Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Supergravity Application in Particle Physics

Download as pdf or txt
Download as pdf or txt
You are on page 1of 46

Supergravity: Application in Particle Physics

Florian Domingo and Michel Rausch de Traubenberg ∗

Abstract We provide a pedagogical introduction to N = 1 supergravity/supersym-


metry in relation to particle physics. The various steps in the construction of a
generic N = 1 supergravity model are briefly described, and we focus on its low
arXiv:2209.12541v1 [hep-th] 26 Sep 2022

energy supersymmetric limit. The conditions for supersymmetry and supergrav-


ity breaking are investigated, and realistic mechanisms suitable for particle physics
identified. We then study the model-building aspects of ‘softly-broken’ supersym-
metric extensions of the Standard Model and discuss several of their phenomeno-
logical features.

Invited chapter for Handbook of Quantum Gravity (Eds. C. Bambi, L. Modesto


and I.L. Shapiro, Springer Singapore, expected in 2023).

Keywords
N = 1 Supergravity/Supersymmetry, Soft-Supersymmetry Breaking, Gravity-Mediation,
Goldstino/Gravitino, MSSM, R-Parity, Electroweak Symmetry, Flavour Problem,
Dark Matter, Collider Phenomenology, Grand Unification.

1 Introduction: Symmetries in Particle Physics

Symmetry principles emerge as a fundamental tool in the description of the laws of


physics. Since particle physics is usually understood in terms of a Quantum Field
Theory (QFT), these principles enable to classify elementary particles and constrain
the form of possible interactions. For instance, elementary particles organise accord-
ing to irreducible representations of the Poincaré group and can be labelled by the
eigenvalues of its two Casimir operators, namely their mass and their spin. Gauge
interactions are associated to a compact Lie group. Spacetime and internal symme-
tries allow the construction of multiple particle physics models; in particular, the
so-called Standard Model of particle physics, which seems in capacity to describe

Florian Domingo, Bethe Center for Theoretical Physics & Physikalisches Institut der Universität
Bonn, Nußallee 12, D-53115 Bonn, Germany, e-mail: domingo@physik.uni-bonn.de ·
Michel Rausch de Traubenberg, Université de Strasbourg, CNRS, IPHC UMR7178, 23 rue du
Loess, Strasbourg, 67037 Cedex, France e-mail: michel.rausch@iphc.cnrs.fr
∗ corresponding author

1
2 Florian Domingo and Michel Rausch de Traubenberg

almost all the currently available high-energy physics data collected in collider ex-
periments.
Since the properties of elementary particles are associated to the underlying sym-
metry group one may wonder what type of symmetries are compatible with the basic
underlying principles. In this respect QFT places severe constraints on the possible
structures. Indeed, if we assume that acceptable symmetries are encoded by Lie al-
gebras, Coleman and Mandula [23] established that the most general form that the
corresponding Lie algebras can take is 2 :

g = Iso(1, 3) × gc

where Iso(1, 3) is the Poincaré algebra and gc a compact Lie algebra. All the gen-
erators of gc commute with those of Iso(1, 3). In other words, all the associated
symmetries, in particular those encoding the gauge interactions, are ‘neutral’ with
respect to spacetime. A fundamental assumption behind the construction of g is that
it contains only generators of a bosonic nature. However, due to the spin-statistics
and Nœther theorems, it is in fact possible to obtain conserved charges of fermionic
nature, closing their own algebra though anticommutating relations and thus leading
to new algebraic structures called Lie superalgebras. Within this latter framework it
is possible to extend the family of Poincaré-compatible symmetries in a non-trivial
way [89, 81], and only supersymmetric extensions 3 of the Poincaré algebra are
then admissible in QFT. The simplest (non-trivial) N = 1 supersymmetric exten-
sion of spacetime symmetries is generated by the generators of the Poincaré algebra
(Lµν = −Lν µ , Pµ , µ , ν = 0, · · · , 3) and one Majorana spinor (Qα , Qα̇ , α , α̇ = 1, 2) 4 .
The non-trivial part of the algebra involving the fermionic charges is
µ
{Qα , Qα̇ } = −2σα α̇ Pµ , (1)
µ
where σα α̇ is defined below — see after Eq.(8). Thus supersymmetry is a non-trivial
extension of the Poincaré algebra.
This discovery subsequently led to the introduction of supersymmetry in particle
physics. Since Qα is of fermionic nature, supersymmetry is a symmetry that maps
a boson into a fermion and vice versa and, as can be proved, supersymmetric mul-
tiplets contain an equal number of bosonic and fermionic degrees of freedom. One
may also promote the invariance under supersymmetry to a local version: since the
anticommutator of two local supersymmetric transformations amounts to a space-
time translation (see Eq.[1]), local supersymmetry in particular contains gravity, and
is thus called supergravity.

2 If we suppose only massless particles a larger symmetry group can be obtained namely the

conformal group.
3 If we suppose only massless particles a larger symmetry structure can be obtained namely the

superconformal algebra.
4 Following the standard conventions, indices for left-handed spinors are denoted by α , β , · · ·,

indices for right-handed spinors are taken to be α̇ , β̇ , · · · and vectors indices are given by µ , ν , · · ·.
Supergravity: Application in Particle Physics 3

The purpose of this chapter is to describe applications of supergravity in particle


physics. To this end, we shall first design a generic theory of fields invariant un-
der N = 1 supergravity. Then, given that exact supergravity is incompatible with the
known particle spectrum, we will study the conditions for supersymmetry and super-
gravity breaking, and present realistic mechanisms suitable for particle physics: we
will more specifically highlight the case of ‘gravity mediation’, where gravitational
effects convey supersymmetry breaking from a ‘hidden’ sector to the observable
one. Finally, we will discuss several phenomenological features of softly-broken
supersymmetric extensions of the Standard Model.

2 How to build an action in supergravity

The first theory of pure supergravity was obtained by Freedman, van Nieuwenhuizen
& Ferrara [70] and Deser & Zumino [31] independently. It took a couple of years to
understand how matter and and Yang-Mills fields could couple to supergravity and
the supergravity coupling of arbitrary interacting matter fields was derived in [24,
25, 26, 27, 7, 8]. The first attempts to build invariant Lagrangians [24, 25, 26, 27]
were performed in the component formalism, or using appropriate tensor calculus
techniques [142, 60, 141, 140, 61]. Then, it was realised that a superspace suited
to supergravity model building could be introduced [151, 153, 152, 85, 1, 160, 6,
16]. The key advantage of this superspace rests with the automatic invariance under
supergravity underlying this formalism, calling all the necessary auxiliary fields and
resulting in a suitable action There exist several books [150, 154, 17, 69, 149, 75,
145, 13, 71, 118, 121, 69, 147, 73] and reviews [122, 123, 42, 14] on the topic.
The purpose of the current section is to provide the reader with the salient points
which lead from general principles to the four-dimensional N = 1 supergravity La-
grangian. The construction closely follows the book of Wess & Bagger [150] and
is based on the recent book and review by one of the authors [147, 45]. In partic-
ular these two last references are complementary since [147] offers a pedagogical
introduction to supergravity, providing many technical details, whilst technical de-
tails are omitted in [45], with a focus on the conceptual scheme that leads to the
four-dimensional N = 1 supergravity Lagrangian.
Four steps can be identified in the derivation of the supergravity-invariant La-
grangian. The first one is purely geometrical and consists in extending the notion of
superspace, already considered in supersymmetry model building [133, 134, 62],
but now in the context of supergravity, i.e., constructing a curved superspace
[151, 153, 152, 85, 17]. This curved superspace is such that, at each superspace
point, a tangent space exists, behaving like the traditional flat superspace of global
supersymmetry. The corresponding structure group is thus the Lorentz group. In par-
ticular, in the Einstein frame one has an invariance under the superdiffeomorphisms
whereas in the tangent or Lorentz frame one has local Lorentz invariance.
Following the standard techniques of general relativity two dynamical variables
(which are indeed two superfields) are introduced. The first dynamical superfield is
4 Florian Domingo and Michel Rausch de Traubenberg

the supervierbein, which connects components in the Einstein frame to components


in the Lorentz (or flat) frame. The second dynamical variable is the superconnec-
tion, which enables to define covariant derivatives. Then, one associates (after com-
puting the (anti)commutators of covariant derivatives) two superfields to the two
dynamical variables: the torsion and curvature tensors. Some constraints are im-
posed to the torsion tensor in order to dispose of overabundant degrees of freedom
[85, 153, 151, 150]. These constraints are of physical importance since, in addition
to drastically reducing the number of degrees of freedom, they allow to explicitly
construct physical models in particle physics that are invariant under supergravity
and supersymmetric transformations at low energy.
Requiring torsion constraints, the Bianchi identities leads to thirteen equations.
Solving these equations allows to express all torsion and curvature tensors in terms
of three basic superfields [86, 135, 150, 41]: a chiral superfield R, a chiral sym-
metric spinor superfield W(αβ γ ) and a real vector superfield Gµ . It was observed
in [96] that the structure of constraints upon the torsion tensors is invariant under
superconformal or Howe-Tucker transformations. Superconformal transformations
are the supergravity analogue of Weyl transformations, or rescaling of the metric, in
general relativity and play a central rôle in the construction of a correctly normalised
action in particles physics.
Having expressed all torsion and curvature tensors in terms of the three basic
superfields, and using the large symmetry due to the supergravity algebra, many
components can be set to zero by means of an appropriate choice of the parameters
[151, 150, 142]. In particular, for the supervierbein and the superfields R and Gµ ,
we have:
 µ 
eµ̃ (x) 12 ψµ̃ α (x) 12 ψ̄µ̃ α̇ (x) R(z)

= − 16 M(x) ,
M  0 δα̃ α 0 
EM̃ (z) = ,
0 0

α
δ α̇ G µ (z) = − 13 bµ (x) ,

where z ≡ (x, θ , θ̄ ) corresponds to a point in a superspace and X represents the
lowest order component of the superfield X in its expansion in terms of the Grass-
mann variables θ , θ̄ 5 (i.e., the fermionic counterpart of the spacetime position x in
the superspace). The fields of the supergravity multiplet are then the helicity-two
graviton eµ̃ µ , the helicity-3/2 gravitino (ψµ̃ α (x), ψ̄µ̃ α̇ (x)) (which is a Majorana
spinor-vector) and two auxiliary fields, M a complex scalar and b µ a real vector.
The last field of the gravity multiplet, the connection is a composite field and can be
expressed in terms of the graviton and the gravitino [31, 70].
If we consider a general transformation, combining an arbitrary superdiffeomor-
phism and a local Lorentz transformation, there is no reason why the gauge fixing
conditions above should be preserved. This means that the set of all possible trans-
formations must be restricted to the subset protecting the previous gauge condition
[151, 152, 150]. This restricted set of transformations is called supergravity trans-
5 Lorentz indices in flat space are taken to be untilded M = ( µ , α , α̇ ), whilst Einstein indices in
ė ). The three entries are related to the vector, the
curved space are taken to be tilded M̃ = (µ̃ , α̃ , α
left-handed spinor and the right-handed spinor counterpart of a point in the superspace.
Supergravity: Application in Particle Physics 5

formations. On the one hand a supergravity transformation is parameterised by a


general transformation in the fermionic direction of the superspace, corresponding
to a local supersymmetric transformation. On the other hand the transformation in
the bosonic direction of the superspace and the local Lorentz transformation are re-
lated to the transformation in the fermionic direction (for the Lorentz transformation
it also involves the auxiliary fields M and bµ ).
Using firstly the constraints on the torsion tensors and secondly the explicit re-
lationship between the torsion or curvature tensors with the supervierbein and the
superconnection (expressed in the gauge (2)) allows to compute the lowest com-
ponents of the basic superfields R, W(αβ γ ) and Gµ together with the lowest com-
ponents of their derivatives with respect to the covariant derivatives [150]. It turns
out that all these lowest components, or the lowest components of derivatives, are
expressed in terms of the supergravity multiplet eµ̃ µ , ψµ̃ α , M, b µ .
Having fixed the gauge, defined what is called supergravity transformations, and
obtained all physical quantities in terms of the supergravity multiplet, one should
then compute the transformation of the supergravity multiplet under supergravity
transformations [152, 150, 142, 60, 65].
The second step in the construction consists in introducing superfields in the
curved superspace. As in supersymmetry, two types of superfields will be consid-
ered in applications of supergravity in particle physics: chiral superfields associ-
ated to ordinary matter and vector superfields associated to Yang-Mills theory. It
is remarkable that the supersymmetry concepts of chiral and vector superfields ex-
tend to the supergravity context. This is in fact a consequence of the torsion con-
straints, which amount to an integrability condition. Chiral and vector multiplets
were independently introduced in supergravity within the tensor calculus approach
[141, 140, 61], the superspace approach [152] or when constructing the first invari-
ant Lagrangians [64, 63, 65, 25, 26, 24, 27].
Superfields live in the curved superspace but we need to compute their covariant
derivatives with respect to the variables with Lorentz indices. For instance introduc-
ing Dα , D α̇ (pay attention to the fact that these are indices in the Lorentz frame and
not indices in the Einstein frame) the covariant derivatives with respect to left and
right-handed spinors, a chiral superfield is defined by the chirality condition 6

D α̇ Φ = 0 (2)


√1 Dα Φ , F = 1 D · D Φ .
ϕ = Φ , χα = 2 4 (3)

where the second line explicitly provides the components in this gauge. They corre-
spond to a scalar field ϕ , a left-handed spinor χ and an auxiliary field F. A vector
superfield is defined by the reality condition V † = V . As in supersymmetry it is
possible to select the Wess-Zumino gauge defined by


V = 0 , Dα V = 0 , D α̇ V = 0 , D · D Φ = 0 , D · D Φ = 0 . (4)

6For the spinor scalar products, we take the usual conventions namely χ · χ = χ α χα for left-handed
spinors and χ̄ · χ̄ = χ̄α̇ χ̄ α̇ for right handed spinors.
6 Florian Domingo and Michel Rausch de Traubenberg

For both chiral or vector superfields, in order to obtain the full supergravity action,
we have to compute several (lowest order components of) derivatives. This com-
putation is performed using the algebra of supergravity (the (anti)commutators of
covariant derivatives expressed in terms of the superfields R,W(αβ γ ) and Gµ ).
At this point, we would like to stress on a fundamental property of chiral su-
perfields. If Φ † is an anti-chiral superfield then it can be shown that [152, 74]
D α̇ D · D − 8R Φ † = 0 or equivalently the superfield

Ξ = D · D − 8R Φ † , (5)

is chiral. This property is central in obtaining a compact chiral expression of the


supergravity action.
Vector superfields can be introduced to describe any Yang-Mills interactions.
Consider a compact Lie algebra gc of dimension n, a unitary representation R and
denote the generators of gc in the representation R by Ta = Ta† , a = 1, · · · , n. If
fab c ∈ R are the structure constants of gc , the Lie brackets take the form:
 
Ta , Tb = i fab c Tc .

