, in the standard model the gauge couplings do not
unify, and new physics is needed to change at some point the beta
functions and restore running to unification. Such new physics could

48
be supersymmetry. In fact, using (211) it is a simple exercise to show
that the beta coefficients in MSSM are bi = (−33/5, −1, 3) instead
of bi = (−41/10, 19/6, 7) in the SM (positive coefficients here mean
asymptotic freedom). One knows the experimental values of gi at MZ
and can evolve them towards larger scales µ using the three equations
(215) for the three gauge couplings. Assuming that all the superpart-
ners lie at the same scale m̃, we can find that they unify to a single
point at µ = MGU T ≈ 1016 GeV providing m̃ ≈ 1 TeV.
I believe this is the best argument for low energy supersymmetry.
It has to be stressed however that the solution is not unique. If for
example only gauginos and higgsinos lied at TeV, while the sfermions
and second Higgs were heavier (such a solution is called split supersym-
metry), nothing would really change. But more split supersymmetry
is also possibile: a light (TeV) wino, intermediate (≈ 108 GeV) gluino
and heavy rest would also unify. The conclusion is however always
similar: at least some partner should be pretty light, although not
always in the reach of the LHC.

12.2 Dark matter


Another good reason to believe in supersymmetry is the dark matter
problem in the universe. As we know, there is no dark matter can-
didate in the SM (a SM with added three right-handed neutrinos to
describe the neutrino mass and mixings have a dark matter candidate
[18]). What is missing is essentially a discrete symmetry to make a
particle stable. This is provided in MSSM with R-parity. The lightest
supersymmetric partner (LSP) is stable due to R-parity. If the spec-
trum is such that the LSP is neutralino (charged dark matter is not
allowed by observation), then it feels weak interactions and is thus a
weakly interacting massive particle (WIMP), which is known to be an
ideal dar matter candidate. It is important to realize here that strictly
speaking supersymmetry is not responsible for that. What is crucial
is the R-symmetry, but its conservation has nothing to do with su-
persymmetry. Also it is not important whether all sparticles are light
(low energy supersymmetry) or we have a kind of split supersymetry:
what is important is just a light (TeV) neutralino.
It is interesting that supersymmetry provides another candidate
on top of the neutralino. As we will se later, the generalization of su-
persymmetry to gravity introduces a spin 3/2 partner of the graviton,
called gravitino. It turns out that gravitino can also be a dark matter

49
candidate.

12.3 The hierarchy problem


Although there are many different ways of describing this problem,
one can approxiamtely say, that the issue here is the stability (or bet-
ter, instability) of the electroweak scale under quantum corrections.
This is visible only in the quantum corrections of some lower dimen-
sional operators, more precisely, in the coefficient (mass squared) of
the bilinear product of scalar fields. The problem is easily seen using
a sharp cutoff regularization. This regularization is however not of-
ten used in field theory, on top of this it can give also wrong results,
especially because it explicitly breaks Lorentz and gauge invariance.
On the contrary, dimensional regularization is the correct tool in the
regularization and renormalization procedure, and so we will use it
here.
To have a feeling of what the hierarchy problems is about, let us
consider a simple toy model with a charged scalar field H, which I
will call the Higgs boson for future use, another charged boson field φ
and a Dirac fermion field ψ. The tree order masses of the three fields
will be denoted by mH0 , mB0 and mF 0 , while the interaction terms
relevant for our discussion are

Lint = −λ |H|2 |φ|2 − yH ψ̄ψ + h.c. . (268)


Now let us calculate the 1-loop correction to the Higgs mass: there
will be one Feynman diagram with the boson φ and one with the
fermion ψ, giving after dimensional regularization and M S renormal-
ization

λ y2
m2H = m2H0 − cB m 2
B0 + cF m2 + ... , (269)
16π 2 16π 2 F 0
where the dots denote the eventual log terms coming after the inte-
gration and are not relevant for our discussion. What is relevant is
the fact that the second and third terms on the righthandside of (269)
are corrections to the mass not proportional to itself, i.e. they correct
the mass mH0 , but are not dependent on mH0 . In particular, they can
give a nonzero correction, even if the original mass mH0 were zero. In
other words, there is no multiplicative renormalization for the scalar
(Higgs) mass.
So we have the following two possibilities:

50
(1) mH0 ≈ 100 GeV (electroweak scale); if there are heavy particles,
mB0,F 0  mH0 , it follows that mH  mH0 . The Higgs mass got
destabilized in this case by quantum corrections.
(2) mH ≈ 100 GeV (this is almost an experimental fact, although,
admittedly, the Higgs boson has not been found yet). In this case,
if mB0,F 0  mH , one needs also a very large tree approximation
mH0  mH to cancel the large contributions due to mB0,F 0 , again a
strange and somewhat unnatural situation.
Of course, strictly speaking, in the standard model alone there is
no problem at all. In fact:
- there is no extra boson;
- mF 0 ≈ y < H >= O(mH0 ).
However, as soon as we have some new physics (at MGU T , MP lanck ),
eq. (269) becomes

2 2 2 2 2
O(MGU T or MP lanck ) + mH0 ≈ mH = O((100GeV) ) , (270)

which gives

O((1016 GeV)2 or (1019 GeV)2 ) + m2H0 ≈ m2H = O((100GeV)2 ) .


(271)
But now we clearly need to fine-tune mH0 ! And not only approx-
imately, but all the 16 or so decimal points! This is very unnatural
and this is what we call the hierarchy problem.
This is typical for scalar (spin=0) fields. In fact for fermions
(spin=1/2) one can easily find out that the one-loop correction to
the mass is

y2
mF = mF 0 + c mF 0 + ... . (272)
16π 2
So the correction is proportional to the tree order value, i.e. the
mass renormalization for fermi fileds is multiplicative. If for example
the tree order mass is zero, then no perturbation will ever generate a
nonzero mass (this is not necessary for nonperturbative corrections,
but these are out of the scope of these lectures and I will not consider
them here), i.e.

mF 0 = 0 → mF = 0 . (273)

51
There is a symmetry here that forbids the appearence of a mass
term. If mF 0 = 0 then the Lagrangian has a U(1)2 chiral symmetry
and it is invariant under ψL,R → exp (iαL,R )ψL,R , which can not be
broken at any perturbative order.
Similarly, in the case of spin 1 particles (for example gluons), it is
the gauge symmetry that makes the mass renormalization multiplica-
tive.
One of the possible solutions to the hierarchy problem is super-
symmetry. Essentially what supersymmetry does is to connect the
Higgs self-coupling λ and the Yukawa coupling y by λ = y 2 , as well
as the boson and fermion masses by mB0 = mF 0 , so that the partic-
ular combination in (269) cancels out. This is equivalent to say that
there are no quadratic divergences in supersymmetric models.
Of course supersymmetry must be broken. In fact there is no ex-
perimental evidence for example for the scalar partner of the electron
with the same mass. Although we still need all the superpartners to
exist, we will put their mass safely high enough. In doing this we must
however be careful not to reintroduce the hierarchy problem. If the
masses were too high, we could integrate them out and thus get the
standard model back with all its hierarchy problems. In other words,
the 1-loop correction to the Higgs mass in a model with λi = yi2 is
X (m2Bi − m2F i )
δm2H = m2H − m2H0 ≈ yi2 . (274)
16π 2
i

Naturalness constraint (no fine-tuning) requires that this square mass


difference should not exceed too much the Higgs mass itself (the ex-
pected one), i.e. (100GeV)2 , from which it follows that

yi2 m2Bi − m2F i . (1TeV)2 . (275)


If we now take the plausible assumption (at least in nature we
expect this to hold) that mBi >> mF i , it follows that

1TeV
mBi . (276)
yi
(but bigger than the experimental lower limit which is approximately
100 GeV). The biggest Yukawa is the top one with yt ≈ 1. This means
that the most constrained is the stop t̃. Also, since we do not want the
supersymmetry breaking terms to break the weak SU(2), the sbottom

