You are on page 1of 122

TESTING NON-STANDARD NEUTRINO

PROPERTIES IN LOW AND HIGH ENERGY


SECTORS

ALEXANDER PARADA VALENCIA

A DISSERTATION PRESENTED TO THE DEPARTAMENTO DE FÍSICA


DEL CENTRO DE INVESTIGACIÓN Y DE ESTUDIOS AVANZADOS DEL
INSTITUTO POLITÉCNICO NACIONAL IN PARTIAL FULFILLMENT OF
THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF SCIENCE IN
PHYSICS

ADVISOR: Dr. OMAR GUSTAVO MIRANDA ROMAGNOLI

February 2014
México, D.F.
Acknowledgements

First I want to thank God Almighty for giving me the faith and granting me the capability
to complete this thesis. I would like to express my sincere gratitude to my advisor, Omar
Gustavo Miranda Romagnoli, for his guidance and support in the course of this work. His
willingness to assist his students has made my postgraduate studies a pleasant experience.
I would also like to thank the Consejo Nacional de Ciencia y Tecnologı́a (CONACyT), for
Ph D scholarship and for the support from the CONACyT grant 166639. To the Centro de
Investigación y de Estudios Avanzados del I.P.N. (Cinvestav). Many thanks to professors
Nora Bretón, David Delepine, Gabriel López, Abdel Pérez Lorenzana, and Arnulfo Zepeda
for reading and reviewing this thesis. I am grateful to the staff of the Physics Department
of CINVESTAV, this work would has not been possible without their help and hospitality.
Finally, I want to express my special thanks to my wife, Diana, her love and patience have
been of great help to accomplish our purposes in life. Likewise, I am grateful to my parents,
Carmen and Hermes, who despite the distance have always supported me.

1
Resumen

En el Modelo Estándar (ME) de fı́sica de partı́culas, los neutrinos no tienen masa ni carga
y sólo interaccionan débilmente. Sin embargo, el mecanismo de oscilaciones de neutrinos
ha mostrado que los neutrinos son masivos. Por tanto, es natural buscar extensiones
al ME a fin de explicar la masa de los neutrinos. En algunas de estas extensiones el
neutrino adquiere propiedades electromagnéticas, como por ejemplo, el momento magnético
del neutrino (MMN). En esta tesis, hemos estudiado el MMN y realizado un análisis
combinado de datos experimentales de reactores y aceleradores para obtener cotas al MMN.
Por otro lado, el origen y la composición de los rayos cósmicos de ultra-alta energı́a (RCUAE)
es actualmente una de las preguntas más intrigantes en la fı́sica de astro-partı́culas. En
este contexto, el estudio de los neutrinos astrofı́sicos podrı́a contribuir significativamente al
entendimiento de los RCUAE. En este trabajo estudiamos también la posibilidad de que el
MMN pueda llevar a efectos de supresión en el flujo esperado de neutrinos provenientes
de fuentes astrofı́sicas. Adicionalmente, la posible interacción entre neutrinos activos (o
estériles) y materia oscura ha sido analizada en la literatura cientı́fica. Mostraremos un
caso en donde estas interacciones podrı́an tener un efecto significativo en los neutrinos de
ultra-alta energı́a.

2
Abstract

In the Standard Model (SM) of Particles Physics, the neutrinos are massless fermions,
they have not charges, and only interact by weak interactions. However, the oscillations
mechanism establish that the neutrinos are massive particles; therefore, it is natural to
extend the Standard Model in order to give an explanation for the neutrino mass. In some
extensions of the SM the neutrino gets electromagnetic properties. In this thesis, we have
mainly focused in the study of the neutrino magnetic moment (NMM), especially, we performed
a combined analysis of experimental data to obtain new bounds on the NMM from reactor and
accelerator experiments.
On the other hand, the origin and composition of ultra-high energy cosmic rays (UHECRs)
are currently a mystery. The astrophysical neutrinos have been considered as a fundamental
tool for looking to answers about the UHECRs. In this context, we present the study of some
neutrino properties where the NMM could have important effects in the expected neutrino flux
from extragalactic sources. Additionally, possible interactions between active neutrinos and
dark matter, or sterile neutrinos and dark matter, have been examined in the literature for
different theoretical models. We show a case where these interactions could have a significant
effect for ultra high energy neutrinos.

3
Contents

1 Introduction 10

2 Neutrinos in the Standard Model 13


2.1 Some properties of the neutrinos . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.1.1 Helicity and chirality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.1.2 Charge conjugation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.2 The Standard Model of particle physics . . . . . . . . . . . . . . . . . . . . . . . . 17
2.3 Massive Neutrinos . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.3.1 Dirac and Majorana mass terms . . . . . . . . . . . . . . . . . . . . . . . . 24
2.4 Neutrino Oscillations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.4.1 Oscillations with two neutrino flavors . . . . . . . . . . . . . . . . . . . . . 31
2.4.2 Neutrino oscillations in matter . . . . . . . . . . . . . . . . . . . . . . . . . 33

3 Constraints on µν from experimental data 38


3.1 Magnetic moment in the minimally extended standard model . . . . . . . . . . . 38
3.1.1 The case of Dirac neutrinos . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.1.2 The case for Majorana neutrinos . . . . . . . . . . . . . . . . . . . . . . . . 41
3.1.3 Parametrization of the Neutrino magnetic moment . . . . . . . . . . . . . 41
3.2 Neutrino-electron elastic scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
3.2.1 Effective magnetic moment . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.3 Experimental measurement of the neutrino magnetic moment . . . . . . . . . . . 48
3.3.1 Neutrino Experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
3.3.2 Neutrino Flux in the detector . . . . . . . . . . . . . . . . . . . . . . . . . . 56
3.3.3 Reactor Experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
3.4 Combined Analysis of reactor and accelerator neutrinos . . . . . . . . . . . . . . . 58
3.4.1 Limits on Effective Magnetic Moment . . . . . . . . . . . . . . . . . . . . . 58

4
CONTENTS 5

4 Ultra-High energy neutrinos 63


4.1 Neutrino Astronomy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
4.1.1 The IceCube Neutrino Observatory . . . . . . . . . . . . . . . . . . . . . . . 65
4.2 Gamma Ray Bursts (GRBs) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
4.3 Waxman-Bahcall theoretical prediction for the high energy neutrinos . . . . . . . 68
4.4 Experimental limits for astrophysical neutrino flux . . . . . . . . . . . . . . . . . 70
4.4.1 ANTARES experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
4.4.2 Pierre Auger experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
4.4.3 IceCube Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
4.5 A reduction in the UHE neutrino flux due to neutrino spin precession . . . . . . 76
4.5.1 Active-sterile conversion in the propagation . . . . . . . . . . . . . . . . . . 78
4.5.2 Survival probability vs distance in the intergalactic medium . . . . . . . . 79
4.5.3 Survival probability vs distance (galactic medium) . . . . . . . . . . . . . . 81
4.6 Cosmological Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
4.6.1 Comoving distance (line-of-sight) . . . . . . . . . . . . . . . . . . . . . . . . 83
4.6.2 Survival probability vs redshift . . . . . . . . . . . . . . . . . . . . . . . . . 85
4.7 Neutrino magnetic moment in extra dimensions . . . . . . . . . . . . . . . . . . . 86

5 Dark matter, sterile neutrinos and resonant effects 89


5.1 Dark Matter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
5.2 Sterile Neutrinos . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
5.2.1 Astrophysical evidence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
5.3 Evidence for sterile neutrinos from oscillation experiments . . . . . . . . . . . . . 97
5.3.1 The accelerator anomaly . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
5.3.2 The Gallium anomaly . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
5.3.3 The Reactor Antineutrino Anomaly . . . . . . . . . . . . . . . . . . . . . . . 100
5.4 A resonant effect in dark matter . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
5.5 Resonant effects of the ν − DM and νs − DM interactions . . . . . . . . . . . . . 105

6 Conclusions 110
List of Figures

3.1 Radiative corrections giving rise to a magnetic dipole moment in the minimally
extended standard model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.2 ν̄ − e scattering. Taken from [1]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.3 Contribution of the SM and electromagnetic to the cross section ν̄ − e as function
of the recoil electron. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.4 Cross section ν − e with and without the term of m3e , for µν = 1. × 10−10 µB . . . . 47
3.5 Nuclear Fission. Taken from [2] . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
3.6 The pp chain of thermonuclear reactions in the Sun. Taken from [3]. . . . . . . . 54
3.7 The CNO cycle of thermonuclear reactions in the Sun. Taken from [3]. . . . . . . 54
3.8 Neutrino flux from the pp and CNO chains as a function of the neutrino energy.
Taken from [3]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
2
3.9 The likelihood χ as a function of the effective magnetic moment µν . . . . . . . . 60

4.1 The cosmic ray spectrum and experiments and some experimental results.
Taken from www.physics.utah.edu/ whanlon/spectrum.html. . . . . . . . . . . . 64
4.2 The IceCube detector. Taken from [4]. . . . . . . . . . . . . . . . . . . . . . . . . . 66
4.3 Long-duration GRBs (left) are thought to originate in the collapse of massive
stars, while short-duration GRBs are likely produced in the merger of two
compact objects. Taken from [5]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

6
LIST OF FIGURES 7

4.4 Upper panel: The upper bound imposed by UHECR observations on the diffuse
muon neutrino flux (νµ and ν̄µ combined) (lower curve: there is no evolution
of the energy production rate, upper curve: there is evolution) predicted by
different models, the upper bound implied by cosmic rays observations and
GRB intensity. The dash-dotted lines are the upper bound given by the
Eq. (4.2). Taken from [7]. Lower panel: The same bound compared with the
atmospheric muon-neutrino background and several upper bounds reported by
experiments such as AMANDA, BAIKAL, RICE, ANITA and Auger. The GZK
curve shows the muon neutrino flux expected from interactions between protons
and micro-wave background. Taken from [8]. . . . . . . . . . . . . . . . . . . . . . 69
4.5 Sum of the 296 individual gamma-ray-burst muon neutrino spectra (red and
blue solid lines) and limits on the flux expected (red and blue dashed lines).
The IceCube IC40 + IC59 limit on the neutrino intensity from 300 GRBs (black
dashed line) and the first ANTARES limit from 2007 using 40 GRBs (grey
dash-dotted line). Taken from [9]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
4.6 Thick lines: Upper limits (at 90% C.L.) to the diffuse flux of UHE neutrinos,
limits from other experiments are also plotted. Thin lines: Expected fluxes
from the theoretical models: ”p, Fermi-LAT” [10], ”p,evol-FRII” and ”Fe,
uniform” [11]. Taken from [12]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
4.7 Comparison of results to predictions based on observed γ-ray spectra. The
predictions are shown in dashed lines and the experimental measurements in
solid lines. Taken from [13]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
4.8 Events number vs deposited energy compared to model predictions. Taken
from [14]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
4.9 Limit on a (νµ + ν̄µ ) astrophysical flux in comparison to theoretical flux
predictions and limits from other experiments. The Black lines: show the
expected atmospheric neutrino flux with and without a prompt component. The
red dashed line: Denotes the Waxman-Bahcall upper bound. The Green dashed
lines: represent several model predictions for astrophysical neutrino fluxes. The
pink solid line: is the experimental limit at 90% C.L. The orange solid line: shows
its sensitivity. Taken from [15]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
4.10 Relation of magnetic field B and size r of astrophysical sources for an efficient
neutrino spin transition νlL → νlR . The curves show different values of the
neutrino magnetic moment [86] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
LIST OF FIGURES 8

4.11 Survival probability (PS ) vs distance for distant sources in the Universe such as
GRBs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
4.12 Survival probability (PS ) vs distance for nearby sources in the Universe like AGNs. 80
4.13 Survival probability (PS ) vs distance in the galactic medium. . . . . . . . . . . . . 81
4.14 Black line: Line-of-sight comoving distance as a function of redshift in a flat
Λ − CDM cosmology. Red dash-line: Hubble length: LH ≈ 3.89 Gpc. Figure
reproduced from [16]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
4.15 Survival probability vs redshift for far sources such as GRBs. . . . . . . . . . . . 85
4.16 Survival probability vs redshift for nearby sources such as AGNs. . . . . . . . . . 86
4.17 Shows the effective magnetic moment for parameters y = 1, M∗ = 104 T eV ,
The effective magnetic moment is plotted as a function of available energy for
n = 1, 2, 4 and 6 dimensions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87

5.1 Rotation curves of NGC 6503. Taken from [17]. . . . . . . . . . . . . . . . . . . . . 90


5.2 Allowed regions in the sin2 (2θnew ) − ∆m2new parameter space from combined
analysis from reactor and gallium experiments, the star represent the best fit.
Taken from [18]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
5.3 Interaction processes that could give a solution to the small-scale problems.
Taken from [19]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
5.4 Mass of the vector intermediary particle vs mass of dark matter candidate (left
panel). Mass of the vector intermediary vs the coupling strength gν (right panel).
Taken from [19]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
5.5 Feynman rule for dm-fermion interactions. Taken from [20]. . . . . . . . . . . . . 104
5.6 DM-ordinary fermions interactions by a new boson U . Taken from [20]. . . . . . 105
5.7 Isocurves for the resonant effect of (anti)neutrinos with ∆m2 = 10−12 −
10−18 eV 2 . We take a constant dark matter density, ρχ = 0.3 GeV cm −3
,
consistent with the average value of the galactic halo. We also show a region
of restricted values reported in the literature [21, 20] (dark blue box) and as
region of values motivated by a solution to the small scale problem [19] (light
blue box). The left panel corresponds to a neutrino energy of E = 1018 eV while
the right one to E = 1015 eV [132]. . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
5.8 Survival probability P (να → να ) for a constant dark matter density, ρχ =
0.3 GeV cm −3 , consistent with the average value of the galactic halo [132]. . . . 109
List of Tables

2.1 Properties of the leptons and quarks. . . . . . . . . . . . . . . . . . . . . . . . . . . 18

3.1 ∆χ2 corresponding to a probability 1 − α in the large data sample limit, for m
parameters. Taken from [22]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
3.2 Limits on the effective magnetic moment. . . . . . . . . . . . . . . . . . . . . . . . 60
3.3 Limits on Λ1 from reactor experiments. . . . . . . . . . . . . . . . . . . . . . . . . 61
3.4 Limits on Λ2 from reactor experiments. . . . . . . . . . . . . . . . . . . . . . . . . 61
3.5 Limits on Λ3 from reactor experiments. . . . . . . . . . . . . . . . . . . . . . . . . 61
3.6 Limits on Λ1 from accelerator experiments. . . . . . . . . . . . . . . . . . . . . . . 62
3.7 Limits on Λ2 from accelerator experiments. . . . . . . . . . . . . . . . . . . . . . . 62
3.8 Limits on Λ3 from accelerator experiments. . . . . . . . . . . . . . . . . . . . . . . 62

5.1 Coupling constants and mass estimates from different models [132]. . . . . . . . 102

9
Chapter 1

Introduction

The Standard Model of Particles Physics has been a very successful theory since its
formulation. It is supported by the large number of experimental results, which confirm
the standard model as a theory of ’almost everything’. Despite the great success achieved
by this theory, there are several problems which are not explained by it. Some of the
most important open questions are related with the dark matter problem, the mechanism
generating neutrino mass, the CP violation in the leptonic sector, the mass hierarchy problem,
and the neutrino nature (Dirac or Majorana), among others. The search for solutions to these
unsolved problems has motivated the proposal of new physics, also known as physics beyond
the standard model.
One of the first motivations for a correction in the Standard Model (SM) of Elementary
Particles came in the neutrino sector as early as the late 50s and early 60s. After many years
of experimental efforts, it has been possible to establish that the neutrinos are massive. Solar,
accelerator, atmospheric and reactor neutrino experiments provide a compelling evidence
regarding the neutrino oscillations. Most of the large amount of collected data by these
experiments can be described in the framework of three neutrinos, where we have three flavor
and three mass eigenstates related by a 3 × 3 mixing matrix.
Although the three neutrino oscillation scheme is currently a well stablished phenomenon,
there are some experimental observations that are not in agreement with the three neutrino
predictions. Some of these anomalies were reported by solar, accelerator, and reactor
neutrinos and open a window to explore another important topic in physics beyond the
standard model: the existence of a sterile neutrino. It has been studied in the literature that
the existence of a fourth or even a fifth sterile neutrino could provide an explanation for the

10
CHAPTER 1. INTRODUCTION 11

anomalies recorded by these experiments. There is also evidence for sterile neutrinos from
Cosmology, where the current measurements report an effective number of neutrino families
greater than three.
On the other hand, the neutrino oscillations imply the existence of the neutrino mass;
therefore, it is necessary to extend, at least minimally, the Standard Model. Some extensions
of the Standard Model implies the existence of other important properties of the neutrino,
for instance, electromagnetic properties; among them, the most studied is the Neutrino
Magnetic Moment (NMM). The study of the NMM started at the end of 1970, when the
first calculations were carried out in a minimal extension of the standard model including
right handed neutrinos. The prediction of the minimally extended standard model for the
NMM is µν = 3.2 × 10−19 (mi /eV )µB , with mi the neutrino mass and µB the Bohr magneton.
Experimental searches for the NMM have been performed for solar, reactor and accelerator
experiments, using electron neutrino or electron antineutrino scattering at low energies. From
the terrestrial experiments the limit on µν is of the order of µν < 10−11 µB , as we will see
in chapter 3. The first stage of this thesis has been dedicated to calculate new limits on
the NMM, based on the experimental data, by making a combined analysis of the available
results. This analysis allows us to obtain more reliable bounds on the NMM, because the
collected information from different experiments, performed in distinct places and conditions,
allow to strength the individual measurements.
Moreover, there has been a growing interest to study astrophysical neutrinos (AN), which
are neutrinos produced in sources far away in the Universe, with extremely high energies,
and that propagate through very long distances, without interact with ordinary matter. This
makes such neutrinos a fundamental tool to explore the Universe in large scales. The Ultra
High Energy (UHE) Astrophysical neutrinos has a direct connection with UHE cosmic rays
(CRs) since it is probable that they come from the same source; but, contrary to the CRs,
the neutrinos are not deviated by the intergalactic and galactic magnetic fields, allowing, at
least in principle, to get first-hand information about the astrophysical sources. Accordingly,
the study of AN could help to solve the most important problems concerning UHECRs, for
instance, what the acceleration mechanism at the source is, which the sources are, and what
the UHECRs composition is, among others. One of the possible sources of UHE neutrinos are
the Gamma Ray Burts (GRBs), which are the most intensive explosions of gamma rays ever
seen in the Universe. The theoretical predictions of the diffuse neutrino flux cover an energy
range from 1012 eV to 1020 eV. Because of the very low prediction for the diffuse neutrino
flux from astrophysical sources, it has been necessary to built very big large detectors for its
CHAPTER 1. INTRODUCTION 12

detection. Currently the most promising experiment in order to detect and identify AN is
the IceCube detector, an experiment located in the South Pole. In a recent result, IceCube
reported the observation of 28 events associated with extraterrestrial neutrinos, the origin of
these neutrinos is unknown and their energies ranges from 30 TeV to 1.14 PeV, with a gap
between 300 TeV and 1.14 PeV. In spite of the observation performed by IceCube, the low
number of events, the range of energy, as well as their sources, continues to be a mystery.
Different mechanisms providing an explanation for the suppression of the neutrino flux from
astrophysical sources have been proposed in the literature, for example, the absorption of
UHE neutrinos by the cosmic neutrino background [23], a reevaluation of the neutrino flux
[24], new information about the physical parameters of the acceleration model [25], and
neutrino decays in invisible particles [16, 26].
In this work, we have studied the implications of a NMM in astrophysical environments. If
the neutrino has a magnetic moment, it could interact with the magnetic fields in the source,
in the intergalactic space or in the galactic medium before being detected on earth; as a result,
this neutrino may suffer a conversion from active to sterile neutrino, which would provide an
explanation for the suppression of the UHE neutrino flux.
Finally, we have studied another important effect that connect two sectors still elusive in
particles physics: the dark matter and sterile neutrinos; in this scenario, an active or sterile
neutrino could interact with the dark matter giving rise to resonant effect, which will convert
active neutrinos into sterile neutrinos.
In the following chapters we will explain in more detail these results and we will also discuss
some possible perspectives for these topics.
Chapter 2

Neutrinos in the Standard Model

The history of the neutrino have its origins in 1896 when a French physicist named Henri
Becquerel observed for the first time the radioactivity of Uranium. Some years later, Ernest
Rutherford was studying this kind of radiation when he realized that it contained two different
products which were called α and β radiation. In 1914, the English physicist James Chadwick
proved that the β− spectrum was continuous, this result was in contrast to α− and γ−rays
which have a unique value in the energy. This was a surprise for the time, but it was
afterwards confirmed by Ellis and Wooster. This result seemed to violate the conservation
of the energy because in a two body decay (proton and electron) a fixed energy was expected.
The missing energy could not be explained with the γ−rays, then there was born the idea of
the existence of a new particle that would explain the undetected energy or as suggested Niels
Bohr: maybe the conservation of the energy is fulfilled only statistically.
A solution to this puzzle was proposed by the Austrian physicist Wolfgang Pauli. He sent a
letter to a physics conference at Tubingen on 4 December of 1930 with a note containing the
following famous words: “’Dear Radioactive Ladies and Gentlemen’, the existence of a neutral
weakly interacting fermion emitted in β−decay could solve the problems” [27]. This neutral
fermion was called neutron by Pauli. Pauli presented for the first time his idea in a talk given
in June 1931 at a meeting of the American Physical Society in Pasadena. After that James
Chadwick discovered in 1932 the neutron and the name of the Pauli particle was changed
to neutrino by Enrico Fermi. The discovery of the neutrino would come more than twenty
years later when Frederick Reines and Clyde Cowan confirmed its detection in 1956. In their
experiment, they used a water tank with dissolved Cd Cl2 with liquid scintillators around it.
The antineutrino source was a nuclear reactor. The fission reactions produce free neutrons,

13
CHAPTER 2. NEUTRINOS IN THE STANDARD MODEL 14

some of which decay producing electron antineutrinos that, after travelling until the detector,
react with protons and give rise to neutrons and positrons in what is called an inverse beta
decay: ν¯e + p → e+ + n. The positron is annihilated with an electron producing two gamma
rays which are detected, while the neutron is captured by a nucleus in the dopant material,
releasing a gamma ray. Both events: positron annihilation and neutron capture coincide
producing a unique signature of the interaction of an electron antineutrino in the detector.
In this chapter we review some of the basic theory that is indispensable to develop a reasearch
project on neutrino physics. The main idea of this chapter will therefore be to provide the
reader with a first insight into neutrino theory.

2.1 Some properties of the neutrinos


We begin this survey by discussing some of the main characteristics of the neutrino particle.
In future subsections the neutrino as a part of the Standard Model will be discussed, as well
as the oscillation mechanism.

2.1.1 Helicity and chirality

In Quantum Field Theory, spin− 12 particles are described by four-component wave functions
ψ(x) (spinors) which obey the Dirac equation:

 
µ ∂
iγ −m ψ =0 (2.1)
∂xµ
The 4 × 4 γ− Dirac matrices are given in the form
   
1 0 0 σi
γ0 =   γi =   (2.2)
0 −1 σi 0
where σi are the Pauli matrices given by
     
0 1 0 −i 1 0
σ1 =   σ2 =   σ3 =  . (2.3)
1 0 i 0 0 −1
Another important matrix is γ5 = iγ0 γ1 γ2 γ3 . The Dirac matrices satisfy the following
anticommutator relations

 α β
γ , γ = 2gαβ , {γ α , γ5 } = 0 (2.4)
CHAPTER 2. NEUTRINOS IN THE STANDARD MODEL 15

with  
+1 0 0 0
 
 0 −1 0 0 
 
gαβ =  . (2.5)
 0 0 −1 0 
 
 
0 0 0 −1
From the equation (2.1), it follows that

 
∂ ∂
i (1 + γ 5 ) + iσ i (1 + γ 5 ) − mγ 0 (1 − γ 5 ) ψ=0 (2.6)
∂x0 ∂xi
 
∂ ∂
i 0 (1 − γ5 ) − iσi (1 − γ5 ) − mγ0 (1 + γ5 ) ψ = 0; (2.7)
∂x ∂xi
we difine the projection operators PL and PR as

1 1
PL = (1 − γ5 ) PR = (1 + γ5 ) (2.8)
2 2
with

PL PR = 0, PL + PR = 1, PL2 = PL , PR2 = PR . (2.9)

Then, we can express the left-handed and right-handed components in terms of the projector
operators

ψL = PL ψ, ψR = PR ψ and PL ψR = PR ψL = 0. (2.10)

The equation of eigenvalues for γ5 have the form:

γ5 ψL,R = ∓ψL,R . (2.11)

The eigenvalues of γ5 are called chirality [28]. We can express any spinor ψ in terms of chiral
projections as follows

ψ = (PL + PR )ψ = PL ψ + PR ψ = ψL + ψR . (2.12)

In this notation, the equations (2.6) and (2.7) are given by

 
∂ ∂
+ iσi ψR = mγ0 ψL . (2.13)
∂x0 ∂xi

 
∂ ∂
0
− iσi ψL = mγ0 ψR (2.14)
∂x ∂xi
CHAPTER 2. NEUTRINOS IN THE STANDARD MODEL 16

For the particular case of m = 0, we obtain

∂ ∂
ψR = −iσi ψR (2.15)
∂x0 ∂xi

∂ ∂
0
ψL = iσi ψL (2.16)
∂x ∂xi
which have the form of the Schrodinger equation, with x0 = t and ~ = 1,

∂ ∂
ψL,R = ±iσi ψL,R . (2.17)
∂t ∂xi
∂ ∂
Using the definitions E = i ∂t and pi = −i ∂x i
, we get:

EψL,R = ∓σi pi ψL,R . (2.18)

Then, a new operator is defined

σ·p
H= (2.19)
|p|
The operator H is known as helicity operator.
According with Eq. (2.18), ψL has helicity H = −1 for particles and H = +1 for antiparticles.
And ψR has helicity H = +1 for particles and H = −1 for antiparticles. When m > 0, it is not
possible to decouple the equations (2.13) and (2.14). Therefore, the eigenfunctions ψL and ψR
can no longer have fixed helicity. In the Standard Model the neutrino is left-handed and the
antineutrino is right-handed.