Associated to the vector superfield V = V a Ta we define the spinor superfield strength


tensor
1 
Wα = − D · D − 8R e2gV Dα e−2gV ,
4
where g is the coupling constant. Observe that Wα is a chiral spinor superfield be-
cause of the ‘projection’ operator D · D − 8R.
Again the algebra
of supergravity

enables to compute (using (4)) Wα , Dβ Wα and D · DWα . It turns out that the
degrees of freedom of a vector superfield are vµ , (λα , λ̄α̇ ) and D, i.e., a vector, a
Majorana spinor and a real scalar.
We can now assume that the chiral superfield Φ is in the representation R of gc ,
i.e., transforms like Φ → e−2giΛ Φ , where Λ = Λ a Ta and Λ a are chiral superfields.
Of course the conjugate of a chiral superfield, which is an anti-chiral superfield and
whose fermionic component contains a right-handed fermion, is in the representa-
tion R, the complex conjugate representation of R, with generators −Ta⋆ = −Tat .
It is possible to couple the chiral superfield Φ to the vector  superfield [150, 147]
in an invariant way by considering X = D · D − 8R Φ † e−2gV (the analogue of
(5)) which is obviously chiral, and is again obtained by computing the lowest order
components of its derivatives up to the order two. This computation is performed us-
ing the algebra of supergravity, but in the so-called Wess-Zumino gauge for which
V 3 = 0.
Up to now we have defined the gravity multiplet (eµ̃ µ , ψµ̃ α , ψ̄µ̃ α̇ , M, b µ ), the
matter multiplets (ϕ i , χ i , F i ), i = 1, · · · , dim R and the Yang-Mills multiplets (vaµ , λαa ,
λ̄α̇a , Da ), a = 1, · · · , dim gc : these can finally be combined to construct a supergravity
action. As seen in (3), a chiral superfied is defined in the curved superspace, but the
Supergravity: Application in Particle Physics 7

constraint (2) and its components (3) involve covariant derivatives with respect to
Lorentz indices. This, of course, will make the computation cumbersome. The idea
of Julius Wess, which can be seen as a tour de force [84], consists in introducing
new Θ −variables carrying a Lorentz index instead of an Einstein index, in such a
way that the expansion of the chiral superfield Φ reduces to [150, 153]
√ 1 √

Φ (x, Θ ) = Φ + 2Θ · (D Φ ) − Θ · Θ (D · D Φ ) = ϕ + 2Θ · χ − Θ ·Θ F .
4
Next, in general relativity the invariant measure of integration involves the de-
terminant of the vierbein. Similarly, in curved superspace the invariant measure of
integration involves the superdeterminant of the supervierbein. However, in the con-
text of a chiral form of the Lagrangian (see below) an adapted invariant measure,
involving the capacity E has to be considered [122, 6, 119].
We now have all the necessary material to compute the supergravity action, ex-
ploiting the notations introduced above. In the fourth and final step we set gc to the
specific algebra corresponding to the desired compact Lie group which determines
the Yang-Mills interactions. The Lie algebra gc can be semisimple (including the
product of simple Lie algebras) or can contain some abelian factors or u(1)−terms.
One introduces dim gc vector superfields associated to gc . Next, selecting a matter
content in a unitary representation R of gc (not necessarily irreducible), we associate
dim R chiral fields in the representation R and dim R anti-chiral fields in the repre-
sentation R. We collectively denote Φ i and Φi†∗ as chiral or anti-chiral superfields
respectively. Further, we introduce three gauge invariant functions: (1) the superpo-
tential W (Φ ), a holomorphic function depending on chiral superfields; (2) the gauge
kinetic function hab (Φ ) (where a, b are gauge indices in the adjoint representation
of gc ), a holomorphic function depending on chiral superfields; and (3) the Kähler
potential K(Φ , Φ † ), a real function depending on Φ and Φ † . The Lagrangian (in
chiral form) takes the form [77, 150]
Z
(
2
3m2p − 1 2 K(Φ ,Φ † e−2gV )
LSUGRA = d Θ E (D ·D − 8R)e 3m p
8
)
1 aα b
+ W (Φ ) + hab (Φ )W Wα + h.c. , (6)
16g2

where m p is the Planck mass. To make it more explicit, we have to expand all fields
in the Θ −variables. Since all superfields have a large number of components, this
computation is lengthy, but not fundamentally complicated. Once the expansion has
been performed, we have to eliminate the auxiliary fields F i , Da , M, b µ through their
equations of motion, which is easy as well. At the end of the computation, it turns
out that the Lagrangian is not canonically normalised. In order to have a canonically
normalised Lagrangian we must perform a dilatation of the vierbein, followed by
a gravitino shift [150], which amount to a redefinition of all the fields and can be
interpreted as a superconformal (or Howe-Tucker) transformation. This is certainly
the most difficult step.
8 Florian Domingo and Michel Rausch de Traubenberg

Due to the technical difficulties associated with the field redefinition, alterna-
tive methods have been proposed in order to compute the final Lagrangian. The
first method [14] is geometrical in nature, enlarging the usual superspace to a
U(1)−superspace. The second approach is based on superconformal methods. In
Poincaré supergravity, the structure group is the Poincaré supergroup. The confor-
mal methods [66, 107, 106, 146, 67, 68] are based on an enlargement of the structure
group to the superconformal group. The book of Freedman & van Proeyen [69] is
devoted to conformal methods in supergravity.
The final Lagrangian contains the Einstein-Hilbert action for general relativity,
the Rarita-Schwinger Lagrangian for the spin 3/2−gravitino, the kinetic terms of
the matter sector (spin 0, 1/2), the kinetic part of the Yang-Mills sector (spin 1, 1/2).
All these fields are naturally coupled and invariant under supergravity transforma-
tions, but also under many other transformations, such as symmetries of the under-
lying Kähler manifold (the complex manifold where chiral fields are living), etc.
Many interacting terms are also generated, such as e.g. four-fermions interactions.
∗ ∗
The scalar potential takes the form (we denote Wi = ∂iW, Ki = ∂i K etc., K i i = ∂ i ∂i K
the Kähler metric of the Kähler manifold and K i i∗ the inverse Kähler metric) 7 :
1
2 K(ϕ ,ϕ † )  ∗ 3 
V = e mp DiW D i W ⋆ K i i∗ − |W | 2
,
m2p

i∗
where DiW = Wi + 1/m2pKiW , with a similar expression for D W ⋆ are covariant
derivatives with respect to the Kähler manifold structure. It is remarkable that the
structure of the final Lagrangian emerges from strictly geometrical and algebraic
principles that are related to the properties of the curved superspace and its associ-
ated supergravity transformations.
If we now study the low energy limit of (6), namely when m p → +∞, the part of
the Lagrangian involving the Kähler potential becomes:
Z Z
1 1
L = −3m2p d 2Θ E R − d2Θ E (D ·D − 8R)K(Φ , Φ † e−2gV ) + h.c. + O( ).
8 m2p

The first piece reduces to [150]


1 1   1 
Lpure 2
SUGRA = e m p R + ε µνρσ ψµ σσ ψ̄νρ − ψ̄µ σ̄σ ψνρ − MM ⋆ + b µ bµ
2 4 3
where R is the scalar curvature, e is the determinant of the vierbein, ψµν is the field
strength of the gravitino and ε µνρσ is the Levi-Civita symbol. This is the Lagrangian
obtained in [70, 31] including the auxiliary fields [142, 60] and describing pure
supergravity.
In the limit m p → +∞, the projector (D ·D − 8R) becomes D · D with D (and D)
the usual covariant derivatives in supersymmetry, Wα becomes Wα the spinor field
strength in supersymmetry [150, 73] and Θ reduce to θ the left-handed Grassmann

7 In (7) we have omitted the usual D−terms (see Sec. 3).


Supergravity: Application in Particle Physics 9

variables considered in supersymmetry. If we now restrict ourselves to a renormal-


isable theory the basic functions reduce to:
1 1
K(Φ , Φ † ) = Φi† Φ i , hab (Φ ) = δabW (Φ ) = αi Φ i + mi j Φ i Φ j + λi jk Φ i Φ j Φ k .
2 6
Therefore, omitting Planck-suppressed terms, we obtain in the limit m p → +∞
Z
( )
1   1
LSUSY = d2 θ − D · D Φ † e−2gV Φ + W + δab W aα Wαb + h.c. (7)
8 16g2

which is the standard action in supersymmetry expressed in a chiral form. After


expanding the fields, we obtain
1 i 1
LSUSY = − δab F aµν Fµν
b
− δab (λ a σ µ Dµ λ̄ b − Dµ λ a σ µ λ̄a ) + δab Da Db
4 2 2
† µ i µ µ
 †
+ Dµ ϕ D ϕ − χσ Dµ χ̄ − Dµ χσ χ̄ + F F (8)
2√ √
− gDa ϕ † Ta ϕ − i 2gλ̄ a · χ̄ Ta ϕ + i 2gϕ † Ta χ · λ a
 1 1 
− mi j (ϕ i F j + χ i · χ j ) − λi jk (ϕ i ϕ j F k + ϕ i χ j · χ k ) + h.c. ,
2 2
a
where Fµν = ∂µ vaν − ∂ν vaµ − g fbc a vbµ vνc , Dµ λ a = ∂µ λ a − g fbc a vbµ λ c are respectively
the field strength tensor and the covariant derivative in the Yang-Mills sector, and
Dµ ϕ i = ∂µ ϕ i + ig(Ta ϕ )i , Dµ χ i = ∂µ χ i +ig(Ta χ )i are the covariant  derivatives in
µ
the matter sector. Finally σα α̇ = σ 0 , σ i and σ̄ µ α̇α = σ 0 , −σ i where σ i , i =
1, 2, 3 are the usual Pauli matrices and σ 0 is the two-by-two identity matrix. For
completeness, we remind our conventions  for the Minkowski metric and Levi-Civita
symbol ηµν = diag 1, −1, −1, −1 and ε0123 = 1. If we eliminate the auxiliary fields
through their equations of motion we get

Fi† = ∂iW ≡ Wi and Da = gϕ † Ta ϕ (9)

and obtain the scalar potential, with its F− and D−terms part (we denote Da =
δab Db , Ta = δab T b , etc.)
1 1 
V = VF + VD = Fi† F i + Da Da = WiW ⋆i + g2 (ϕ † Ta ϕ )(ϕ † T a ϕ ) = αi + mi j ϕ i
2 2
1  1 ∗iℓm † †  1 2 ab †
∗i ∗iℓ †
j k
+ λi jk ϕ ϕ α + m ϕℓ + λ ϕℓ ϕm + g δ (ϕ Ta ϕ )(ϕ † Tb ϕ )
2 2 2
and the Yukawa interactions between spinors and scalar fields
1
LYuk = − Wi j χ i · χ j + h.c. = −λi jk ϕ i χ j · χ k + h.c. (10)
2
10 Florian Domingo and Michel Rausch de Traubenberg

For explicit computations in the supersymmetry context the reader may refer to
the book by one of the author [73]. We further stress that the supergravity action is
derived with all details in [147, 45]. As a complement to this section one may consult
[29] in the same volume, where simple supergravity in components is presented,
[111] which provides a superspace formulation of supergravity, or [110], related to
Supergravity-Matter couplings in projective superspace.

3 Supergravity breaking

In the previous section we have described the basic steps allowing to construct a
Lagrangian coupling matter and Yang-Mills theory to supergravity. Its limit for
mp → ∞, i.e., the supersymmetric theory, has been simultaneously obtained. This
derivation opens the path to applications of the supergravity/supersymmetry formal-
ism to particle physics. At the centre of any such application is the gain in regular-
ity of a QFT protected by supersymmetry, taking the form of non-renormalisation
theorems [87]. Corresponding particle physics models may then remain perturba-
tive over the ‘Grand Desert’ separating the electroweak from the Planck scales,
thus preserving the general structure of the studied model over a huge hierarchy
of scales: this is the type of applications in which we specialise below. However,
this same advantage of regular quantum corrections also finds use in the study of
the general properties of Yang-Mills theories at the quantum level (‘integrability’)
[94, 39, 52]. In particular, implementing supersymmetry/supergravity in particle
physics enables the protection of the mass of scalar fields from quantum correc-
tions: a massless scalar field remains massless at any order of perturbation theory.
We will argue in the following paragraph that supersymmetry needs to be broken
for phenomenological applications; still, this breaking can be enforced in such a
‘soft’ way that this nice property under renormalisation is preserved. In this fashion,
softly broken supersymmetry provides a technical solution to the Hierarchy Prob-
lem [155, 32, 157, 108, 131]. Nevertheless, supersymmetry breaking terms should
remain comparable to the electroweak scale so that supersymmetry actively pro-
tects the Higgs mass from comparatively low scales: observable effects at colliders
are then expected.
Due to the fermionic nature of the generators of supersymmetry/supergravity
transformations, a multiplet contains both bosonic and fermionic degrees of free-
dom, which find a natural embedding in the chiral or the vector superfields of Sec. 2.
In particular, if one wants to construct a supersymmetric version of the Glashow-
Weinberg-Salam model, each standard particle must be associated to a supersym-
metric partner: a scalar to each fermion and a fermion to each boson (gauge and
Higgs). Noticing at this point that the operator Pµ Pµ is a Casimir operator, parti-
cles and their superpartners are necessarily degenerate in mass if supersymmetry
is unbroken. More specifically, this means that if supergravity/supersymmetry were
an exact symmetry of Nature, a massless neutral fermion (associated to the photon)
or a light charged scalar (associated to the electron) should have been observed by
Supergravity: Application in Particle Physics 11

now. Which is obviously not the case. Thus, supersymmetry/supergravity must be


broken at the scale of current particle physics experiments.

3.1 Mechanisms of supersymmetry and supergravity breaking

We first recall the classical mechanisms achieving a (global) breaking of supersym-


metry. One way to induce a spontaneous breaking consists in designing a field con-
figuration such that the equations of motions for the auxiliary fields F or D are in-
compatible with the trivial solution, i.e., at least one of the auxiliary fields develop a
vacuum expectation value (vev). Correspondingly, the scalar potential (10) becomes
strictly positive at the minimum in a spontaneously broken supersymmetric theory.
The first mechanism (O’Raifeartaigh) of supersymmetry breaking achieves F 6= 0
and involves at least three chiral superfields [126]. The second mechanism (Fayet-
Iliopoulos) induces D 6= 0 through a vector superfield associated to a U(1)−gauge
symmetry, allowing for a Fayet-Iliopoulos term in the Lagrangian [59, 54]. Both
mechanisms fail to produce a spectrum compatible with the experimental situation,
as at least some of the ‘exotic’ particles would be lighter than their standard part-
ners. Thus, the Standard Model cannot be straightforwardly embedded in a sponta-
neously broken supersymmetric framework. For some details related to these two
mechanisms see e.g. [73].
However, this picture could play a rôle if a less simplistic structure is intro-
duced: supersymmetry is spontaneously broken in a hidden sector by one of the
two mechanisms described above; then this breaking is transmitted to an observ-
able (or matter) sector, which contains a supersymmetric extension of the Standard
Model, by the interaction connecting both sectors. The nature of this interaction
determines three classes of mediation models. In the first type, known as gravity-
mediated [22, 11, 97, 125, 46, 3, 127], the supersymmetry breaking is commu-
nicated to the observable sector via gravitational interactions. In the second type,
called gauge-mediation, supersymmetry is broken by a singlet field (called the spu-
rion), which couples (via the superpotential) to another field, the messenger, be-
longing to a non-trivial representation of the gauge group. Quantum loop correc-
tions involving the messenger then break supersymmetry in the observable sec-
tor [123, 33, 36, 30, 57, 34, 120, 2, 35, 37, 38, 79, 5]. In the third type, called
anomaly mediation, the introduction of compensating fields having a conformal
(super-Weyl) anomaly induces supersymmetry breaking by purely quantum effects
[129, 5, 9]. Concerning general properties of broken supergravity we refer the reader
to [155, 156, 59, 99]. Here, we will restrict ourselves to the specific case of gravity-
induced supersymmetry breaking in Sec. 3.3.
As seen in (7) the scalar potential in supergravity, hence its minimum, is no
longer positive (as in global supersymmetry). This turns out to be an advantage,
since, supergravity being a theory of gravitation, one expects that the cosmologi-
cal constant, hence the potential at the minimum vanishes. Furthermore, as will be
established in Sec. 3.2 after supergravity breaking the gravitino becomes massive.
12 Florian Domingo and Michel Rausch de Traubenberg

The assumption of an vanishing cosmological constant, though phenomenologically


motivated (as we just stated), appears as a fine tuning problem in general, except in
theories called no-scale supergravity (see Sec. 3.4) where geometrical properties
of the Kähler manifold lead to a vanishing potential. This condition of a vanishing
cosmological constant then promotes the gravitino mass to the status of an order
parameter.

3.2 The Goldstino

Let us first consider a supersymmetric theory with the Lagrangian (8) and suppose
that it is broken by some mechanism among those briefly described in Sec. 3.1, i.e.,
that some of the auxiliary fields F i , Da and some of the scalar fields ϕ i develop a
vev 8 . We thus obtain the fermion mass matrix (see (8) and (10))
 

† !
1 i √ a χ j

W i j −g ϕ Tb i
Lmass = − χ i 2λ MF √ b , MF ≡ .
2 i 2λ −g ϕ † Ta j 0

Using the minimisation of the potential (10), the gauge invariance of the superpo-
tential and (9) we have

(





i 
∂i V = Wi j F i + g ϕ † Ta i Da = 0
F = 0 , (11)
=⇒ MF
δaW ⋆ = W ⋆i δ ϕi† = −F i (ϕ † Ta )i = 0 − Da

and thus a massless field emerges, called the Goldstino or the Goldstone fermion:
1 
† i i
 1
i

ΨG = √ Fi χ − √ Da λ a = √ Fi† χ i − Da λ a . (12)
2 2 2 2

It is the analogue for broken supersymmetry of the Golstone boson in broken gauge
theories [98, 132]. However, since such a massless particle was not observed, again,
the absence of a massless fermion in Nature pleads against a spontaneous breaking
of global supersymmetry.
The inconvenience of a massless Goldstino is lifted in supergravity by a phe-
nomenon comparable to the Brout-Englet-Higgs mechanism for the electroweak
symmetry. We derive it below, assuming for simplicity, and since this does not mod-
ify the conclusion, a canonical Kähler potential and a trivial gauge kinetic function
as in (7). The interaction terms for fermions are the same as in supersymmetry mu-
tatis mutandis the term Wi j , which becomes Di D jW with Di , the covariant derivative
with respect to the Kähler manifold (for notations see [147]), and the presence of an
exponential factor:

8 We do not consider fermion condensates in this analysis


Supergravity: Application in Particle Physics 13

ϕ†ϕ
1 √
m2p
e−1 Lferm = − e Di D jW χ i · χ j − i 2gλ̄ a · χ̄ Ta ϕ + h.c. (13)
2
The scalar potential consist of (7) (which is in fact the F−term and the N−term of
the potential 9 ) and the usual D−term of supersymmetry. After solving the equation
of motion for the auxiliary fields one obtains (see e.g. [147]).