52
b̃, superpartner of the lefthanded bottom quark, must also have the
same mass as its SU(2) partner t̃. So this gives approximately

mt̃,b̃ . 1TeV , (277)


while for the other sfermions the constraints are milder (i.e. they can
be heavier).
One comment at the end. Although all these requirements for no
fine-tuning are amusing and bring into the game some constraints to
the otherwise almost completely free parameters of the theory, one
should not take them too seriously. After all, who can say which
exactly is an admissible fine-tuning and which not? Is a 40% correction
to the Higgs mass ok? Is a 7 times bigger tree order mass compared to
the 1-loop mass a sign of fine-tuning? Nobody can give a firm answer,
so one should have a common sense in judging these issues. Otherwise
there is a danger to end in a paradoxical situation.

13 Spontaneous supersymmetry break-


ing
As we can see, the bad side of the MSSM, even assuming R-parity) is
a big proliferation of unknown parameters, something like 100 or so.
The desire to simplify things adding a new symmetry (supersymmetry)
clashed with the facts of nature, forcing us to add the new terms (232)-
(236). Theoretically it would be thus much better to have a model of
calculating or deriving the unknown coefficients. This will be studied
in detail in the remaining part of this section. We will present two
different scenarios.

13.1 MSSM is not enough


The best option would be to break supersymmetry spontaneously, as
we do for example in gauge theories. We will see that this is not easy
to do in two examples.
In supersymmetric theories the hamiltonian is proportional to the
sum of the squares of the generators of supersymmetry algebra. So the
energy is zero iff susy is preserved, and positive if susy is spontaneously
broken. We saw in the previous lecture that the potential (energy) is
made from two terms, the F-term (198) and the D-term (199). We
will consider here only the so called F-term breaking:

53
∂W
6= 0 for at least some i . (278)
∂φi
Notice that such a requirement automatically implies a massless
fermion. In fact, from the minimization of (198), the second derivative
of the superpotential in the field space direction defined by (278) must
vanish. This is expected from the Goldstone theorem. In the case of
spontaneously broken internal symmetry a massless Goldstone boson
appears. Here the spontaneously broken supersymmetry generator is
a Grassmann object, thus a fermionic zero mode follows. This object
is called the goldstino.
Although W is a gauge singlet, in the SM φi are not. Thus the
only field that could break supersymmetry without breaking some
unwanted extra gauge symmetry is one of the two Higgses Hu or Hd .
Imagine we do it with Hd . Then suppose that we are able to properly
change the superpotential so that
 
∗ ∂W
FHd ≡ 6= 0 . (279)
∂Hd
To see why this cannot work, let us concentrate on the mass of a
down squark or charged slepton f˜:

W = yf f h0d f c + ... (280)


From (198) and (279) the mass terms for the sfermion in the po-
tential are

yf2 vd2 yf FH∗ d f˜


  
V = f˜∗ f˜c

, (281)
yf FHd yf2 vd2 f˜c∗

where vd is the vev of Hd . The eigenvalues satisfy the mass sum rule

m2f˜ + m2f˜c − 2m2f = 0 . (282)


This is in contradiction with what we know from low energy physics:
there is for example no scalar with quantum numbers of the electron
and a smaller (or equal) mass.
It is possible to show that any spontaneously broken global su-
persymmetric model has similar unacceptable mass sum rules at tree
order. There are two possible ways out: either one goes local, i.e. to
supergravity, or one transmits the information of susy breaking not at

54
tree order, but at one loop. In both cases the mass sum rules change
and unwanted constraints get relaxed.
Either way, the mechanism should roughly look as follows: a sec-
tor with no interaction with the SM fields (and thus called the hidden
sector) is responsible for the spontaneous breaking of supersymme-
try. The information that susy is broken in this hidden sector gets
thus transmitted to our SM sector either by 1/MP l suppressed higher
dimensional terms (as in supergravity) or through an intermediate
(messenger) field, that couples to both SM and hidden sector fields.
In this scenario loops with external SM fields and internal messen-
ger fields (with susy breaking couplings and/or masses) transmit the
information on susy breaking to our sector.
Let us see now in more details the above mechanisms of susy break-
ing mediation.