2.1.2 Charge conjugation

If a particle is identical to its antiparticle, it is called Majorana particle. The latter requires
that all additive quantum numbers (for example: charge, strangeness, baryon number, lepton
number, etc) are the same for particles and antiparticles. If a neutrino is a Majorana particle,
the lepton number is not conserved. The charge conjugation operator Ĉ is defined as follows

Ĉ : ψ → ψ c = C ψ̄ T , C = iγ2 γ0 , ψ̄ = ψ † γ0 (2.20)

The matrix C satisfy the following properties:

C † = C T = C −1 = −C, Cγµ C −1 = −γµT (2.21)


CHAPTER 2. NEUTRINOS IN THE STANDARD MODEL 17

When Ĉ act over the chiral field, it flips its chirality,

Ĉ : ψL → (ψL )c = (ψ c )R , ψR → (ψR )c = (ψ c )L (2.22)

Therefore, the antiparticle of a left-handed fermion is a right-handed one. The operator Ĉ is


different from C, the first one flips all the quantum numbers including the chirality and the
helicity, the second one changes all the quantum numbers but not the chirality and helicity.

2.2 The Standard Model of particle physics


The Standard Model (SM) is defined as a quantum field theory which describes the strong,
electromagnetic, and weak interactions of elementary particles. From a mathematical point
of view, it is a gauge theory based on the local symmetry group SU (3)c × SU (2)L × U (1)Y ,
where the subscripts C, L and Y denote color, left-handed chirality and weak hypercharge,
respectively. The interactions are fixed by the local symmetry group, with three independent
unknown parameters that are the coupling constants of SU (3)C , SU (2)L and U (1)Y groups.
These parameters are determined from the experiments. Each subgroup of the symmetry
group contains gauge bosons which correspond to the generators of the subgroup. For instance,
the SU (3)C group, which describes the Quantum Chromodynamics, has eight generators that
mediate the strong interactions.
The SU (2)L × U (1)Y group that is responsible for electroweak interactions has four gauge
bosons, corresponding to the three generators of the SU (2)L subgroup, W i , and to the
generator of the U (1)Y subgroup, B. After spontaneous symmetry breaking, these bosons
mixed and three of them become massive: the bosons W ± and Z, while there is another boson
that remains massless, the photon, γ.
Moreover, the groups SU (3)C and SU (2)L × U (1)Y are not mixed, for that reason, the
electroweak interactions can be studied separately from strong interactions. On the contrary,
there is mixing between the SU (2)L and U (1)Y groups, therefore the electromagnetic and
weak interactions must be studied together.
We are going to review briefly the electroweak interactions, based on the symmetry group
SU (2)L × U (1)Y , which determines, among others, the interactions of neutrinos.

Electroweak Model

The symmetry group SU (2)L is called weak isospin group which has three generators (Ia , a =
1, 2, 3) that satisfy the angular momentum commutation relations: [Ia , Ib ] = iεabc Ic . Here εabc
CHAPTER 2. NEUTRINOS IN THE STANDARD MODEL 18

is the Levi-Civita tensor. The symmetry group U (1)Y is known as the hypercharge group.
where the hypercharge Y , isospin I3 and charge operator Q are connected by the Gell-Mann
Nishijima relation

Y
Q = I3 + . (2.23)
2
This relation unify the electromagnetic and weak interactions, in the electroweak model
SU (2)L × U (1)Y , the elementary particles are grouped in doublets and singlets. The doublets
are chiral left-handed fields while the singlets are right-handed fields,

   
um νm
L − doublets : qmL =   , lmL =   (2.24)
dm e−
m
L L
R − singlets : umR , dmR , e−
mR .

Here, m = 1, 2, 3 labels the family. The values of the weak isospin, hypercharge, and electric
charge are listed in Table 2.1.

I I3 Y Q
 
νeL 1/2 0
Lepton Doublet LL =   1/2 -1
eL −1/2 −1
Lepton Singlet e 0 0 -2 -1
R 
uL 1/2 2/3
Quark Doublet QL =   1/2 1/3
dL −1/2 −1/3
uR 4/3 2/3
Quark Singlet 0 0
dR −2/3 −1/3

Table 2.1: Properties of the leptons and quarks.

Here, for simplicity, we will consider only the first generation of chiral leptons fields eR , eL ,
and νeL . The Lagrangian for the free Dirac fields can be written as
 
νeL
L = (ν̄eL , ēL )(iγ µ ∂µ )   + ēR iγ µ ∂µ eR . (2.25)
eL
As we mentioned before, in the case of SU (2)L we have three generators and, therefore, there
are three vectors fields called Wµ1 , Wµ2 , Wµ3 . The Lagrangian including the W −fields can then
be written as
CHAPTER 2. NEUTRINOS IN THE STANDARD MODEL 19

 
1 ν eL
L = T r(Wµρ W µρ ) + (ν̄eL , ēL )iγ µ (∂µ + igWµ )   + ēR iγ µ ∂µ eR . (2.26)
2 eL

We define the fields Wµ± as

1
Wµ± = √ (Wµ1 ∓ iWµ2 ) (2.27)
2
and obtain

 
σ ν eL
L = −g(ν̄eL , ēL )iγ µ Wµ   (2.28)
2 eL

3
√ +
 
1 W µ 2W µ ν eL
= −g(ν̄eL , ēL )iγ µ  √  
2 2Wµ− −Wµ3 eL
gh √ √ i
= − Wµ3 (ν̄eL γ µ νeL − ēL γ µ eL ) + 2Wµ+ ν̄eL γ µ eL + 2Wµ− ēL γ µ νeL ,
2

where σ are the Pauli matrices. Besides the two charges fields, Wµ± , we have two massless
neutral vectors fields Wµ3 , Bµ , the combination of these fields describes the weak neutral
currents and the electromagnetism.
Using the definition of the two orthogonal linear combinations Zµ and Aµ :

1
gWµ3 − g ′ Bµ

Zµ = p (2.29)
g 2 + g ′2
1
g ′ Wµ3 + gBµ ,

Aµ = p (2.30)
g 2 + g ′2

and difining the weak mixing angle,

g′ g
sin θW = p , cos θW = p , (2.31)
g2 + g ′2 g2 + g ′2
the result is

Zµ = cos θW Wµ3 − sin θW Bµ (2.32)

Aµ = sin θW Wµ3 + cos θW Bµ . (2.33)

If we replace the fields Wµ3 , Bµ by Zµ , Aµ in (2.28), the result obtained is


CHAPTER 2. NEUTRINOS IN THE STANDARD MODEL 20

g
− √ Wµ+ ν̄eL γ µ eL + Wµ− ēL γ µ νeL

L = (2.34)
2
 
p 1 1
− g 2 + g ′2 Zµ ν̄eL γ µ νeL − ēL γ µ eL + sin2 θW (ēL γ µ eL + ēR γ µ eR )
2 2

gg
+ p Aµ (ēL γ µ eL + ēR γ µ eR ) .
g 2 + g ′2

The fields Wµ , Zµ and Aµ correspond to the charged, neutral and electromagnetic currents,
respectively. In order to reproduce the electromagnetism, we choose

gg ′
p = e. (2.35)
g2 + g ′2
From the Eq. (2.31) it is obtained a relation between the electromagnetic charge and the weak
mixing angle

e
sin θW = . (2.36)
g
Consequently the Lagrangian for the electroweak interactions is written as

 
µ 1 1 µ
Wµ+ ν̄eL γ µ eL + Wµ− ēL γ µ νeL +

L = −e Aµ Jem +√ Z µ JN C (2.37)
2 sin θW sin θW cos θW

With the electromagnetic and weak neutral currents given by

µ
Jem = (ēL γ µ eL + ēR γ µ eR ) = ēγ µ e (2.38)
µ 1 1
JN C = ν̄eL γ µ νeL − ēL γ µ eL + sin2 θW Jem
µ
(2.39)
2 2

Spontaneous symmetry breaking and the Higgs mechanism

In the Standard Model of particle physics, the Z and W gauge bosons and the fermions receive
their mass through the Higgs Mechanism. Consider the following classical Lagrangian

L = (∂µ Φ)† (∂ µ Φ) − µ2 Φ† Φ − λ(Φ† Φ)2 , (2.40)

where Φ(x) is a complex scalar field. In principle, L is invariant under the group U (1) of global
transformations. The first term correspond to the kinetic energy and is always positive. The
interaction potential is
CHAPTER 2. NEUTRINOS IN THE STANDARD MODEL 21

V (Φ) = µ2 Φ† Φ + λ(Φ† Φ)2 . (2.41)

If µ2 > 0 and λ > 0, there is a minimum configuration at the origin. Moreover, if µ2 < 0, the
minimum is located at

µ2 v2
ρ = ΦΦ† = − = , (2.42)
2λ 2
p √ p
with v = −µ2 λ. The Eq. (2.42) describes a whole ring with radius |Φ| = v/ z = −µ2 /1λ
in the complex plane, hence there are infinitely many ground states which show no longer the
original symmetry. This means that symmetry is broken spontaneously, where the degeneracy
of the ground state is associated with the spontaneous symmetry breaking. But how is it
achieved? Basically we have to introduce a doublet of complex scalars fields, which contain
one charged and one neutral field:
 
φ+ φ1 + iφ2 φ3 + iφ4
Φ= , where φ† = √ and φ0 = √ . (2.43)
φ 0 2 2

The expression for the Higgs Lagrangian is given by

LHiggs = (∂µ Φ)† (∂ µ Φ) − µ2 Φ† Φ − λ(Φ† Φ)2 . (2.44)

As we mentioned before, the minimum of the potential V (φ) is given for µ2 < 0 when Φ† Φ =
√ p
−µ2 /2λ = v 2 /2. Thus, the minimum lies on a circle with hΦi ≡ v/ 2 = −µ2 /2λ. There is a
degeneration in the ground state; it has not orientation defined in the isospin space.
This ground state is degenerate; its orientation in two-dimensional isospin space is not
defined. We can select any value between [0, 2π], choosing a particular field configuration,
 
1  0 
φ0 = √ . (2.45)
2 v
Expanding the field Φ around the vacuum,
 
1  0
Φ= √ , (2.46)
2 v + H(x)
is possible to obtain masses for the particles through a coupling to the vacuum expectation
value (vev) of the Higgs field. We consider the fermions first. We couple the Higgs doublet
with a fermion doublet and a singlet, which is known as a Yukawa coupling:
CHAPTER 2. NEUTRINOS IN THE STANDARD MODEL 22

 
νeL
LY uk = −ge ēR Φ†   + h.c. (2.47)
eL
   
ν eL
= −ge ēR Φ†0   + (ν¯e , ēL )φ0 eR 
eL
 
1 1
= −ge ēR √ veL + ēL √ veR
2 2
1
= −ge v √ (ēR eL + ēL eR )
2
v
= −ge √ ēe
2

where ge is a coupling constant. The mass term is given by the last expression, where the

electron mass is me = ge v/ 2. The idea to obtain mass for other leptons and quarks is exactly
the same. Finally it is the way to give mass to the fermions in the Standard Model.
A special case are the neutrinos, because in the particle content of the Standard Model there
are not right-handed neutrino singlet states (νR ); in consequence one cannot write couplings
like (2.47). As we will see later, at present there is strong evidence for massive neutrinos and,
therefore, it is necessary to generate the masses for them, for instance, adding right-handed
neutrino singlets or using Higgs triplets.
The gauge bosons acquire mass in an analog way. In this case, we substitute the normal
derivative for the covariant derivative in the Lagrangian (2.44) giving a result (see [3, 28]),

g2 v2 e2 v 2
m2W = = (2.48)
4 4 sin2 θW

(g 2 + g ′2 )v 2 e2 v 2
m2Z = = 2 , (2.49)
4 4 sin θW cos2 θW
with the relation

mW
= cos θW . (2.50)
mZ
According with the relation (2.48) we obtain an estimate for the vacuum expectation value v,


v = ( 2GF )−1/2 ≈ 246GeV. (2.51)

The spontaneous symmetry breaking produces then an important result: the existence of a
new scalar particle, called the Higgs boson, with a mass of mH , where
CHAPTER 2. NEUTRINOS IN THE STANDARD MODEL 23

√ p
mH = 2λv 2 = −2µ2 . (2.52)

Here µ2 is a negative parameter. Despite being a great achievement of the theory to explain
the origin of the masses for the elementary fermions, it can be noticed that the SM does not
predict the mass for the elementary particles including the Higgs boson. Instead, the particle
masses are free parameters of the theory which are determined experimentally.

Search and discovery of the Higgs boson

The first search for the Higgs boson was performed in the 1990s by the Large
Electron-Positron Collider (LEP), at CERN; the searches at LEP allowed to exclude
values lower than 114.4 GeV at 95 % CL [29]. Fermilab, in United States, also took part in
this search with the Tevatron experiment, which excluded the mass range 162 − 166 GeV at
95 % CL [30]. The Large Hadron Collider (LHC) experiment was built with the main purpose
of validating the existence of the Higgs Boson. First direct searches at the LHC were carried
out with data from proton-proton collisions at an integrated luminosity of 5.1f b−1 with 7 TeV
of center of mass energy [31].
The energy of the centre of mass was increased to 8 TeV in 2012, with the addition of an
integrated luminosity of the order of 5.3f b−1 , which was recorded by the two experiments
dedicated to measuring the Higgs mass: Compact Muon Solenoid (CMS) and A Toroidal
LHC Apparatus (ATLAS). This allowed to significantly enhance the sensitivity regarding
the searches of the Higgs boson. In July 2012, ATLAS and CMS experiments reported
the observation of a new boson with a mass of approximately 125 GeV. These results were
published in the papers [32] and [33] respectively.

We can summarize the observations as follows: CMS experiment performed searches for the

Standard Model (SM) Higgs boson in proton-proton collisions at energies of s = 7 and 8 TeV
and with integrated luminosities of up to 5.1f b−1 at 7 TeV and 5.3f b−1 at 8 TeV; the search
was carried out in five decay modes: γγ, ZZ, W + W − , τ + τ − , and bb̄. CMS observed and excess
of events, with 5.0 standard deviations, at a mass close to 125 GeV, indicating the production
of a new particle. In the decay modes γγ and ZZ the excess was most important; fitting
these signals, the mass obtained was of 125.3 ± 0.4(stat.) ± 0.5(syst.) GeV. On the other hand,
ATLAS experiment performed a search for the SM Higgs boson in proton-proton collisions

with integrated luminosities of approximately 4.8f b−1 collected at s = 7 TeV in 2011 and
CHAPTER 2. NEUTRINOS IN THE STANDARD MODEL 24


5.8f b−1 at s = 8 TeV in 2012. Individual searches in the channels H → ZZ (∗) → 4l, H → γγ
and H → W W (∗) → eνµν in the 8 TeV data were combined with previous results of searches
for H → ZZ (∗) , W W ∗ , bb̄, and τ + τ − in the 7 TeV data and results from improved analysis of
the H → ZZ (∗) → 4l and H → γγ channels in the 7 TeV data. The observation showed a clear
evidence of a boson with mass 126.0 ± 0.4(stat.) ± 0.4(syst) GeV (with 5.9 σ) for ATLAS [32],
and a mass of 125.3 ± 0.4(stat.) ± 0.5(syst) GeV (with 5.8 σ) for the CMS experiment [33].
Finally, it may be important to notice that the behaviors and properties of this particle, so
far as examined since July 2012, are fully compatible with the SM Higgs, which indicates its
discovery.

2.3 Massive Neutrinos


It is currently well established from oscillation experiments that neutrinos have a mass,
although it has not been possible to measure their absolute value yet. The experimental
results give us information about the squared-mass differences and the mixing angles
but we do not exactly know which is the value of the neutrino mass. The best fits of the
current experimental data of the oscillation parameters produces splittings of the order of
∆m221 = 7.6 × 10−5 eV 2 and ∆m231 = 2.5 × 10−3 eV 2 [34]. The origin of the smallness of the
neutrino mass is unknown. As we have already mentioned, the neutrinos in the Standard
Model were proposed to be masless and the physics of models that could explain the neutrino
mass pattern constitutes one of the most active topics in particle physics.

There are plenty of theoretical models that attempt to give an explanation for the smallness
of the neutrino mass. Among them, the most popular is the see-saw-mechanism, which we will
briefly comment later. In what follows, we will give a brief discussion of the different types of
neutrino mass terms.

2.3.1 Dirac and Majorana mass terms

We can start our discussion by showing the well known Dirac Lagrangian

 
µ ∂
L = ψ̄ iγ − mD ψ, (2.53)
∂xµ
where the first term represents the kinetic energy and the second the Dirac mass term, which
has the typical form L = mD ψ̄ψ. This last term, as well as the complete Lagrangian must be
CHAPTER 2. NEUTRINOS IN THE STANDARD MODEL 25

real in order to be Lorentz Invariant (LI) and Hermitian, therefore, mD must be real (m∗D =
mD ). Using the relations (2.8), (2.9), and (2.10), we obtain the following expressions for two
arbitrary spinors ψ and φ [28],

ψ̄L φL = ψ̄PR PL φ = 0, ψ̄R φR = 0, (2.54)

and, therefore,

ψ̄φ = (ψ̄L + ψ̄R )(φL + φR ) = ψ̄L φR + ψ̄R φL . (2.55)

The Dirac term expressed in its chiral components is as follows

L = mD (ψ̄L ψR + ψ̄R ψL ) with ψ̄R ψL = (ψ̄L ψR )† . (2.56)

We can notice that, for producing a mass term, we need both left-handed and right-handed
neutrinos. But, as we saw in the previous subsections, in the Standard Model there are only
left-handed neutrinos, that is why the neutrinos in the SM are massless. It is possible to
have other spinor combinations that behave as Lorentz scalars. If we include a conjugate field
ψ c [28], we can build, for instance, the invariant term ψ̄ c ψ. This invariant (and its hermitical
conjugate) form an additional, Majorana, mass term and is given by

1 1
L= (mM ψ̄ψ c + m∗M ψ̄ c ψ) = mM ψ̄ψ c + h.c. (2.57)
2 2
where mM represents the Majorana mass. In terms of the Weil spinors,

c
ψL,R = (ψ c )R,L = (ψL,R )c , (2.58)

we get two hermitian mass terms:

1 1
LL = c
mL (ψ̄L ψR c
+ ψ̄R ψL ) = c
mL ψ̄L ψR + h.c. (2.59)
2 2
1 1
LR = c
mR (ψ̄L c
ψR + ψ̄R ψL )= c
mR ψ̄L ψR + h.c. (2.60)
2 2

On the other hand, the Majorana fields are written as

c c
φ 1 = ψL + ψR , φ 2 = ψR + ψL (2.61)

where ψL,R are interaction eigenstates and φ1,2 are mass eigenstates. With the previous
elements we can rewrite the expressions (2.59) and (2.60) as
CHAPTER 2. NEUTRINOS IN THE STANDARD MODEL 26

1 1
LL = mL φ̄1 φ1 , LR = mR φ̄2 φ2 . (2.62)
2 2
The combination of the mass terms of Dirac (2.56) and Majorana (2.62) produces a general
expression which is given by

c c c c
2L = mD (ψ̄L ψR + ψ̄L ψR ) + mL ψ̄L ψR + mR ψ̄L ψR + h.c. (2.63)
  
mL mD ψc
= c 
(ψ̄L , ψ̄L )   R  + h.c.
mD mR ψR
= Ψ̄L M ΨcR + Ψ̄cR M ΨL

where,
     
mL mD ψL ψL
M =  ΨL =  =  (2.64)
c
mD mR ψL (ψR )c
with
   
(ψL )c c
ψR
(ΨL )c =  =  = ΨcR . (2.65)
ψR ψR
c
Here ψL and ψR are fields, which could be coupled with left-handed W ± and Z bosons; but the
c
fields ψR and ψL do not have couplings with these bosons. Following [28], we use the notation:
c c c
ψL = ν L , ψR = νR for active neutrinos and ψR = NR , ψL = NLc for sterile neutrinos. The
expression (2.63) becomes then in

2L = mD (ν̄L NR + N̄Lc νRc


) + mL ν̄L νRc
+ mR N̄Lc NR + h.c. (2.66)
  
c
m L m D ν R
= (ν̄L , N̄Lc )    + h.c. (2.67)
mD mR NR

After diagonalizing the M matrix, the following mass eigenstates are obtained:

c c c
ψ1L = cos θψL − sin θψL ψ1R = cos θψR − sin θψR (2.68)
c c c
ψ2L = sin θψL + cos θψL ψ2R = sin θψR + cos θψR (2.69)

with the mixing angle:


CHAPTER 2. NEUTRINOS IN THE STANDARD MODEL 27

2mD
tan θ = , (2.70)
mR − mL
and the mass eigenvalues,

 
1
q
m̃1,2 = 2 2
(mL + mR ) ± (mL − mR ) + 4mD . (2.71)
2
In order to get positive masses, we use m̃k = ǫk mk with mk = |m̃k | and ǫk = ±1(k = 1, 2) [28].
Now we are going to express the Lagrangian in terms of the mass eigenstates, then we define
c
two independent Majorana fields as in (2.61), φk = ψkL + ǫk ψkR , in particular,

c c c
φ1 = ψ1L + ǫ1 ψ1R = cos θ(ψL + ǫ1 ψR − sin θ(ψL + ǫ 1 ψR ) (2.72)
c c c
φ2 = ψ2L + ǫ2 ψ2R = sin θ(ψL + ǫ2 ψR + cos θ(ψL + ǫ2 ψR ); (2.73)

and for the conjugate field,

φck = (ΨkL )c + ǫk ψkL = ǫk (ǫk ψkR


c
+ ψkL ) = ǫk ψk , (2.74)

indicating that ǫk is the CP eigenvalue of the Majorana neutrino described by φk . Finally, an


analogous expression to (2.66) is obtained

2L = m1 φ̄1 φ1 + m2 φ̄2 φ2 . (2.75)

There are four important cases [3, 28]:

1. If mL = mR = 0(θ = 45) we have m1,2 = mD , ǫ1,2 = ∓1 with two degenerate Majorana


states:

1 c c 1
φ1 = √ (ΨL − ψR − ψL + ψR ) = √ (ψ − ψ c ) (2.76)
2 2
1 c c 1
φ2 = √ (ΨL + ψR + ψL + ψR ) = √ (ψ + ψ c ). (2.77)
2 2

With the above Majorana fields, it is possible to construct a Dirac field ψ:

1
√ (φ1 + φ2 ) = ψL + ψR = ψ, (2.78)
2

and the corresponding mass term has the form:


CHAPTER 2. NEUTRINOS IN THE STANDARD MODEL 28

1
L= mD (φ̄1 + φ̄2 )(φ1 + φ2 ) = mD ψ̄ψ. (2.79)
2

From the last expressions we can see that we have a pure Dirac field, thus a Dirac field
can be seen as a field composed of two degenerate Majorana fields. Whereby the Dirac
case may be described as a particular solution of the most general case represented by
Majorana neutrinos.

2. We have another interesting case if mD ≫ mL , mR (θ ≈ 45) : According to the


expression (2.71), the states φ1,2 are almost degenerated with m1,2 ≈ mD , this neutrino
is known as pseudo-Dirac neutrino.

c
3. The case of mD = 0(θ = 0) : The result is m1,2 = mL,R and ǫ1,2 . Then φ1 = ψL + ψR and
c
φ 2 = φ 2 = ψR + ψL , this corresponds to the pure Majorana case.

4. Lastly, if mR ≫ mD , and mL = 0(θ = (mD /mR ) ≪ 1) : Based on (2.71) we get two mass
eigenvalues:

m2 m2D
 
m1 = mν = D and m2 = mN = mR 1 + 2 ≈ mR (2.80)
mR mR

and ǫ1,2 = ∓1. The respective Majorana fields are

c c
φ 1 ≈ ψL − ψL φ 2 ≈ ψL + ψR . (2.81)

Where N is a heavy neutrino and ν is a very light neutrino, this is the well known see-saw
mechanism [35].