† †
1 ϕ ϕ 1 ϕ ϕ

i 2 m2
i ⋆
a



2 m2

F =e p D W , D = g ϕ Ta ϕ , N = −3e p W

and the Golstino of (12) emerges as a massless field, as in supersymmetry.


However, under a supergravity transformation we have

δε χ i = 2ε F i + · · · , δε λ a = iε Da + · · ·

where the dots indicate terms that are irrelevant for our analysis. In particular, the
Goldstino transforms as:

1

δε ΨG = ε Fi† F i + Da Da + · · · (14)
2
which involves the F− and D−parts of the scalar potential. As the prefactor of the
symmetry parameter ε is non-zero for broken supergravity, a local choice of ε allows
to purge the Lagrangian from the Goldstino field (similarly to the Goldstone boson
in broken gauge theory) [25, 24, 27]. In this process, the gravitino acquires a mass

1 D 21 mK2p E
m3 = We . (15)
2 m2p

The irrelevance of the Goldstino as a dynamical field can also be exhibited by an


explicit diagonalisation of the mass matrix, in which a field redefinition makes it dis-
appear [150, 69, 147]. Depending on the mediation mechanism, the gravitino mass
can take very different values. For instance in gauge-mediation it is typically very
light, because m p −suppressed, (and it can be the lightest supersymmetric particle).
We shall see in the next subsection that in gravity mediation it determines the super-
symmetry breaking scale (under assumption of a vanishing cosmological constant)
and thus remains comparable to the electroweak scale, i.e., ∼100GeV−1TeV.

3.3 Gravity induced supersymmetry breaking

In this section we focus on gravity-induced supersymmetry breaking. The chiral su-


perfields are denoted as X I = (Z i , Φ a ), i = 1, · · · , m, a = 1, · · · , n where Z (resp.
Φ ) are the superfields in the hidden (resp. observable) sector (the corresponding

9 The auxiliary field N is related in a simple way to the field M after an change of variables.
14 Florian Domingo and Michel Rausch de Traubenberg

scalars fields are denoted xI = (zi , ϕ a )). We do not specify the observable sector
here, which corresponds to a supersymmetric extension of the Standard Model
(see Sec. 4). Of course the model also contains vector superfields associated to
the Yang-Mills interactions, but these fields are irrelevant for the derivation of the
supersymmetry-breaking terms in the observable sector, which we perform in this
subsection. In addition, we assume that supergravity is broken in the hidden sector
by an O’Raifeartaigh mechanism. More specifically, we suppose that some of the
fields of the hidden sector develop a vev of the order of the Planck mass while vevs
for the fields in the observable sector would be of the order of the electroweak or
GUT scales, and thus subleading.
Before expanding the scalar potential (7) in the vicinity of its minimum for
m p → ∞ (now in the presence of vevs of order m p , contrarily to the discussion
of Sec. 2 with unbroken supergravity), it is necessary to specify the behaviour of
the terms connecting hidden and observable sectors in this limit: these are assumed
to remain finite when m p → ∞, which considerably restricts the possible form of
the Kähler potential and superpotential, as was demonstrated by Soni and Weldon
in [139]. The form retained by Soni and Weldon leads to a soft-breaking of super-
symmetry in the observable sector for the limit m p → ∞, as phenomenologically
desirable in traditional applications of supersymmetry to particle physics, so that
only UV divergences of logarithmic type appear in quantum corrections, thus pro-
tecting electroweak scalar masses against corrections from the UV spectrum (Hier-
archy Problem). Nevertheless, we note that supergravity breaking admits solutions
beyond those of Soni & Weldon, with in particular possible hard-breaking terms.
Models of this type were analysed in [117] and are not deprived of phenomenologi-
cal qualities. Nevertheless, we will focus below on the more commonly studied case
of soft-breaking.
In [78, 15] the Kähler potential and superpotential were chosen as
 
b Z † ) + Φ †∗ Λ a∗ a (Z, Z † )Φ a + Γk (Z, Z † )gk2 (Φ ) + h.c.
K(Z, Z † , Φ , Φ † ) = K(Z, a
b (Z) + Wk (Z)gk3 (Φ ) .
W (Z, Φ ) = W

Remembering the relative mass dimensions of these quantities it is further assumed


that the realisation of the Kähler potential in the hidden (resp. observable) sector
scales like 10 Kb ∼ m2p (resp. M 2 ), while the superpotential is of order W
b ∼ Mm2p
(resp. M 3 ), with M ≪ m p . There are a priori no restrictions on the form of the
functions gk2 and gk3 and the general form of the scalar potential after supergravity
breaking can be found in [78, 15].
For the current analysis, however, we assume a renormalisable theory in the ob-
servable sector, and specify

10 This assumption is a consequence of the results of Soni & Weldon [142].


Supergravity: Application in Particle Physics 15
1 
b Z † ) + Φ †∗ Λ a∗ a (Z, Z † )Φ a +
K(Z, Z † , Φ , Φ † ) = K(Z, Γab (Z, Z † )Φ a Φ b + h.c.
a
2
b 1 1
W (Z, Φ ) = W (Z) + mab (Z)Φ Φ + λabc (Z)Φ a Φ b Φ c
a b
(16)
2 6
i.e., the Kähler potential (resp. superpotential) is a quadratic (resp. cubic) function of
the observable fields while the dependence of Kb and W b or Λ a∗ a , Γab , λabc , mab on the
hidden fields remains arbitrary. Expanding the scalar potential is not intrinsically
difficult, but the computation is lengthy because, as always in supergravity, there
are numerous terms involved. We shall emphasise a few steps of this calculation
(further technical details are given in [147]). Choosing an O’Raifeartaigh breaking
mechanism, at least one of the F i −fields in the hidden sector develops a vev:
D 1 Kb
∗ E
i
F = e
2 m2
p bi I ∗ D I W
K b⋆ =6 0, (17)





(note that K ∼ K b and W ∼ W b ). The order parameter of supersymmetry
breaking is then given by


j∗
i
b i F .
m4susy = Fj†∗ K (18)

1. Inversion of the Kähler metric: Writing the Kähler metric and its inverse


  ∗ ∗ 
AB ∂a ∂ a K ∂i ∂ a K
K= = ∗ ∗
CD ∂a ∂ i K ∂i ∂ i K
 
−1 (A − BD−1C)−1 −(A − BD−1C)−1 BD−1
K = ,
−D−1C(A − BD−1C)−1 D−1C(A − BD−1C)−1 BD−1 + D−1

we have A = A0, B = m−1 −1 −2


p B1 ,C = m p C1 , D = m p D2 +D0 , with A0 , B1 ,C1 , D0 , D2
(not explicitly given here) of order m0p , so that D−1 = D−1 −2
0 (1 + m p D0 D2 )
−1 and
−1 −1 −1 −2 −1 −1
(A − BD C) = A0 (1 − m p A0 B1 D C1 ) are perturbatively inverted.
2. One should compute all covariant derivatives DI W = ∂I W + m−2 p W ∂I K , and all
J∗
terms exp(m−2 ⋆ I ∗ ∗ ∗ ∗ ∗
p K)DI W D W K J∗ for (I, J ) = (i, j ), (a, j ), (i, b ) and (a, b )
corresponding to the hidden-hidden, observable-hidden, hidden-observable, obs-
ervable-observable sectors. We stress that these expressions all involve the in-
verse Kähler metric.

3. Assume that the scalar fields z in the hidden sector are frozen to their vev z ∼
m p and introduce the gravitino cd mass (15). Note that, since in the expansion of

DI W K i J∗ D J W ⋆ − 3m−2 2 2
p |W | there are terms of order m p (see the cosmological
constant below in Vsoft ), an additional contribution emerges from the expansion
of the exponential factor (the term 1/m2p coming from the observable sector −see
(16)).
Collecting all terms, the scalar potential takes the form [15, 147]
16 Florian Domingo and Michel Rausch de Traubenberg


V = m2pΛ + ∂aWm (Λ −1 )a a∗ ∂ a Wm⋆ + Vsoft (19)

to which the usual D−term (see (10)) should be added. The first term Λ (with a
summation applying only to the fields of the hidden sector)
1
Λ= b j∗ i ihF i i − 3|m 3 |2 = 1 m4susy − 3 m 3 2 ,
hFj†∗ ihK
2
mp 2 mp2 2

is the cosmological constant. It is commonly accepted that, for phenomenologically


acceptable theories, Λ ≈ 0. Imposing this condition admittedly amounts to consid-
erable fine-tuning. Assuming a vanishing cosmological constant leads to the relation
2
m4susy = 3m2p m 3 . (20)
2

The second term of (19) is the usual F−term of an unbroken supersymmetric


theory with superpotential Wm given by

∗ 
Wm = 16 λ̂abc Φ a Φ b Φ c + 12 m̂ab + m 3 hΓab i − Fi†∗ h∂ i Γab i Φ a Φ b , (21)
2
b
hKi b
hKi
2m2p 2
λ̂abc = e hλabc i and m̂ab = e 2m p hmab i . (22)

Finally the last term explicitly breaks supersymmetry:


 ∗ ∗
 1 1 
2
Vsoft = ϕa†∗ m 3 Sa a + Λ hΛ a a i ϕ a + Aabc ϕ a ϕ b ϕ c + Bab ϕ a ϕ b + h.c. ,
2 6 2
where
b
"
1 hKi
2 m2
i 1
Aabc = e p F b λabc i + h∂i λabc i
h∂i Kih (23)
m2p
#
 
−1 d ∗ a∗
− h(Λ ) a ∂iΛ a λdbc i + (a ↔ b) + (a ↔ c) ,

for the trilinear terms


" #
b
1 hKi  
2 m2
i 1 ∗
Bab = e p F b
h∂i Kihm −1 c a
ab i + h∂i mab i − h(Λ ) a∗ ∂i Λ b mac i + (a ↔ b)
m2p
b
1 hKi
† 2 m2

−m 3 e hmab i + (2|m 3 |2 + Λ )hΓab i − m†3 Fi⋆∗ h∂ i Γab i
p (24)
2 2 2
" #

i  ∗

−1 c ∗ a
+m 3 F h∂i Γab i − h(Λ ) a ∂iΛ bΓac i + (a ↔ b)
2
" #

i ⋆ ∗
 ∗ ∗

− F Fi∗ h∂i ∂ i Γab i − h(Λ −1 )c a∗ ∂i Λ a b ∂ i Γac i + (a ↔ b) .,
Supergravity: Application in Particle Physics 17

for the bilinear terms, and


∗ ∗ 1 D ∗ ∗ ∗ ∗ ∗
E
Sa a = hΛ a a i + 2
hFi†∗ i ∂ i Λ a bΛ −1 )b b∗ ∂i Λ b a − ∂i ∂ i Λ a a hF i i (25)
|m 3 |
2





b ∼
If (20) is satisfied, F i ∼ m p m 3 , using (15), (17) and the scalings W ∼ W



2
Mm2p , K ∼ K b ∼ m2p and z ∼ m p , one arrives at


i D ∗ 1 c⋆  i E
F = m p m†3 ∂ j b + m p ln W K
K b j∗ = m p m†3 ρ̂ i (26)
2 mp m3p 2

with ρ̂ i ∼ m0p , and all soft supersymmetric scales are thus controlled by the grav-
itino mass, which is then required to be of the order of the electroweak scale (as an
implementation of supersymmetry as a solution of the Hierarchy Problem).
As announced, the potential Vsoft contains terms that break supersymmetry ex-
plicitly, but in a soft way. This means that the loop-quantum corrections associated
to these terms lead to at most logarithmic ultraviolet divergences. The last soft-
breaking term is the gaugino mass and is associated to a non-trivial gauge kinetic
function hab 6= δab (see e.g. [147]). We thus have four types of soft-supersymmetric
breaking terms:

• squared masses for all the scalar fields (see the Sa b −terms);
• masses for all gauginos (not derived here);
• trilinear couplings among the scalars (see the Aabc −terms);
• bilinear couplings among the scalars (see the Bab -terms).
The trilinear terms are directly related to the cubic part of the superpotential whilst
the bilinear terms have two origins: (1) one related to the quadratic part of the super-
potential and (2) one related to the Γ −term in the Kähler potential. The contribu-
tion of the latter type is very interesting and allows to solve the famous µ −problem
of the Minimal-Supersymmetric-Standard-Model (see also the Γab −contribution in
(21)) [78]. (See Sec. 4.) In the presence of an additional linear term αa Φ a in the
superpotential (with Φ a representing fields of the observable sector that are singlet
under all gauge groups), the gravity-mediated supersymmetry breaking mechanism
would have generated additional soft-breaking terms of the form Ca ϕ a + h.c. The
soft supersymmetric terms were classified in [76]. Terms of a similar form can be
generated with the gauge- and anomaly-mediation mechanisms.
At the term of this construction, we recall that
the effective Kähler potential for

the observable fields Φ a still reads Keff = Φa†∗ Λ a a Φ a , so that the kinetic terms
for the associated scalars and fermions are not canonically normalised. The last step
for an application to particle physics thus amounts to a field re-definition Φ a → Φ bi,
enforcing the normalisation Keff = Φ b†Φ b i [147]. This operation further transforms
i
the effective couplings applying to the Φ b i basis, as compared to the expressions
in Eqs.(19-25). In particular, the effective superpotential Wm (Φ a ) → Wmeff (Φ b i ) then
contributes to the scalar

potential in the canonical form of Eq.(10), instead of that

of Eq.(19) involving (Λ −1 )a a .
18 Florian Domingo and Michel Rausch de Traubenberg

As a summary, let us briefly recapitulate the main steps entering the construction
of a supergravity model exploitable in particle physics:
1. specify the observable gauge interactions and matter content, respectively de-
scribed by vector and chiral superfields;
2. introduce the (gauge invariant) basic functions of Eq.(16) fixing the interplay
between observable and hidden sectors;
3. induce spontaneous supersymmetry breaking in the hidden sector and take the
limit m p → ∞, which results in an effective field theory for the observable sector
including soft supersymmetry-breaking terms — see Eqs.(19-25);
4.
re-define
the fields of the observable sector (via diagonalisation and rescaling of
Λ ) so that their kinetic terms are canonical.
We will directly exploit these results in Sect. 4 when setting up a realistic supersym-
metry-inspired model of particle physics, the Minimal Supersymmetric Standard
Model.

3.4 No-scale supergravity

As mentioned in the previous section, gravity-mediated supersymmetry breaking


suffers from a fine tuning problem, related to the cosmological constant, which is
artificially set to zero. In [28, 47, 112], a class of models, known as no-scale super-
gravity, was derived, for which the vanishing of the cosmological constant emerges
as a property of the underlying Kähler manifold where the scalar fields are liv-
ing. Furthermore, the gravitino mass is generated dynamically. In fact, the scalar
potential vanishes identically in such models (one speaks of flat directions in the
potential). In particular, the gravitino mass is non-vanishing, although classically
undetermined, and its values emerges through quantum corrections.
Let us denote as Φ I , I = (0, i), i = 1, · · · , n the chiral superfields (Φ 0 being in the
hidden sector and Φ i in the observable sector). We first notice that it is possible
to unify the Kähler potential and the superpotential into the so-called generalised
Kähler potential

|W |2
G (Φ , Φ † ) = K(Φ , Φ † ) + m2p ln .
m6p

In fact, this transformation corresponds to a super-conformal transformation, or a


Kähler transformation, (see e.g. [147]). With this new function, the scalar potential
takes the form
1
2 G ∗

V = m2p e m p Gi G i i∗ G i − 3m2p ,

∗ ∗ ∗ ∗
with Gi = ∂i G , G i = ∂ i G and G i j = ∂i ∂ i G the Kähler metric and G i i∗ its inverse.
Choosing
Supergravity: Application in Particle Physics 19
" #
φ 0 + φ0†∗ h(φ i , φi†∗ )
G = −3m2p ln − + m2pF(Φ i ) + m2pF ∗ (Φi†∗ )
m2p m2p

where h, which contributes to the kinetic part of the scalar ϕ i and the fermions χ i , is
an unspecified function and F is a function related to interactions in the observable
sector (in fact one can show that W = m3p eF with W the superpotential), a direct

computation leads to Gi G i i∗ G i = 3m2p Generalised Kähler potentials of the above
type lead to no-scale models. Such models have been studied in the context of SU(5)
Grand Unified theories in [47, 112] or in the context of the Standard Model [48].