13.2 Gravity mediation


In the case of local supersymmetry (supergravity), one needs to intro-
duce the gravity multiplet, which essentially means the spin 2 graviton
(2 d.o.f. on shell) plus the spin 3/2 gravitino (also 2 d.o.f. on shell).
Analogous to the case of spontaneously breaking of a local symmetry,
where the would-be goldstone boson gets eaten by the longitudinal
component of the vector boson, here the gravitino eats the goldstino,
acquiring a nonzero mass, m3/2 . This mass turns out to be the typical
scale, and soft masses will be proportional to it.
The potential in the supergravity case becomes (ϕi denote both
SM fields φi and hidden sector fields X)

" ! #
∂W ∗ ∂K W ∗|W |2

K/M∗2 ∂W ∂K W −1 i

VF = e + K + − 3 ,
∂ϕi ∂ϕi M∗2 j ∂ϕ∗j ∂ϕ∗j M∗2
M∗2
√ (283)
where M∗ = MP l / 8π ≈ 2 × 1018 GeV is the so called reduced Planck
i
mass and K −1 j is the inverse matrix of ∂ 2 K/∂ϕi ∂ϕ∗j . The Kähler
potential K(ϕ, ϕ∗ ) is often assumed canonical, Kcan = i ϕ∗i ϕi , but
P
since supergravity is anyway nonrenormalizable, it can be actually
modified with higher dimensional polinomials.
Assuming that it is a singlet field X that breaks susy, the order
parameter in supergravity is defined as

55
∂W ∂K W
FX∗ = + . (284)
∂X ∂X M∗2
It must be nonzero, so to break supersymmetry. In flat spacetime
one can parametrize this breaking by the gravitino mass:

FX = 3m3/2 M∗ . (285)
As we said before, this field X should have small enough couplings
with the SM fields φi . This is most easily obtained, assuming

W (X, φ) = W (X) + WSM (φ) . (286)


Then all the couplings between the two sectors are through 1/MP l
suppressed operators.
One then needs the following requirements at X = hXi:

W = m3/2 M∗2 (zero cosmological constant) , (287)


∂V
=0 (extremum of the potential) , (288)
∂X
∂2V ∂2V
≤ (no tachyons) . (289)
∂X 2 ∂X∂X ∗

As an example take the so-called Polonyi superpotential:

W (X) = aX + b (290)
and a canonical Kähler. With properly chosen constants a and b, so
to satisfy (284), (287)-(289), it is possible to determine all the soft
parameters in (232)-(236) in terms of few parameters at the Planck
scale, which are at the moment still compatible with any experimental
constraint.
We were thus able to reduce the 100 and so parameters to very few
of them. It has to be kept in mind however, that this is no more than
just a reparamentrization of the original soft terms. In fact, there
is absolutely no real reason to believe that the superpotential should
look like (290) and even less that the Kähler potential is canonical. In
fact, taking for example

X ∗ Xφ∗i φj
K = X ∗ X + φ∗i φi + cij , (291)
MP2 l

56
gives the sfermion soft mass squares
  |FX |2
m2f˜ = (δij + cij ) , (292)
ij MP2 l
which not only introduces new couplings (c’s), but may very easily be
in contradiction with experiment due to possible large contribution to
the flavour changing neutral currents. Although there are some ways
of making these c’s small in the infrared via running, this requires
extra physics, and is certainly not a minimalist’s approach. It is safe
to conclude that supergravity cannot explain the structure, less the
particular values of the soft parameters.