In the above discussion, we did focus in one neutrino flavor, for the generalization to n flavors
see [36].
CHAPTER 2. NEUTRINOS IN THE STANDARD MODEL 29

2.4 Neutrino Oscillations


Neutrinos are massive particles. Neutrino flavor states can be described in terms of mixed
massive states. In order to describe this mixing, two bases are defined: the flavor and the
mass base. In a standard three family picture, the neutrinos are produced at the source as
flavor eigenstates that are a combination of three mass eigenstates. Experimentally, it is
observed that a neutrino produced in the source with definite flavor, for example νµ , could
be registered in a detector, after travelling some distance, as a neutrino of flavor νe . This
phenomenon is known as neutrino oscillation as was noted for the first time by Pontecorvo in
1957 [36].
Basically, the idea is as follows: Assume there is a pion beam which decay into muons and
neutrinos, these neutrinos are called muon neutrinos νµ ; additionally, at some distance from
the source, there is a detector which records the neutrino interactions and suppose that an
interaction with an electron is registered in this final state. Since we know that an electron
is produced by νe and not by a νµ , we inmediatly would wonder what has happened?. The
answer is that the muon neutrino was transformed into an electron neutrino while it went
from the source to the detector. How can we explain this transformation?. The answer lies in
the quantum mechanics and we will try to describe it below.
We start by assuming that there are n flavor eigenstates defined by |να i with hνβ |να i = δαβ .
The flavor eigenstates are connected with the mass eigenstates νi (hνi |νj i = δij ) by the unitary
mixing matrix U (see for example [28]):

X X †
X

|να i = Uαi |νi i |νi i = Uiα |να i = Uαi |να i (2.82)
i α α

with
X X
U †U = 1 ∗
Uαi Uβi = δαβ ∗
Uαi Uαj = δij . (2.83)
i α

The matrix U is called the Pontecorvo-Maki-Nakagawa-Sakata (PMNS) mixing matrix. For



antineutrinos we change Uαi by Uαi , then:

X

|ν̄α i = Uαi |ν̄i i. (2.84)
i

If we have n neutrino flavors and n massive neutrinos, the matrix U is a n × n matrix which
is parameterized by n(n − 1)/2 Euler angles and n(n + 1)/2 phases. The number of physical
phases, for Dirac neutrinos, is (n − 1)(n − 2)/2, which are responsible of CP violation in the
lepton sector. For instance, in the case of n = 3, we have one CP violation called the Dirac
CHAPTER 2. NEUTRINOS IN THE STANDARD MODEL 30

phase. However, if we assume that neutrinos are Majorana particles the mixing matrix U
contains n(n − 1)/2 CP violation phases [46]. U can be written in the form

U = V P, (2.85)

where the matrix V contains (n − 1)(n − 2)/2 Dirac phases, and P is a diagonal matrix which
includes the additional (n − 1) Majorana phases α21 , α31 , ...., αn1 ,

α21 α31 αn1


P = diag(1, ei 2 , ei 2 , ...., ei 2 ). (2.86)

In the case of n = 3, there are 3 CP phases, one Dirac and two Majorana.
Let us suppose that neutrinos of a given flavor, α, are produced in a source located at x =
0(t = 0). As we mentioned before, the neutrinos produced in the source, να , are expressed as
coherent superposition of mass eigenstates νi . The neutrinos are emitted with a momentum
p. They will evolve with time as

X X
|ν(x, t)i = Uαi e−iEi t |νi i = ∗ ipx −iEi t
Uαi Uβi e e |νβ i. (2.87)
i i,β

For relativistic neutrinos, p ≫ mi , we will have

m2i m2
q
Ei = m2i + p2i ≃ pi + ≃E+ i, (2.88)
2pi 2E
where E ≈ p is the neutrino energy.
The transition amplitude from a να state to a νβ state after time t is given by

X
A(α → β; t) = hνβ |ν(x, t)i = ∗
Uβi Uαi eipx e−iEi t , (2.89)
i

and using (2.88) we obtain:

m2 L
X  
A(α → β; t) = ∗
Uβi Uαi exp −i i = A(α → β; L), (2.90)
i
2 E
where L = x = t is the distance from the source to the detector. The oscillation probability
will be given by the square modulus of the transition amplitude

XX
P (α → β; L) = |A(α → β; t)|2 = ∗
Uαi Uαj ∗
Uβi Uβj e−i(Ei −Ej )t (2.91)
i j
!
X X ∆m2ij L
= |Uαi |2 |Uβi
∗ 2
| + 2Re ∗
Uαi Uαj ∗
Uβi Uβj exp −i
i j>i
2E
CHAPTER 2. NEUTRINOS IN THE STANDARD MODEL 31

with

∆m2ij = m2i − m2j . (2.92)

In the Eq. (2.91), if L ≫ 2E/∆m2ij , the oscillatory term is averaged and the oscillation
probability corresponds to the first term of (2.91), which is a constant.
In order to have oscillations, it is necessary that the following conditions are satisfied:

• ∆m2ij 6= 0.

• The off-diagonal elements of the U matrix must be non-vanishing in order to have mixing
between neutrinos.

∗ ∗
In the case of antineutrinos, we must make the change Uαj → Uαj and Uβj → Uβj . The
expression for the probability, Eq. (2.91), is not modified with the previous changes and, then,

P (ᾱ → β̄; t) = P (α → β; t). (2.93)

As expected, the oscillation probability is CPT invariant. Moreover, from the unitarity of the
matrix U , we have

X X
P (α → β; t) = P (ᾱ → β̄; t) = 1. (2.94)
β=e,µ,τ β=e,µ,τ

In the case that CP (or T) is conserved, the mixing matrix U becomes real, and the following
relation is satisfied:

P (α → β; t) = P (β → α; t). (2.95)

2.4.1 Oscillations with two neutrino flavors

The most simple case of oscillations corresponds to two neutrinos. We have two flavor
eigenstates, two mass eigenstates and a 2 × 2 mixing matrix. For instance, in the case of
νe and νµ , the unitary transformation (2.82) is given by
    
νe cos θ sin θ ν1
 =   (2.96)
νµ − sin θ cos θ ν2
and the transition probability, Eq. (2.91), is as follows
CHAPTER 2. NEUTRINOS IN THE STANDARD MODEL 32

P (νe → νµ ) = P (νµ → νe ) = P (ν̄e → ν̄µ ) = P (ν̄µ → ν̄e ) (2.97)


∆m2 L
 
2 2
= sin (2θ) × sin = 1 − P (νe → νe ) (2.98)
4E

where the CP invariance is assumed. The oscillation length is defined as the distance for
 2 
which the argument of sin2 ∆m
4E
L
becomes π (see for example ref. [36]),

4πE 2.48(E/M eV )
Losc = 2
≡ m. (2.99)
∆m (∆m2 eV 2 )
Oscillatory behaviour is important when L ∼ Losc .

The case of three generations

The framework of three generations of neutrinos can describe very well all existing data on
neutrinos oscillation with the exception of some anomalies which we will discuss later. In this
case, the mixing matrix U is given by [22]

 
c12 c13 s12 c13 s13 e−iδ
  α α
i 21 i 31
U =  −s12 c23 − c12 s23 s13 eiδ c12 c23 − s12 s23 s13 eiδ s23 c13  × diag(1, e 2 , e 2 ), (2.100)
 
 
s12 s23 − c12 s23 s13 eiδ −c12 s23 − s12 c23 s13 eiδ c23 c13

where cij = cos θij , sij = sin θij , with θij = [0, π/2] as the mixing angle. Likewise, δ = [0, 2π] is
the Dirac CP violation phase and α21 , α31 are two Majorana CP violation phases.
In the case of 3-neutrino families, there are only two independent neutrino mass squared
differences, namely ∆m221 6= 0 and ∆m231 6= 0. The oscillation probability has the form:

3
!
X
2
∆m2ij L
P (να → νβ ) = δαβ − 4 Re(Kαβ,ij ) sin (2.101)
i>j=1
4E
3
! !
X ∆m2ij L ∆m2ij L
+ 4 Im(Kαβ,ij ) sin cos
i>j=1
4E 4E

with

∗ ∗
Kαβ,ij = Uαi Uβi Uαj Uβj . (2.102)

Assuming that ∆m232 = ∆m2atm ∼ 2.5 × 10−3 eV 2 , and ∆m221 = ∆m2sol ≪ ∆m231 ≈ ∆m232 =
∆m2atm , the oscillation probabilities between the different flavors are given by
CHAPTER 2. NEUTRINOS IN THE STANDARD MODEL 33

∆m2atm L
 
Pνµ →ντ = sin2 (2θ13 ) cos2 θ13 sin2 (2.103)
4E
∆m2atm L
 
Pνe →νµ = sin2 (2θ13 ) sin2 θ23 sin2 (2.104)
4E
∆m2atm L
 
2 2 2
Pνe →ντ = sin (2θ13 ) cos θ23 sin (2.105)
4E

2.4.2 Neutrino oscillations in matter

When a neutrino goes through matter, it undergoes to interactions with the particles in the
medium. We can represent this interaction by an effective potential which produces changes
in the masses and mixing angles of the neutrinos. Moreover, when a neutrino cross the matter
may suffer resonant oscillations; when these oscillations take place, the νe ’s may be converted
into νµ ’s (ντ ’s or νx ’s) as they go through the resonance region. This is known as Mikheyev,
Smirnov and Wolfenstein (MSW) effect [131, 3, 28, 37], which we will discuss in the next
sections.
This effect provided a solution to the solar neutrino problem [36], since matter effects into the
Sun on the oscillations of solar neutrinos are quite important. The MSW effect may also play
a significant role in astrophysical environments and we will present an application of it in the
chapter 5.

Evolution equation in matter

Neutrinos of different flavors (νe , νµ , ντ ) can interact by the weak interaction with matter
particles such as protons, neutrons and electrons. This interaction is given by Z bosons
interchange (neutral current interactions) for all flavors, or by W bosons interchange (charged
current interactions) for electron neutrinos.
The effective Hamiltonian describing the neutrino-electron interaction in matter is as follows
[37],

e GF
HC = √ [ēγµ (1 − γ5 )e] [ν̄e γ µ (1 − γ5 )νe ] . (2.106)
2
The effective potential VCC is written as

Z
e
hνe | d~x HC (~x)|νe i. (2.107)

The solution of the last integral is as follows [36],


CHAPTER 2. NEUTRINOS IN THE STANDARD MODEL 34


(Vνe )CC ≡ VCC = 2GF Ne . (2.108)

Here, Ne is the electron number density in the medium. In a similar way, the effective

potential due to neutral current interactions is (Vνα )N C = −GF Nn / 2, where Nn is the
neutron number density and α = e, µ, or τ . Therefore, the effective potential that the neutrino
’feel’ going through the matter is

√ √
   
Nn Nn
Vνe = 2GF Ne − , Vνµ = Vντ = 2GF − . (2.109)
2 2
Since neutrinos and antineutrinos have opposite isospins, the sign of Vνα changes for
antineutrinos [36].
For the case of two generations, the evolution equation in the mass eigenstates basis, in the
absence of matter, has the following form
   
d E1 0 ν1
i |νm i = Hm |νm i, where Hm =  ; |νm i =  , (2.110)
dt 0 E2 ν2
where θ0 is the vacuum mixing angle. On the other hand, the equation of motion for the weak
eigenstates |να i is as follows [36]
    
δm2 δm2
d  |νe i   − 4E cos 2θ0 4E sin 2θ0 |νe i
i =  . (2.111)
dt |νµ i δm2 δm2
4E sin 2θ0 4E cos 2θ0 |νµ i
Turning back to the case of matter effects, we will find that the evolution equation describing
the νe ↔ νµ oscillation is given by [36],

      
2
δm2
d  |νe i  m
|νe i − δm
4E cos 2θ0 + (Vνe )CC 4E sin 2θ0 |νe i
i = MW =  .
dt |νµ i |νµ i δm2
sin 2θ0 δm2
cos 2θ0 |νµ i
4E 4E
(2.112)
In general, the number density Ne depends on the position, but as a first approximation we
will consider the case of constant density. After diagonalizing the Hamiltonian of the eq.
(2.112), the following result is obtained:

|νe i = |ν1m i cos θm + |ν2m i sin θm (2.113)

|νµ i = −|ν1m i sin θm + |ν2m i cos θm ,

where θm is the mixing angle in the matter which is given by


CHAPTER 2. NEUTRINOS IN THE STANDARD MODEL 35

tan 2θ0 ∆m2


tan 2θm = , with VR = cos 2θ0 (2.114)
1 − VVCC
R
2E
or

sin2 (2θ0 )
sin2 (2θm ) =  2 . (2.115)
2
cos2 (2θ0 ) 1 − VVCC
R
+ sin (2θ 0 )
m m
The states |ν1,2 i have energies, E1,2 , such that

" 2 #1/2
∆m2 VCC ∆M 2
E2m − E1m = 1− cos2 2θ0 + sin2 2θ0 ≡ . (2.116)
2E VR 2E
The oscillation probability of νe → νµ in matter, with constant density Ne = constant, is
written as

 
2ν 2 2 L 2π
Pm (νe → νµ ) = sin 2θm sin π , with Lm = . (2.117)
Lm (E2m − E1m )
Here, Lm is the oscillation length in matter. According with (2.115), the oscillation probability
has resonant behavior because of the dependence of sin2 2θm on Ne . The oscillation probability
has a maximum value at VCC = VR when sin2 2θm = 1. As we can notice, the sign of VCC is
crucial; for example, if it is negative, there is no resonant behavior.
The maximum probability is achieved when

√ ∆m2
2GF Ne = cos 2θ0 (2.118)
2E
and this is called the MSW resonance condition.

Adiabaticity Condition

In the last section, we studied the case of constant density, but in general the matter density
in the medium is variable. In a medium with non-uniform density, it is not always possible
to find an analytic solution for the evolution equation (2.112), and, therefore, in most of the
cases we will be forced to use numerical solutions. However, assuming that the matter density
changes slowly (adiabatically) from a high density zone to a low density zone, we can get an
approximate analytic solution. If the density varies, then the mixing angle θm depends on
time. From the equation connecting the flavor and mass eigenstates,
      
|νe i |ν1m i cos θm sin θm |ν1m i
  = U (θm )  =  , (2.119)
|νµ i |ν2m i − sin θm cos θm |ν2m i
CHAPTER 2. NEUTRINOS IN THE STANDARD MODEL 36

and using
     
d  |νe i |ν1m i |ν̇1m i
 = U̇ (θm )   + U (θm )  , (2.120)
dt |νµ i |ν2m i |ν̇2m i
we get the evolution equation for the mass eigenstates in matter [36]
     
|ν̇1m i m
|ν1m i |ν1m i
i  = U † (θm )MW U (θm )   − iU † (θm )U̇ (θm )   (2.121)
|ν̇2m i |ν2m i |ν2m i
or
    
|ν̇1m i E1m (t) −iθ̇m (t) |ν1m i
i =  . (2.122)
|ν̇2m i ˙ (t)
iθ m E2m (t) |ν2m i
In the equation (2.122), we can see that the effective Hamiltonian is not diagonal due to
the mixing angle θm , which is not constant. In a first place, we will take small variations
in the mixing angle (|θ̇| ≪ |E1m − E2m |), this is known as the adiabatic approximation. The
adiabaticity condition can be written as [38],

2
∆m
1 2|θ̇| 2E sin 2θ0
≡ m = |V̇CC | ≪ 1. (2.123)
γ |E1 − E2m | |E1m − E2m |3
Then, we could have adiabatic conversions when the following condition is satisfied

γ ≫ 1. (2.124)

In order to understand in more detail the adiabatic case, we use the following description: the
flavor eigenstates in terms of the mass eigenstates are given by

|νe i = |ν1m i cos θm + |ν2m i sin θm (2.125)

|νµ i = −|ν1m i sin θm + |ν2m i cos θm (2.126)

or

|ν1m i = |νe i cos θm − |νµ i sin θm (2.127)

|ν2m i = |νe i sin θm + |νµ i cos θm . (2.128)

Consider now an electron neutrino, νe , which is produced in matter with a very high density
(VCC ≫ VR ), the neutrino state is described by the equation (2.125). According with the
CHAPTER 2. NEUTRINOS IN THE STANDARD MODEL 37

equation (2.115), the mixing angle θm is approximately 90o . This means that |νe i is practically
a pure state |ν2m i, namely, |νe i is born as |ν2m i. Then the neutrino propagates towards the
regions with smaller density and the mixing angle (θm ) decreases, in the resonance point
(VCC = VR ), the mixing angle is maximal (θm = 45o ). As the neutrino goes through the matter
declining density, θm continues decreasing until reaching the value θm = θ0 when VCC ≪ VR .
The final result is that a neutrino in a state |ν2m i is converted into ν2m = νe sin θ0 + νµ cos θ0 and
the transition probability is

P (νe → νµ ) = cos2 θ0 . (2.129)

In this way, may there be an adiabatic conversion from νe to νµ for small values of the vacuum
mixing angle.

Non-Adiabatic Process

We have another case when the medium density does not vary slowly. In this scenario,
|θ̇| ∼ |E1m − E2m |, and according to the expression (2.122), it is possible to have transitions
between the eigenstates |ν1m i and |ν2m i. These transitions are due to the off-diagonal elements
in eq. (2.122) which are no longer negligible. The process is known as a non-adiabatic
conversion. Again, considering the case of two generations, an electron neutrino born in a
pure state |ν2m i, reaches the resonance region, but now it is converted from |ν2m i to |ν1m i with
a conversion probability:

 π 
PLZ = exp − γ . (2.130)
4
This expression is known as the Landau-Zener Probability [36]. As we mentioned before
γ ≫ 1 for an adiabatic process and γ ≪ 1 for a non-adiabatic process. As a result, PLZ = 0 and
PLZ = 1 respectively.
On the other hand, the survival probability for a non-adiabatic process is given by

1
P (νe → νe ) = [1 + (1 − 2PLZ ) cos 2θ0 cos 2θm ] . (2.131)
2
In this work, we have dealt with the most simple case of oscillations which corresponds to
two generations of neutrinos. A treatment with three families neutrino may be found in the
references [36] and [38].
Chapter 3

Constraints on µν from
experimental data

In the Standard Model the neutrino was proposed to be massless. However, neutrino
oscillations experiments clearly show that neutrinos have mass and are also mixed (see for
example [93]). This is an important fact because we have to extend the Standard Model
to explain the neutrino masses. In several extensions of the Standard Model, neutrinos
also acquire electromagnetic properties through quantum loops, such as charge, charge
radius, anapolar moment, magnetic and electric dipole moments (see for example [39] for
a nice discussion on this topic). In this chapter, we are going to study theoretical and
phenomenological aspects of the magnetic moment of the neutrino. The first calculations
of the neutrino dipole moments within the minimal extension of the SM with right-handed
neutrinos were performed in [40, 41].

3.1 Magnetic moment in the minimally extended


standard model
For charged fermions, the interaction Lagrangian between a fermion and the electromagnetic
field contains the term:

Lint = eQψγα ψAα , (3.1)

where e is the positron charge, and Q is the charge of the particle (in units of e) involved

38
CHAPTER 3. CONSTRAINTS ON µν FROM EXPERIMENTAL DATA 39

in the interaction [39]. In the case of neutrinos, the previous term is absent and, therefore,
the interaction arises only from radiative corrections. For uncharged particles such as the
neutrino, we can write the effective interaction in analogy with this term as follows

Lef f = ψOα ψAα = jα Aα . (3.2)

The next step is to determine the form of this interaction. Taking the matrix element of Lef f
between two one-particle states, we get

hp′ , s′ |jα |p, si = us′ (p′ )Γα (q)us (p), (3.3)

where q = p − p′ . From the hermicity conditions and taking into account the electromagnetic
current conservation [1], we can write the most general form for Γα , which is also consistent
with Lorentz covariance,

 
/q
fQ (q 2 ) + fA (q 2 )γ5 − σαρ q ρ fM (q 2 ) + ifE (q 2 )γ5 ,
   
Γα (q) = γ α − qα 2 (3.4)
q
where fM (q 2 ) is called the magnetic form factor, fE (q 2 ) is the electric form factor, fQ (q 2 ) is
the charge form factor, and fA (q 2 ) is the anapole form factor. In particular, fM (0) = µ is the
magnetic dipole moment, fE (0) = d is the electric dipole moment. In the most general case,
the previous form factors are represented by matrices.

3.1.1 The case of Dirac neutrinos

The main contribution to the neutrino magnetic moment in the minimal extension of the
Standard Model comes from one-loop Feynman diagrams which are shown in the Fig. (3.1).
The evaluation of these diagrams leads to the transition matrix [42],

T = −iǫµ q ν ν̄j (p′ )σµν (A + Bγ5 )νi (p), (3.5)

where q = p′ − p is the transferred momentum and ǫ is the polarization,

 2 !
eGF X † ml
A = − √ (mi + mj ) Uli Ujl F , (3.6)
8 2π 2 mW
l
 2 !
eGF X † ml
B = − √ (mi − mj ) Uli Ujl F , (3.7)
8 2π 2 mW
l

with
CHAPTER 3. CONSTRAINTS ON µν FROM EXPERIMENTAL DATA 40

3 x2 log x
 
1 3 2
F (x) = − (2 − 5x + x ) + (3.8)
(1 − x)2 4 2 1−x
3 3
≃ − + x, (x ≪ 1). (3.9)
2 4

Now we use the general expression for the transition amplitude in terms of the form factors,

T = −iǫµ ν̄(p′ )[F2 (q 2 ) − G2 (q 2 )γ5 ]σµν q ν ν(p), (3.10)

then, the magnetic moment is defined by

µ = F2 (0) = A (3.11)

and, in a similar way, the electric dipole moment is

d = G2 (0) = B. (3.12)

In this work we will not consider the electric moment; therefore, we take B = 0.

γ γ

ℓ ℓ W W

ν ν ν ν
W ℓ

Figure 3.1: Radiative corrections giving rise to a magnetic dipole moment in the minimally
extended standard model.

Then, for a Dirac neutrino we have diagonal and transition magnetic moments [39]. The
diagonal magnetic moment (i = j) is given by
 
 2
3eGF mi  1 X ml
µD
ii ≃ √ 1− |Uli |2  , (3.13)
8 2π 2 2 m W
l=e,µ,τ

where the second term is negligible. We can notice that, in the minimal extension of the
Standard Model, the magnetic moment is directly proportional to the neutrino mass,

m 
i
µD
ii ≃ 3.2 × 10
−19
µB , (3.14)
eV
CHAPTER 3. CONSTRAINTS ON µν FROM EXPERIMENTAL DATA 41

e
with µB = 2me the Bohr magneton. For the transition moment (i 6= j), the contribution from
the first term of (3.9) vanishes in the summation over l due to the orthogonality condition of
U . Therefore, we are left with the term proportional to 34 x, which is given by

3eGF X  ml 2
µD
ij ≃ √ (mi + mj ) Uli∗ Ulj . (3.15)
32 2π 2 mW
l=e,µ,τ

With the evaluation of the constants we obtain


mi + m j
 X  ml  2
µD
ij ≃ 4 × 10 −23
µB Uli∗ Ulj . (3.16)
eV mτ
l=e,µ,τ

From the equations (3.16) and (3.13), we observe that the transition magnetic moment is more
suppressed than the diagonal moment.

3.1.2 The case for Majorana neutrinos

In this case, the Feynman diagrams are similar to those of Fig. (3.1), but now we have
both a charged lepton and an antilepton propagating in the internal loop. Consequently, the
transition matrix is given by the following expression[42],

−iǫµ ν̄j (p′ ) [F2 (q 2 )ji − F2 (q 2 )ij ] − [G2 (q 2 )ji − G2 (q 2 )ij ]γ5 σµν q ν νi (p)

Tji = (3.17)

= −iǫµ ν̄j (p′ )[2iImF2 (q 2 )ji − 2ReG2 (q 2 )ij γ5 ]σµν q ν νi (p). (3.18)

From the equation (3.17), we can see that µM


ii = 0. Hence, Majorana neutrinos do not have

diagonal magnetic moments. For the transition moment, it is possible to prove that

µM D
ij = 2µij (3.19)

Provided that the relative CP phase of νi and νj is odd [1, 42].