4 Supergravity and supersymmetry in particle physics

The formalism of the previous sections has provided us with a framework orig-
inally embedded in an N = 1 supergravity theory, but whose observable sector
eventually reduces to an N = 1 supersymmetric model with explicit, albeit soft,
supersymmetry-breaking terms. The avowed purpose behind this construction rested
with the double wish of, first, embedding the obviously non-supersymmetric spec-
trum of particle physics (‘Standard Model’), at energies comparable to the elec-
troweak scale, second, exploiting the technical protection of scalar masses against
quantum corrections from the UV spectrum (Grand Unification, Quantum Gravity,
etc.) in supersymmetric theories. The resulting hybrid indeed amounts to a non-
supersymmetric model at energies below the scale of the supersymmetry-breaking
terms emerging in the observable sector, Msoft (≈ m 3 in gravity mediation), but it
2
retains the properties of a supersymmetric theory at energies above that scale. We
shall now fulfil our program by explicitly introducing the Standard Model fields in
this framework. The reader interested in a detailed discussion of supersymmetric
extensions of the Standard Model will read with profit [43, 44].

4.1 The Minimal Supersymmetric Standard Model

Observing, first, that particle physics at energies comparable to the electroweak


scale is well described by the Standard Model, second, that Msoft should be rela-
tively close to the electroweak scale in order to efficiently shield the Higgs squared
mass — radiative corrections still involve the hierarchy between the electroweak and
soft scales —, it is meaningful to attempt an as-economical-as-possible embedding
of the Standard Model in terms of new-physics fields. The product of this operation
is known as the Minimal Supersymmetric Standard Model (MSSM) [55, 56, 58,
123, 90], and it will be the main model under discussion in this section. Following
this principle of minimality, the gauge group remains unchanged as compared to the
Standard Model, GSM = SU(3)c × SU(2)L × U(1)Y , but its implementation in a su-
20 Florian Domingo and Michel Rausch de Traubenberg

persymmetric context calls for the introduction of the full set of vector superfields,
gauge bosons and associated gauginos, in the adjoint representation:
Ta
SU(3)c → Ĝ = (Gaµ , g̃a ) = (8, 1, 0) = (gluons, gluinos)
2 ee
σ i
SU(2)L → Ŵ = (Wµi , w̃i ) = (1, 3, 0) = (W − bosons, winos) (27)
2 ee
U(1)Y → B̂ = (Bµ , b̃) = (1, 1, 0) = (B − boson, bino) .
ee
In the notation (d3 , d2 , q1 ), d3 , d2 and q1 correspond to the dimension of the SU(3)c
representation, thate e e e representation and the hypercharge under which
of the SU(2) L
the fields transform. The objects Ta and σi respectively denote the Gell-Mann and
Pauli matrices. Superfields are written with a hat and supersymmetric partners with
a tilde.
Similarly, the fermionic matter content of the Standard model calls for the intro-
duction of three generations ( f = 1, 2, 3) of chiral superfields describing the quarks
and leptons, as well as their supersymmetric counterparts, the squarks and sleptons.
Writing only the left-handed fields (the right-handed ones are deduced by hermitian
conjugation; the superscript c is just a notation to distinguish SU(2)L singlets from
the doublets):
 f  f !  
f ue u 1 cf cf cf 2
Q̂ = , = (3 , 2 , ) , Û = e
u , u = (3̄, 1, − ) ,
def L d f L ee 6 R R
ee 3
  1
D̂c f = deRc f , dRc f = (3̄, 1, ) , (28)
ee 3
 f  f !  
νe ν 1 cf cf
L̂ f = f , f = (1, 2, − ) , Ê c f = eeR , eR = (1, 1, 1) .
ee L e L
ee 2 ee

Finally, the Higgs field requires another chiral supermultiplet Ĥ = (1, 2, − 12 ), which
it shares with the higgsino. Yet, the above field content would lead to e non-vanishing
e
chiral anomalies, hence to a non-renormalisable model, due to the addition of the
electroweakly charged higgsino to the fermion sector of the Standard Model. In ad-
dition, the holomorphicity of the superpotential forbids the use of conjugate fields
when writing the Yukawa terms, so that top and bottom masses cannot be gener-
ated with a single Higgs superfield. This prompts for the introduction of two Higgs
supermultiplets with opposite hypercharges instead:
 0 !!  + !!
HD e
h0D 1 HU e
hU+ 1
ĤD = , e− = (1, 2, − ) , ĤU = , e0 = (1, 2, ) .
HD− hD ee 2 HU0 hU ee 2
(29)
For the Higgs fields we have further indicated their U(1)em −charges.
The chiral superfields Q̂ f , Û c f , D̂c f , L̂ f , Ê c f , ĤU , ĤD represent, for the MSSM,
the observable fields denoted as Φ b i at the end of the construction of Sec. 3.3 —
i.e., after enforcing a canonical normalisation of the kinetic terms. Following the
Supergravity: Application in Particle Physics 21

results of Eqs.(19-25) the interactions in the observable sector of the MSSM can be
described by
• an effective
(by assumption
renormalisable) superpotential — we stress that
the matrix (Λ −1 )a a∗ appearing in Eq.(19) becomes trivial after the field re-
definition restoring canonical kinetic terms —, constrained by the symmetries,

′ ′ f f′ ′ ′ ′
WMSSM =µ ĤU · ĤD + Yuf f Q̂ f · ĤÛ c f − Yd Q̂ f · ĤD D̂c f − Yef f L̂ f · ĤD Ê c f
+ µi ĤU · L̂i + λi jk L̂i · L̂ j Ê c k + λi′jk L̂i · Q̂ j D̂c k + λi′′jkÛ c i D̂c j D̂c k (30)

where Φ a · Φ b ≡ (Φ a )1 (Φ b )2 − (Φ a )2 (Φ b )1 stands for the SU(2)L -invariant


product (with the indices 1, 2 referring to the weak isospin);
• a set of soft-breaking terms of similar form, with bilinear and trilinear scalar
couplings B’s and A’s as in Eqs.(23,24);

• squared mass terms m2 aa φ̂a†∗ φ̂ a — determined by Eq.(25) — where off-diagonal
elements only occur for scalar fields with identical gauge quantum numbers;
• gaugino mass terms 12 M1 b̃ · b̃ + 21 M2 w̃i · w̃i + 21 M3 g̃a · g̃a + h.c.
Also the D-term contribution to the scalar potential should be restored in Eq.(19).
Then, given the arbitrariness of the functions K, b Λ and Γ in the hidden sector —
as well as the possibility to call to various breaking mechanisms beyond gravity
mediation —, it is evident from Eqs.(23-25) that almost any choice of soft-breaking
parameters can be a posteriori justified by a judicious Ansatz. The simplistic case
of trivial functions leads to universality conditions (the flavor index is omitted):
 
cMSSM
Vsoft = m20 |HU |2 + |HD |2 + |Q|2 + |U c |2 + |Dc |2 + |L|2 + |E c |2
 
+ 21 M 1 b̃ · b̃ + w̃i · w̃i + g̃a · g̃a + h.c. (31)
2
h   i
+ b µ HU · HD + a Yu Q · HU U − Yd Q · HD Dc − Ye L · HD E c + h.c. .

This choice, (when the second line of (30) is not considered) more predictive than
realistic, is only marginally compatible with experimental results today [12, 91].
A discrete symmetry, R-parity [53], with charge assignment 1 to all Standard
Model fields and −1 to their supersymmetric partners, is often introduced to elimi-
nate the terms of the second line of Eq.(30), as well as the associated soft-breaking
couplings. These terms explicitly violate lepton- or baryon-number and lead to a
distinctive phenomenology [10]. We discard them till further notice and work with
RpC
the reduced superpotential WMSSM defined by the first line of Eq.(30). An alterna-
tive charge assignment, actually equivalent, is that of a matter parity, with all quark
and lepton superfields carrying a charge −1, while the Higgs superfields transform
trivially.
Let us pause at this point and look at the spectrum in the absence of electroweak-
symmetry breaking. Protected by GSM , the quarks, leptons and gauge bosons remain
massless, as expected. Squarks and sleptons receive squared masses ∼ Msoft 2 from
22 Florian Domingo and Michel Rausch de Traubenberg

the quadratic soft terms: the absence of observed resonances at the Large Hadron
Collider tends to push Msoft above the TeV scale. The gauginos take masses of order
Msoft from the soft-breaking parameters M1,2,3 , while higgsinos receive a mass µ
from the superpotential. Several squared-mass scales intervene in the Higgs sector,
µ 2 from the superpotential, B ∼ µ Msoft and m2HU,D ∼ Msoft
2 from the soft-breaking

scalar potential: their interplay must be examined more closely to deduce the con-
ditions of emergence for electroweak-symmetry breaking.

4.2 The µ -problem

It was realised early on [109] that the presence of a supersymmetry-conserving mass


term µ in the superpotential of Eq.(30) was phenomenologically problematic. This
parameter being a priori unrelated to supersymmetry breaking, a natural choice for
µ would involve some high-energy scale, Planck or Grand Unification, much above
Msoft . However, in such a case, the Higgs potential is dominated by the squared-
mass term |µ |2 (|HD |2 + |HU |2 ), which no soft contribution can balance, and the
electroweak symmetry cannot be broken. Failing to have µ large, we may set it to 0,
such a choice being protected by the emergence of a U(1) symmetry. Nevertheless,
given that the higgsinos take their mass from µ and that none was observed at the
Large Electron-Positron collider, µ ≥ 100 GeV, which invalidates this alternative.
The non-renormalisation theorems [87] also forbid to generate the µ parameter ra-
diatively. In the aftermath, the unnatural choice µ ≈ Msoft must be retained: this is
the µ -problem.
A first solution consists in relating the emergence of µ to the supersymmetry-
breaking mechanism: this is the proposal by Giudice and Masiero [78]. In this ap-
proach, the µ parameter is absent in the original superpotential of the supergravity
model (all parameters are then dimensionless in the Kähler potential and the super-
potential (16)), while a term Γ (Z, Z † )ĤU · ĤD appears in the Kähler potential — thus
explicitly breaking the U(1) symmetry and avoiding the appearance of an associated
Goldstone boson when Higgs fields take a vev. Then the ‘observable’ µ parameter
appears in the limit m p → ∞ through the Γ contributions to the bilinear terms in
Eq.(21), and is naturally of the order of the gravitino mass. A variant [20] consists
in writing a term m12 λ (Z)Wb (Z)HU · HD in the superpotential: it can be related to a
p
choice Γ (Z, Z † ) = λ (Z) + λ ⋆ (Z † ) by a Kähler transformation.
Another solution consists in generating the µ parameter directly in the observable
sector from the vev of an additional singlet superfield Ŝ = (1, 1, 0) [54]. The corre-
sponding model is known as the Next-to-Minimal supersymmetric e e Standard Model
(NMSSM) [50, 51]. The simplest choice to forbid the µ term in the superpotential
amounts to a Z3 symmetry, allowing only for cubic terms:

Z3 κ 3
WNMSSM = λ ŜĤU · ĤD + Ŝ + Yukawas, (32)
3
Supergravity: Application in Particle Physics 23

where ‘Yukawas’ stands for the first line in (30) without the µ −term.
Then the
minimisation of the scalar potential generally

allows S to develop a vev S , which
results in an effective µ parameter µeff = λ S . We stress that the only scale entering
the scalar potential is the supersymmetry-breaking one, thus naturally relating the
electroweak-symmetry breaking scale to the latter. In addition, there are numerous
phenomenological applications of the new singlet fields, on which we will oppor-
tunistically comment in due time. On the other hand, the Z3 symmetry proves prob-
lematic from the perspective of cosmology, where it causes a Domain-Wall problem
[148]. Alternative choices using R-symmetries have been studied in [113].

4.3 Radiative corrections and renormalisation group evolution

Meaningful predictions in the MSSM (or any supersymmetric extension of the Stan-
dard Model) imply processing this framework as a Quantum Field Theory, i.e., being
able to calculate quantum corrections. To this end, it is necessary to regularise ultra-
violet divergences appearing in loop diagrams and redefine bare parameters so that
such divergences cancel out at the level of observable quantities (renormalisation).
The most popular approach to regularisation consists in performing calculations in
D = 4 − 2ε spacetime dimensions instead of 4. However, this ‘naive’ dimensional
regularisation explicitly violates supersymmetry in that vector fields become 4 − 2ε
dimensional while their fermionic counterparts remain 4-dimensional, hence intro-
ducing a mismatch between bosonic and fermionic degrees of freedom. This issue is
addressed in dimensional reduction [136, 143], where vector fields formally retain
4 dimensions through the introduction of ‘epsilon scalars’, living in the 2ε dimen-
sions. It is then possible to renormalise the model, e.g. through modified minimal
subtraction of the ultraviolet divergences, leading to the DR scheme. Alternatively,

the DR scheme [101] allows to decouple the epsilon-scalars. We stress that it is
not absolutely imperative to work in a framework respecting supersymmetry. One
then forfeits the relations between parameters (e.g. in the scalar and fermion interac-
tions) that are guaranteed by supersymmetry. Thus, MS, for example, remains a le-
gitimate choice. Nevertheless, to ensure that this non-supersymmetric treatment de-
scribes a (softly-broken) supersymmetric model, it is ultimately necessary to define
its renormalised parameters through their connection to those of a supersymmetry-
conserving scheme.
In a context involving vastly different scales, such as electroweak physics on
one side and gravity-mediated supersymmetry-breaking on the other, it is prefer-
able to resum the ultraviolet logarithms developing between the two scales through
the use of the renormalisation group evolution, rather than work with parameters
defined at a widely different energy from that of the physical process under study.
This coming-and-going between scales is therefore needed, both for testing the pre-
dictions of a given high-energy model, e.g. of supersymmetry breaking, on particle
physics, or in view of inferring the structure of the high-energy theory from the low-
energy phenomenology. The beta functions of gauge couplings and superpotential
24 Florian Domingo and Michel Rausch de Traubenberg

parameters are known up to four and three loop, respectively [102, 103, 104, 105].
On the other hand, the renormalisation group equations for the soft-breaking param-
eters are generically known at two-loop order [114, 158], although methods have
been proposed and implemented in special cases to include higher orders. As could
be anticipated, the running of gauge and superpotential parameters involves only
supersymmetry-conserving parameters in their beta functions, while the equations
for the soft parameters include both soft-breaking and supersymmetry-conserving
parameters. The impact of the running for well separated scales is considerable in
general and a set of degenerate soft mass parameters at m p , as in Eq.(31), — or at a
Grand Unification scale at which several fields are connected by the extended gauge
symmetry — would produce a hierarchical spectrum at the TeV scale.
As an interesting consequence of the modified matter content of the MSSM,
with new physics fields in the TeV range, the gauge couplings of GSM show an
approximate convergence when running them up from the electroweak scale towards
a unification scale of MGUT ≈ 1016 GeV [4]. This feature would allow for a Grand
Unification of GSM in a single step.