13.3 Gauge mediation


The problem with the previous example was, that gravity is not actu-
ally flavour blind (the masses of fermions of different generations are
very different!), which can be parametrized by a nontrivial matrix c
above. This led people to use as mediators of susy breaking gauge
interactions, which are known to be flavour blind
As before, we have to forbid that supersymmetry breaking is trans-
mitted to the SM particles at tree order, so no SM field can satisfy
(278). Thus imagine that we have a gauge singlet field X with a non
zero F-term
 
∗ ∂W
FX = 6= 0 . (293)
∂X
Can it be coupled to any of the standard model fields? Allowing
only renormalizable couplings the only possibility is to transform the
dimension 2 term (222) to a dimension 3 term [19]

Wµ = λX XHu Hd . (294)
The higgsino mass appears from the vev of X,

µ = λX hXi . (295)
The sum rule (282) applied to the Higgs supermultiplets is not
dangerous in this case, since no mass of this multiplet has been mea-
sured yet. On the other side, the sfermions do not couple directly to
X, so they are not influenced by it at tree order and no mass sum rule
thus applies. Unfortunately the one loop correction to the masses of

57
the sfermions gets negative and proportional to the relevant Yukawas
[20]. This would destabilize the stop, breaking SU(3).
In other words, one cannot couple only the Higgs to the X field,
but must use another pair of multiplets, call them Φ and Φ̄:

W = W (X) + λX X Φ̄Φ + WSM . (296)


These new multiplets are not gauge singlets, so they can inter-
change gauge bosons (and gauginos) with the SM fields. Thus at
leading order in a small FX /MΦ2 ratio, where MΦ = λX hXi is the
susy preserving Φ mass, the gaugino mass gets a 1-loop contribution

αi λX FX
mλ = ci , (297)
4π MΦ
while the sfermions get a two-loop contribution
 α 2 |λ F |2
i X X
m2f˜ = di , (298)
4π MΦ2
with ci , di (i = 1, 2, 3) depending on the representation under the
SM gauge group SU(3)×SU(2)×U(1). Similar relations are possible
also for the other soft parameters. Although this seems to solve the
flavour problem mentioned above, one must be careful. In fact, there
is no gaurantee that the Φ and Φ̄ do not couple or mix with the SM
fields. Typically these mediators have the quantum numbers of the
sfermions. This mixing can in principle spoil the flavour blindness of
the mediation [20].
A very useful way of getting the soft terms in gauge mediation
is to solve the renormalization group equations in the presence of
the messengers of supersymmetry breaking. Then one promotes the
messenger mass M = λX X to a chiral superfield, with X = hXi +
θθFX . Finally, using the formulae (F = λX FX )

1 ∂ log αa−1 (X) F


mλa = − (299)
2 ∂ log X X=M M
∂ log Zf (X, X † ) FF†
m2f˜ = − (300)
∂ log X∂ log X † X=M M M †
∂ F
Aijk = log (Zi (X)Zj (X)Zk (X)) (301)
∂ log X X=M M
valid at the messenger scale M . For details see [16].

58
14 Exercises
Solve (at least !) one exercise from group 1 and one from group 2.
Group 1
1. Derive (36) when applied on ψα and F . In doing that use the
Fierz identity (valid for arbitrary spinors ψ, ξ, η)

ψα (ξη) + ξα (ηψ) + ηα (ψξ) = 0 (302)

which you derive from the well known Jacobi identity

αβ γδ + αγ δβ + αδ βγ = 0 . (303)

This in practice means that

(ψ1 ψ2 ) (ψ3 ψ4 ) + (ψ1 ψ3 ) (ψ2 ψ4 ) + (ψ1 ψ4 ) (ψ2 ψ3 ) = 0 (304)

Notice that only pairs are important, never the order (ηχ = χη).
Then derive eq. (52) using the known relations for the γ matrices
and eq. (79). Finally show explicitly that (98) can bring (97)
into the form (99).
2. Write in components (in terms of φ and ψ only) the most gen-
eral renormalizable multifield Wess-Zumino model, i.e. general-
ize (145). Write in components the one chiral superfield Wess-
Zumino model for the most general form of K(Φ, Φ† ) = Φ† Φ+
higher terms, and W (Φ). Hint: expand K and W around Φ = φ.
3. Consider the simplest possible supersymmetric theory with U(1)
gauge invariance. Such a theory is called supersymmetric QED.
Argue why such a theory cannot be consistent with one single
chiral superfield. Write down the simplest consistent and renor-
malizable quantum field theory. Find the minima of the poten-
tial and compute the spectrum of the theory as a function of the
model parameters, ξ included.
4. Consider a SU(N) gauge model with the superpotential

W = yΦΣΦ (305)

with Φ, Σ and Φ the antifundamental, adjoint and fundamental


representations of SU(N). Compute the unknown parameters in
the 1-loop RGE equations and solve them analytically.