3.1.3 Parametrization of the Neutrino magnetic moment

In the last section, we saw that the form of the neutrino magnetic moment is different
depending whether we have Majorana or Dirac Neutrinos. We can express these magnetic
moments in a more general way using matrix notation. For our purposes, we are going to use
the parametrization of the previous matrices in both the flavor and the mass basis as in [43].
CHAPTER 3. CONSTRAINTS ON µν FROM EXPERIMENTAL DATA 42

Matrix of magnetic moments for Dirac neutrinos

Based on the equations (3.13) and (3.15), the matrix of magnetic moments in the flavor and
mass basis are defined as follows
   
λee λeµ λeτ λ11 λ12 λ13
   
λαβ =  λeµ λµµ λµτ  , λij =  λ12 λ22 λ23  . (3.20)
   
   
λeτ λµτ λτ τ λ13 λ23 λ33
The left matrix is written in the flavor basis and the right one in the mass basis. As it
was referred above, for Dirac neutrinos we have diagonal and transition magnetic moments,
which means that, when a Dirac neutrino interacts with an electromagnetic field, could either
changes its flavor or remains the same.

Matrix of magnetic moments for Majorana neutrinos

In the case of Majorana neutrinos, taking into account the antisymmetric properties of the
transition matrix, we can write it in terms of the Levi-Civita tensor, in the flavor basis: λαβ =
εαβγ Λγ , while in the mass basis we will have λ̃ = λjk = εjkl Λl [43]. Thus,
   
0 Λτ −Λµ 0 Λ3 −Λ2
   
λ =  −Λτ 0 Λe  λ̃ =  −Λ3 0 Λ1  (3.21)
   
   
Λµ −Λe 0 Λ2 −Λ1 0
The magnetic moment is defined in the mass and flavor basis as a vector of three components:
Λ = (Λα ) and Λ̃ = (Λj ) respectively. In our discussion, we assume that neutrinos have
Majorana nature as motivated from different theories beyond the Standard Model [44, 45, 46].

3.2 Neutrino-electron elastic scattering


We will consider the elastic scattering ν +e− → ν +e− of an antineutrino carrying an energy Eν
with an electron at rest in the laboratory frame. We have two contributions to the scattering
cross section: one which is derived from the Standard Model, and the other one due to the
electromagnetic interaction,

   
dσ dσ dσ
= + . (3.22)
dT dT SM dT µ
In the first place, we are going to derive the weak-interaction cross section. The physical
process is showed in the Fig. (3.2), the interaction is produced by charged and neutral current.
CHAPTER 3. CONSTRAINTS ON µν FROM EXPERIMENTAL DATA 43

ν¯e ν¯e
ν¯e ν¯e
W−
NC Z0 +

e− CC e−

e− e−

Figure 3.2: ν̄ − e scattering. Taken from [1].

The scattering amplitude is obtained by applying the Feynman rules in the diagram of the
Fig. (3.2).

GF  µ
ēγ (1 − γ 5 )νe [ν̄e γµ (1 − γ5 )e]

M = √
 2 µ
ν̄e γ (1 − γ 5 )νe [ēγµ (gV − gA γ5 )e] ,

+

with the standard coupling constants gV and gA given by


 2sin2 θ + 1/2 f or νe ,
W
gV = (3.23)
 2sin2 θ − 1/2 f or νµ , ντ
W

 1/2 f or νe ,
gA = . (3.24)
 −1/2 f or ν , ν
µ τ

Rearranging terms, we will have,

GF 
M = √ ν̄e γ µ (1 − γ 5 )νe [ēγµ (gV′ − gA


γ5 )e] , (3.25)
2
where gV′ = gV + 1 and gA

= gA + 1.
Now we use the Fermi golden rule,

" ! !#
ν̄−e ~|M |2 d3 p~′ν d3 p~′e
dσ = q × (2π)4 δ 4 (pν + pe − p′ν − p′e )
2 2 (2π)3 2E3 (2π)3 2E4
4 (pν · pe ) − (me mν )
(3.26)
CHAPTER 3. CONSTRAINTS ON µν FROM EXPERIMENTAL DATA 44

We should also consider that, after the corresponding computation we will obtain,

|M |2 = 64G2F gL
 2 2
(pν · pe )(p′ν · p′e ) + gR (pν · p′e )(pe · p′ν ) − me gL gR (pν · p′v ) ,

where gL = 12 (gV − gA ), gR = 12 (gV + gA ). Writing the cross section as a function of Lorentz


transformations invariants, known as Mandelstam’s variables, we have:

u = (pν − p′e )2 = (pe − p′ν )2 , s = (pν + pe )2 = (p′ν + p′e )2 , t = (pν − p′ν )2 . (3.27)

If we work in the Laboratory frame, we will have,

u = m2e − 2me Eν′ , s = m2e − 2me Eν , t = 2me (Eν′ − Eν ), (3.28)

and using the result [42]:

d3 p~′ ν d3 p~′ e 1 dEe′


Z Z
4 4 ′ ′
3 ′ 3 ′
(2π) δ (p ν + p e − p ν − p e ) = , (3.29)
(2π) 2Eν (2π) 2Ee 8π Eν
the final expression for the cross section ν̄ − e is as follows,

2me G2F 2
     
dσ 2 T me T
= gR + gL 1 − − gL gR 2 (3.30)
dT SM π Eν Eν
Where,


 sin2 θ + 1/2 f or ν̄e ,
W
gR = (3.31)
 sin2 θ − 1/2 f or ν̄µ , ν̄τ
W
n
gL = sin2 θW f or all f lavors. (3.32)

For neutrinos, we exchange gL and gR .

Electromagnetic cross section

The electromagnetic interaction is represented in the Fig. (3.2), and the scattering amplitude
is given by [47]

ie 
ν̄(k2 )µ12 σαβ q β ν(k1 ) × [ē(p2 )γ α e(p1 )] ,

Mem = 2
(3.33)
q
i
where σαβ = 2 (γα γβ − γβ γα ), q β = (k2 − k1 )β . Adding over spins, we have

X 1 e2 2 β η
|Mem |2 = µ q q T r [k/2 σαβ k/1 σρη ] × T r [p/2 γ α p/1 γ ρ ] , (3.34)
spins
4 q 4 12
CHAPTER 3. CONSTRAINTS ON µν FROM EXPERIMENTAL DATA 45

ℓ ℓ

ν ν

where,

T r [k/2 σαβ k/1 σρη ] = −4 {k2α k1ρ gβη − k2α k1η gβρ + k2β k1η gαρ − k2β k1ρ gαη (3.35)

+ (k2 · k1 ) (gαη gβρ − gαρ gβη ) + k2ρ k1α gβη − k2ρ k1β gαη + k2η k1β gαρ − k2η k1ρ gβρ }

ρ
T r [p/2 γ α p/1 γ ρ ] = 4 (pα
2 p1 − (p2 · p1 )g
αρ
+ pρ2 pα
1)

As a result:

X 2e2 µ212
|Mem |2 = {[(p′ν · pe ) + (pν · pe )] × [(p′ν · p′e ) + (pν · p′e )] − (p′e · pe )(p′ν · pν )} . (3.36)
spins
p′ν · pν

In terms of the Mandelstam’s Variables,


( 2 )
4e2 µ212 s − m2e u − m2e t2 m2 t
X  
|Mem | = 2
− + − e . (3.37)
spins
t 2 2 4 2

Using the result (3.28), we obtain:

2e2 µ2ν 4m2e Eν Eν′


X  
2 3
|Mem | = + me . (3.38)
spins
me Eν − Eν′

Here, we can take two different approaches: the first one corresponds to the standard
approximation m3e = 0, which is considered by most authors, and the other one, would be
to consider m3e 6= 0. In the first case, the expression for the electromagnetic cross section (in
the laboratory frame L) is the

2
πα2
   
dσ 1 − T /Eν µν
= 2 , (3.39)
dT EM me T µB
where T = Ee′ − me = Eν − Eν′ is the recoil energy of neutrino, α is the fine structure constant
e
and µB = 2me is the Bohr’s Magneton.
CHAPTER 3. CONSTRAINTS ON µν FROM EXPERIMENTAL DATA 46

If we consider m3e 6= 0, we obtain:

2
πα2
    
dσ 1 − T /Eν me µν
= + . (3.40)
dT EM m2e T 4Eν2 µB

2.0<Eν<8.0 MeV
(dσ/dT)(10 cm MeV )
4
-1

10
3
10
2
10
1
10
2

0
10
-1
10
-45

-2
10
-3
10
-4
10
-5
10 Modelo Estandar
-6 -10
10
-7
EM(µν=10 µB)
10
-8
10 -2 -1 0 1
10 10 10 10
T(MeV)
Figure 3.3: Contribution of the SM and electromagnetic to the cross section ν̄ − e as function
of the recoil electron.

The scattering cross section with the Standard Model and electromagnetic contributions is
shown in Fig. (3.3). As we can see the electromagnetic cross section is bigger than the
Standard Model one at low energies. In order to consider the effect of the term m3e , we have
taken, as an example, neutrinos coming from the reaction 7 Be + e− → 7
Li + νe + 0.862 MeV
in the Sun. The result is shown in the Fig. (3.4), where we can notice that the contribution of
the term with m3e 6= 0 is negligible in this case.
CHAPTER 3. CONSTRAINTS ON µν FROM EXPERIMENTAL DATA 47

Seccion Eficaz (Eν=0.862MeV)


4
10

(dσ/dT)em(10 cm MeV )
-1
dσ/dT
3
3 +m e
2 10

2
10
-45

1
10

0
10

-1
10 -3 -2 -1 0
10 10 10 10
T(MeV)

Figure 3.4: Cross section ν − e with and without the term of m3e , for µν = 1. × 10−10 µB .

3.2.1 Effective magnetic moment

In ν − e scattering experiments the neutrino is created as a flavor neutrino, this neutrino is a


linear combination of the mass eigenstates. Hence, the magnetic moment which is measured
in these experiments is an effective magnetic moment that takes into account the neutrino
mixing [1]. For that reason, it is necessary to generalize the expression of the magnetic
moment to consider these effects. In this sense, the expression for the electromagnetic cross
section (3.39) in terms of the effective magnetic moment is as follows

πα2
   
dσ 1 − T /Eν
= 2 2 µ2ef f , (3.41)
dT EM me µ B T
where [43] ,

† † †
(µF 2 †
ef f ) = a− λ λa− + a+ λλ a+ . (3.42)

Here, a− and a+ are the negative and positive helicity amplitudes in the detector and λ = µ−id
is the transition moment matrix, with µ as the matrix of magnetic moments and d = 0 for our
purposes. The λ matrix expressed in the mass and flavor basis is given by the expressions
(3.20) and (3.21) for Dirac and Majorana neutrinos respectively.
The effective magnetic moment, Eq. (3.42), does not depend on the chosen basis. In what
follows, we will consider both the flavor basis and the mass eigenstate basis. In this case,
a± and λ = (λαβ ) denote the quantities in the flavor basis and ã± and λ̃ = (λjk ) in
CHAPTER 3. CONSTRAINTS ON µν FROM EXPERIMENTAL DATA 48

the mass basis. Likewise, in the case of oscillations, the two basis are connected by the
Pontecorvo-Maki-Nakagawa-Sakata matrix (PMNS), denoted as U [43],

ã− = U † a− , ã+ = U T a+ , λ̃ = U T λU. (3.43)

In a similar way to Eq. (3.42), the expression for the effective magnetic moment in the mass
basis is written as

† † †
(µM 2 †
ef f ) = ã− λ̃ λ̃ã− + ã+ λ̃λ̃ ã+ , (3.44)

where λ̃ is represented by the expression on the right side of the Eq. (3.21).

3.3 Experimental measurement of the neutrino magnetic


moment
The usual method used for the experimental test of the neutrino magnetic moment is through
the neutrino-electron scattering proccess. The measurements are performed with low energy
neutrinos in accelerator, reactor and solar neutrinos. Studies have been carried out around
the globe for almost 40 years, always with the expectation of observing a value of the magnetic
moment above the theoretical prediction of the Standard Model, (3.14). We present a summary
of the experiments performed with this purpose since 1976 until today:

• The scattering ν̄ − e was observed at first time by Savannah River Laboratory(SRL) in


1976 by Reines, Gurr, Sobel [48].

• In 1989, Vogel and Engel carried out an analysis of SRL’s data, and they calculate the
limit µνe ≤ (2 − 4) × 10−10 µB [49].

• In 1991, the Krasnoyarsk experiment in Russia obtained the new limit µνe ≤ 2.4 ×
10−10 µB [50].

• In 1993, the LAMPF experiment in Los Alamos National Laboratory got the limits µνe ≤
10.8 × 10−10 µB , µνµ ≤ 7.4 × 10−10 µB [51].

• In 1993, the Rovno experiment in Ukraine obtained the limit µνe ≤ 1.9 × 10−10 µB [52].

• In 2001, the LSND collaboration reported the limits µνe ≤ 1.1 × 10−9 µB , µνµ ≤ 6.8 ×
10−10 µB [53].
CHAPTER 3. CONSTRAINTS ON µν FROM EXPERIMENTAL DATA 49

• In 2005, the MUNU experiment obtained the limit µνe ≤ 9 × 10−11 µB [54].

• In 2010, the TEXONO collaboration got the limit µνe ≤ 2.2 × 10−10 µB [55].

• In 2012, the GEMMA experiment in Russia reported the limit µνe ≤ 2.9 × 10−11 µB [56].

3.3.1 Neutrino Experiments

The previous experiments, among others, have measured the neutrino-electron scattering and
also have calculated limits to the neutrino magnetic moment. As we have already mentioned,
these experiments include mainly reactor, accelerator and solar neutrinos. In what follows we
give a brief explanation of these different techniques.

Effective magnetic moment for reactor neutrinos

This type of experiments use a nuclear power plant as an antineutrino source. The
antineutrinos produced are detected by the antineutrino-electron interaction in a detector
located at a given distance L known as baseline. The production mechanism is the nuclear
fission, where the nucleus of an atom splits into smaller parts (lighter nuclei), this splitting
produces free neutrons and gamma rays releasing a large amount of energy. The free neutrons
collide producing other nuclei or decay by n → p+ +e− + ν̄e , this process is known as Beta decay.

Figure 3.5: Nuclear Fission. Taken from [2]

235 238 239 241


The fission of four isotopes: U, U, P u, and P u produces most of the power of the
reactor. The antineutrino flux decreases rapidly with distance which represents a problem
for neutrino oscillation experiments. From the previous list, Krasnoyarsk, Rovno, MUNU,
TEXONO, GEMMA are reactor experiments.
CHAPTER 3. CONSTRAINTS ON µν FROM EXPERIMENTAL DATA 50

For this type of experiments we, first of all, calculate the effective magnetic moment in the
flavor basis. For this purpose, we use the expressions (3.42) and (3.21) for the form of the
matrix of magnetic moments in the Majorana case. We can note that the result depends on
the kind of neutrinos in the experiment, for instance, in reactor experiments, the neutrinos
produced are electron-antineutrinos, that is why the helicity amplitudes are the following
   
1 0
   
a+ =  0  , a− =  0  . (3.45)
   
   
0 0
Then, substituting the expressions (3.45) and (3.21) in (3.42), we obtain

(µF 2 2 2
R ) = |Λµ | + |Λτ | . (3.46)

Note that, in this case, the effective magnetic moment does not depend on the Λe component.
For that reason, if we take into account only reactor neutrinos, the result is not complete. For
this reason, if we would like to obtain a more detailed analysis, we would need to include and
combine reactor, accelerator, and solar neutrino experiments.

In a similar way, we calculate the term for the effective moment in the mass basis, using the
relation (3.44), the result obtained is the following

(µM
R )
2
= |Λ|2 − s212 c213 |Λ2 |2 − c212 c213 |Λ1 |2 − s213 |Λ3 |2 (3.47)

− 2s12 c12 c213 |Λ1 ||Λ2 |cosξ − 2c12 c13 s13 |Λ1 ||Λ3 |cos(β − δ)

− 2s12 c13 s13 |Λ2 ||Λ3 |cos(β − ξ − δ).

We can see that the effective moment in the mass basis depends on the mixing angles, θ12 and
θ13 , on the Dirac phase, δ, and on ξ = ϕ2 − ϕ1 , β = ϕ3 − ϕ1 , which are the relative phases
introduced by the presence of the magnetic moment. The expression (3.47) includes the mixing
angle θ13 , which was measured by experimentos such as Daya Bay in China, RENO in Korea
and Double Chooz in France [57, 58, 59]. Here we would like to stress that this result, that
includes the mixing angle θ13 , has never been computed before, as far as we know. The closest
reference could be that of some of our collaborators [43], where the mixing angle θ13 was
CHAPTER 3. CONSTRAINTS ON µν FROM EXPERIMENTAL DATA 51

considered to vanish. In this sense, our result showed in (3.47), can be considered as entirely
new.

Effective magnetic moment for accelerator neutrinos

In this kind of experiments, a particles beam, commonly protons is hit with a target. The
collision produces pions and kaons which decay giving rise to muon and antimuon neutrinos.
There is a magnetic horn which focus the pions and kaons, depending of the polarity of the
horn, the neutrinos beam is composed by νµ or ν̄µ . In the case of a νµ beam, the decays are as
follows:

π + , K + → µ+ + ν µ , µ+ → e+ + νe + ν̄µ (3.48)

The contribution of ν̄µ and νe in the second process is about 1% [3]. The typical energy of
the neutrinos is of the order of a few GeV, but depending on the energy of the protons it may
increase.
In this case, there are three different interactions in the detector, νe e− → νe e− , νµ e− → νµ e−
and ν̄µ e− → ν̄µ e− . In order to calculate the effective magnetic moment for the electron and
muon neutrino, we separate the distinct cases in the following way:

1. Electron neutrino-electron scattering (νe e− → νe e− ). For this process, the helicity


eigenstates in the detector are the following:

   
1 0
   
a− =  0  , a+ =  0  , (3.49)
   
   
0 0

then, using the expressions (3.42) and (3.21), we obtain for the effective magnetic
moment in the flavor basis,


(µνFe e )2 = |Λτ |2 + |Λµ |2 . (3.50)

Moreover, the electromagnetic interaction does not distinguish between neutrinos and
antineutrinos. Therefore, the expression for the effective magnetic moment in the mass
basis is the same as for reactor experiments,
CHAPTER 3. CONSTRAINTS ON µν FROM EXPERIMENTAL DATA 52


νe e 2
(µM ) = |Λ|2 − s212 c213 |Λ2 |2 − c212 c213 |Λ1 |2 − s213 |Λ3 |2 (3.51)

− 2s12 c12 c213 |Λ1 ||Λ2 | cos ξ − 2c12 c13 s13 |Λ1 ||Λ3 | cos(β − δ)

− 2s12 c13 s13 |Λ2 ||Λ3 | cos(β − ξ − δ),

with δ, the Dirac phase and ξ = ϕ2 − ϕ1 , β = ϕ3 − ϕ1 , the relative phases. For θ13 = 0,
we obtain:


νe e 2
(µM ) = |Λ|2 − c212 |Λ1 |2 − s212 |Λ2 |2 − 2c12 s12 |Λ1 ||Λ2 | cos ξ (3.52)

2. Muon neutrino-electron scattering (νµ e− → νµ e− ). In this case, the helicity eigenstates


in the detector are given by

   
0 0
   
a− =  1  , a+ =  0  , (3.53)
   
   
0 0

In a similar way to the case 1, we obtain for the effective magnetic moment in the flavor
basis,

ν e− 2
(µFµ ) = |Λτ |2 + |Λe |2 , (3.54)

and the corresponding expression in the mass basis is

ν e− 2
(µMµ ) = c223 (c212 − 2c12 s12 s13 cos δ + s212 s213 + c213 )|Λ1 |2

+ c223 (s212 + 2c12 s12 s13 cos δ + c212 s213 + c213 )|Λ2 |2 + c223 (1 + s213 )|Λ3 |2

2c223 s12 c12 c213 |Λ1 ||Λ2 | cos ξ + s13 |Λ1 ||Λ2 |(c212 cos(ξ − δ) − s212 cos(ξ + δ))

+

+ c13 |Λ1 ||Λ3 |(s12 cos β + c12 s13 cos(β − δ))

+ c13 |Λ2 ||Λ3 |(s12 s13 cos(γ − δ) − c12 cos γ)] . (3.55)

For θ13 = 0, the result is the following

ν e− 2
(µMµ ) = c223 (1 + c212 )|Λ1 |2 + c223 (1 + s212 )|Λ2 |2 + c223 |Λ3 |2 + 2c223 [s12 c12 |Λ1 ||Λ2 | cos ξ

+ s12 |Λ1 ||Λ3 | cos β − c12 |Λ2 ||Λ3 | cos γ] , (3.56)


CHAPTER 3. CONSTRAINTS ON µν FROM EXPERIMENTAL DATA 53

where θ12 , θ13 and θ23 are the mixing angles of the PMNS matrix. The results (3.55)
and (3.56) are also completely new.

3. Antimuon neutrino-electron scattering (ν̄µ e− → ν̄µ e− ). In this case, the expressions


obtained for the effective magnetic moment in the flavor and mass basis are the same as
the muon neutrino case.

Solar Neutrinos

The nuclear reactions in the Sun produce electron neutrinos with energies of the order of 1
MeV. The estimation for the neutrino flux from the sun is around 6 × 1010 cm−2 s−1 [3]. As we
showed in some previous pictures the ν − e cross section is very small ∼ 10−45 cm2 M eV −1 ,
then although the flux is large, it is challenging to detect neutrinos produced in the Sun. For
that reason, there have been strong efforts to built detectors located underground, where
the rock is a shielding for charged particles such as muons. The first experiment which
detected neutrinos from the Sun was the Homestake experiment in 1960’s. After that, other
experiments came, such as Kamiokande, GALLEX/GNO and SAGE, Super-Kamiokande,
SNO and Borexino [28].

Energy production in the Sun Core

In the Sun, two main thermonuclear reactions are carried out: the pp chain and the CNO
cycle. They are displayed in the Figs. (3.6) and (3.7), respectively [28, 3]. These reactions
convert protons and electrons into 4 He nuclei by the following process

4p + 2e− →4 He + 2νe + Q (3.57)

Here, Q, is the released energy, which is [3],

Q = 4mp + 2me − m4 He = B(4, 2) + 2me − 2(mn − mp ) = 26.731 M eV, (3.58)

mp , mn , and me are the masses of the proton, neutron and electron, respectively.
One way to get information about the solar interior is by the electron neutrinos produced
there. The neutrino flux as a function of the energy is shown in Fig. (3.8).
CHAPTER 3. CONSTRAINTS ON µν FROM EXPERIMENTAL DATA 54

(pp) p + p → 2 H + e+ + νe p + e− + p → 2 H + νe (pep)

❳❳❳ ✘✘
99.6% ❳❳❳ ✘✘✘ 0.4%
❳❳❳ ✘✘✘
❳❳
❳✘✘✘✘

2
H + p → 3 He + γ

✘✘❳❳❳❳
✘✘✘ ❳❳❳
✘✘✘✘✘ ❳❳❳
✘✘✘ ❳❳❳
85% ✘ ✘✘✘ ❳❳❳ 2 × 10−5 %

❄ ❄
3
He + 3 He → 4 He + 2 p 3
He + p → 4 He + e+ + νe
15%
ppI ❄ (hep)
3
He + 4 He → 7 Be + γ

✏##
✏✏ ##
✏✏ ##
99.87% ✏✏✏✏ ## 0.13%
#
❄ ❄
(7 Be) 7
Be + e− → 7 Li + νe 7
Be + p → 8 B + γ

❄ ❄
7
Li + p → 2 4 He 8
B → 8 Be∗ + e+ + νe (8 B)

ppII

8
Be∗ → 2 4 He

ppIII

Figure 3.6: The pp chain of thermonuclear reactions in the Sun. Taken from [3].

12
C + p → 13 N + γ ✲ 13 N → 13 C + e+ + ν (13 N)
e


✛✘
✲ ❄
15
N + p → 12 C + 4 He CN ❄ 13 C + p → 14 N + γ
✚✙
✻ ✛

99.9%

15
( O) 15
O→ 15
N + e+ + νe ✛ 14
N + p → 15 O + γ

0.1%

15
N + p → 16 O + γ 17
O + p → 14 N + 4 He


16
O + p → 17 F + γ ✲ 17 F → 17 O + e+ + ν (17 F)
e

Figure 3.7: The CNO cycle of thermonuclear reactions in the Sun. Taken from [3].

Effective magnetic moment for solar neutrinos

Finally, we will get the expression for the effective magnetic moment in the mass basis for the
case of solar neutrinos. According to the equation (3.44)

j † †
(µM 2 k
ef f ) = (ã− ) (λ̃ λ̃)jk (ã− ). (3.59)
CHAPTER 3. CONSTRAINTS ON µν FROM EXPERIMENTAL DATA 55

Figure 3.8: Neutrino flux from the pp and CNO chains as a function of the neutrino energy.
Taken from [3].