4.4 Electroweak symmetry breaking and the MSSM Higgs sector

The tree-level scalar potential for the Higgs fields in the MSSM reads:
h i
VH = (m2HD + |µ |2)|HD |2 + (m2HU + |µ |2 )|HU |2 + m212 HU · HD + h.c.
g2 + g22 h i g2 − g2 g2
+ 1 |HD |4 + |HU |4 + 2 1
|HD |2 |HU |2 − 2 |HU · HD |2 (33)
8 4 2
where g1 and g2 are the electroweak gauge couplings, from which one can in-
fer the D-term origin of the quartic operators. The mass-terms m2HD,U ∼ Msoft 2 and

m212 ∼ Msoft µ (substituting the ‘B’ notation) denote the quadratic and bilinear soft-
breaking terms respectively. Any complex phase in m212 can be absorbed through a
re-definition of the (super)fields, so that the tree-level Higgs potential of the MSSM
is automatically CP-conserving — this is no longer necessarily the case in extended
models such as the NMSSM.
The
of Eq.(33) — see the details in e.g. [40, 73] — leads to the solu-

minimisation
tion ( HD0 , HU0 ) = v(cos β , sin β ) respecting the electromagnetic U(1) symmetry
under some conditions among the parameters. √ The parameter v should be identified
to the electroweak symmetry-breaking vev (2 2GF )−1/2 , where GF is the Fermi
constant measured in muon decays. Then, after examining the quadratic terms in
the vicinity of this minimum, a triplet of Goldstone bosons G0 , G± decouples, leav-
2m212
ing a CP-odd mass eigenstate A0 with squared mass MA2 ≡ sin 2β and a charged one
g22 2
H ±,with MH2 ±
= MA2 2
+ MW
and ≡ 2
MW 2v .
Two states h0
and H 0 remain in the
CP-even sector and mix according to the mass matrix (MW,Z are the masses of the
Supergravity: Application in Particle Physics 25

electroweak gauge bosons):


 2 2 
2 MA sin β + MZ2 cos2 β −(MA2 + MZ2 ) sin β cos β g21 +g22 2
MCPE ≡ ; MZ2 ≡ 2 v .
−(MA2 + MZ2 ) sin β cos β MA2 cos2 β + MZ2 sin2 β
(34)
In this equation the minimisation of the Higgs potential has been explicitly applied.
The reader may consult [73] for an explicit computation of the various mass matrix
terms.
An effective Standard Model is obtained in the limit MA ≫ MZ , where h0 exactly
identifies with the electroweak partner of the Goldstone bosons while (H 0 , A0 , H ± )
form a heavy (degenerate) doublet. In view of the measured properties of the Higgs
particle at 125.25 GeV, as well as the phenomenological constraints on non-standard
Higgs doublets, this decoupling scenario appears as the most realistic one. Never-
theless, light non-standard Higgs states are still very compatible with collider data
as long as they are dominantly singlet, as happens in e.g. the NMSSM, due to their
suppressed coupling to standard matter.
The lightest eigenvalue of Eq.(34) satisfies m2h0 ≤ min(MZ2 , MA2 ) cos2 2β . Given
MZ ≈ 91 GeV, mh0 seems much below the observed 125.25 GeV. However, this rep-
resents no fundamental incompatibility, because the connection between mh0 and
MZ is a tree-level relation enforced by supersymmetry, but, the latter being broken
at the electroweak scale, it is not preserved by radiative corrections. The estimated
upper bound on the physical mass Mh0 before the Higgs discovery amounted to
Mh0 ≤ 135 GeV in the MSSM (strongly dependent on the mass of the top quark mt )
[40]. Nevertheless, tan β ≫ 1 is favoured in order to maximise the tree-level contri-
bution. Further tree-level effects are possible in extensions of the MSSM, such as a
contribution of the F-term λ — see Eq.(32) — in the NMSSM, or a mass uplift via
mixing of h0 with a lighter CP-even singlet. Still, radiative corrections to the mass
of the Standard-Model-like Higgs reach a considerable relative size, so that a good
control of the higher orders is needed to reduce the uncertainties. The correspond-
ing calculations are reviewed in [138]. The leading corrections are controlled by the
Yukawa coupling of the top:

3mt4 h MT̃2 Xt2 Xt4


i
∆ m2h0 ≈ ln mt2
+ MT̃2
− 12MT̃4
(35)
4 π 2 v2
where MT̃ ∼ Msoft ≫ mt represents the average mass of the scalar partners of the
top, while Xt ∼ Msoft parameterises their mixing, which is induced by electroweak-
symmetry breaking. The first term corresponds to an ultraviolet logarithm and shows
that, in the presence of a hierarchy between the standard and the supersymmetric
2
sectors, the expansion parameter in the perturbative series is not simply 43m t
π 2 v2
but
2
3mt2 Msoft
4π 2 v2
ln mt2
. This implies the continuing emergence of large effects at higher or-
2 M2 4 M2
ders, e.g. of the form απS3mv2t ln2 msoft mt
2 , π 4 v4 ln
2 soft
mt2
at two-loop. These ultraviolet log-
t
arithms can be resummed using the effective field theory techniques, the impact of
this resummation being already substantial for Msoft ≈ 1 TeV. Non-logarithmic and
26 Florian Domingo and Michel Rausch de Traubenberg

electroweak corrections are naturally also needed for precision predictions. As such,
the higher-order uncertainty on the Higgs mass prediction may still remain above
1 GeV in scenarios with TeV-scale supersymmetric sectors.
Of course, the Higgs mass is only one electroweak observable among many. For
a decoupling scenario, the Higgs couplings are expected to be Standard-Model-like,
which is consistent with the current experimental status: narrower determinations
in the future may place indirect constraints on the non-Standard spectrum. Leaving
the Higgs sector, the relations among the electroweak parameters — fine structure
constant, Fermi constant, W - and Z-masses and the gauge couplings to fermions —
constrain the non Standard spectrum. The reader may refer to [92, 93] for analyses
in the MSSM. Nevertheless, for models with only doublet (and singlet) vevs and a
heavy (TeV-like) new-physics sector, such observables are expected to be in good
agreement with the Standard Model.
Another precision observable of great interest (though only loosely related to
electroweak-symmetry breaking) is the anomalous magnetic moment of the muon,
where a historically durable deviation with the Standard Model prediction was re-
cently confirmed by the Fermilab Muon g − 2 experiment. The discrepancy is of
comparable magnitude with the standard electroweak contributions to this observ-
able and could thus hint at new-physics effects at comparable scales. The status of
this observable in the MSSM is reviewed in [144]: gauginos, higgsinos and / or slep-
tons are then expected well below the TeV scale in order to account for the measured
anomaly, to which these particles contribute already at one-loop order.
Let us finally comment on the question of the stability of the electroweak-
symmetry-breaking vacuum. Indeed, the scalar potential of the MSSM is not re-
stricted to Eq.(33), but also includes squarks and sleptons. Vacuum expectation
values of charged or coloured fields may lead, for a given choice of Lagrangian
parameters, to minima deeper than the electroweak-symmetry-breaking one, thus
endangering the phenomenological consistency of an expansion in the vicinity of
that specific minimum — meta-stability may still legitimize the point in parameter
space. Approximate analytical constraints on the parameters were derived at tree-
level early on. A full analysis of vacuum stability is nevertheless considerably more
involved, due to the large dimensionality of the space of scalar fields, and even more
so when including radiative corrections. We refer the reader to [95] for a recent dis-
cussion and a list of references.

4.5 The Supersymmetry flavour problem

The flavour structure of the Standard Model is particularly simple in that only the
misalignment between the Yukawa coupling matrices in the quark sector allows for
flavour transitions governed by the Cabibbo-Kobayashi-Maskawa matrix V CKM and
mediated by the electroweak charged current. This feature is lost when considering
a supersymmetric extension of the Standard Model, because the soft-breaking terms
in the scalar potential introduce new sources of flavour violation that are poten-
Supergravity: Application in Particle Physics 27

tially independent from the Yukawa couplings: these are the quadratic mass terms
′ f f′
m2F̃ f f (F̃ = Q̃, Ũ c , D̃c , L̃, Ẽ c ) and the trilinear couplings (A)u,d,e . Even though an
alignment of the soft terms with the Yukawa structure would be engineered at, e.g.,
m p by the supersymmetry-breaking mechanism, this feature is not preserved by the
renormalisation group evolution, so that the sfermion couplings at Msoft would still
be misaligned with the fermion couplings.
Flavour violation in the quark sector is tested with remarkable accuracy in low-
energy transitions such as meson oscillations — e.g. K − K̄, Bd − B̄d , Bs − B̄s — or
rare meson decays — e.g. K → πν ν̄ , B̄ → Xs γ , Bs → µ + µ − , B+ → τ + ντ . Flavour vi-
olation in the squark sector then typically contributes at the loop level, with squarks
and gauginos / higgsinos running in the loop. In particular, the neutral squark in-
teraction with gluinos may induce a flavour transition in association with the strong
coupling. If the corresponding flavour structure is completely unrelated to V CKM ,
experimental bounds on e.g. meson oscillation parameters constrain the flavour-
violating spectrum up to scales typically reaching beyond 100 TeV [100]. Slightly
milder limits — the strong coupling does not contribute at leading order — emerge
in the lepton sector from observables such as µ → eγ .
As a consequence of the tight pattern of flavour violation from an experimen-
tal perspective, globally consistent with a Standard Model interpretation, one must
either renounce new physics close to the TeV scale and set Msoft ≥ 100 TeV, or
assume that the flavour imprint of the Yukawa couplings also applies to the soft
supersymmetry-breaking terms at Msoft ≈ 1 TeV. This latter phenomenological hy-
pothesis is known as Minimal Flavour Violation — see [100] and references therein.
In its simplest application, the squark and slepton sectors are exactly aligned with
their fermionic counterparts, so that their mass matrices are block-diagonal in the
basis of SM flavours and V CKM is the only source of flavour violation in their
charged interactions; flavour transitions can no longer be induced by neutral media-
tors such as the gluino. Still, flavour observables continue to constrain new-physics
effects. The latter, now following the V CKM structure, are mediated by loops in-
volving charged Higgs and quarks, or charged gauginos / higgsinos and squarks,
resulting in limits on the masses of non-standard particles in the TeV range. Small
deviations from a strict alignment can of course be considered as well.
This phenomenological picture of alignment at low energy is not really satisfac-
tory from the perspective of model-building and calls for interpretations. The most
popular strategy consists in producing the soft supersymmetry-breaking terms via
the gauge-mediation mechanism [80], with comparatively light mediators (‘mes-
sengers’) at about 10 − 100 TeV: gauge interactions would then act in accordance
with the existing Yukawa structure in the observable sector and the low scale of
the mediation would forbid any significant deviation from alignment to develop
via running effects. The situation is more critical in gravity mediation, where the
generation of the soft terms involves physics at m p : a solution to the flavour prob-
lem then implies to ‘guess’ the physics behind the flavour structure. A possible
approach consists in regarding all flavour structures (Yukawas, soft terms) as spuri-
ons, i.e., as relics of fields (‘flavons’) governing the dynamics behind the breaking
28 Florian Domingo and Michel Rausch de Traubenberg

of the flavour-symmetry group. The prototype of such constructions is the Froggatt-


Nielsen model, using a horizontal U(1) symmetry [72, 88].
CP-violation raises an issue comparable to that of flavour to the MSSM. Even un-
der the assumption of a flavour-aligned sfermion structure, several phases of phys-
ical meaning (beyond that in V CKM ) are a priori allowed in the soft sector, in as-
sociation with the gaugino mass terms M1,2,3 and the trilinear couplings Au,d,e . On
the one hand, new sources of CP-violation are desirable in order to account for the
baryon-antibaryon asymmetry of the Universe. On the other, CP-violating phases
are tightly constrained by the absence of any experimental evidence for electric
dipole moments in nuclei and atoms [49]. Constructions similar to those addressing
the flavour problem can be designed [124].

4.6 Dark Matter phenomenology

A distinctive feature of R-parity conserving supersymmetry is the stability of the


lightest R-odd (i.e., ‘supersymmetric’) particle: indeed, as there exists no lighter
final state with R-parity −1 (by definition), this particle cannot decay if R-parity is
conserved. Then, according to the usual understanding of the thermal history of the
Universe, a stable particle coupled to the Standard Model would leave thermal relics
when it drops out of equilibrium in the cooling Universe (‘freeze out’) [82]. As such,
the lightest supersymmetric particle may contribute to the Dark Matter relic density.
Given strong astrophysical constraints on the abundance of charged particles, the
most promising candidates in the MSSM sector are the (neutral) scalar partners
of the neutrinos (sneutrinos) and the neutralinos, i.e., the neutral (and uncoloured)
gauginos and higgsinos, which mix after electroweak-symmetry breaking.
All the MSSM Dark-Matter candidates belonging to a non-trivial SU(2)L mul-
tiplet (winos, higgsinos, sneutrinos) tend to annihilate very efficiently in the early
Universe, so that they would leave negligible thermal relics unless their mass is in
the TeV range or above. On the contrary, a singlet fermion such as the bino — or the
singlino, i.e., the fermionic component of the singlet superfield, in the NMSSM —
tends to lead to excessive relics if left to itself. An excess in Dark Matter is a priori
more problematic than a shortage, as other sectors (beyond the MSSM) could be
invoked as complementary (or essential) sources of Dark Matter. However, several
mechanisms can be called upon to boost the annihilation cross-sections of singlet
fermions: sizable mixing with the other electroweakly charged neutralinos, resonant
annihilation in a Higgs (or Z) ‘funnel’, presence of a comparatively light t-channel
mediator, e.g. a slepton, existence of a heavier but almost degenerate R-odd particle,
still abundant at the moment of freeze out and helping in the depletion of R-odd
particles via ‘co-annihilation’ processes [83].
The wish to explain all of the observed Dark Matter with a comparatively light
MSSM candidate, as well as theoretical prejudice on the spectra, made the case of
a bino-dominated Dark matter, dependent on the previous mechanisms, a popular
one in the latest few decades. Nevertheless, one should be aware of the numerous
Supergravity: Application in Particle Physics 29

caveats behind this hypothesis. First, the lightest R-odd particle of the MSSM sector
need not be the lightest R-odd particle in absolute: obvious competitors would be a
lighter gravitino, typically in a gauge-mediation context, or any other exotic particle,
e.g. an axino, partner of an axion introduced to address the strong CP problem.
Then, small R-parity violating effects, negligible or not in collider physics, may
render the lightest supersymmetric particle unstable on cosmological scales. Finally,
non-thermal effects or as yet unknown shortcomings in the formulation of standard
cosmology may ruin the thermal picture. Below, we will forget about these warnings
and focus on the phenomenology of a weakly-interacting thermal explanation of
cold Dark Matter.
We have already copiously discussed the Dark Matter relic density: this quantity
is extracted from the measured properties of the Cosmic Microwave Background,
corresponding to the radiation trace of the epoch of atom formation (recombination).
Its evolution up to this day has been affected through gravitational interactions by
the matter / energy content of the Universe. Other evidence for Dark Matter, such
as the rotation curves of galaxies or gravitational lensing, can convince us of the
existence of Dark Matter in the current Universe, and more specifically in the Solar
System. Two main experimental strategies are pursued in the attempt at detecting
its by-definition faint interactions with Standard-Model matter. The first one, ‘Di-
rect Detection’, consists in searching for the recoil of heavy nuclei in elastic col-
lisions with Dark Matter particles. Such experiments steadily progress in covering
the available plane mass vs. cross-section down to the ‘neutrino floor’, at which the
competition of neutrinos (of Sun, Earth or Cosmic origin) shall raise a challenge for
further investigation. The alternative strategy, ‘Indirect Detection’, looks for Dark
Matter annihilation currently taking place in regions of space where Dark Matter is
expected to be dense, inside massive bodies or near the galactic centre. Typical sig-
nals would be energetic neutrinos or gamma-rays, detectable either in Earth-, air- or
space-based experiments. In both direct and indirect detection strategies, the identi-
fication of a signal involving Dark Matter necessarily depends on the astrophysical
modelisation of Dark Matter densities and fluxes in the investigated regions, so that
the consequences for particle physics might be tempered. To this day, no robust
discovery of a Dark Matter signal in direct or indirect detection has been reported.
Independently of whether the lightest R-odd particle of the MSSM represents a
sizable component of Dark Matter, one can also attempt to produce it at colliders
from Standard Model matter, which leads us to the collider phenomenology. A more
detailed overview of supersymmetric Dark Matter may be found in e.g. [21].

4.7 Collider phenomenology

The traditional investigation path in experimental particle physics consists in accel-


erating electrons and / or protons to very high energies before colliding them: the
high kinetic energy of the projectiles may then convert into interactions involving
very massive particles, thus providing access to physical effects beyond electromag-
30 Florian Domingo and Michel Rausch de Traubenberg

netism and nuclear forces. In this fashion, new physics resonances could be directly
produced, provided the kinematical threshold is reached and the cross-section ×
integrated luminosity is sufficiently large to make such events probable enough in
collisions, and detectable with the available experimental sensitivity. An alternative
strategy, which we mentioned in previous subsections, consists in precisely mea-
suring the properties of known particle to attempt and detect effects ascribable to
physics beyond the Standard Model. We here focus on the direct production ap-
proach.
Under the assumption of R-parity conservation, one expects the supersymmetric
particles to be produced in pairs in collisions of standard matter. The resonances
may then either decay to a lighter R-odd particle through radiation of standard parti-
cles, or escape the detector if they are stable or long-lived. For ‘usual’ MSSM spec-
tra, the strongly or electroweakly interacting R-odd particles decay promptly, via
a cascade, down to the lightest supersymmetric particle. In some cases, however,
e.g. if the decay has very little available phase space, or if couplings to the light-
est supersymmetric particle are feeble (or suppressed by the recourse to e.g. very
massive mediators), a R-odd particle may be long-lived at the scales of the collider
experiment (where large boost factors can substantially lengthen the apparent life-
time). The most studied scenario is that of a prompt decay into a stable and massive
neutralino, which escapes the experiment without being detected, similarly to a neu-
trino, resulting in a sizable drain of energy and momentum in the apparent balance
of the transition. This ‘missing energy’ signature underlies the classical strategy for
searches of supersymmetric particles at colliders.
The production of supersymmetric pairs typically occurs through s-channel ex-
change of a gauge boson or t- and u-channel exchanges of a R-odd particle. At
lepton colliders, the production process is electroweak, while it can be strong or
electroweak at hadron colliders. The expected final states typically involve missing
energy, plus jets (strongly interacting matter), plus leptons. A skillful prescription
of cuts is usually necessary to distinguish such final states from the Standard Model
background, for instance QCD processes at a hadron collider, with radiated W ’s and
Z’s producing the leptons and missing energy. The absence of discovery at the Large
Electron-Positron collider placed limits on the masses of the supersymmetric parti-
cles in the 100 GeV range. Similarly, the Large Hadron Collider has been excluding
vast areas of the parameter space available to R-odd particles, especially those that
are produced in strong processes (squarks, gluinos): a typical lower bound on their
mass would be one to a few TeV. For uncoloured particles (sleptons, gauginos, hig-
gsinos), the smaller (electroweak) production cross-sections result in milder limits,
not exceeding a few 100 GeV. Final states with multiple leptons (plus missing en-
ergy) are the usual targets in this case. Of course, such experimental constraints are
almost never generic, but generally apply to a specific type of spectra, so that their
transposition to different scenarios usually requires the recourse to extrapolations or
estimates (‘recast’).
More exotic signatures are also looked for. For instance, long-lived particles de-
caying on length scales comparable to the detector size may produce leptons or jets
with a point of origin distinct from the interaction point (‘displaced leptons’ and
Supergravity: Application in Particle Physics 31

‘displaced vertices’). A long-lived (or stable) charged particle could leave identifi-
able tracks. For larger lifetimes, the deployment of detectors placed at a few 10−100
meters from the interaction points has been planned at the Large Hadron Collider.
However, in the absence of distinctive characteristics such as a large missing en-
ergy or the existence of long-lived heavy particles, the identification of new res-
onances at high energy colliders might prove difficult in general; such a scenario
would nevertheless be regarded as highly exotic. Finally, reconstructing the nature
of the hypothetically discovered new fields and their classification in a supersym-
metric spectrum would call for a long-term, as yet unforeseeable, effort.
The extended Higgs sector of supersymmetric extensions of the Standard Model
offers another direction for investigations. The heavy doublet states H 0 , A0 and H ±
are R-even, hence need not be produced in pairs. Typical production modes at the
Large Hadron Collider are comparable to those of the Standard Model Higgs boson
and involve gluon-gluon fusion, (electroweak) vector boson fusion, assisted produc-
tion with bottom and / or top quarks. At an e+ e− collider, these heavy Higgs bosons
would have to be produced in pairs or in association with an h0 , from a Z-boson ex-
change in the s-channel. The dominant decay channels involve fermion pairs, with
typical searches in τ + τ − , µ + µ − , bb̄ or t t¯ (τ ν̄τ or bt¯ for H − ). Bosonic decay chan-
nels are suppressed in the MSSM, although H 0 → h0 h0 may be detectable, provided
H 0 is light enough (though with mass above threshold). In the NMSSM, however,
Higgs-to-Higgs cascade decays involving singlet states may be dominant and over-
shadow the fermionic modes: these processes are also actively looked for.
A short summary of the search for supersymmetry at the Large Hadron Collider
can be found in [19].