59
Group 2
1. Derive all the components of the mass matrices MR2 , MI2 , M±2 ,
MC̃ , MÑ , Mũ2 , Md2˜, Mẽ2 , Mν̃2 , Mg̃ from the MSSM Lagrangian.
Show explicitly that (a) MR2 has non-negative eigenvalues, (b) the
zero eigenvalues predicted by the Nambu-Goldstone theorem and
(c) that there are only two different eigenvalues in the chargino
mass matrix.
2. Calculate explicilty the sfermion soft masses and A terms in
the Polonyi supergravity model with canonical Kähler potential.
How could gaugini get a mass term (i.e. which operator has for
its θθθθ component the gaugino mass)?
3. Calculate the soft terms of the MSSM assuming as messengers
Φ (Φ) to have the same quantum numbers as the quark Q (con-
jugate of Q). Suppose the scale M is at 106 GeV. What is the
minimal value of F allowed by data? Is unification of couplings
still obtained?

References
[1] H. E. Haber and G. L. Kane, Phys. Rept. 117 (1985) 75.
[2] H. P. Nilles, Phys. Rept. 110 (1984) 1.
[3] J. Wess and J. Bagger, “Supersymmetry And Supergravity”,
Princeton University Press (1992).
[4] S. P. Martin, arXiv:hep-ph/9709356.
[5] J. P. Derendinger, “Lecture Notes On Globally Supersymmet-
ric Theories In Four-Dimensions And Two-Dimensions,” in Pro-
ceedings of the Hellenic School of Particle Physics, Corfu, Greece,
September 1989, edited by G. Zoupanos and N. Tracas; also avail-
able at:
http://www.unine.ch/phys/hepth/Derend/SUSY nd.pdf.
[6] M. Dine, arXiv:hep-ph/9612389.
[7] M. F. Sohnius, Phys. Rept. 128 (1985) 39.
[8] J. D. Lykken, arXiv:hep-th/ 9612114.
[9] G. F. Giudice and R. Rattazzi, Phys. Rept. 322 (1999) 419
[arXiv:hep-ph/9801271].

60
[10] D. J. H. Chung, L. L. Everett, G. L. Kane, S. F. King,
J. D. Lykken and L. T. Wang, Phys. Rept. 407 (2005) 1
[arXiv:hep-ph/0312378].
[11] Y. Shadmi, arXiv:hep-th/0601076.
[12] M. Dine, arXiv:0901.1713 [hep-ph].
[13] S. Weinberg, “The quantum theory of fields. Vol. 3: Supersym-
metry,” Cambridge, UK: Univ. Pr. (2000) 419 p.
[14] S. J. Gates, M. T. Grisaru, M. Rocek and W. Siegel, Front. Phys.
58 (1983) 1 [arXiv:hep-th/0108200] (only pdf-no ps).
[15] H. K. Dreiner, H. E. Haber and S. P. Martin, arXiv:0812.1594
[hep-ph].
[16] G. F. Giudice and R. Rattazzi, Nucl. Phys. B 511 (1998) 25
[arXiv:hep-ph/9706540].
[17] M. Carena and H. E. Haber, Prog. Part. Nucl. Phys. 50 (2003)
63 [arXiv:hep-ph/0208209].
[18] T. Asaka, M. Shaposhnikov and A. Kusenko, Phys. Lett. B 638
(2006) 401 [arXiv:hep-ph/0602150].
[19] G. R. Dvali and M. A. Shifman, Phys. Lett. B 399 (1997) 60
[arXiv:hep-ph/9612490].
[20] M. Dine, Y. Nir and Y. Shirman, Phys. Rev. D 55 (1997) 1501
[arXiv:hep-ph/9607397].

61

You might also like