Since there are not antineutrinos in the detector, ã+ = 0. From the Eq. (3.21) we get

(λ̃† λ̃)jk = |Λ|2 δjk − Λ∗k Λj . (3.60)

In the case of solar neutrinos, the helicity state in the source is given by aT− = (1, 0, 0).
The neutrinos develop oscillations from the Sun to the detector and their helicity state will
change. Therefore, the effective magnetic moment depends on the oscillation probability as
follows [43],

(µM
ef f )
2
= (ãj− )∗ (|Λ|2 δjk − Λ∗k Λj )(ãk− ) (3.61)
X X j

= Pej |Λ|2 − h(ã− )∗ (ãk− )iΛ∗k Λj .
j jk

The probability that a neutrino state produced in the Sun, created initially as an electron
neutrino, νe , will contain a mass eigenstate component νj is


Pej = h|ãj− |2 i (j = 1, 2, 3). (3.62)

Taking sin θ13 = s13 , cos θ13 = c13 , and Ue3 = s13 , the following result is obtained [43]
CHAPTER 3. CONSTRAINTS ON µν FROM EXPERIMENTAL DATA 56


Pe3 = s213 , and 3ν
Pej = c213 Pej

(j = 1, 2) 2ν
with Pe1 2ν
+ Pe2 = 1. (3.63)

Using the above expressions, we obtain:

h(ãj− )∗ (ãk− )iΛ∗k Λj


X
(µM 2 2
ef f ) = |Λ| − (3.64)
jk

2
h(ãj− )∗ (ãk− )iΛ∗k Λj
X
(µM
ef f )
2
= |Λ| − 2
h|ã3− |2 i|Λ3 |2 − (3.65)
j,k=1

and finally,

2
X
(µM 2 2 2 2 2 2
ef f ) = |Λ1 | + |Λ2 | + c13 |Λ3 | − c13

Pej |Λj |2 (3.66)
j=1

3.3.2 Neutrino Flux in the detector

The event expected ratio in a neutrino detector is given by a convolution of the neutrino
spectrum, cross section, and the detector resolution function as follows

Eνmax Tmax (Eν ) Ti+1



Z Z Z
Rexpt = κ dEν dT dT ′ λ(Eν ) (Eν , T, µ)R(T, T ′ ). (3.67)
Eνmin Tmin Ti dT

Here, κ is a constant depending on ρe (the electron density number per kg of target mass) and
exposure time, λ(Eν ) denotes the neutrino spectrum, (dσ/dT ) is the ν − e total cross section
which is given by equation (3.22). The resolution function for the most of experiments has the
form

−(T − T ′ )2
 
C
R(T, T ′ ) = exp (3.68)
σ 2σ 2
For each of the experiments, we compare the expected and measured event rates using the
method of least squares

N
( )
bin 2
2
X [Rmeas (i) − Rexpt (i)]
χ = (3.69)
i=1
(∆(i))2

Where Nbin is the number of bins of the experiment.


CHAPTER 3. CONSTRAINTS ON µν FROM EXPERIMENTAL DATA 57

3.3.3 Reactor Experiments

In the experiments studied here, with the exception of GEMMA, the limit on the neutrino
magnetic moment is obtained by searching for an excess of events, which are registered
by comparing the measured and expected event ratios. The treatment used by GEMMA
experiment is different and will describe it in more detail below.

GEMMA Experiment

GEMMA is the most recent experiment (2012) which have obtained the world best reactor
limit to the neutrino magnetic moment: µν < 2.9 × 10−11 µB . The GEMMA experiment
is located in the Kalinin Nuclear Power Plant (KNPP) in Russia. The detector consists of
germanium of high purity which is placed at a distance of 13.9 m from the reactor core. Layers
of copper, lead and plastic scintillators are covering it [56]. The GEMMA experiment followed
a strategy different from the rest of the experiments studied: In order to get a recoil electron
spectrum, they compared the spectra measured when the reactor was in operation (ON) and it
was shut down (OFF). The experiment contains 3 phases: The Phase-I consists of 5184 hours
for the reactor ON period and 1853 hours for the reactor OFF period. In the phase II, 6798
ON-hours and 1021 OFF-hours were analyzed, and the Phase-III, 6152 ON-hours and 1653
OFF-hours were considered [56].
Then, the energy spectrum for the ON period is SON and SOF F for the OFF period which are
normalized by the times TON and TOF F , respectively,

SON SOF F
= + md Φν (W + X(EM )). (3.70)
TON TOF F
In the last term of the equation (3.70), md is the fiducial detector mass, Φν is the antineutrino
flux in the reactor, W represents the contribution to the cross section of the Standard Model
and EM , the electromagnetic contribution, which is proportional to the magnetic moment
value X,

 
µν
X= (3.71)
10−11 µB
In order to obtain an upper limit on X parameter, a comparison between SON and SOF F is
performed. After this, the result obtained is µν < 2.9 × 10−11 µB , which is currently the lower
limit on neutrino magnetic moment from reactor experiments.
CHAPTER 3. CONSTRAINTS ON µν FROM EXPERIMENTAL DATA 58

3.4 Combined Analysis of reactor and accelerator


neutrinos
According with the previous section, there are different experimental limits reported around
the world for the neutrino magnetic moment. We have performed a combined analysis of the
data obtained by these experiments, which allows us to get more reliable bounds because
we can reduce the error and gain robustness. This analysis has been performed using
experimental data from reactor and accelerator experiments. The corresponding analysis to
solar neutrinos has not been finished yet, but it will be included in the final paper which will
be sent to publication soon.

3.4.1 Limits on Effective Magnetic Moment

The first step in this work was to reproduce the experimental limits reported by the
experiments. Once we done it, our work was focused in combine the experiments so as to
get a global limit. In the next section, we will briefly describe the statistical method which
was used for calculate the bounds on the neutrino magnetic moment.

Statistical Method: Least squares

Consider N independent measurements yi at known points xi , and we assume the


measurement yi have a Gaussian behavior with mean F (xi ; θ) and known variance σi2 , where
θ are unknown parameters. The idea is to find an estimation for these parameters θ. Then
the likelihood function is given by the following expression [22],

N
X (yi − F (xi ; θ))2
χ2 (θ) = . (3.72)
i=1
σi2

The estimators of θ̂ are defined by the minimum of the equation ( 3.72). Now we expand χ2 (θ)
about θ̂ so as to find the contour in parameter space which is defined by

χ2 (θ) = χ2 (θ̂) + 1 = χ2min + 1. (3.73)

In principle, the Gaussian function is symmetric between θ and θ̂, then the contours of χ2
cover the real values with a certain probability. That is known as the confidence region, which
is defined as follows
CHAPTER 3. CONSTRAINTS ON µν FROM EXPERIMENTAL DATA 59

χ2 (θ) ≤ χ2min + ∆χ2 . (3.74)

In the table (3.1), it is shown values of ∆χ2 for the different number of parameters m.

(1 − α)(%) m=1 m=2 m=3


68.27 1.00 2.30 3.53
90. 2.71 4.61 6.25
95. 3.84 5.99 7.82
95.45 4.00 6.18 8.03
99. 6.63 9.21 11.34
99.73 9.00 11.83 14.16

Table 3.1: ∆χ2 corresponding to a probability 1 − α in the large data sample limit, for m
parameters. Taken from [22].

Limits on the effective magnetic moment

Our results for the different experiments discussed in this chapter are shown in Fig. (3.9) and
in Table (3.2). We can note that the more constrained limit is given by GEMMA experiment
(µνe ≤ 0.29 × 10−10 µB ), and the higher limit was found by LSND Collaboration (µνe ≤ 11 ×
10−10 µB ).

Limits on the transition magnetic moments: reactor neutrinos

As we mentioned before, the limit reported by the experiments corresponds to an effective


magnetic moment, namely, this limit does not give us information about the transition
magnetic moments. Using the expressions (3.47) and (3.51) we have calculated the bounds
on one of the components of the magnetic moment vector Λ. The results are presented in the
tables (3.3), (3.4) and (3.5). For the mixing angles, we have used the best fit reported in the
reference [34], with sin2 θ13 = 0.0246, sin2 θ12 = 0.320 and sin2 θ23 = 0.427.

Limits on the transition magnetic moments: accelerator neutrinos

The tables (3.6), (3.7) and (3.8) display the bounds obtained for the components |Λ1 |, |Λ2 | and
|Λ3 |, using accelerator neutrinos. These limits corresponds to the νe e− scattering, with θ13 = 0
CHAPTER 3. CONSTRAINTS ON µν FROM EXPERIMENTAL DATA 60

15
14 LAMPF
LSND
13 Krasnoyarsk
12 Rovno
11 TEXONO
GEMMA
10 MUNU
9 Combinado(Total)
8
2
χ

7
6
5
4
3
2
1
0
2e-12 4e-12 7e-12 1e-11 3e-11 6e-11 1e-10 2e-10 5e-10 9e-10
µν(µB)
Figure 3.9: The likelihood χ2 as a function of the effective magnetic moment µν .

Experiment Reported Limit Calculated Limit C.L.


ACCELERATORS
LAMPF µνe ≤ 10.8 × 10−10 µB µνe ≤ 7.4 × 10−10 µB 90%
−10 −10
LSND µνe ≤ 11 × 10 µB µνe ≤ 10 × 10 µB 90%
LSND µνµ ≤ 6.8 × 10−10 µB µνµ ≤ 7.4 × 10−10 µB 90%
REACTORS
KRASNOYARSK µν̄e ≤ 2.4 × 10−10 µB µν̄e ≤ 2.75 × 10−10 µB 90%
ROVNO µν̄e ≤ 1.9 × 10−10 µB µν̄e ≤ 2.05 × 10−10 µB 95%
MUNU µν̄e ≤ 0.9 × 10−10 µB µν̄e ≤ 0.86 × 10−10 µB 90%
−10 −10
TEXONO µν̄e ≤ 2.2 × 10 µB µν̄e ≤ 1.9 × 10 µB 90%
GEMMA µν̄e ≤ 0.29 × 10−10 µB µν̄e ≤ 0.29 × 10−10 µB 90%
COMBINED - µνe ≤ 0.29 × 10−10 µB 90%
COMBINED-GEMMA - µνe ≤ 0.87 × 10−10 µB 90%

Table 3.2: Limits on the effective magnetic moment.

and θ13 6= 0, respectively. In a similar way, it is posible to calculate the bounds to the transition
magnetic moments from the νµ e− interaction.
The limits on the total magnetic moments Λ and µM
ef f will be obtained once the solar

experiments be included in this analysis.


CHAPTER 3. CONSTRAINTS ON µν FROM EXPERIMENTAL DATA 61

EXPERIMENT |Λ1 | ≤ EXPERIMENT |Λ2 | ≤


−10
KRASNOYARSK 4.75 × 10 µB KRASNOYARSK 3.32 × 10−10 µB
ROVNO 3.29 × 10−10 µB ROVNO 2.31 × 10−10 µB
MUNU 1.48 × 10−10 µB MUNU 1.03 × 10−10 µB
TEXONO 3.29 × 10−10 µB TEXONO 2.3 × 10−10 µB
COMBINED 1.49 × 10−10 µB COMBINED 1.04 × 10−10 µB

Table 3.3: Limits on Λ1 from reactor Table 3.4: Limits on Λ2 from reactor
experiments. experiments.

EXPERIMENT |Λ3 | ≤
KRASNOYARSK 2.79 × 10−10 µB
ROVNO 1.94 × 10−10 µB
MUNU 8.7 × 10−11 µB
TEXONO 1.93 × 10−10 µB
COMBINED 8.7 × 10−11 µB

Table 3.5: Limits on Λ3 from reactor


experiments.
CHAPTER 3. CONSTRAINTS ON µν FROM EXPERIMENTAL DATA 62

EXPERIMENT |Λ1 | ≤ EXPERIMENT |Λ2 | ≤


θ13 = 0 θ13 = 0
−9
LAMPF 1.3 × 10 µB LAMPF 8.97 × 10−10 µB
LSND 1.77 × 10−9 µB LSND 1.21 × 10−9 µB
θ13 6= 0 θ13 6= 0
LAMPF 1.28 × 10−9 µB LAMPF 8.92 × 10−10 µB
LSND 1.72 × 10−9 µB LSND 1.21 × 10−9 µB

Table 3.6: Limits on Λ1 from Table 3.7: Limits on Λ2 from


accelerator experiments. accelerator experiments.

EXPERIMENT |Λ3 | ≤
θ13 = 0
LAMPF 7.4 × 10−10 µB
LSND 1.0 × 10−9 µB
θ13 6= 0
LAMPF 7.49 × 10−10 µB
LSND 1.01 × 10−9 µB

Table 3.8: Limits on Λ3 from


accelerator experiments.
Chapter 4

Ultra-High energy neutrinos

A cosmic ray is a particle or heavy nuclei which is produced in some galactic or extra galactic
source. These cosmic rays travel through space to reach the Earth. That is why they are
called cosmic messengers and its study is a very important tool to obtain information about
the Universe. The High Energy Cosmic Rays (HECR) are composed meanly by protons and
nuclei of heavier elements whose nature is yet unknown. Their energy spectrum ranges
from 100 MeV to several 1020 eV. Despite having been discovered more than 100 years ago,
cosmic rays still pose several important questions such as What is their origin?, What is their
composition?, What is the acceleration mechanism in the source?, among others.
Cosmic rays interact in the Earth’s atmosphere producing ionization and showers of secondary
charged particles. Some of these showers, with energies below GeV, die out in the high
atmosphere and, therefore, escape from detection in ground experiments. For that reason,
balloons and satellites are used in order to detect these low energy cosmic rays [60]. Showers
above GeV can be detected from the ground, using optical telescopes. When the energy is
above 1015 eV, the secondaries particles are quite numerous and, besides the optical detection,
Cherenkov detectors can be used to reconstruct the events.
The Cosmic rays were discovered in 1912, by Victor Hess [61]. As we mentioned before,
despite of 100 years of research, the origin of the UHECR is currently unknown. We know
that particles below 100 MeV must come from the Sun, since particles coming from outside
the solar system are shielded by the Sun wind. For energies above 100 MeV, the spectrum of
cosmic rays behaves as ∝ E −γ , where γ is an spectral index. At the energy E = 4 × 1015 eV
(known as “the knee”) the flux goes from γ = 2.7 to γ = 3.0. Above of E = 5 × 1018 eV (called
“the ankle”), the flux changes to an index γ = 2.8, see Fig. (4.1).

63
CHAPTER 4. ULTRA-HIGH ENERGY NEUTRINOS 64

Figure 4.1: The cosmic ray spectrum and experiments and some experimental results. Taken
from www.physics.utah.edu/ whanlon/spectrum.html.

On the other hand, it is estimated that cosmic rays with energies above of 6 × 1019 eV,
interacting with photons of the Cosmic Microwave Background, lost most of their energy
for distances bigger than 20M pc, this is known as the Greissen-Zatsepin-Kusmin (GZK)
cutoff [63]. Some experiments, such as HiRes and Pierre Auger, have observed the suppression
of the cosmic ray flux, consistent with the GZK effect [64, 65, 66, 67]. The cosmic rays are
charged, then, they are deflected by the cosmic magnetic fields. In this sense, the neutrinos
are an ideal tool to get information about the original sources of the cosmic rays, since they
can travel cosmological distances without suffer any deviation in their way from the source to
the Earth.

4.1 Neutrino Astronomy


In the Standard Model, neutrinos interact only through weak interactions, they can not be
deviated by cosmic magnetic fields, their mass is very small to be affected by gravity, and
they can hardly be absorbed by matter. These features make the neutrinos strong candidates
CHAPTER 4. ULTRA-HIGH ENERGY NEUTRINOS 65

to explore the Universe. There are however some problems, for instance, the neutrino flux
arriving from distant astrophysical sources is expected to be low; this has made necessary the
construction of big detectors, such as IceCube. The detection principle of the High Energy
neutrino detectors was first proposed by Markov in the 60’s [68]. The interaction of the
neutrinos in the detector produces Cherenkov light, this radiation is emitted when a charged
lepton travels with a velocity higher than the light velocity in the medium. The radiation
emitted is collected by photosensors installed in the detector. The amplitude and timing of
the signal recorded by the photosensors allow to know the arrival direction, the neutrino type
and its energy. In order to decrease the background of atmospheric charged particles, these
detectors are placed deep underground.
At present, the are several detectors of UHE neutrinos around the globe, with the IceCube
experiment being the most successful up to now. This detector is located at the South Pole
and its detection material is the ice itself [69]. Other important detectors are the Pierre
Auger experiment [12] in Argentina, the ANTARES detector deployed in the Mediterranean
Sea in the Toulon cost (France) [70].

4.1.1 The IceCube Neutrino Observatory

IceCube is currently the most sensitive detector of high energy neutrinos. The construction of
IceCube was motivated by the already mentioned questions about the origin of cosmic rays.
The IceCube experiment, located at the geographic South Pole is a high energy neutrino
detector covering a volume of a 1 km3 of ice. The installation of IceCube, with all its
components was completed in December 2010. The main purpose of IceCube is the detection of
high energy neutrinos from astrophysical sources via the Cherenkov light of charged particles
generated in neutrino interactions with the ice [4].
The IceCube detector is composed of an array of 86 strings and 5160 light detectors at a
depth within 1450 m and 2450 m [13]. Each string contain 60 light detectors (Digital Optical
Modules (DOMs)) which consist of 10 photomultiplier tubes (PMT) each. The function of the
PMTs is record the Cherenkov light produced by charged particles in the ice, the signal
collected by the PMTs is processed by a digitalization system. The IceTop component is
located on the surface with 81 stations which contains two ice-filled tanks each. Regarding the
detection of Ultra High-energy neutrinos from astrophysical sources, the IceCube is the most
promising experiment. So far, IceCube Collaboration have reported the possible observation
of a few events from yet unidentified extragalactic sources. In the next section, we are going
CHAPTER 4. ULTRA-HIGH ENERGY NEUTRINOS 66

฀ ฀

฀ ฀ ฀


฀ ฀ ฀ ฀
฀ ฀ ฀ ฀


฀ ฀
฀ ฀

Figure 4.2: The IceCube detector. Taken from [4].

to describe the most energetic sources observed in the Universe (GRBs) and the neutrino
production by these sources.

4.2 Gamma Ray Bursts (GRBs)


The Gamma-ray bursts are known as the most intense explosions which are observed in the
Universe. These electromagnetic events can release more energy in 10 seconds than the Sun
will emit in all its lifetime. The discovery of the GRBs was reported in 1973 [71]; for twenty
years, these explosions were considering as a mere curiosity, but today they are essential
tools for the study of early structure formation, the evolution of chemical elements and the
high-redshift stars and galaxies.
Something important to remark is that the GRBs may play a revolutionary role regarding
to the astronomic messengers, in this case we talk about the neutrinos. These astrophysical
objects produce neutrinos which go through the space without interact with ordinary matter
in its way to the earth. This represent an advantage over ordinary cosmic rays that are
deflected by the magnetic fields in the source and the interstellar space. Therefore, the study
of neutrinos produced in GRBs and in general, in any astrophysical source is a fundamental
tool in order to answer questions related with the cosmic rays and with the search for physics
CHAPTER 4. ULTRA-HIGH ENERGY NEUTRINOS 67

beyond the Standard Model.

Figure 4.3: Long-duration GRBs (left) are thought to originate in the collapse of massive stars,
while short-duration GRBs are likely produced in the merger of two compact objects. Taken
from [5].

The GRBs are frequently available signposts throughout the Universe, since several of
them are detected each day. They are classified according to their duration Fig. (4.3) in
(1) Long-duration GRBs (> 2s) and (2) Short-duration GRBs (< 2s). Different models are
proposed in order to explain the GRB radiation, among them we have the models of fireball
and cannon ball [6]. Here, we are going to describe two main models proposed in the literature:

1. The fusion of two neutrons stars, or a neutron star and a black hole of a stellar mass.
In this model, the (ν, ν̄) annihilation produce γ rays, but this does not explain the most
powerful GRBs.

2. Hypernova. A GRB could be caused by the explosion of a very powerful supernova.


Moreover, in this type of models is also assumed the collapse of a massive nucleus, the
formation of a black hole with a mass Mbh ∼ 20M⊙ surrounded by a massive disk, the
fast accretion and, as a result, a GRB.
CHAPTER 4. ULTRA-HIGH ENERGY NEUTRINOS 68

4.3 Waxman-Bahcall theoretical prediction for the high


energy neutrinos
The GRBs are intense sources of Ultra-High Energy neutrinos, the protons are accelerated
by the magnetic fields and collide with the radiation in these objects producing resonances
which give rise to pions and kaons. The pions and kaons decay into neutrinos resulting into a
neutrino flux that should be detected in the ground,

p+γ → ∆+ → n + π + , p + π0 (4.1)

π+ → µ+ + ν µ

µ+ → e+ + νe + ν̄µ

The estimated energy for neutrinos from GRBs is of the order of TeV and EeV, which, in
principle could be detected by very big detectors such as IceCube. The identification of these
neutrinos would provide strong evidence about the source and acceleration mechanism.
Currently there are different progenitors models of GRBs which study the neutrino
production. These models are divided into two main categories, the first category considers
the fusion of two compact objects, such as two neutron star [72], a black hole with a black
hole [73] or a black hole with a Helium star or with a white dwarf [74]. The second category
involves the death of a massive star [75], where a black hole is created promptly with an
accretion disk which feeds a strong relativistic jet. Long GRBs are usually attributed to the
second category of progenitors, while short GRBs are attributed to the first category.
CHAPTER 4. ULTRA-HIGH ENERGY NEUTRINOS 69

Figure 4.4: Upper panel: The upper bound imposed by UHECR observations on the diffuse
muon neutrino flux (νµ and ν̄µ combined) (lower curve: there is no evolution of the energy
production rate, upper curve: there is evolution) predicted by different models, the upper
bound implied by cosmic rays observations and GRB intensity. The dash-dotted lines are the
upper bound given by the Eq. (4.2). Taken from [7]. Lower panel: The same bound compared
with the atmospheric muon-neutrino background and several upper bounds reported by
experiments such as AMANDA, BAIKAL, RICE, ANITA and Auger. The GZK curve shows the
muon neutrino flux expected from interactions between protons and micro-wave background.
Taken from [8].
CHAPTER 4. ULTRA-HIGH ENERGY NEUTRINOS 70

According with the result presented in [7], Imax is defined as the muon neutrino intensity,
where νµ s and ν̄µ s are combined:

c 2 dṄCR
Imax ≈ 0.25ξZ tH E (4.2)
4π CR dECR
≈ 1.5 × 10−8 ξZ GeV cm−2 s−1 sr−1 .

Then, the expected neutrino intensities are given by

c 2 dNνµ 1
Eν2 Φνµ = E = Imax , with Φνe ≈ Φνµ ≈ Φν̄µ . (4.3)
4π ν dEν 2
Here, ξZ describes the possible contribution of high redshift sources of high-energy cosmic
rays and includes the effect of the redshift in neutrino energy; tH ≈ 1010 yr, is the Hubble
time.
If the protons are accelerated in internal shocks, the neutrinos produced will arrive at
Earth simultaneously with the photons of the burst proper and will have an energy 1015 −
1016 eV [76]. If accelerated in external shocks, they will arrive to the Earth simultaneously
with the photons of the afterglow and will have an energy bigger than 1017 eV [77]. The results
showed in the Fig. (4.4) provide an upper limit to the high energy neutrino flux from GRBs [7].
The estimated detection rate of extra-galactic astrophysical neutrinos in a 0.1 km2 detector is
∼ 100 events per-steradian per year [78].

4.4 Experimental limits for astrophysical neutrino flux


Several experiments around the world are trying to measure the neutrinos from GRBs
and from other astrophysical sources. Among them, we have Frejus [79], MACRO [80],
AMANDA-II [81], Baikal [82], ANTARES [83], Pierre Auger and IceCube [13]. At present,
we do not have a clear and significant signal about these neutrinos. Next we will show
the experimental measurements performed by three important experiments: ANTARES in
France, Pierre Auger in Argentina and IceCube in South Pole.

4.4.1 ANTARES experiment

Astronomy with a Neutrino Telescope and Abyss environmental RESearch (ANTARES) is a


telescope located in the Mediterranean Sea (in the coast of Toulon, France) at a depth of 2.4
Km. This detector is composed by 12 vertical strings (anchored to the seabed), each with a
CHAPTER 4. ULTRA-HIGH ENERGY NEUTRINOS 71

length of 450 m and equipped with 25 triplets of PMTs. The distance between strings is of 70
m. The instrumented volume is ∼ 0.01km3 [9].

Figure 4.5: Sum of the 296 individual gamma-ray-burst muon neutrino spectra (red and
blue solid lines) and limits on the flux expected (red and blue dashed lines). The IceCube
IC40 + IC59 limit on the neutrino intensity from 300 GRBs (black dashed line) and the first
ANTARES limit from 2007 using 40 GRBs (grey dash-dotted line). Taken from [9].

The ANTARES experiment performed a search for muon neutrinos from 296 gamma-ray
bursts occurring between the end of 2007 and 2011. No events passed the selection criteria
and the limits derived are shown in the Fig. (4.5).