4.8 Supersymmetric Grand Unification

The MSSM or its singlet extension (NMSSM) are the simplest models fulfilling a
softly-broken supersymmetric embedding of the Standard Model. However, about
any idea in non-supersymmetric particle physics, e.g. axions, vector quarks, right-
handed neutrinos, can be transposed to the supersymmetry / supergravity-breaking
framework of section 3. Here we shall provide a few insights concerning Grand
Unification in this context. Two coincidences of the Standard Model (or the MSSM)
particularly motivate the idea of a unification of gauge interactions within a less
disparate group. The first one is the cancellation of the chiral anomalies, acciden-
tally resulting from the matter content. The second one is the quantisation of the
electric charge, i.e., the fact that all elementary particles take charges that are mul-
tiple of 1/3: the hypercharge being a U(1) symmetry, there is no deep reason why
the various hypercharges in the Standard Model should appear in rational propor-
tions. Reciprocally, the protection of radiative corrections by supersymmetry bene-
fits grand unified constructions through the stabilisation of scale hierarchies and the
reduced risk of encountering a non-perturbative regime (Landau pole). Finally, we
emphasise that the apparent convergence of the Standard Model gauge couplings
32 Florian Domingo and Michel Rausch de Traubenberg

at MGUT ≈ 1016 GeV in the MSSM favours a one-step unification (SU(5), SO(10),
etc.), with a ‘Grand Desert’ between the electroweak and Grand Unification scales,
rather than the multistep path of left-right symmetry and Pati-Salam. For reviews of
the group theoretical concepts involved in Grand Unification, we refer the reader to
(e.g.) [137, 130, 159, 18], and to [116, 115, 128] concerning their application in a
supersymmetric context.
The simplest embedding of GSM in a compact simple Lie group employs an
SU(5) gauge symmetry, thus involving no reduction of rank in spontaneous sym-
metry breaking from SU(5) to GSM . One may view the first three rows of the funda-
mental SU(5) representation as transforming under the fundamental representation
of SU(3)c and the two last, q under the fundamental representation of SU(2)L . Fi-
24 3 1 1 1 1 1

nally, the generator t = 5 diag − 3 , − 3 , − 3 , 2 , 2 commutes with all SU(3)c
and SU(2)L generators and can be identified, up to a proportionality constant, to
Y 24
the hypercharge:q Y = C24 t . The conventional normalisation of the hypercharge
Y = 5
leads to C24 3 (because all generators in the fundamental representation satisfy
Tr(ta tb ) δ ab
= /2). This connection explains the quantisation of the electric charge.
Then, studying the branching rules of SU(5) ⊃ GSM , one observes:

5 = (3̄, 1, 31 ) ⊕ (1, 2, − 12 ) , 5 = (3, 1, − 13 ) ⊕ (1, 2, 21 ) ,


e ee ee e ee ee
10 = (3̄, 1, − 23 ) ⊕ (1, 1, 1) ⊕ (3, 2, 16 ) , (36)
f ee ee ee
24 = (8, 1, 0) ⊕ (1, 3, 0) ⊕ (1, 1, 0) ⊕ (3, 2, − 56 ) ⊕ (3̄, 2, 56 ) .
ee ee ee ee ee
Consequently, the chiral superfields containing the Standard model fermions can be
f f
embedded in (three generations of) a 5F ⊕ 10F representation of SU(5) (see (28)),
e f
while HU,D are identified with the doublet components of 5S ⊕ 5S (see (29)). In the
adjoint representation (24) of SU(5), one finds an embedding e e for all the Standard
f
Model gauge (super)fields, with the additional (3, 2, − 6 ) ⊕ (3̄, 2, 65 ) called lepto-
5

quarks. The latter take mass at the scale of the breaking e e SU(5)e→ e G and convey
SM
new physics effects, such as lepton- and baryon-number violation. These should re-
main small at the electroweak scale for a phenomenologically realistic model, so that
the breaking scale needs satisfy MGUT ≫ MZ . In fact MGUT is most naturally chosen
as the scale at which the standard gauge couplings converge, i.e., MGUT ≈ 1016 GeV
for a MSSM field content with Msoft ≈ 1 TeV. The field content described above
does not allow the breaking SU(5) → GSM (i.e., would break GSM simultaneously
with SU(5)), so that, as generic in grand unified models, further fields belonging
to representations of larger dimensions are needed: the simplest choice consists in
introducing 24S in the adjoint representation of SU(5) and taking a vev proportional
f
to t24 . This field content is again (accidentally) anomaly-free.
We may now write the superpotential for the SU(5) model thus designed:
Supergravity: Application in Particle Physics 33

m λ
WSU(5) = Tr(24S )2 + Tr(24S )3 + µ (5S )i (5S )i + β (5S )i (24S )i j (5S ) j
2 f 3 f e e e f e
+ Y5 (5F )i (10F )i j (5S ) j + Y10ε i jkℓm (10F )i j (10F )kℓ (5S )m (37)
e f e f f e
where the indices i, . . . , m correspond to the transformation under the (anti)funda-
mental SU(5) — the generation indices are omitted —and ε i jkℓm is the five-
dimensional Levy-Civita symbol. The matter-parity, i.e., invariance under 5F , 10F →
e f
−5F , −10F , has been implicitly required. The SU(5)-breaking minimum < 24S >=
me f
λ diag(2, 2, 2, −3, −3) does not break supersymmetry and is therefore degener-
ate with the SU(5)-conserving one as long as supersymmetry-breaking terms are
not introduced. The symmetry breaking GSM → SU(3)c × U(1)em is achieved by
the doublet vevs of 5S ⊕ 5S . The two last terms in the first line of Eq.(37) gen-
e e
erate an effective µ -term µeff = µ − 3mλ β for HU,D (embedded in 5S ⊕ 5S ), while
e e
the supersymmetric mass associated to the colour-triplets H3 , H 3 of 5S ⊕ 5S is
e e
µ + 2mλ β . From the phenomenological perspective, it is necessary to impose the
fine-tuning |µeff | ≪ |µ |, | 3mλ β |, making the doublets light and the colour-triplets su-
perheavy: while this requirement is non-natural, it remains technically natural once
set, due to the protection by supersymmetry — contrarily to the situation in non-
supersymmetric SU(5). Further model-building ingredients can also naturally pro-
tect the doublet mass term.
The terms in the second line of Eq.(37) are at the origin of the Yukawa cou-
plings in the MSSM. Explicit decomposition immediately provides that Y5 gener-
ates Yukawa couplings for the down-type quarks and leptons while Y10 gives the up-
type Yukawa. For the third generation, the resulting unification of down and lepton
masses at the GUT scale (mGUT b = mGUT
τ ) roughly yields mb ∼ 3mτ at low energy,
which is in acceptable agreement with the measurements. This does not work for the
lighter generations. Noting however that 5 ⊗ 10 = 5 ⊕ 45 and 10 ⊗ 10 = 5 ⊕ 45 ⊕ 50,
e f e f f f e f f
one can introduce additional ‘Higgs’ superfields in a 45S ⊕45S representation, which
allow for further terms of Yukawa type and containf g accepting a vev. The
doublets
Clebsch-Gordan coefficients at MGUT then produce mµ = 3mGUT GUT
s (for the second
generation; in the absence of contributions from the 5S ⊕ 5S terms), resulting in the
e e
phenomenologically acceptable mµ ∼ ms . Thus, models involving both 5S ⊕ 5S and
e e
45S ⊕ 45S are potentially viable explanations of the unification of lepton and down-
f f
type masses. Similarly, the SU(5) symmetry implies the unification (at MGUT ) of
numerous soft terms, such as the quadratic terms for all fields belonging to the
same representation. As long as supersymmetry-breaking is communicated to the
observable sector via an SU(5) singlet, the gaugino masses also unify (at MGUT )
M1GUT = M2GUT = M3GUT , leading to the popular hierarchy M3 ≈ 3M2 ≈ 6M1 at low
energy.
Beyond SU(5), the immediate benefit of considering unification groups of higher
rank, such as SO(10) or E6 , rests with the automatic cancellation of chiral anoma-
lies. The branching rules of SO(10) ⊃ SU(5) × U(1):
34 Florian Domingo and Michel Rausch de Traubenberg

16 = (5, −3) ⊕ (10, 1) ⊕ (1, −5) , 10 = (5, 2) ⊕ (5, −2) ,


f e f e f e e
45 = (24, 0) ⊕ (10, 4) ⊕ (10, −4) ⊕ (1, 0) (38)
f f f f e
120 = (5, 2) ⊕ (5, −2) ⊕ (10, −6) ⊕ (45, 2) ⊕ (45, −2)
g e e f f f
126 = (1, −10) ⊕ (5, −2) ⊕ (10, −6) ⊕ (15, 6) ⊕ (45, 2) ⊕ (50, −2)
g e e f f f f
suggest an embedding of all quark and lepton superfields within three generations
of 16Ff , also containing SU(5) singlets amounting to right-handed neutrinos. Then,
f
observing that 16 ⊗ 16 = 10 ⊕ 120 ⊕ 126, one can write Yukawa couplings employ-
f f f g g
ing Higgs fields in the real 10S , 120S or the complex 126S ⊕ 126S representations.
This leaves ample maneuvering g to accommodateg
f space g masses, which
the fermion
a single Higgs representation would fail to explain due to phenomenologically un-
physical Yukawa unifications. In particular, in order to generate a ‘Majorana mass
term’ for the right-handed neutrinos and allow for a Type-I seesaw, the residual U(1)
symmetry needs to be broken. This can be achieved with a 16S or a 126S vev, with
the difference that the second choice preserves a residual Zf g
2 -symmetry, which can
be understood as R-parity. The 5S ⊕ 5S Higgs fields of SU(5) are embedded within a
e e SU(5) can be broken with the 24 contained
10S representation of SO(10). Finally, S
f
within the 45S representation of SO(10) (or a 54 or a 210). Soft-breakingf terms a
priori unifyf f since
more completely at MGUT than in SU(5), g all MSSM fields are col-
lected within 16Ff and 10S . An even more ambitious unification pattern is possible
f f f
in E6 , with both 16F and 10S of SO(10) collected within a 27 of E6 , which we will
not discuss here.f f f
Matter stability is usually presented as the weak point of grand unified theories.
Indeed, the interactions mediated by the (3, 2, − 56 ) ⊕ (3̄, 2, 56 ) gauge bosons con-
tribute to effective dimension 6 baryon- and e elepton-number
e e violating low-energy
operators, which can mediate proton decay, via channels such as p → e+ π 0 . The
4
typical associated lifetime scales like τ p ∼ MGUT /(αU2 MP5 ), where MP is the proton
2
mass and αU ≡ gU /(4π ) with gU the gauge coupling at the unification scale. For
a non-supersymmetric Grand Unification, the predicted lifetime tends to be incom-
patible with the corresponding experimental limit. In a supersymmetric context, the
larger MGUT apparently provides some breathing space. However, the existence of
comparatively light scalar partners to the standard fermions alters the validity of the
naive analysis, since it allows the emergence at Msoft of baryon- and lepton-number
operators of dimension 5 involving such fields, which in turn, below Msoft , alter the
−2
MGUT suppression of the contributions to dimension 6 operators mediating proton
decay. Due to the structure of these operators, the proton decay modes that they
funnel typically involve matter of higher-generation content, such as kaons, muons,
muon and τ neutrinos. In practice, a loop suppression factor alleviates the contribu-
tions via such intermediate operators of dimension 5, so that these do not necessarily
ruin the viability of supersymmetric Grand Unification: a detailed analysis is needed
in each case. In fact, purely scalar dimension 4 operators at Msoft can also mediate
proton decay, but contribute only at two-loop. In all this discussion, we have as-
Supergravity: Application in Particle Physics 35

sumed the existence of a matter parity, which constrains baryon- and lepton-number
violating effects to appear in operators of higher dimensions (or involving scalar
partners). The pattern of violation is modified if R-parity is not satisfied at Msoft ,
which we consider in the following section.

4.9 R-parity-violating phenomenology

So far, we have forbidden the terms in the second line of Eq.(30), and their soft-
breaking counterparts, by requesting the model to satisfy R-parity. The original mo-
tivation behind this choice was to forbid the emergence of baryon- (in the last term
of this equation: ‘UDD’) or lepton-number violating effects (in the three previous
terms: ‘LH’, ‘LLE’ and ‘LQD’). As we noticed in the discussion concerning Grand
Unification, however, R-parity does not forbid the emergence of non-renormalisable
operators violating baryon- and lepton-number, as one would expect if the MSSM
is only a low-energy effective field theory of some higher-energy dynamics; thus
it does not fully protects matter stability. We may therefore relax the assumption
of R-parity conservation altogether, which results in significant alterations of the
phenomenology. We refer the reader to Ref.[10].
Baryon-number violating processes, such as nucleon decay, obviously represent
the first challenge to R-parity violation, as new physics at Msoft potentially con-
tributes to such phenomena. Here, we stress that proton decay usually requires both
lepton- and baryon-number violation, because leptons (electron, muon, neutrinos)
are the only available fermions below the proton mass (in the absence of exotic par-
ticles) and that the angular momentum should be preserved in the transition. New
discrete symmetries, such as baryon triality, lepton triality or proton hexality, can
be invoked to forbid nucleon decays. Pure baryon-violating processes such as di-
nucleon decay or neutron-antineutron oscillations place further limits on couplings
of UDD-type, but they usually depend on the pattern of flavour violation as well,
due to the antisymmetry of the baryon-number violating UDD operators with re-
spect to down-quark generation indices. In the presence of a light exotic state, usu-
ally a bino, below the proton mass, proton decay can occur without lepton-number
violation. The baryon-number violating couplings thus appear as more immediately
constrained and are often explicitly eliminated by the call to a symmetry.
As we implied above, the bino state may be light if it is unstable, which is the
case if R-parity is violated: cosmological and astrophysical bounds are indeed no
longer applicable. If the bino is long-lived, its decays might be observed in far de-
tectors, as currently deployed at the Large Hadron Collider. More generally, collider
phenomenology may significantly change in the presence of R-parity violating ef-
fects, as supersymmetric particles can now be produced as single resonances while
the missing energy signature is no longer guaranteed. Searches with single leptons,
jets and little missing energy may then prove viable alternative strategies. The light-
est supersymmetric particle may still be a Dark Matter candidate if it is sufficiently
long-lived to persist on cosmological timescales. In such a case, it needs to be neu-
36 Florian Domingo and Michel Rausch de Traubenberg

tral and satisfy conditions relative to its abundance at the time of recombination.
Alternatively, a charged or coloured particle with short lifetime is eligible as lightest
supersymmetric particle for collider physics, without implications for cosmology.
Given that R-parity is no longer present to distinguish between standard fields
and their supersymmetric partners, the latter usually mix, such as leptons with
uncoloured gauginos and higgsinos, or sleptons with Higgs bosons. This mixing
should generally remain subleading so that standard particles retain their usual prop-
erties. It also entails some degree of ambiguity in the definition of the superfields:
a convenient, though not mandatory choice consists in requesting that the sneutrino
fields do not take vevs. In addition, the emergence of neutrino masses is an inter-
esting consequence of the mixing of neutrinos with neutral gauginos and higgsinos.
At tree level, only one mass, scaling like (µi MZ /µ )2 /M2 develops; comparison with
observed limits place bounds on |µi /µ |, |µi /M2 | of the order 10−5 . Loop corrections
involving couplings of LLE or LQD type produce a second neutrino mass, opening
the possibility for an explanation of neutrino oscillations in terms of pure R-parity
violating effects.
The a priori independent flavour- and CP-violating pattern of the trilinear R-
parity violating couplings can be constrained in low-energy observables. Such cou-
plings indeed contribute at tree-level (typically via a sfermion exchange) to pre-
cisely known observables, such as lepton and meson decays or particle-antiparticle
oscillations, quark- or lepton-flavour transitions, the unitarity triangle, the anoma-
lous magnetic dipole moments of leptons, the electric dipole moments of leptons
and atoms, neutrinoless double beta decay, etc. A comprehensive coverage of these
bounds is usually performed under simplifying assumptions, such as the dominance
of a single R-parity violating coupling, or a pair of couplings, or in numerical form,
as the diversity of available terms is otherwise difficult to efficiently tackle.
From the perspective of Grand Unification, we have mentioned how the conser-
vation / violation of R-parity could be related to SO(10) breaking. In SU(5), the
R-parity violating superpotential summarizes to:

RpV f 1 SU(5) f
WSU(5) = κ f (5F )i (5S )i + λ f gh (5F )i (5gF ) j (10hF )i j . (39)
e e 2 e e f
In this case, trilinear couplings are unified and the limits from proton decay makes
such couplings negligible at the electroweak scale. The coupling constants κi are
also phenomenologically very small due to their implications for neutrino physics.
As a summary of this phenomenological overview of supersymmetry / supergrav-
ity in particle physics, we have seen that very diverse types of effects could emerge
within such theories, due to the rich field content and the a priori uncontrolled pat-
tern of couplings inherited from their construction. Such models thus raise consider-
able challenge to model building for ‘natural’ explanations of hierarchies, although
these are now stabilised by supersymmetry (i.e., technically natural). The most sim-
plistic model-building assumptions are usually incompatible with the experimental
situation, leading to an increase in complexity in the field content and the parame-
ter space, where one had originally wished a predictive and symmetry-constrained
Supergravity: Application in Particle Physics 37

structure. Conversely, this diversity makes supersymmetry-inspired models a good


laboratory to study and motivate non-standard effects in particle physics.

References

[1] Akulov V, Volkov D, Soroka V (1975) Gauge fields on superspaces with dif-
ferent holonomy groups. JETP Lett 22:187–188
[2] Alvarez-Gaumé L, Claudson M, Wise MB (1982) Low-energy supersymme-
try. Nucl Phys B207:96, DOI 10.1016/0550-3213(82)90138-9
[3] Alvarez-Gaumé L, Polchinski J, Wise MB (1983) Minimal low-energy su-
pergravity. Nucl Phys B221:495, DOI 10.1016/0550-3213(83)90591-6
[4] Amaldi U, de Boer W, Furstenau H (1991) Comparison of grand unified the-
ories with electroweak and strong coupling constants measured at LEP. Phys
Lett B 260:447–455, DOI 10.1016/0370-2693(91)91641-8
[5] Arkani-Hamed N, Giudice GF, Luty MA, Rattazzi R (1998) Supersymme-
try-breaking loops from analytic continuation into superspace. Phys Rev
D58:115005, DOI 10.1103/PhysRevD.58.115005, hep-ph/9803290
[6] Arnowitt RL, Nath P, Zumino B (1975) Superfield densities and action prin-
ciple in curved superspace. Phys Lett B56:81, DOI 10.1016/0370-2693(75)
90504-3
[7] Bagger J, Witten E (1982) The gauge invariant supersymmetric nonlinear
sigma model. PhysLett B118:103–106, DOI 10.1016/0370-2693(82)90609-8
[8] Bagger JA (1983) Coupling the gauge-invariant supersymmetric non-
linear sigma model to supergravity. NuclPhys B211:302, DOI 10.1016/
0550-3213(83)90411-X
[9] Bagger JA, Moroi T, Poppitz E (2000) Anomaly mediation in super-
gravity theories. JHEP 0004:009, DOI 10.1088/1126-6708/2000/04/009,
hep-th/9911029
[10] Barbier R, et al (2005) R-parity violating supersymmetry. Phys Rept 420:1–
202, DOI 10.1016/j.physrep.2005.08.006, hep-ph/0406039
[11] Barbieri R, Ferrara S, Savoy CA (1982) Gauge models with sponta-
neously broken local supersymmetry. Phys Lett B119:343, DOI 10.1016/
0370-2693(82)90685-2
[12] Bechtle P, et al (2015) Killing the cMSSM softly. PoS EPS-HEP2015:139
[13] Becker K, Becker M, Schwarz J (2007) String Theory and M-Theory:
A Modern Introduction. Cambridge: Cambridge University Press., DOI
10.1017/CBO9780511816086
[14] Binetruy P, Girardi G, Grimm R (2001) Supergravity couplings: a geo-
metric formulation. PhysRept 343:255–462, DOI 10.1016/S0370-1573(00)
00085-5, hep-th/0005225
[15] Brignole A, Ibanez LE, Munoz C (1998) Soft supersymmetry breaking terms
from supergravity and superstring models. Adv Ser Direct High Energy Phys
18:125–148, DOI 10.1142/9789812839657 0003, hep-ph/9707209
38 Florian Domingo and Michel Rausch de Traubenberg

[16] Brink L, Gell-Mann M, Ramond P, Schwarz JH (1978) Supergravity as geom-


etry of superspace. PhysLett B74:336, DOI 10.1016/0370-2693(78)90671-8
[17] Buchbinder I, Kuzenko S (1998) Ideas and Methods of Supersymmetry
and Supergravity: Or a Walk Through Superspace. Studies in High Energy
Physics, Cosmology and Gravitation. Bristol: Institute of Physics (IOP).
[18] Campoamor-Stursberg R, Rausch de Traubenberg M (2019) Group Theory
in Physics. WSP, DOI 10.1142/11081
[19] Canepa A (2019) Searches for Supersymmetry at the Large Hadron Collider.
Rev Phys 4:100033, DOI 10.1016/j.revip.2019.100033
[20] Casas JA, Muñoz C (1993) A Natural solution to the mu problem. Phys Lett
B 306:288–294, DOI 10.1016/0370-2693(93)90081-R, hep-ph/9302227
[21] Catena R, Covi L (2014) SUSY dark matter(s). Eur Phys J C 74:2703, DOI
10.1140/epjc/s10052-013-2703-4, 1310.4776
[22] Chamseddine AH, Arnowitt RL, Nath P (1982) Locally supersymmetric
grand unification. Phys Rev Lett 49:970, DOI 10.1103/PhysRevLett.49.970
[23] Coleman SR, Mandula J (1967) All possible symmetries of the S matrix.
PhysRev 159:1251–1256, DOI 10.1103/PhysRev.159.1251
[24] Cremmer E, Julia B, Scherk J, Ferrara S, Girardello L, van Nieuwenhuizen
P (1978) Super-Higgs effect in supergravity with general scalar interactions.
Phys Lett B79:231, DOI 10.1016/0370-2693(78)90230-7
[25] Cremmer E, Julia B, Scherk J, Ferrara S, Girardello L, van Nieuwen-
huizen P (1979) Spontaneous symmetry breaking and Higgs effect in super-
gravity without cosmological constant. Nucl Phys B147:105, DOI 10.1016/
0550-3213(79)90417-6
[26] Cremmer E, Ferrara S, Girardello L, Van Proeyen A (1982) Coupling super-
symmetric Yang-Mills theories to supergravity. Phys Lett B116:231, DOI
10.1016/0370-2693(82)90332-X
[27] Cremmer E, Ferrara S, Girardello L, Van Proeyen A (1983) Yang-Mills the-
ories with local supersymmetry: Lagrangian, transformation laws and super-
Higgs effect. Nucl Phys B212:413, DOI 10.1016/0550-3213(83)90679-X
[28] Cremmer E, Ferrara S, Kounnas C, Nanopoulos DV (1983) Naturally van-
ishing cosmological constant in N=1 supergravity. PhysLett B133:61, DOI
10.1016/0370-2693(83)90106-5
[29] Dall’Agata G, Zagermann M (2022) In this volume
[30] Derendinger JP, Savoy CA (1982) Gaugino masses and a new mechanism for
proton decay in supersymmetric theories. Phys Lett B118:347, DOI 10.1016/
0370-2693(82)90201-5
[31] Deser S, Zumino B (1976) Consistent supergravity. Phys Lett B62:335, DOI
10.1016/0370-2693(76)90089-7
[32] Dimopoulos S, Georgi H (1981) Softly Broken Supersymmetry and SU(5).
Nucl Phys B 193:150–162, DOI 10.1016/0550-3213(81)90522-8
[33] Dimopoulos S, Raby S (1981) Supercolor. Nucl Phys B192:353, DOI 10.
1016/0550-3213(81)90430-2
Supergravity: Application in Particle Physics 39

[34] Dine M, Fischler W (1982) A phenomenological model of particle physics


based on supersymmetry. Phys Lett B110:227, DOI 10.1016/0370-2693(82)
91241-2
[35] Dine M, Nelson AE (1993) Dynamical supersymmetry breaking at
low energies. Phys Rev D48:1277, DOI 10.1103/PhysRevD.48.1277,
hep-ph/9303230
[36] Dine M, Fischler W, Srednicki M (1981) Supersymmetric technicolor. Nucl
Phys B189:575, DOI 10.1016/0550-3213(81)90582-4
[37] Dine M, Nelson AE, Shirman Y (1995) Low energy dynamical supersym-
metry breaking simplified. Phys Rev D51:1362, DOI 10.1103/PhysRevD.51.
1362, hep-ph/9408384
[38] Dine M, Nelson AE, Nir Y, Shirman Y (1996) New tools for low energy
dynamical supersymmetry breaking. Phys Rev D53:2658, DOI 10.1103/
PhysRevD.53.2658, hep-ph/9507378
[39] Dixon LJ (2015) A Brief Introduction to Modern Amplitude Methods. In:
Theoretical Advanced Study Institute in Elementary Particle Physics: Jour-
neys Through the Precision Frontier: Amplitudes for Colliders, pp 39–97,
DOI 10.1142/9789814678766 0002
[40] Djouadi A (2008) The Anatomy of electro-weak symmetry breaking. II. The
Higgs bosons in the minimal supersymmetric model. Phys Rept 459:1–241,
DOI 10.1016/j.physrep.2007.10.005, hep-ph/0503173
[41] Dragon N (1979) Torsion and curvature in extended supergravity. Z Phys
C2:29–32, DOI 10.1007/BF01546233
[42] Dragon N, Ellwanger U, Schmidt MG (1987) Supersymmetry and supergrav-
ity. ProgPartNuclPhys 18:1, DOI 10.1016/0146-6410(87)90008-1
[43] Drees M, Godbole R, Roy P (2004) Theory and phenomenology of sparticles:
An account of four-dimensional N=1 supersymmetry in high energy physics.
World Scientific Publishing Co Pte Ltd
[44] Dreiner HK, Haber HE, Martin SP (2022) From Spinors to Supersymmetry.
Cambridge Monographs on Particle Physics, Nuclear Physics & Cosmology,
Cambridge University Press
[45] Ducrocq R, Rausch de Traubenberg M, Valenzuela M (2021) A pedagogical
discussion of N = 1 four-dimensional supergravity in superspace. Mod Phys
Lett A 36(16):2130015, DOI 10.1142/S0217732321300159, 2104.06671
[46] Ellis JR, Nanopoulos DV, Tamvakis K (1983) Grand unification in simple
supergravity. Phys Lett B121:123, DOI 10.1016/0370-2693(83)90900-0
[47] Ellis JR, Kounnas C, Nanopoulos DV (1984) No-scale supersymmetric
GUTs. NuclPhys B247:373–395, DOI 10.1016/0550-3213(84)90555-8
[48] Ellis JR, Lahanas A, Nanopoulos DV, Tamvakis K (1984) No-scale super-
symmetric standard model. PhysLett B134:429, DOI 10.1016/0370-2693(84)
91378-9
[49] Ellis JR, Lee JS, Pilaftsis A (2008) Electric Dipole Moments in the
MSSM Reloaded. JHEP 10:049, DOI 10.1088/1126-6708/2008/10/049,
0808.1819
40 Florian Domingo and Michel Rausch de Traubenberg

[50] Ellwanger U, Rausch de Traubenberg M, Savoy CA (1997) Phenomenol-


ogy of supersymmetric models with a singlet. Nucl Phys B492:21–50,
DOI 10.1016/S0550-3213(97)80026-0,10.1016/S0550-3213(97)00128-4,
hep-ph/9611251
[51] Ellwanger U, Hugonie C, Teixeira AM (2010) The Next-to-Minimal super-
symmetric Standard Model. Phys Rept 496:1–77, DOI 10.1016/j.physrep.
2010.07.001, 0910.1785
[52] Elvang H, Huang Yt (2013) Scattering Amplitudes 1308.1697
[53] Farrar GR, Fayet P (1978) Phenomenology of the Production, Decay, and De-
tection of New Hadronic States Associated with Supersymmetry. Phys Lett
B 76:575–579, DOI 10.1016/0370-2693(78)90858-4
[54] Fayet P (1975) Supergauge invariant extension of the Higgs mechanism and
a model for the electron and its neutrino. Nucl Phys B90:104, DOI 10.1016/
0550-3213(75)90636-7
[55] Fayet P (1976) Supersymmetry and weak, electromagnetic and strong inter-
actions. PhysLett B64:159, DOI 10.1016/0370-2693(76)90319-1
[56] Fayet P (1977) Spontaneously broken supersymmetric theories of weak,
electromagnetic and strong interactions. PhysLett B69:489, DOI 10.1016/
0370-2693(77)90852-8
[57] Fayet P (1978) Massive gluinos. Phys Lett B78:417, DOI 10.1016/
0370-2693(78)90474-4
[58] Fayet P (1984) Supersymmetric theories of particles and interactions. Phys-
Rept 105:21, DOI 10.1016/0370-1573(84)90113-3
[59] Fayet P, Iliopoulos J (1974) Spontaneously broken supergauge symmetries
and goldstone spinors. Phys Lett B51:461, DOI 10.1016/0370-2693(74)
90310-4
[60] Ferrara S, van Nieuwenhuizen P (1978) The auxiliary fields of supergravity.
PhysLett B74:333, DOI 10.1016/0370-2693(78)90670-6
[61] Ferrara S, van Nieuwenhuizen P (1978) Tensor calculus for supergravity.
PhysLett B76:404, DOI 10.1016/0370-2693(78)90893-6
[62] Ferrara S, Wess J, Zumino B (1974) Supergauge multiplets and superfields.
Phys Lett B51:239, DOI 10.1016/0370-2693(74)90283-4
[63] Ferrara S, Gliozzi F, Scherk J, Van Nieuwenhuizen P (1976) Matter couplings
in supergravity theory. NuclPhys B117:333, DOI 10.1016/0550-3213(76)
90401-6
[64] Ferrara S, Scherk J, van Nieuwenhuizen P (1976) Locally supersymmetric
Maxwell-Einstein theory. PhysRevLett 37:1035, DOI 10.1103/PhysRevLett.
37.1035
[65] Ferrara S, Freedman D, van Nieuwenhuizen P, Breitenlohner P, Gliozzi F,
Scherk J (1977) Scalar multiplet coupled to supergravity. PhysRev D15:1013,
DOI 10.1103/PhysRevD.15.1013
[66] Ferrara S, Kaku M, Townsend P, van Nieuwenhuizen P (1977) Unified field
theories with U(N) internal symmetries: Gauging the superconformal group.
NuclPhys B129:125, DOI 10.1016/0550-3213(77)90023-2
Supergravity: Application in Particle Physics 41