4.4.2 Pierre Auger experiment

The Pierre Auger experiment is located at the Mendoza province (Argentina). It is a hybrid
UHECR detector which combine an array of particle detectors at ground level, and 24
fluorescence telescopes housed in four buildings. The surface detector array is composed
by water Cherenkov detectors in the form of cylinders of 3.6 m diameter and 1.2 m height,
each containing 12 tonnes of purified water. Charged particles collide with the nuclei in
the water and produce Cherenkov light which is collected by three 9-inch photomultiplier
tubes (PMT) at the top surface and in optical contact with the water [12]. There are ∼ 1600
water stations arranged in a triangular grid with 1.5 km spacing between them, covering a
surface of ∼ 3000km2 , at an approximate altitude of 1400 m above sea level. Pierre Auger, like
ANTARES, has made searches for a diffuse neutrino flux, but no events have been found. In
CHAPTER 4. ULTRA-HIGH ENERGY NEUTRINOS 72

Fig. (4.6) the experimental limits and several predictions of different theoretical models are
shown.

Figure 4.6: Thick lines: Upper limits (at 90% C.L.) to the diffuse flux of UHE neutrinos, limits
from other experiments are also plotted. Thin lines: Expected fluxes from the theoretical
models: ”p, Fermi-LAT” [10], ”p,evol-FRII” and ”Fe, uniform” [11]. Taken from [12].

4.4.3 IceCube Results

IceCube is currently the detector that, in principle, satisfies the conditions to detect UHE
neutrinos from astrophysical sources. In this work, we describe some of its main results.

Limit on the neutrino flux from GRBs

These results were obtained while the IceCube detector was under construction, using the 40-
and 59-string configurations of the detector. The data were taken from April 2008 to May
2009 and from May 2009 until May 2010, respectively. Around 230 GRBs were observed in
these two periods of time [84]. In this result, IceCube reported an upper limit on the neutrino
flux associated with gamma-ray bursts which is at least a factor of 3.7 below the theoretical
prediction of Waxman and Bahcall, see Fig. (4.7).
CHAPTER 4. ULTRA-HIGH ENERGY NEUTRINOS 73

10–8
Waxman and Bahcall 10
IC-40
IC40 Guetta et al.

E 2Φ ν (GeV cm–2 s–1 sr–1)


IC40+59 Combined
limit
IC40+59 Guetta
et al.

10–9
10

104 105 106 107


Neutrino energy (GeV)

Figure 4.7: Comparison of results to predictions based on observed γ-ray spectra. The
predictions are shown in dashed lines and the experimental measurements in solid lines.
Taken from [13].

Evidence for extraterrestrial Neutrinos at the IceCube detector

In a recent result presented in the reference [14], The IceCube collaboration have reported the
observation of two electron-neutrino events. Their energies were in the PeV range: 1.04 PeV
and 1.14 PeV, Moreover, 26 events between 30 and 300 TeV also were observed in the same
measurement. These results correspond to a search performed between May 2010 and May
2012, with the following configurations: 86 strings (between May 2011 and May 2012) and 79
strings (between May 2010 and May 2011), with a total combined live time of 662 days. The
flavors, directions and energies of these events are not in agreement with those expected for
atmospheric neutrinos, and there is strong evidence that they are of extraterrestrial origin.
It is possible to identify seven events with charged current muon neutrinos whereas the
remaining 21 events correspond to other type of neutrino interactions.
In the Fig. (4.8), we can see the observed events as a function of energy, the blue color denotes
the atmospheric neutrino background and the red color correspond to the atmospheric muon
background. The green line shows the expected atmospheric neutrino flux, and the magenta
line corresponds to the experimental 90% limit.
These 28 reported events correspond to a 4.1σ excess with respect to the background
and are interpreted as evidence for an astrophysical all-flavor component of Eν2 Φν =
(3.6 ± 1.2) × 10−8 GeV cm−2 s−1 sr−1 with indications for a cutoff at ∼ 2P eV . This latter suggests
a suppression of the neutrino spectrum above this energy [85]. The intensity measured is
CHAPTER 4. ULTRA-HIGH ENERGY NEUTRINOS 74

Figure 4.8: Events number vs deposited energy compared to model predictions. Taken
from [14].

very close to the theoretical limit of Waxman-Bahcall: Eν2 Φν < 2 × 10−8 GeV cm−2 s−1 sr−1 [8].
The best-fit flux of the high-energy analysis corresponds to a muon neutrino flux of
Eν2 Φ(Eν ) = 1.2 × 10−8 GeV cm−2 s−1 sr−1 .

What is the origin of these neutrinos?

There are several questions to answer regarding to this result, maybe the most important
one is whether these neutrinos are Galactic or extragalactic in origin. In words of Francis
Halzen: “It is too early to speculate. Moreover, the current information is insufficient to identify
the sources of these events which is illustrated by the wide speculation in the literature” [85].
IceCube was designed to operate for 20 years; therefore, we expect that the results of the next
years will shed light on this puzzle.

Search for a diffuse flux of astrophysical muon neutrinos

In this section, we briefly describe the last results published by IceCube (December 2013)
regarding the limits on the diffuse neutrino flux from astrophysical sources. These results are
discribed in the reference [15]. The IceCube collaboration performed a search for high-energy
neutrinos with data collected by the IceCube detector from May 2009 to May 2010, when the
CHAPTER 4. ULTRA-HIGH ENERGY NEUTRINOS 75

array was running in its 59-strings configuration.

Figure 4.9: Limit on a (νµ + ν̄µ ) astrophysical flux in comparison to theoretical flux predictions
and limits from other experiments. The Black lines: show the expected atmospheric
neutrino flux with and without a prompt component. The red dashed line: Denotes the
Waxman-Bahcall upper bound. The Green dashed lines: represent several model predictions
for astrophysical neutrino fluxes. The pink solid line: is the experimental limit at 90% C.L.
The orange solid line: shows its sensitivity. Taken from [15].

A global fit of the reconstructed energies and directions of observed events was carried out,
taking into account possible contributions from astrophysical and atmospheric backgrounds,
and also systematic uncertainties. The results are shown in Fig. (4.9), the best fit yields an
astrophysical signal flux (for νµ + ν̄µ ) of

Eν2 Φ(Eν ) = 0.25 × 10−8 GeV cm−2 s−1 sr−1 , (4.4)

and the best fit of the prompt atmospheric flux is zero. The experimental limit at 90%
confidence level is Eν2 Φ(Eν ) ≤ 1.44 × 10−8 GeV cm−2 s−1 sr−1 . This limit is a factor of 1.5 above
the Waxman-Bahcall upper bound.
CHAPTER 4. ULTRA-HIGH ENERGY NEUTRINOS 76

The Fig. (4.8) presents a comparison of the upper limit of this analysis with theoretical flux
predictions and limits from other experiments. It is worth to examine in detail this figure,
especially the results concerning the diffuse muon neutrino results. The Icecube reported limit
is already close to the Waxman-Bahcall prediction (approximately a 40% above) while from the
expected sensitivity we could see that, in future, it will be possible to have a constraint below
this Waxman-Bahcall limit. It will be interesting, therefore, to wait for future IceCube results
in order to confirm Waxman-Bahcall predictions, or to get stronger constraints.
In summary, so far, IceCube has reported the observation of 28 neutrino events: 2
electron-neutrino events with energies of 1.04 PeV and 1.14 PeV, respectively, and 26 events
(including all-flavors of neutrinos) with energies between 30 and 300 TeV. In principle, these
events could be galactic or extragalactic, which is currently a subject of intense study.
The non-observation of neutrinos with energies above 1016 eV (Ultra-High Energy Neutrinos
(UHEN)) by IceCube might indicate that the UHEN may be suppressed in its way from the
source to the earth.
Although we need to wait for more data in order to have a definite understanding of UHE
neutrino fluxes, we could, for the moment, study possible mechanisms that could give an
explanation to the non-observation of neutrinos from GRBs, or to the suppression of UHEN
from astrophysical sources.
With this motivation in mind, we are going to address this topic in more detail below. This will
not only offer a possible answer to the problem of astrophysical neutrinos, but also opens a
window to explore another very important topic: the existence of physics beyond the Standard
Model.

4.5 A reduction in the UHE neutrino flux due to neutrino


spin precession
Neutrinos with diagonal or transition magnetic moment (µ) propagating in transverse
magnetic fields (B⊥ ) will undergo spin precession or spin-flavor precession, respectively. Both
types of precession can have important astrophysical implications.
Regarding the case of GRB’s, the particular case of spin flip conversion could be of great
interest [86]. In particular, we can study the possibility that active neutrinos, produced in a
GRB (or in another astrophysical object, such as an Active Galactic Nucleus (AGN)), may be
converted into a right handed sterile neutrino, due to the presence of strong magnetic fields
CHAPTER 4. ULTRA-HIGH ENERGY NEUTRINOS 77

in this kind of objects.


The spin precession mechanism had been studied long time ago in the context of the solar
neutrino problem [87]. The evolution equation for this mechanism is given by
    
ν̇lL 0 µν B ⊥ ν lL
i =  . (4.5)
ν̇lR µν B ⊥ 0 ν lR
It is possible to compute, from this equation, the corresponding conversion probability from
νlL → νxR , which is given by [87]

Z r 
2 ′ ′
P (νlL → νlR ; r) = sin µν B⊥ (r )dr (4.6)
0
where, µν corresponds to the neutrino magnetic moment, B⊥ (r) stands for transverse
magnetic field, and, finally, r is the distance along the neutrino path.
It is already known [88, 89] that a constant magnetic field may fully convert an active neutrino
flux into an sterile one when
π
µν B ⊥ r ≈ . (4.7)
2
It is expected that astrophysical objects have magnetic fields in its interior. The magnitude of
such magnetic fields will depend on the particular characteristics of these objects. However,
the wide range of sizes for these objects, naturally implies that an extense range of magnetic
field intensities can be found in the Universe. Therefore, there will not be a surprise that
we can find astrophysical objects that may satisfy Eq. (4.7), provided that the neutrino has
a non zero magnetic moment. In Fig. (4.10), the regions in the B-r plane satisfying Eq. (4.7)
are shown, for diagonal NMM 10−12 µB , 10−14 µB , 10−15 µB . In this case we have based our
calculations in two different models of magnetic fields, the left figure corresponds to the
model displayed in [90] and the right one is according to [91]. The astrophysical objects
that lies in a given curve may experience a neutrino conversion into a sterile state, if the
neutrino has the corresponding neutrino magnetic moment value. In this picture, a relatively
small neutrino magnetic moment, of the order µν = 3 ∗ 10−12 µB could produce an efficient
conversion into sterile states in the case of GRBs, as can be seen from the left figure. This is
an interesting phenomenon, considering the recent limit on GRBs neutrinos reported by the
IceCube collaboration discussed in the previous sections [84]. A higher value of the neutrino
magnetic moment, of the order of 10−14 µB , could produce the same conversion for an AGN.
Besides the conversion effect that may be produced in the GRB due to the intense magnetic
fields, it might also be possible, at least in principle, that the galactic, or the intergalactic,
magnetic field may also produce a spin conversion. Despite being a rather weak magnetic
CHAPTER 4. ULTRA-HIGH ENERGY NEUTRINOS 78

15
Neutron Stars -12
µν = 3*10 µΒ
-14
µν = 1*10 µB
10 -15 NS
µν = 1*10 µB
−12
12 µν=10 µB
Log B [Gauss]

−14
GRB µν=10 µB

log(B [gauss])
5 −15
White Dwarfs µν=10 µB
GRBs
AGNs
6 AGN

0 Jets
0
SNR
-5 -6 Galactic disk
Galactic halo
-10 -12
0 5 10 15 20 0 5 10 15 20
Log R [km] log(r [km])

Figure 4.10: Relation of magnetic field B and size r of astrophysical sources for an efficient
neutrino spin transition νlL → νlR . The curves show different values of the neutrino magnetic
moment [86]
.

field, there is a long distance that must be traveled by the neutrino flux.
The above discussion has been part of an already published work [86], where we emphsize the
plausibility of efficient transitions of active neutrinos into sterile ones, thanks to the presence
of a non-zero magnetic moment, in combination with the strong magnetic fields, expected to
exist both in the GRBs as well as in the AGNs. A possible perspective of this thesis will be to
study in more detail this mechanism, especially if future measurements report stronger limits
for the corresponding fluxes. In the next part of this chapter we will show different directions
that a future work may take in this context.

4.5.1 Active-sterile conversion in the propagation

Besides the source, there are at least two more places where the neutrino might undergo a
conversion into a sterile state, depending on the value of the magnetic field: the intergalactic
and the galactic region.

Intergalactic and Galatic Magnetic Fields

An active neutrino travels, on its way from the source to the earth, long distances through the
intergalactic medium, so it may interact with the intergalactic magnetic field and be turned
CHAPTER 4. ULTRA-HIGH ENERGY NEUTRINOS 79

into a sterile neutrino by the same mechanism outlined above. Another possibility could be the
interaction with the magnetic field in the Milky Way before to arrive to the earth. Additionally,
we can also have an active-sterile conversion in the galactic medium for neutrinos traveling
from Milky Way’s AGN to the earth.
As reported by [103], the knowledge about the intergalactic and galactic magnetic fields has
been increasing in the last years although, according to [104], our knowledge about the cosmic
magnetic fields is still limited. Particularly, magnetic fields at cosmological scales (above
1 Mpc) remain obscure. However, currently, bounds have been established which give us
an idea of the magnitude of these fields. This allow us to make some estimates regarding
the conversion probability of active neutrinos into sterile neutrinos when they interact with
such magnetic fields. For instance, methods based on the rotation measure in the radio band
constraint the intensity of the intergalactic magnetic field (IGMF) below B < 10−11 − 10−9
Gauss (G) [104]. In the same work, the value of the intergalactic magnetic field is constrained
to be larger than B = 5 × 10−15 G.
On the other hand, regarding galactic magnetic fields (GMFs), it is estimated, from
measurements of radio synchrotron, that the averaged magnetic field in the Milky Way is
about B ∼ 6µG(0.6nT ) [103], extending to higher values near of its black hole: B ∼ 1mG.

4.5.2 Survival probability vs distance in the intergalactic medium

After the discussion in the previous sections, it is possible to compute the neutrino conversion
probability as a function of distance from the source. According to equation (4.6),

PS = sin2 (µν B⊥ R). (4.8)

In the Fig. (4.11), we show the survival probability as a function of the distance, for different
values of the neutrino magnetic moment. As can be seen from this figure, we are interested
in very distant sources, such as Gamma Ray Bursts, of the order of Giga-parsec (Gpc). We
have taken values for the intergalactic magnetic field according to those reported in [104]. As
we can see in the Fig. (4.11), there could be a total or partial conversion for neutrinos from
sources located at distances of the order of 1Gpc or more. Alike, we can note that the values
for the magnetic moment are in agreement with the experimental bounds.
A similar computation can be done for closer sources, like the case of AGNs. Fig. (4.12) displays
the survival probability as a function of the distance in this case. It is estimated that these
sources are located at distances of the order of Mega-parsec (Mpc). Again, we can notice that
CHAPTER 4. ULTRA-HIGH ENERGY NEUTRINOS 80

1
0,9
0,8
0,7
0,6
PS

0,5
0,4
−15 -10
µν = 7∗10 µB, B = 10 G, R = 2.5 Gpc
0,3 µν = 1∗10
−12
µΒ, Β = 10
−12
G, R = 1.7 Gpc
−11
0,2 µν = 2∗10
-13
µΒ, B = 10 G, R = 0.9 Gpc

0,1
0 -2 -1 0
10 10 10
R (Gpc)

Figure 4.11: Survival probability (PS ) vs distance for distant sources in the Universe such as
GRBs.

could be a total or partial conversion for sources situated in the range between 1 and 100 Mpc,
for values of the magnetic moment consistent with current experimental limits.

1
0,9
0,8
0,7
0,6
PS

0,5
0,4
−12 -10
µν = 1∗10 µB, B = 10 G, R = 17 Mpc
0,3 µν = 5∗10
−11
µΒ, Β = 10
−11
G, R = 3.5 Mpc
−13
0,2 µν = 2∗10
-10
µΒ, B = 10 G, R = 86 Mpc

0,1
0 -2 -1 0 1 2
10 10 10 10 10
R (Mpc)

Figure 4.12: Survival probability (PS ) vs distance for nearby sources in the Universe like
AGNs.
CHAPTER 4. ULTRA-HIGH ENERGY NEUTRINOS 81

4.5.3 Survival probability vs distance (galactic medium)

Another possibility to the active-sterile conversion is that the neutrino interacts with the
magnetic field in the Milky Way. In this case, the neutrino flux produced in the neighborhood
of the Active Galactic Nuclei of the Milky Way might be suppressed by this interaction.
Similarly, an extragalactic neutrino would undergo a transformation when crossing the Milky
Way’s magnetic field.
The Fig. (4.13) shows the survival probability as a function of distance in Kpc. Taking the
average magnetic field B = 6µG and a magnetic moment of the order of µν ∼ 10−14 µB , the
length of conversion is R ∼ 8Kpc which is in agreement with the average distance from the
Earth to the galactic center.

1
0,9
0,8
0,7
0,6
PS

0,5
0,4
0,3
−14 -6
µν = 3.5∗10 µB, B = 6*10 G, R = 8 Kpc
0,2
0,1
0 -2 -1 0 1
10 10 10 10
R (Kpc)
Figure 4.13: Survival probability (PS ) vs distance in the galactic medium.

An important point to remark in the analysis of the last two sections is that the conversion
from active neutrino to sterile neutrino does not depend on the neutrino energy. In principle,
we could have a suppression for neutrinos of any energy, provided that its magnetic moment,
and the conditions for the magnetic field and distance, are satisfied.
In this work we have taken a first approximation, assuming an average magnetic field. As a
future perspective in this thesis, we could consider a more realistic scenario by introducing a
magnetic field profile.
CHAPTER 4. ULTRA-HIGH ENERGY NEUTRINOS 82

4.6 Cosmological Effects


When we talk about objects at distances R ∼ 100 Mpc or more, the cosmological effects, such as
the expansion of the Universe, start to play an important role. In this case, a new parameter
appears: the redshift z. We are going to assume now that the neutrinos under study are
produced at a source with a redshift z. Over cosmological distances, we need to express the
distance between source and detector in terms of redshift, L(z).
The expansion of the universe means that the proper physical distance between a pair of well
separated astrophysical objects is increasing with time, that is, the objects are receding from
each other [105]. A gravitationally bound system such as the Local Group is not expanding,
the homogeneous expansion law refers to objects far enough apart for the local irregularities to
be ignored. The proper physical distance l(t) between a pair of well-separated objects changes
with time as

l(t) = lo a(t). (4.9)

Where lo is the distance at rest (constant) and a(t) is the universal expansion factor. The time
derivative of this expression is the rate of recession of one astrophysical object as measured
by an observer,


v = l˙ = lo ȧ = l = Hl. (4.10)
a
Here, H is the Hubble’s parameter. The recession causes a redshift in the spectrum of the
light from one object received by an observer. For small recession speed, v, this is the ordinary
first-order Doppler shift, where the observed wavelength λo differs from the wavelength λe at
emission (as measured by an observer at rest at the source) by the fractional amount

λo v Hl
z= −1= = . (4.11)
λe c c
The Friedman-Lemaitre model is characterized by dimensional parameters, such as Hubble’s
constant, Ho , and dimensionless parameters, such as the product, Ho to , of Hubble’s constant
and the age of the universe, to . The expansion rate, H(t) = ȧ/a, is, in general, a function of
time; the present value is written as the Hubble’s constant, Ho = 100 h kms−1 M pc−1 , where
h is a dimensionless quantity. The Universe at redshifts z ≤ 1000 is thought to be dominated
by matter such as stars and gas, with pressure small compared to the mass density. In this
case, the mass density varies as ρb ∝ a(t)−3 , and we can write the cosmological equation for
the expansion rate H as
CHAPTER 4. ULTRA-HIGH ENERGY NEUTRINOS 83

 2  2
2 ȧ ż
= Ho2 ΩM (1 + z)3 + ΩR (1 + z)2 + ΩΛ .

H = = (4.12)
a 1+z
Where the matter density, the curvature, and the cosmological constant parameter are,
respectively,

8πGρo 1 Λ
ΩM = , ΩR = , ΩΛ = . (4.13)
3Ho2 (ao Ho R)2 3Ho2
These parameters completely determine the geometry of the Universe, if it is homogeneous,
isotropic, and matter-dominated. Because of the Universe expansion, the celestial objects
changes constantly their positions, varying the distance between them. In this case, there are
several ways to measure the distance between two points, for example, Comoving distance
(line-of-sight), Comoving distance (transverse), Angular diameter distance, Luminosity
distance, Parallax distance. For a detailed description of the above see [106]. For our purposes,
we will only deal with Comoving distance (line-of-sight).

4.6.1 Comoving distance (line-of-sight)

Following Peebles [105], at low redshifts, we can write the expansion parameter as a function
of world time t as

a(t) = ao 1 − Ho (to − t) − qo Ho2 (to − t)2 /2 + · · · ,



(4.14)

where to − t is the lookback time (the time measured back from the present world time,
to ). Hubble’s constant is Ho = ȧo /ao , and the dimensionless acceleration (or deceleration)
parameter is

äo ao ΩM
qo = = , (4.15)
ȧ2o ΩΛ
with 1 + z = ao /a(t). We can rewrite equation (4.14) as an expression as a function of the
redshift,

qo 2
Ho (to − t) = z − (1 + )z + · · · (4.16)
2
The most general expression for the time evolution has the form,

a a inf
da da dz
Z Z Z
Ho t(z) = Ho = = E(z), (4.17)
0 ȧ 0 aE(z) z 1+z
CHAPTER 4. ULTRA-HIGH ENERGY NEUTRINOS 84

where E(z) is given by the square root of the Eq. (4.12). Hence, if we take ΩR = 0 (assuming
a flat ΛCDM cosmology) in the above expression, we obtain the total line-of-sight comoving
distance

z
dz ′
Z p
L(z) = LH , h(z) = Ho ΩM (1 + z)3 + ΩΛ , (4.18)
0 (1 + z ′ )h(z ′ )
with LH = c/Ho ≈ 3.89 Gpc, the Hubble length, (the speed of light multiplied by the Hubble
time). The Hubble time tH = Ho−1 ≈ 13.8 Gyr provides an estimate for the age of the Universe
by assuming that the Universe has always expanded at the same rate as it is expanding today.
Here, we choose ΩM = 0.27 and ΩΛ = 0.73 according with [107].
3
10

2
10
L(z) [Gpc]

1
10

0
10

-1
10

-2
10

-3
10 -3 -2 -1 0 1 2 3
10 10 10 10 10 10 10
Z
Figure 4.14: Black line: Line-of-sight comoving distance as a function of redshift in a flat
Λ − CDM cosmology. Red dash-line: Hubble length: LH ≈ 3.89 Gpc. Figure reproduced
from [16].

The comoving distance L(z) is shown in Fig. (4.14), we can see that this distance is limited
by the Hubble length LH ≈ 3.89 Gpc. This is very important, because this length represents
the horizon beyond which ultra-relativistic particles can not be seen. Therefore, the neutrinos
that could be detected in terrestrial detectors would come from sources located at most at 3.89
Gpc from Earth.
CHAPTER 4. ULTRA-HIGH ENERGY NEUTRINOS 85

4.6.2 Survival probability vs redshift

As we have discussed, for very distant objects, the redshift is the observable measurement
that we should consider. Therefore, we have calculated the survival probability as a function
of the redshift for different sources. In accordance with [16, 108] the GRBs have redshift z > 1
and AGNs display redshift 0.1 < z < 1.0. Now the expression for the probability will be as
follows

PS = sin2 (µν B⊥ L(z)), (4.19)

where L(z) is given by (4.18).

1
-14 −11
G , µν = 6∗10 µB
0,9 B = 1*10
-13 −12
B = 1*10 G , µν = 5∗10 µB
0,8 -12
B = 1*10 G , µν = 4∗10
−13
µB

0,7
0,6
PS

0,5
0,4
0,3
0,2
0,1
0 -2 -1 0 1 2
10 10 10 10 10
Z
Figure 4.15: Survival probability vs redshift for far sources such as GRBs.

The Figs. (4.15) and (4.16) show the survival probability for far sources and nearby sources
respectively. Here, we have taken values corresponding to the intergalactic magnetic field,
since the redshift for galactic sources is negligible. We can note that the values for the
neutrino magnetic moment are in agreement with the experimental bounds.
CHAPTER 4. ULTRA-HIGH ENERGY NEUTRINOS 86

1
0,9
0,8
0,7
0,6
PS

0,5
0,4
0,3 -10
B = 10 G, µν = 10
−11
µΒ
-10 −12
0,2 B = 10 G , µν = 8∗10 µ B
-10 −12
B = 5*10 G , µν = 5∗10 µB
0,1
0 -2 -1 0
10 10 10
Z

Figure 4.16: Survival probability vs redshift for nearby sources such as AGNs.

4.7 Neutrino magnetic moment in extra dimensions


Kalusa and Klein were the first to suggest that there may be extra dimensions in Nature [109,
110]. Theories with extra dimensions provide a natural mechanism to generate neutrino
masses, mν = 10−3 eV , where the existence of a right-handed neutrino, νR , is postulated as
a bulk fermion coupled to the SM left-handed neutrino field, νL .
Following [111], the neutrino-Higgs interaction is given by

y yM∗
√ H ν̄L νR = ν̄L νR , (4.20)
M∗ V n MP
where M∗ and MP are, respectively, the string and the Planck scales, H denotes the Higgs
doublet, y is the Yukawa coupling, and νR is the bulk fermion. The state νR is a linear
combination of a tower of Kaluza-Klein (KK) states.
Spontaneous electroweak symmetry breaking generates a Dirac mass given by

yv
mD = p . (4.21)
M∗n Vn
In this scheme, there will also be an electromagnetic interaction coupling, written in terms of
mass eigenstates, as [111]
CHAPTER 4. ULTRA-HIGH ENERGY NEUTRINOS 87

λρ
µ(k) Ψ̄νkR
m σλρ Ψν m F
oL
, (4.22)

where νoL is the kth state of lowest mass, and the transition magnetic moment is given by

 
3eGF k
µ(k) =√ θk , (4.23)
2(4π)2 R
with, θk , the mixing angle,


tan 2θk ≈
. (4.24)
k − ξ2

Where ξ is considered to be small and defined as ξ = 2mD R, and R is the compactification
radius.
0
10
-2
10 n=1
-4 n=2 4
10 n=4 M* = 10 TeV
-6
10 n=6
-8
10 X X X X
-10 X
10
µeff(µB)

-12
10 X X
-14 X
10
10
-16
X (Eν(eV) , µ(µΒ))
-18 X (4E+14 , 1E-14)
10
-20 X (2E+16 , 1E-9)
10 X (4E+15 , 2E-10)
-22 X (3E+14 , 1E-10)
10
-24 X (1E+15 , 1E-9)
10 X (1E+14 , 1E-13)
-26 X (9E+15, 1E-9)
10
-28 X (1E+15 , 2E-13)
10
-30
10 -2 -1 0 1 2 3 4 5 6
10 10 10 10 10 10 10 10 10
Eν(TeV)
Figure 4.17: Shows the effective magnetic moment for parameters y = 1, M∗ = 104 T eV , The
effective magnetic moment is plotted as a function of available energy for n = 1, 2, 4 and 6
dimensions.

Active neutrinos could transform to kth sterile bulk neutrino, due to the magnetic moment
interaction, at a rate proportional to N αµ2(k) , where N is the number of available bulk neutrino
modes and α is the fine structure constant.
The total effective magnetic moment, µef f , is a sum over the full multiplicity of the available
KK modes [111],
CHAPTER 4. ULTRA-HIGH ENERGY NEUTRINOS 88


µef f = N µ(k) (4.25)
n/2 n/2
10−5 Eν
 
p 1T eV
= 4 × 10−8 µB y An × .
2π(10M eV ) M∗

We can see that neutrino magnetic moment depends on the neutrino energy, on the number of
extra dimensions and on the string scale, M∗ ; An is the volume of the positive hemisphere of an
n-dimensional unit sphere. The case of UHE neutrinos has already been discussed previously
by [112]. In this thesis, we also show our computations of the neutrino magnetic moment for
extra dimensions in the case when M∗ ∼ 104 TeV. The results are shown in Fig. (4.17). In
this figure we show that different combinations of the neutrino energy and of the number of
extra dimensions, n, could lead to neutrino magnetic moments that are in agreement with
the current constraints and that may also produce the neutrino conversion discussed in the
previous sections.
As a conclusion, in this chapter we have made a phenomenological analysis where it is shown
that a non zero neutrino magnetic moment could produce an efficient active-sterile conversion
in the UHE regime for astrophysical objects, such as GRBs and AGNs. As a perspective,
we have outlined possible models that explain such a neutrino magnetic moments; we have
also discussed in this chapter the possibility of having an efficient conversion for galactic and
intergalactic magnetic fields.
Chapter 5

Dark matter, sterile neutrinos


and resonant effects

In this chapter we will outline a mechanism that could connect two interesting, though elusive,
topics: dark matter and sterile neutrinos. In order to discuss this mechanism it will be
necessary to briefly review these two topics separately.

5.1 Dark Matter


Since the publication of Principia in 1687, by Isaac Newton, people have been making a big
effort to explain the motion of the astrophysical objects through the Gravitation Theory. The
deviations of the celestial bodies from the expected trajectories are a very effective method to
understand the behavior of the Universe on a macroscopic scale. The observation of anomalies
in the motion of planets raised questions that could indicate either a limit in the range of
validity of the gravitational laws or the presence of an unseen object (see for example [113]).
In the case of Uranus, the second possibility proved to be the correct one, by the discovery of
Neptune in 1846 by J.G. Galle. Moreover, it was attempted to explain the Mercury anomalies
by the same method, but the explanation failed. The answer was found in the Einstein’s
theory of General Relativity, which is a more refined treatment of the Newton’s laws. With the
current dark matter problem, we are facing a question similar to the old problems about the
motion of planets. In this case, some anomalies are recorded in large astrophysical systems,
whose sizes cover galactic and cosmological scales. In order to explain these anomalies, it has
been proposed the existence of a large amount of unseen matter known as Dark Matter; it can

89
CHAPTER 5. DARK MATTER, STERILE NEUTRINOS AND RESONANT EFFECTS 90

be mentioned that the deviation from the laws of General Relativity has also been considered
as a possible solution to the dark matter problem [113].
The existence of a large amount of matter in the Universe, which has not been seen, is a
problem which has remained since 1930s. Some of the evidence for this matter comes from
the dynamic of galaxies and galaxy clusters.
The observation of the rotation curves of galaxies provides the most plausible and direct
evidence for the existence of the dark matter and is based on the study of the circular velocities
of stars and gas around the galactic center. In the case of the Milky Way which is an spiral
galaxy, the Earth is at an approximated distance of ∼ 8.5Kpc from its center. The stars and
gas cover a distance of about 10 Kpc. According the Virial theorem the circular velocity is
r
GM (r)
v(r) = . (5.1)
r
At large distances, the rotations curves show a flat behavior as we can see in the Fig. (5.1).

Figure 5.1: Rotation curves of NGC 6503. Taken from [17].

According to Eq. (5.1), and assuming that all the mass in the Galaxy is composed by the visible
stars and gas, the rotation curve should decline at radii larger than approximately 10 kpc. But
the experimental observation does not agree with the prediction (See for example Fig. (5.1))
and indicates that the velocity remains constant for large radii, giving rise to a flat curve.
This implies that M (r) ∝ r for r ≫ 10 kpc. Therefore, apart from the stars and gas, there are
additional ’dark’ matter contributing to the total mass distribution.
CHAPTER 5. DARK MATTER, STERILE NEUTRINOS AND RESONANT EFFECTS 91

If we consider a matter distribution with spherical symmetry, the mass is given by

Z r
M (r) = 4π ρ(r′ )r′2 dr′ , (5.2)
0

where ρ(r′ ) is the mass density profile. Based on the behavior of the rotation curve, we
can get an estimate for the dark matter distribution in the galaxy. For small r, v(r) ∝ r,
M (r) ∝ r3 and, as a consequence, ρ(r) ∼ const. On the other hand, for large r, v(r) ∼ const ,
M (r) ∝ r and the density, ρ(r) ∝ 1/r2 . Thousands of individual galaxies form Galaxy clusters.
The mass of these objects is determined by several methods such as the virial theorem,
weak gravitational lensing, and the study of the emission of X-rays [113]. When the virial
theorem was used for determining the velocities of a galaxy clusters, the results indicated
the observation of much more matter than the expected [17]. It is not possible to explain the
observations from galaxy clusters assuming a purely baryonic stars and gas content. For that
reason, it is again supposed the existence of additional matter which has a much larger mass
than the observed.

On other hand, based on the Einstein’s theory of General Relativity, if the light propagates
along a geodesic and passes a near intense gravitational field, it is deviated from its original
trajectory. This effect is known as gravitational lensing. The mass of a galaxy cluster can
be derived from the distortion of objects suffered by strong gravitational fields [114]. The
deflection angle can be written as

 1/2
GM
α∼ , (5.3)
dc2
where M denotes the cluster mass, d the impact parameter, c the speed of light, and G
the gravitational constant. Therefore, measuring the impact parameter and the deflection
angle, it is possible to calculate the mass M of the cluster, which is, again, much larger than
the observed baryonic mass Mb . Another evidence for the existence of dark matter is the
Cosmic Microwave Background (CMB), which is the radiation present in the universe since
its formation. As it is described by theoretical models, when the photons decoupled from the
rest of the matter, they began to propagate through the Universe until today. George Gamow
predicted the existence of this radiation in 1948 and it was inadvertently discovered by Arno
Penzias and Robert Wilson in 1965 [17]. Currently, we know that CMB precisely follows the
radiation pattern of a black body with a temperature T = 2.726o K.
Over time, measurements of the CMB radiation in large scale structures have been made, the
CHAPTER 5. DARK MATTER, STERILE NEUTRINOS AND RESONANT EFFECTS 92

results provide strong evidence regarding the abundance of dark matter and its nonbaryonic
nature. If, as we have stressed above, the dark matter is of non-baryonic nature, then we are
facing a big question:What is the nature of dark matter?. The answer to this question has been
a topic of intense study for many years. Throughout this search, possible candidates to dark
matter have been proposed, ranging from neutrinos to supersymmetric particles. According
with [115] in order to be considered a good dark matter candidate, a particle should answer
all the questions listed below

• Is it in agreement with the relic density?

• Is it cold?

• Is it neutral?

• Is it consistent with the predictions of BBN?

• Does not it affect the stellar evolution?

• Constraints on self-interactions are consistent with it?

• Is it compatible with the experimental results about direct searches, gamma-ray and
astrophysical bounds?

• Is it possible to detect it?

Several candidate particles have been proposed to explain the dark matter problem that, at
least in principle, also try to answer the previuos questions. We show below an incomplete list
of them.
Weakly Interacting Massive Particles (WIMPs)
The WIMPs are particles similar to neutrinos, they interact by the weak force and gravity
and, perhaps, through interactions weaker than weak interactions. The WIMPs have
analogous properties to the neutrinos, but they are more massive and slower than the latter.
The velocities of the WIMPs would be low because of their large masses, this fact would make
them to clump together. Therefore, these particles are considered as main candidates to be
Cold Dark Matter, which are sub-relativistic particles. The mass suggested for the WIMPs is
of the order of Mχ ∼ 10 − 103 GeV [17].

Axions
The existence of the axion is predicted by a solution to the CP violation problem. They
CHAPTER 5. DARK MATTER, STERILE NEUTRINOS AND RESONANT EFFECTS 93

are considered as low-mass pseudoscalars like the neutral pions, but their masses and
interactions are suppressed in comparison with the pions [116]. These particles have also
been contemplated as possible dark matter candidates. Limits on the axion mass of the order
of m ≤ 0.01 eV are obtained from measurements of the dynamics of the supernova 1987A.

Sterile Neutrinos
A sterile neutrino is a particle that does not interact by weak interactions as active neutrino
does. Since long time ago these particles have been considered as possible dark matter
candidates [117]. Sterile neutrinos with masses < 10keV have been ruled out by the results
of WMAP [113].

Supersymmetric candidates
* Sneutrinos
In Supersymmetry, the sneutrinos are defined as the superpartners of the neutrinos of the
Standard Model. The sneutrinos could contribute to relic density of the universe if their mass
is in the range of 550 to 2300 GeV.
* Neutralinos
In the Minimal Supersymmetric Standard Model (MSSM) the binos (B̃) and winos (W̃3 ) are
defined to be the superpartners of the Standard Model bosons B and W3 . On the other hand,
the superpartners of the neutral Higgs bosons H10 and H20 are called higgsinos (H̃10 and H̃20 ).
The previous states are mixed into four Majorana mass eigenstates, given by χ̃01 , χ̃02 , χ̃03 and
χ̃04 , which are known as neutralinos [115, 113].

* Gravitinos
In a supersymmetric theory, the superpartners of the graviton are called gravitinos. They are
theoretically well motivated since they could be stable.

Kaluza-Klein states
As we saw in chapter 3, the Kaluza-Klein states appear in theories with extra dimensions
and could also be dark matter candidates.

Superheavy dark matter


The superheavy dark matter candidates also known as Wimpzillas are particles with masses
of the order of mDM > 1010 GeV [115, 113].
CHAPTER 5. DARK MATTER, STERILE NEUTRINOS AND RESONANT EFFECTS 94

5.2 Sterile Neutrinos


A sterile neutrino is defined as an SU (2) singlet which does not couple by standard
interactions. These particles are motivated, from the experimental point of view, by some
anomalies that appear in experiments of solar, accelerator and reactors neutrinos. Moreover,
sterile neutrinos are present in many theoretical models, which include physics beyond
the Standard Model, for instance, the see-saw mechanism that gives an explanation to the
smallness of the neutrino mass. According to the expression (2.63), a sterile neutrino is
described by the Lagrangian:

D−M 1 
Lmass = −mD ν̄L νR − mL νLC νL + mR νR

R + h.c. (5.4)
2
where mD is a Dirac mass, and mL and mR are the Majorana masses. The fields νR and νLC
represent sterile neutrinos.
Defining  
νL
ν=  (5.5)
C
νR
The equation (5.4) can be rewritten as
 
D−M 1 mL mD
Lmass = − ν C Mν + h.c with M= . (5.6)
2 mD mR
In accordance with the Eq. (5.6) there are several special cases:

1. The pure Majorana case: mD = 0. In this instance, there is no mixing between the active
and sterile states, and the sterile neutrino decouples (unless there are new interactions).

2. The pure Dirac case: mL = mR = 0. This leads to two degenerate Majorana neutrinos
which can be combined to form a Dirac neutrino with a conserved lepton number.

3. The seesaw limit, mR ≫ mD,L . In this case we have a heavy state (m2 = mR ) and a light
one (m1 ≃ mL − m2D /mR ). The mL = 0 case correspond to the see-saw mechanism

4. The pseudo-Dirac limit, mD ≫ mL , mR , which leads to a small shift in the mass


eigenvalues m1,2 = mD ± (mL + mR )/2.

5. The active-sterile mixed case, where mD ∼ mR or mD ∼ mL . In this instance, there is


significant mixing between active and sterile neutrinos.
CHAPTER 5. DARK MATTER, STERILE NEUTRINOS AND RESONANT EFFECTS 95

5.2.1 Astrophysical evidence

Although the experimental hints reported on sterile neutrinos, there is no confirmation of its
existence. However, it is interesting for us to briefly review the experimental evidence. We
start this review by enumerating the astrophysical evidence as follows.
In the early Universe, left-handed Standard Model neutrinos were in thermal equilibrium
with the other particles of the Standard Model through weak interactions. The interactions
decreased due to the expansion of the Universe and the cooling of the primordial plasma.
When the temperature of the plasma was of the order of 1 MeV, the interaction rate was lower
than the expansion rate and, as a consequence, neutrinos were decoupled from the primordial
plasma [118, 18].
At T ∼ 0.2MeV the electron-positron annihilation took place. This process caused the
reheating of the photons, but the neutrinos did not feel this reheating because they had
already been decoupled. Therefore, the temperature at which emerged the neutrinos is
slightly lower than that of the photons. The estimation for the neutrino temperature is given
by

 1/3
4
Tν = Tγ , (5.7)
11
where Tν denotes the neutrino temperature and Tγ corresponds to the photon temperature
after the e+ e− annihilation processes [118, 18]. The total neutrino energy density in the
radiation era (after e+ e− annihilation) is given by

7π 2 4
ρν = Nef f T . (5.8)
120 ν
Here, Nef f corresponds to the effective number of neutrino families. The number of active
neutrino flavors that participate in weak interactions of the Standard Model has been
determined from the Z 0 decay width to be 2.984 ± 0.008 [119]. In agreement with Eq. (5.8), the
effective number of neutrino families can be determined from Cosmology, because the energy
density affects the expansion rate of the Universe during the radiation era. The expansion
rate H(t) is written as

8πG
H 2 (t) ≃ (ργ + ρν ). (5.9)
3
The photon density ργ is well determined from measurements of CMB temperature. Limits on
H(t) in the early universe can be traduced in constraints on ρν and, according with Eq. (5.8),
the previous limits are interpreted as bounds on Nef f . Therefore, if a process alters the
CHAPTER 5. DARK MATTER, STERILE NEUTRINOS AND RESONANT EFFECTS 96

neutrino abundance or changes the expansion rate, this would alter the value of Nef f .
There is evidence about sterile neutrinos in Cosmology, for instance, from abundances of light
elements from BBN, from anisotropies of the CMB, and from the distribution of large-scale
structures. We will briefly describe the observations from Big-Bang Nucleosynthesis and
from anisotropies of the CMB.

Big Bang Nucleosynthesis (BBN)


Sterile neutrino and primordial elemental abundances
Protons and neutrons are the main components of the light elements, such as Hellium. The
formation of these elements begin with the synthesis of deuterium (D) at a temperature of
approximately O.1 MeV, subsequently, nuclear reactions give rise to the production of heavier
nuclei. Most of the neutrons initially available are found in bounded states of 4 He; the
resulting fraction of protons, in terms of neutrons and 4 He is

4n4 He
Yp = ∼ 0.25, (5.10)
nn + np
where nx denotes the number density of neutrons and protons. The neutrinos play an
important role in this picture; in the first place, the electron neutrinos are involved in the
weak interaction of charged currents which determine the production rate of neutrons and
protons as follows

ν̄e + p ↔ n + e+ , (5.11)

νe + n ↔ p + e− . (5.12)

On the other hand, the neutrinos alter the expansion rate of the Universe (Eq. (5.9)) during
the Big-Bang Nucleosynthesis. If the expansion rate changes, the neutron-to-proton ratio is
altered as well, which is traduced in a modification of the light elements abundance, meanly
of 4 He. In conclusion, the Big-Bang Nucleosynthesis can be altered by changing the Hubble
expansion rate which directly affects the effective number of neutrinos Nef f . There are
different experimental measurements for the Nef f from BBN, for example, in the Ref. [118],
+0.47
it is reported the value Nef f = 3.71−0.45 , considering that there is no lepton asymmetry. And
if the He abundance is fixed to the value got from CMB measurements, the neutrino effective
+0.66
number decreases to Nef f = 3.53−0.63 [118].
Current Constraints on Nef f from the anisotropies in the CMB
The Cosmic Microwave Background was discovered by Penzias and Wilson in 1965 [120].
CHAPTER 5. DARK MATTER, STERILE NEUTRINOS AND RESONANT EFFECTS 97

Since then, the people have been looking for the anisotropies in the CMB. In 1977 Smoot
and collaborators reported convincing evidence relating with a dipole anisotropy. Years later
in 1992, the Cosmic Background Explorer (COBE) presented results for anisotropies to higher
orders [121]. After COBE and the Wilkinson Microwave Anisotropy Probe (WMAP), Planck is
the third generation of space mission focused to measurements of the anisotropies in CMB.
The main goal of the Planck Collaboration is to measure the polarization and the temperature
of the anisotropies with a sensitivity of the order of µK [121].
On the other hand, the effective number of neutrinos can be constrained through the
measurements of CMB anisotropies [122]. The combination of experimental results from
Planck, WMAP polarization (WP) and from analysis with high multipoles (l ≥ 40) yields [121],

+0.68
Nef f = 3.36−0.64 (with 95% C.L.) (5.13)

Combining Planck, WP, High-l and Baryon Acoustic Oscillations (BAO), the result is given by

+0.54
Nef f = 3.30−0.51 (with 95% C.L.) (5.14)

Moreover, as we saw above, the effective number Nef f is correlated with the Hubble constant
H0 . Then, the combination of direct measurements of H0 with Planck, WP and high-l leads to

+0.50
Nef f = 3.62−0.48 (with 95% C.L.) (5.15)

Finally, the inclusion of BAO in the previous analysis produces the following values for Nef f ,

+0.48
Nef f = 3.52−0.45 (with 95% C.L.) (5.16)

For a detailed description of the above results, see [121].

5.3 Evidence for sterile neutrinos from oscillation


experiments
Since the discovery of neutrino oscillation phenomenon, many oscillation experiments have
been carried out around the globe, this has allowed measuring with sufficient precision
the neutrino parameters, such as mass splittings and mixing angles. The three neutrino
framework accounts for almost all data from solar, reactor, atmospheric and accelerator
experiments. However, there are some results which can not be explained considering three
CHAPTER 5. DARK MATTER, STERILE NEUTRINOS AND RESONANT EFFECTS 98

neutrino families. These discrepancies between measured and expected data are known as
anomalies. In order to understand these anomalies different scenarios have been proposed,
among them, the most popular explanation comes from considering the existence of additional
neutrinos which would be sterile neutrinos.
The Liquid Scintillator Neutrino Detector (LSND) was the first experiment that reported an
anomaly in its observations. A search for ν̄µ → ν̄e oscillations using ν̄µ from µ+ decay at
rest led to a > 3σ evidence for oscillations with ∆m2 > 0.2eV 2 [123]. Some years later,
the MiniBooNe experiment was constructed to test the LSND results, making searches in
the neutrino and antineutrino channels. MiniBooNe reported the observation of an excess
of 78.4 ± 28.5 events (2.8σ) in the energy range 200 < EνQE < 1250 MeV [124]. The results
obtained are not in agreement with the parameter space reported by LSND in the neutrino
mode; however, the data for the antineutrino channel is consistent with the LSND signal with
∆m2 ∼ 1eV2 [124].
On the other hand, we have evidence for oscillations to sterile neutrinos from the
measurements of the neutrino flux from radioactive sources in SBL reactor experiments.
In what follows we will briefly describe the anomalies from reactor, accelerator and solar
experiments.

5.3.1 The accelerator anomaly

In a typical oscillation experiment, the observed event rate is compared with the predicted
event rate; if these two rates are different one from each other, this is known as an anomaly.
The results are interpreted in terms of new mixing parameters ∆m2 and θ. In the case of
accelerator experiments, this anomaly gives ∆m2 ∼ 1eV2 as a possible solution. As already
mentioned, the first evidence for sterile neutrinos was collected by the LSND experiment
which was confirmed by the MiniBooNe experiment, being both accelerator experiments.
However, the KARMEN experiment has not reported an event excess, although it has not
rule out the complete parameter regions allowed by LSND. The combination of LSND,
MiniBooNe, KARMEN and ICARUS data gives as a result a parameter region centered around
(∆m2 ∼ 0.5eV 2 , sin2 2θ ∼ 5 × 10−3 ) [125, 126].

5.3.2 The Gallium anomaly

In the case of solar neutrinos, experiments such as GALLEX [127] and SAGE [128] have
performed different tests related with the calibration of their detectors. During the tests,
CHAPTER 5. DARK MATTER, STERILE NEUTRINOS AND RESONANT EFFECTS 99

Figure 5.2: Allowed regions in the sin2 (2θnew ) − ∆m2new parameter space from combined
analysis from reactor and gallium experiments, the star represent the best fit. Taken
from [18].

51 37
the detectors were exposed to artificial radioactive sources of Cr and Ar, the final results
reported the observation of an experimental number of events lower from expected.
The neutrino production is given by electron capture as follows

e− +51 Cr → 51
V + νe (5.17)

e− +37 Cr → 37
Cl + νe (5.18)

while the emitted neutrinos are detected by the reaction

νe +71 Ga →71 Ge + e− . (5.19)

The average ratio of measured to predicted events is R = 0.86 ± 0.05. This is known as the
Galium Anomaly [129].
If this anomaly is interpreted in terms of neutrino oscillations, gives as a result the mixing
parameters sin2 θ ≥ 0.07 and ∆m2 ≥ 0.35eV2 at 99% CL [129].
CHAPTER 5. DARK MATTER, STERILE NEUTRINOS AND RESONANT EFFECTS 100

5.3.3 The Reactor Antineutrino Anomaly

Another motivation for studying sterile neutrinos is given by some refined calculations
performed about of antineutrino spectra in reactors. In this case, the anomaly is manifested
by a new evaluation of the neutrino flux produced in reactor experiments, which yields an
increase in the flux by about 3% [130]. With this reevaluation, R = Nobs /Npred = 0.943 ± 0.023,
which is known as the reactor antineutrino anomaly [130].
Once more, the oscillations to sterile neutrinos could provide an explanation to the gallium
and reactor anomalies, a combined analysis is showed in Fig. (5.2) and the following best fit
parameters are obtained: ∆m2new = 0.5eV2 and sin2 2θnew ∼ 0.14 [18].