[67] Ferrara S, Grisaru MT, van Nieuwenhuizen P (1978) Poincaré and confor-
mal supergravity models with closed algebras. NuclPhys B138:430, DOI
10.1016/0550-3213(78)90389-9
[68] Fradkin E, Tseytlin AA (1985) Conformal supergravity. PhysRept 119:233–
362, DOI 10.1016/0370-1573(85)90138-3
[69] Freedman DZ, Van Proeyen A (2012) Supergravity. Cambridge Univ. Press,
DOI 10.1017/CBO9781139026833
[70] Freedman DZ, van Nieuwenhuizen P, Ferrara S (1976) Progress toward a
theory of supergravity. Phys Rev D13:3214–3218, DOI 10.1103/PhysRevD.
13.3214
[71] Freund PG (1986) Introduction to Supersymmetry. Cambridge University
Press
[72] Froggatt CD, Nielsen HB (1979) Hierarchy of Quark Masses, Cabibbo
Angles and CP Violation. Nucl Phys B 147:277–298, DOI 10.1016/
0550-3213(79)90316-X
[73] Fuks B, Rausch de Traubenberg M (2011) Super-
symétrie : exercices avec solutions. Ellipses, URL
http://editions-ellipses.fr/supersymetrie-exercices-avec-solutions-p-7697.h
[74] Gates J S J, Stelle K, West PC (1980) Algebraic origins of superspace con-
straints in supergravity. NuclPhys B169:347, DOI 10.1016/0550-3213(80)
90037-1
[75] Gates S, Grisaru MT, Rocek M, Siegel W (1983) Superspace, or one thousand
and one lessons in supersymmetry. FrontPhys 58:1, hep-th/0108200
[76] Girardello L, Grisaru MT (1982) Soft breaking of supersymmetry. NuclPhys
B194:65, DOI 10.1016/0550-3213(82)90512-0
[77] Girardi G, Grimm R, Muller M, Wess J (1984) Superspace geometry and
the minimal, non minimal, and new minimal supergravity multiplets. ZPhys
C26:123, DOI 10.1007/BF01572550
[78] Giudice GF, Masiero A (1988) A natural solution to the µ −problem in su-
pergravity theories. Phys Lett B206:480–484, DOI 10.1016/0370-2693(88)
91613-9
[79] Giudice GF, Rattazzi R (1999) Theories with gauge-mediated supersym-
metry breaking. Phys Rept 322:419, DOI 10.1016/S0370-1573(99)00042-3,
hep-ph/9801271
[80] Giudice GF, Rattazzi R (1999) Theories with gauge mediated supersymmetry
breaking. Phys Rept 322:419–499, DOI 10.1016/S0370-1573(99)00042-3,
hep-ph/9801271
[81] Golfand Y, Likhtman E (1971) Extension of the algebra of Poincaré group
generators and violation of P invariance. JETP Lett 13:323–326
[82] Gondolo P, Gelmini G (1991) Cosmic abundances of stable particles: Im-
proved analysis. Nucl Phys B 360:145–179, DOI 10.1016/0550-3213(91)
90438-4
[83] Griest K, Seckel D (1991) Three exceptions in the calculation of relic abun-
dances. Phys Rev D 43:3191–3203, DOI 10.1103/PhysRevD.43.3191
[84] Grimm R (2011) private communication
42 Florian Domingo and Michel Rausch de Traubenberg

[85] Grimm R, Wess J, Zumino B (1978) Consistency checks on the superspace


formulation of supergravity. PhysLett B73:415, DOI 10.1016/0370-2693(78)
90753-0
[86] Grimm R, Wess J, Zumino B (1979) A complete solution of the Bianchi
identities in superspace. NuclPhys B152:255, DOI 10.1016/0550-3213(79)
90102-0
[87] Grisaru MT, Siegel W, Rocek M (1979) Improved Methods for Supergraphs.
Nucl Phys B 159:429, DOI 10.1016/0550-3213(79)90344-4
[88] Grossman Y, Nir Y (1995) Lepton mass matrix models. Nucl Phys B 448:30–
50, DOI 10.1016/0550-3213(95)00203-5, hep-ph/9502418
[89] Haag R, Lopuszanski JT, Sohnius M (1975) All possible generators
of supersymmetries of the S-matrix. NuclPhys B88:257, DOI 10.1016/
0550-3213(75)90279-5
[90] Haber HE, Kane GL (1985) The search for supersymmetry: probing
physics beyond the standard model. PhysRept 117:75–263, DOI 10.1016/
0370-1573(85)90051-1
[91] Han C, Hikasa Ki, Wu L, Yang JM, Zhang Y (2017) Status of cMSSM in
light of current LHC Run-2 and LUX data. Phys Lett B769:470–476, DOI
10.1016/j.physletb.2017.04.026, 1612.02296
[92] Heinemeyer S, Hollik W, Stockinger D, Weber AM, Weiglein G (2006)
Precise prediction for M(W) in the MSSM. JHEP 08:052, DOI 10.1088/
1126-6708/2006/08/052, hep-ph/0604147
[93] Heinemeyer S, Hollik W, Weber AM, Weiglein G (2008) Z Pole Observ-
ables in the MSSM. JHEP 04:039, DOI 10.1088/1126-6708/2008/04/039,
0710.2972
[94] Henn JM (2021) What Can We Learn About QCD and Collider Physics from
N = 4 Super Yang–Mills? Ann Rev Nucl Part Sci 71:87–112, DOI 10.1146/
annurev-nucl-102819-100428, 2006.00361
[95] Hollik WG, Weiglein G, Wittbrodt J (2019) Impact of Vacuum Stability Con-
straints on the Phenomenology of Supersymmetric Models. JHEP 03:109,
DOI 10.1007/JHEP03(2019)109, 1812.04644
[96] Howe PS, Tucker RW (1978) Scale invariance in superspace. PhysLett
B80:138, DOI 10.1016/0370-2693(78)90327-1
[97] Ibañez LE (1982) Locally supersymmetric SU(5) grand unification. Phys Lett
B118:73, DOI 10.1016/0370-2693(82)90604-9
[98] Iliopoulos J, Zumino B (1974) Broken supergauge symmetry and renormal-
ization. Nucl Phys B76:310, DOI 10.1016/0550-3213(74)90388-5
[99] Intriligator KA, Seiberg N (2007) Lectures on supersymmetry break-
ing. ClassQuantGrav 24:S741–S772, DOI 10.1088/0264-9381/24/21/S02,
hep-ph/0702069
[100] Isidori G, Nir Y, Perez G (2010) Flavor Physics Constraints for Physics Be-
yond the Standard Model. Ann Rev Nucl Part Sci 60:355, DOI 10.1146/
annurev.nucl.012809.104534, 1002.0900
Supergravity: Application in Particle Physics 43

[101] Jack I, Jones DT, Martin SP, Vaughn MT, Yamada Y (1994) Decoupling of
the epsilon scalar mass in softly broken supersymmetry. PhysRev D50:5481–
5483, DOI 10.1103/PhysRevD.50.R5481, hep-ph/9407291
[102] Jack I, Jones DRT, North CG (1996) N=1 supersymmetry and the three loop
anomalous dimension for the chiral superfield. Nucl Phys B 473:308–322,
DOI 10.1016/0550-3213(96)00269-6, hep-ph/9603386
[103] Jack I, Jones DRT, North CG (1996) N=1 supersymmetry and the three loop
gauge Beta function. Phys Lett B 386:138–140, DOI 10.1016/0370-2693(96)
00918-5, hep-ph/9606323
[104] Jack I, Jones DRT, North CG (1997) Scheme dependence and the NSVZ Beta
function. Nucl Phys B 486:479–499, DOI 10.1016/S0550-3213(96)00637-2,
hep-ph/9609325
[105] Jack I, Jones DRT, Pickering A (1998) The connection between DRED
and NSVZ. Phys Lett B 435:61–66, DOI 10.1016/S0370-2693(98)00769-2,
hep-ph/9805482
[106] Kaku M, Townsend P (1978) Poincaré supergravity as broken superconfor-
mal gravity. PhysLett B76:54, DOI 10.1016/0370-2693(78)90098-9
[107] Kaku M, Townsend P, van Nieuwenhuizen P (1978) Properties of conformal
supergravity. PhysRev D17:3179, DOI 10.1103/PhysRevD.17.3179
[108] Kaul RK, Majumdar P (1982) Cancellation of Quadratically Divergent Mass
Corrections in Globally Supersymmetric Spontaneously Broken Gauge The-
ories. Nucl Phys B 199:36, DOI 10.1016/0550-3213(82)90565-X
[109] Kim JE, Nilles HP (1984) The mu Problem and the Strong CP Problem. Phys
Lett B 138:150–154, DOI 10.1016/0370-2693(84)91890-2
[110] Kuzenko S (2022) In this volume
[111] Kuzenko S, Tartaglino Mazzucchelli G (2022) In this volume
[112] Lahanas A, Nanopoulos DV (1987) The road to no-scale supergravity. Phys-
Rept 145:1, DOI 10.1016/0370-1573(87)90034-2
[113] Lee HM, Raby S, Ratz M, Ross GG, Schieren R, Schmidt-Hoberg K, Vau-
drevange PKS (2011) Discrete R symmetries for the MSSM and its singlet
extensions. Nucl Phys B 850:1–30, DOI 10.1016/j.nuclphysb.2011.04.009,
1102.3595
[114] Martin S, Vaughn M (1994) Two loop renormalization group equations for
soft supersymmetry breaking couplings. PhysRev D50:2282, DOI 10.1103/
PhysRevD.50.2282,10.1103/PhysRevD.78.039903, hep-ph/9311340
[115] Mohapatra RN (1986) Unification and Supersymmetry: The frontiers of
Quark - Lepton Physics. Springer, Berlin, DOI 10.1007/978-1-4757-1928-4
[116] Mohapatra RN (1999) Supersymmetric grand unification: An Update. In:
ICTP Summer School in Particle Physics, pp 336–394, hep-ph/9911272
[117] Moultaka G, Rausch de Traubenberg M, Tant D (2019) Low Energy Su-
pergravity Revisited (i). Int J Mod Phys A34(01):1950004, DOI 10.1142/
S0217751X19500040, 1611.10327
[118] Muller (1989) Consitent Supergravity Theories. Springer-Verlag
[119] Muller M (1982) The density multiplet in superspace. Z Phys C16:41, DOI
10.1007/BF01573745
44 Florian Domingo and Michel Rausch de Traubenberg

[120] Nappi CR, Ovrut BA (1982) Supersymmetric extension of


the SU(3)×SU(2)× U(1) model. Phys Lett B113:175, DOI
10.1016/0370-2693(82)90418-X
[121] Nath P (2016) Supersymmetry, Supergravity, and Unifica-
tion. Cambridge Monographs on Mathematical Physics, Cam-
bridge University Press, DOI 10.1017/9781139048118, URL
http://www.cambridge.org/academic/subjects/physics/theoretical-physics-and-m
[122] van Nieuwenhuizen P (1981) Supergravity. Phys Rept 68:189, DOI 10.1016/
0370-1573(81)90157-5
[123] Nilles HP (1984) Supersymmetry, supergravity and particle physics. Phys
Rept 110:1, DOI 10.1016/0370-1573(84)90008-5
[124] Nir Y, Rattazzi R (1996) Solving the supersymmetric CP problem with
Abelian horizontal symmetries. Phys Lett B 382:363–368, DOI 10.1016/
0370-2693(96)00571-0, hep-ph/9603233
[125] Ohta N (1983) Grand unified theories based on local supersymmetry. Prog
Theor Phys 70:542, DOI 10.1143/PTP.70.542
[126] O’Raifeartaigh L (1975) Spontaneous symmetry breaking for chirals scalar
superfields. Nucl Phys B96:331, DOI 10.1016/0550-3213(75)90585-4
[127] Polonyi J (1977) Generalization of the massive scalar multiplet coupling to
the supergravity. Hungary Central Inst Res - KFKI-77-93 (unpublished)
[128] Raby S (2017) Supersymmetric Grand Unified Theories: From
Quarks to Strings via SUSY GUTs, vol 939. Springer, DOI
10.1007/978-3-319-55255-2
[129] Randall L, Sundrum R (1999) Out of this world supersymmetry breaking.
NuclPhys B557:79–118, DOI 10.1016/S0550-3213(99)00359-4
[130] Ross GG (1985) Grand Unified Theories. Frontiers in Physics, Westview
Press
[131] Sakai N (1981) Naturalness in Supersymmetric Guts. Z Phys C 11:153, DOI
10.1007/BF01573998
[132] Salam A, Strathdee J (1974) On Goldstone fermions. PhysLett B49:465–467,
DOI 10.1016/0370-2693(74)90637-6
[133] Salam A, Strathdee JA (1974) Super-gauge transformations. Nucl Phys
B76:477, DOI 10.1016/0550-3213(74)90537-9
[134] Salam A, Strathdee JA (1975) Superfields and Fermi-Bose symmetry. Phys
Rev D11:1521, DOI 10.1103/PhysRevD.11.1521
[135] Siegel W (1978) Solution to constraints in Wess-Zumino supergravity for-
malism. NuclPhys B142:301, DOI 10.1016/0550-3213(78)90205-5
[136] Siegel W (1979) Supersymmetric Dimensional Regularization via Dimen-
sional Reduction. Phys Lett B 84:193–196, DOI 10.1016/0370-2693(79)
90282-X
[137] Slansky R (1981) Group Theory for Unified Model Building. Phys Rept
79:1–128, DOI 10.1016/0370-1573(81)90092-2
[138] Slavich P, et al (2021) Higgs-mass predictions in the MSSM and beyond. Eur
Phys J C 81(5):450, DOI 10.1140/epjc/s10052-021-09198-2, 2012.15629
Supergravity: Application in Particle Physics 45

[139] Soni SK, Weldon HA (1983) Analysis of the supersymmetry breaking in-
duced by N=1 supergravity theories. Phys Lett 126B:215–219, DOI 10.1016/
0370-2693(83)90593-2
[140] Stelle K, West PC (1978) Relation between vector and scalar multiplets
and gauge invariance in supergravity. NuclPhys B145:175, DOI 10.1016/
0550-3213(78)90420-0
[141] Stelle K, West PC (1978) Tensor calculus for the vector multiplet coupled to
supergravity. PhysLett B77:376, DOI 10.1016/0370-2693(78)90581-6
[142] Stelle KS, West PC (1978) Minimal auxiliary fields for supergravity. Phys
Lett B74:330, DOI 10.1016/0370-2693(78)90669-X
[143] Stockinger D (2005) Regularization by dimensional reduction: consistency,
quantum action principle, and supersymmetry. JHEP 03:076, DOI 10.1088/
1126-6708/2005/03/076, hep-ph/0503129
[144] Stockinger D (2007) The Muon Magnetic Moment and Supersym-
metry. J Phys G 34:R45–R92, DOI 10.1088/0954-3899/34/2/R01,
hep-ph/0609168
[145] Terning J (2006) Modern Supersymmetry: Dynamics and Duality. Interna-
tional Series of Monographs on Physics 132. Oxford: Oxford University
Press.
[146] Townsend P, van Nieuwenhuizen P (1979) Simplifications of conformal su-
pergravity. PhysRev D19:3166, DOI 10.1103/PhysRevD.19.3166
[147] Rausch de Traubenberg M, Valenzuela M (2020) A Supergravity Primer:
From Geometrical Principles to the Final Lagrangian. World Scientific, DOI
10.1142/11557
[148] Vilenkin A (1985) Cosmic Strings and Domain Walls. Phys Rept 121:263–
315, DOI 10.1016/0370-1573(85)90033-X
[149] Weinberg S (2000) The Quantum Theory of Fields, Vol 3. Cambridge Uni-
versity Press
[150] Wess J, Bagger J (1992) Supersymmetry and Supergravity, 2nd edn. Prince-
ton University Press
[151] Wess J, Zumino B (1977) Superspace formulation of supergravity. Phys Lett
B66:361, DOI 10.1016/0370-2693(77)90015-6
[152] Wess J, Zumino B (1978) The component formalism follows from the
superspace formulation of supergravity. PhysLett B79:394, DOI 10.1016/
0370-2693(78)90390-8
[153] Wess J, Zumino B (1978) Superfield Lagrangian for supergravity. PhysLett
B74:51, DOI 10.1016/0370-2693(78)90057-6
[154] West PC (1986) Introduction to Supersymmetry and Supergravity. Singapore,
Singapore: World Scientific
[155] Witten E (1981) Dynamical breaking of supersymmetry. Nucl Phys
B188:513, DOI 10.1016/0550-3213(81)90006-7
[156] Witten E (1981) Lecture notes on supersymmetry. Lecture given at ICTP,
Trieste
[157] Witten E (1981) Mass Hierarchies in Supersymmetric Theories. Phys Lett B
105:267, DOI 10.1016/0370-2693(81)90885-6
46 Florian Domingo and Michel Rausch de Traubenberg

[158] Yamada Y (1994) Two loop renormalization group equations for soft SUSY
breaking scalar interactions: Supergraph method. PhysRev D50:3537–3545,
DOI 10.1103/PhysRevD.50.3537, hep-ph/9401241
[159] Yamatsu N (2015) Finite-Dimensional Lie Algebras and Their Representa-
tions for Unified Model Building 1511.08771
[160] Zumino B (1978) Supergravity and superspace. In Cargese 1978, Proceed-
ings, Recent Developments In Gravitation, 405-459

You might also like