5.4 A resonant effect in dark matter


As we saw in chapter 2, neutrinos propagating through a material medium may suffer an
enhancement effect in the oscillation probability, the so called MSW effect [131]. The standard
MSW effect takes into account the interaction of active neutrinos with electrons and quarks.
We have studied [132] a possible consequence of the MSW effect in a different regime. If we
consider that active and sterile neutrinos can mix, then, astrophysical neutrinos propagating
from distant objects might interact with the dark matter along its trajectory to the Earth and
undergo an analogous MSW effect. If there is indeed an interaction between sterile neutrinos
and dark matter, there is a possibility of having a dark matter density inducing a resonant
effect, just as in the case of ordinary matter. We will describe in the rest of this chapter how
this mechanism could work and potential regions where the resonance could take place.
We can consider first the neutrino evolution equation that describes active and sterile
neutrinos in the presence of ordinary and dark matter potentials. Restricting ourselves to
the case with only one family of active neutrinos, να=e,µ,τ and one family of sterile neutrinos,
νs , we will have, for a neutrino energy, E [132],
   
d  να  1 να
i = Mα  , (5.20)
dt νs 4E νs
where,  
−∆m2i4 cos 2θ0 + Vνα f + Vνα χ ∆m2i4 sin 2θ0
 
Mα =  . (5.21)
 
 
∆m2i4 sin 2θ0 ∆m2i4 cos 2θ0 + Vνs χ
CHAPTER 5. DARK MATTER, STERILE NEUTRINOS AND RESONANT EFFECTS 101

Here, ∆m2i4 = m24 − m2i , θ0 , is the vacuum mixing angle relating the active and the
sterile neutrino; Vνα f = VνCC
αf
+ VνNα fC , is the interaction potential of the active neutrino with
ordinary fermions; Vνα χ is the potential for an interaction between active neutrinos and dark
matter. Finally, Vνs χ accounts for the potential arising from the interaction between sterile
neutrinos and dark matter. This kind interactions naturally appears in extensions of the
Standard Model. In particular, several models consider that many dark particle candidates,
including sterile neutrinos, may take part in the dark matter sector and interact among
themselves [133, 134, 135]. In principle, we might also include in Eq. (5.21) the potential
due to the interaction of sterile neutrinos with ordinary fermions, Vνs f , that we will consider
to vanish; previous work, where this last potential has been studied, can be found in Ref. [136].
We can derive, from Eq. (5.20), the resonance condition [132]

∆m2i4 cos 2θ0 = 2E(Vνα f + Vνα χ − Vνs χ ) . (5.22)

These potential can be written as

1 g2 √
Vνα f = (Nα − Nn /2) = 2GF (Nα − Nn /2) ; (5.23)
4 m2W
g να g χ
Vνα χ ∼ Nχ = G′να Nχ = ενα χ GF Nχ ; (5.24)
m2I
g νs g χ
Vνs χ ∼ Nχ = G′νs Nχ = ενs χ GF Nχ , (5.25)
m2I
where, Nα is the lepton number density, Nn is the neutron number density, and Nχ is the dark
matter number density accounting for the possible dark matter particles that may interact
with the neutrinos. mW is the W boson mass, g, the Standard Model coupling constant, and
gνα , gνs , and gχ are the coupling constants of the given particle (active or sterile neutrino,
or dark matter candidate, respectively). The presence of an intermediary gauge boson, with
mass mI , is also required for this mechanism. On the other hand, ενα,s χ stands for the coupling
strength in terms of GF , the Fermi constant.
The resonance condition can now be written as [132]


∆m2i4 cos 2θ0 = 2EGF [ 2(Nα − Nn /2) + (ενα χ − ενs χ )Nχ ] . (5.26)

From this equation, it is possible to compute the region where a resonant effect can take place.
In order to perform such a computation wi will consider that the main contribution to the dark
matter density, Nχ , comes from a single dark matter particle candidate with mass mχ . Then,
the expression for the density will be given by Nχ = ρχ /mχ . With this expression we can
compute an estimate for ενα,s χ . Before computing our results, we could study in more detail
CHAPTER 5. DARK MATTER, STERILE NEUTRINOS AND RESONANT EFFECTS 102

(gχ )(gν )
Ref (mI /M eV )2 εν e χ εν s χ mχ (eV)
−6 −1
(0.7)(10 −10 )
Aarssen et. al. [19] 10−2 −100 0 105 − 1015 1012
(1)(1)
Mirror [133, 134] (30MW )2 0 10−3 109
10−6
Fayet [21, 20] < (100 )2 < 105 0 107
−3
10
Mangano et. al. [137] < 100 < 108 0 107

Table 5.1: Coupling constants and mass estimates from different models [132].

the possible estimated values of gνs and gχ . We can do these estimates by considering recent
models where an intermediary boson, with a light mass, mI , has been included [19, 137, 21,
20, 138, 139]. We show a list of such models in Table 5.1, where we have also shown the
corresponding ενα,s χ values; in compiling this table, we have focused in models with an extra
neutrino species, that could, therefore, have a contribution in our neutrino potential. In order
to give a better description of these models we will introduce a description of them in the next
subsections.

1. Neutrino-dark matter interactions and small-scale problems

In order to find a solution to the structure formation problems at small scales, van den Aarssen
and collaborators [19] have considered the interactions of neutrinos (active or sterile) with
dark matter and the dark matter self-interactions. Despite the success of the standard ΛCDM
cosmological model, its predictions are not in agreement with the observations at, small,
galactic scales [19].

Figure 5.3: Interaction processes that could give a solution to the small-scale problems. Taken
from [19].

In order to present a solution that accounts for these problems, interacting DM models have
been examined, which contain Yukawa-like interactions between the DM particles, mediated
by a light messenger. The interaction processes are shown in the Fig. (5.3) [19]; the left
CHAPTER 5. DARK MATTER, STERILE NEUTRINOS AND RESONANT EFFECTS 103

diagram represent the annihilation of dark matter into neutrinos, the middle one shows the
dark matter self-interactions and the right one displays the interaction between neutrinos
and dark matter.

Figure 5.4: Mass of the vector intermediary particle vs mass of dark matter candidate (left
panel). Mass of the vector intermediary vs the coupling strength gν (right panel). Taken
from [19].

The Fig. (5.4) [19] shows the parameters region that would give a solution to the current
small-scale problems. This solution is possible for DM ≥ 600 GeV and an intermediary boson
with a mass between 0.05 MeV and MeV (left panel). In the same Figure, different values
of the mass cutoff (cutoff in the primordial power spectrum of density perturbations) are
displayed in the plane MV vs gν .
In summary, the problems with structure formation to small scales could be solved
simultaneously by considering a (sub)MeV vector messenger particle which is weakly coupled
to active neutrinos or to sterile ones.

2. Mirror Model

The existence of a shadow world or mirror universe has been motivated by superstring
theories and sterile neutrinos. This mirror world would have particles and interactions
identical to the visible world but interacting with the later via gravity and, perhaps, other
very weak forces. In the past, a mirror model was proposed so as to explain the LSND
anomaly [141, 142]. Moreover, mirror baryons have been considered as dominant dark matter
of the Universe [143], and in a more recent work, experimental data of DAMA, CoGeNT and
CRESST-II could simultaneously be explained considering mirror metals [144].
In string theories, the GSM × G′SM group is considered, where the group GSM = SU (3) ×
CHAPTER 5. DARK MATTER, STERILE NEUTRINOS AND RESONANT EFFECTS 104

SU (2) × U (1) represents the Standard Model with quarks qi , uci , dci , leptons li , eci and a Higgs
doublet φ, and G′SM = [SU (3) × SU (2) × U (1)]′ denotes the mirror analogue with the particle
content qi′ , u′c ′c ′ ′c ′
i , di , li , ei and φ [143].

Then, in this framework, a sterile neutrino is a mirror neutrino which interact with standard
model particles only by gravity, or by another force, weaker than the standard weak
interaction.

3. Dark matter couplings

Figure 5.5: Feynman rule for dm-fermion interactions. Taken from [20].

Instead of considering a particular model, it is also possible to study in a phenomenological


approach the possible couplings of dark matter with ordinary particles.
There are several ways through which dark matter could be coupled to ordinary matter
particles [20, 21], for instance, an exotic scalar dark matter particles could couple ordinary
fermions, f , with new exotic fermions, F . The mass of F s should be above ∼ 100GeV s in
order to be compatible with new particles searches. In this case, the Feynman rule is given
by Fig. (5.5), where PL and PR represent the left and right-handed projector, respectively. The
Yukawa couplings are given by

δ(Cl f¯L FR + Cr f¯R FL ) + h.c. (5.27)

Where δ denotes the spin-0 dark matter field, and Cl and Cr are its Yukawa couplings to the
left-handed and right-handed ordinary fermion fields.

Another possible interaction between dark matter particles and ordinary matter may
be produced through Z interchanges. Since these dark matter candidates have not been seen
in Z decays, they should be heavy for this case [20, 21].
Finally, instead of the Z and W gauge bosons, there could be a new neutral boson U which
mediates the interaction between dark matter particles and ordinary fermions. In this case,
CHAPTER 5. DARK MATTER, STERILE NEUTRINOS AND RESONANT EFFECTS 105

Figure 5.6: DM-ordinary fermions interactions by a new boson U . Taken from [20].

the dark matter candidates could be, in principle, light particles. The Feynman rules are
shown in the Fig. (5.6), where, CU is the coupling for a scalar dark matter field, CUl and
CUr are the coupling constants for a Dirac dark matter field, while fUl and fUr represent the
couplings of the U boson with ordinary fermion fields [20].
It is possible to constraint the interaction between neutrinos and dark matter by applying this
phenomenological approach to the cosmological observations. Such effects have been studied
by Mangano and collaborators [137]. Moreover, interactions of dark matter with neutrinos
have been proposed by several authors [145, 146, 147]. In these scenarios, the Lagrangian
that describes the interaction DM-neutrinos is given by

Lint = igψ (ψ ∗ ∂ µ ψ − ψ∂ µ ψ ∗ )Uµ + gψ2 ψ ∗ ψUµ U µ + gν ν̄L γ µ νL Uµ , (5.28)

where Uµ denotes an intermediate vector-boson field Uµ , ψ is the DM field, gψ and gν are the
couplings DM − U and DM − ν, respectively [137]. In the range of neutrino temperature
T ≤MeV, the thermally averaged ψ − ν scattering cross section is as follows

T2 T2
hσdm−ν i ∼ gψ2 gν2 4 = g2 4 (5.29)
mU mU
Where, g = gψ gν . The bound on these DM-neutrino coupling is shown in Table 5.1.

5.5 Resonant effects of the ν − DM and νs − DM


interactions
After the previous discussion, we can see that the coupling of active neutrinos with dark
matter is strongly constrained (gχ gν ∼ 10−6 [21, 20]). Moreover, recent studies of a possible
CHAPTER 5. DARK MATTER, STERILE NEUTRINOS AND RESONANT EFFECTS 106

ν − DM through the Z gauge bosons have also found restrictive limits [139]. For this reason
we will consider in the following only the case of an interaction between sterile neutrinos and
dark matter, despite our phenomenological analysis could be easily translated to the more
restricted case of an interaction between dark matter and active neutrinos [132].
We will search now physical processes that may be sensitive to the neutrino potential, VνS χ .
There are already plenty of constraints for the mixing between active and sterile neutrinos,
for example, from primordial big bang nucleosynthesis [152, 154, 155, 156, 140, 157]. In
this case, the production rate of sterile neutrinos will be given by considering the interaction
rate of active neutrinos times the oscillation conversion probability to sterile neutrinos, Γνs =
(hPνα →νs i)Γνα . This conversion probability, hPνα →νs i, will depend on the interaction potential
√  2

of neutrinos and is given by V = 2GF Nγ L − A MT 2 , where L is the fermion asymmetry,
W

Nγ accounts for the photon number density, and A is a numerical factor [154]. It is important
to notice that this potential has been computed considering only the interaction of active and
sterile neutrinos with the ordinary matter; therefore, a computation of the contribution of the
interaction potential for sterile neutrinos with dark matter may be of interest [132].
We now turn our attention to the impact of this neutrino potential in the UHE regime, for
objects coming from extragalactic sources [159, 160, 161]. Currently, different constraints
on the UHE neutrino flux have been reported, for instance, from Icecube [84], Pierre Auger
Collaboration [162], and ANTARES [163]. Besides, Icecube has reported the detection of
two neutrino events in the energy range of PeV [164]; it is expected that, with the future
accumulation of registed events in IceCube and Auger, with neutrino energies around 1015 eV
and 1017 eV, respectively, there would be a better understanding of the extragalactic neutrino
spectrum. On the other hand, the previously reported constraints on UHE neutrino fluxes
have motivated the proposal of models that predict low neutrino fluxes [165]. Besides, it has
also lead to the proposal of mechanisms that could induce a decrease in the neutrino detection
through the existence of new physics [166, 167, 168, 169, 170, 26]. In particular, solutions
involving pseudo-Dirac neutrinos have been extensively discussed [168, 169, 171, 172, 173,
174].
We have studied [132] the cases where our neutrino potential could induce a resonant effect
for neutrinos in UHE regimes. This could be of interest since, in future, experimental data
might collect enough events as to observe a possible distortion in the neutrino spectrum.
By using Eq. (5.26) we have obtained the values of ∆m2 , εT = |(ενα χ − ενs χ )|, and mχ that
may lead to a resonant effect [132]. In doing this computations we have consider θ0 ≈ 0.
Our results are shown in Fig. (5.7) where we show, for different values of ∆m2 , the resonance
CHAPTER 5. DARK MATTER, STERILE NEUTRINOS AND RESONANT EFFECTS 107

20 20
10 10
15 νl 15 νl
10 10
10 10
10 10
εT

εT
5 5
10 10
2 -12 -18 2
0 2 -12
∆m =10 -10 eV
-18 2 0 ∆m =10 10 eV
10 10
-5 -5
10 6 7 8 9 10 11 12 13 14 10 6 7 8 9 10 11 12 13 14
10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10
mχ(eV) mχ(eV)

Figure 5.7: Isocurves for the resonant effect of (anti)neutrinos with ∆m2 = 10−12 − 10−18 eV 2 .
−3
We take a constant dark matter density, ρχ = 0.3 GeV cm , consistent with the average
value of the galactic halo. We also show a region of restricted values reported in the
literature [21, 20] (dark blue box) and as region of values motivated by a solution to the
small scale problem [19] (light blue box). The left panel corresponds to a neutrino energy of
E = 1018 eV while the right one to E = 1015 eV [132].

region in the plane εT vs mχ . It is important to notice that (ενα χ − ενs χ ) changes sign according
to the coupling we are considering, either with sterile or with active neutrinos; therefore the
resonant effect will take place only for neutrinos (if ενα χ 6= 0) or for antineutrinos (if ενs χ 6= 0).
We have considered, first, the dark matter in our galactic halo, where the electron potential is
given by Ve = 5 × 10−39 eV [175] and, therefore, Ne = 3 × 10−6 eV 3 . We have taken the average
value ρχ = 0.3 GeV · cm−3 [144]. As we have already mentioned, we have considered different
values of ∆m2 ; we show in Fig. (5.7) a tilted band corresponding to ∆m2 = 10−12 − 10−18 eV 2 ,
a region already discussed before for this regimes, but for just-so oscillation scheme [168, 169,
171, 172, 173, 174]. In the left panel of Fig. (5.7) we show the case of a neutrino energy,
Eν = 1018 eV. The right panel corresponds to the case of Eν = 1015 eV. We also show the
region of parameters proposed in the work of Aarssen et. al [19] (light box), remaped in
this case in the plane εT -mχ . As we already discussed, there are also restrictions for the
case of an interaction between active neutrinos and dark matter; we show in this figure the
constraints reported by [21, 20]. It can be noticed from the plot that for higher masses of the
dark matter particle higher couplings are necessary in order to have a resonance [132]. It
is very interesting to note from Fig. (5.7) that an overlap is present between the tilted ∆m2
region and the parameters proposed, in a different context by Aarssen et. al. [19]. On the
other hand, comparing with the reported constraints for a light dark matter candidate, we
can observe that some values of εT are allowed.
CHAPTER 5. DARK MATTER, STERILE NEUTRINOS AND RESONANT EFFECTS 108

In a previous work, it has been discussed that a just-so oscillation could produce an UHE
neutrino flux reduction [168]. We would like to stress that, for the same mass difference
region, we could have a resonant effect if the correct coupling to a dark matter candidate
exist. In this way, the mentioned recent proposal would not need the typical fine tuning of the
just-so oscillations but, instead, it may have a more natural resonant effect. It is also worth of
remark that, if this effect takes place in our galactic dark matter halo, then, independently of
the original source of the neutrino flux, there would be an energy region where the resonant
conversion into a sterile neutrino would make the active neutrino flux to vanish. A future
perspective for us in this direction will be to consider a specific dark matter density profile,
instead of its average value, example, a Navarro Frenk White [176, 177] profile.
Finally, we will refine our previous computation by considering a non zero mixing angle. In
this case, the survival probability will be given by [132]

Losc
 
2 2 0
P (να → να ) = sin (2θm ) sin π osc , (5.30)
Lm
where,
sin2 (2θ0 )
sin2 (2θm ) =  2 (5.31)
(Vνα f +Vνα χ −Vνs χ )
cos2 (2θ0 ) 1 − VR + sin2 (2θ0 )
∆m2
and VR = 2E cos(2θ0 ). The matter oscillation length, Losc
m , will be

Losc
0
Losc
m = r  2 (5.32)
(Vνα f +Vνα χ −Vνs χ ) 2
cos2 (2θ 0) 1− VR + sin (2θ0 )

For this case, we have calculated the survival probability, considering different values of
sin2 (2θ0 ). We have taken ∆m2 = 7 × 10−13 eV 2 , εT = 3 × 1011 , and mχ = 2 × 1010 eV. After
computing the survival probability, it is possible to find the resonant energy region [132],
which will be around 8 × 1014 eV. This probability is shown in Fig. (5.8). We can see that a
mixing around sin2 (2θ0 ) = 0.25 gives a maximal conversion.
Another important quantity to compute is the effective oscillation length. In our case, this
length is approximately approximately equal to the galactic halo dark matter region. This
can be seen by computing the matter oscillation length at the resonance point:

4πE
Losc
m = 2 ∼ 1018 km. (5.33)
sin (2θ0 )∆m2
This value corresponds approximately to the Milky way dark matter radius of r =
200kpc [179].
Galactic
1,0
0,9
2
0,8 sin (2θ0) = 0.25
2
P(να→να) sin (2θ0) = 0.15
0,7 2
sin (2θ0) = 0.05
0,6 2
sin (2θ0) = 0.01
0,5 2
sin (2θ0) = 0.008
0,4 2 −13 2
∆m = 7×10 eV
0,3 εT = 3×10
11

10
0,2 mχ = 2×10 eV
0,1 14
E = 8×10 eV
0,0
14 15 16
10 10 10
Eν(eV)

Figure 5.8: Survival probability P (να → να ) for a constant dark matter density, ρχ =
0.3 GeV cm −3 , consistent with the average value of the galactic halo [132].

As a summary of this chapter, we would like to stress that we have study a new mechanism
that could induce a supression in the neutrino flux of extragalactic objects. Although this
mechanism might be difficult to test with future IceCube data, it is interesting that the effect
opens a window to connect two ellusive sectors: dark matter and sterile neutrinos.

109
Chapter 6

Conclusions

The neutrino oscillations mechanism allows to establish that the neutrinos are massive. In the
original Standard Model of Particle Physics the neutrino is massless. This has led to propose
new models that attempt to give an explanation to the neutrino mass. In some extensions
of the SM, the neutrino also gets electromagnetic properties such as the magnetic moment
which is directly related with the mass. In particular, the minimally extended standard model
predicts a value for the neutrino magnetic moment of the order of µν = 3.2 × 10−19 (mi /eV )µB .
On the other hand, terrestrial experiments measurements give the bound µν < 10−11 µB .
In the first stage of this thesis we have obtained the expressions for the effective magnetic
moment in the mass basis, including the mixing angle θ13 , which had not been considered
previously. We performed a combined analysis of the experimental data collected from
measurements of the NMM, which allowed us get bounds on the effective magnetic moment,
both in the flavor as well as in the mass basis; the limit obtained in the flavor basis is
µνe < 2.9 × 10−11 µB .
A next step in this analysis will be to include the Solar neutrino data. With the inclusion of
these data, it will be possible to perform an analysis that will give more information about the
different parameters of the neutrino magnetic moment matrix.
On the other hand, neutrino magnetic moment may also plays an important role in
astrophysical environments; in order to extend our study regarding the NMM, we have
studied possible interactions of the NMM with magnetic fields, which could produce a
conversion from active to sterile neutrino. In particular, we show that a nonzero NMM could
produce a reduction of the UHE neutrino flux coming from gamma ray burst.
Most of the experimental data from oscillations experiments can be described in a three

110
neutrino framework, but some anomalies have been reported which have led to propose the
existence of sterile neutrinos. Additionally, the observation of anomalies in large astrophysical
systems as well as in galactic and cosmological scales has also led to infer the existence of
dark matter. In this thesis, we have explored the possible existence of an effect that may
connect sterile neutrinos, dark matter, and astrophysical neutrinos. If the sterile neutrinos
exist, they could induce a suppression of the neutrino flux from extragalactic sources, such
as GRBs; in this scenario, active neutrinos interact with dark matter particles producing a
resonant effect like the MSW effect, which would cause that active neutrinos are converted
into a sterile one. The mechanisms proposed form part of a purely phenomenological study
about the non-standard properties of the neutrino, the forthcoming results about astrophysical
neutrinos from experiments such as IceCube could rule out or confirm this kind of scenarios.

111
Bibliography

[1] C. Broggini, C. Giunti and A. Studenikin, Adv. High Energy Phys. 2012, 459526 (2012)
[arXiv:1207.3980 [hep-ph]].

[2] http://www.atomicarchive.com/Fission/Fission1.shtml

[3] Fundamentals of Neutrino Physics and Astrophysics, Carlo Giunti and Chung W. Kim
(Oxford University Press 2007)

[4] H. Kolanoski [for IceCube Collab.], arXiv:1209.5610v1 (2012)

[5] L. Amati, J. -L. Atteia, L. Balazs, S. Basa, J. Becker Tjus, D. F. Bersier, M. Boer and
S. Campana et al., arXiv:1306.5259 [astro-ph.CO].

[6] G. S. Bisnovatyi-Kogan, Phys. Part. Nucl. 37, 647 (2006) [astro-ph/0701461].

[7] E. Waxman and J. Bahcall, Phys. Rev. D 59, 023002 (1998)

[8] E. Waxman, arXiv:1101.1155 [astro-ph.HE].

[9] S. Adrin-Martnez, A. Albert, I. A. Samarai, M. Andr, M. Anghinolfi, G. Anton, S. Anvar


and M. Ardid et al., A and A 559, A 9 (2013) [arXiv:1307.0304 [astro-ph.HE]].

[10] M. Ahlers, L. A. Anchordoqui, M. C. Gonzalez-Garcia, F. Halzen and S. Sarkar, Astropart.


Phys. 34, 106 (2010) [arXiv:1005.2620 [astro-ph.HE]].

[11] K. Kotera, D. Allard and A. V. Olinto, JCAP 1010, 013 (2010) [arXiv:1009.1382
[astro-ph.HE]].

[12] P. Abreu et al. [Pierre Auger Collaboration], Adv. High Energy Phys. 2013, 708680 (2013)
[arXiv:1304.1630 [astro-ph.HE]].

[13] IceCube Collaboration, NATURE 484, 351 (2012).

112
[14] M. G. Aartsen et al. [IceCube Collaboration], Science 342, 1242856 (2013)
[arXiv:1311.5238 [astro-ph.HE]].

[15] M. G. Aartsen et al. [ The IceCube Collaboration], arXiv:1311.7048 [astro-ph.HE].

[16] P. Baerwald, M. Bustamante and W. Winter, JCAP 1210, 020 (2012) [arXiv:1208.4600
[astro-ph.CO]].

[17] S. Matarrese, M. Colpi, V. Gorini and U. Moschella, 2011Dark Matter and Dark Energy
(Springer Dordrecht Heidelberg London New York).

[18] K. N. Abazajian, M. A. Acero, S. K. Agarwalla, A. A. Aguilar-Arevalo, C. H. Albright,


S. Antusch, C. A. Arguelles and A. B. Balantekin et al., arXiv:1204.5379 [hep-ph].

[19] L. G. van den Aarssen, T. Bringmann and C. Pfrommer, Phys. Rev. Lett. 109, 231301
(2012) [arXiv:1205.5809 [astro-ph.CO]].

[20] C. Boehm and P. Fayet, Nucl. Phys. B 683, 219 (2004) [hep-ph/0305261].

[21] P. Fayet, Phys. Rev. D 75, 115017 (2007) [hep-ph/0702176 [HEP-PH]].

[22] J. Beringer et al. (Particle Data Group), Phys. Rev. D86, 010001 (2012) and 2013 partial
update for the 2014 edition.

[23] C. Lunardini, E. Sabancilar and L. Yang, JCAP 1308, 014 (2013) [arXiv:1306.1808
[astro-ph.HE]].

[24] S. Hummer, P. Baerwald and W. Winter, Phys. Rev. Lett. 108, 231101 (2012)
[arXiv:1112.1076 [astro-ph.HE]].

[25] S. Gao, K. Kashiyama and P. Mszros, Astrophys. J. 772, L4 (2013) [arXiv:1305.6055


[astro-ph.HE]].

[26] L. Dorame, O. G. Miranda and J. W. F. Valle, arXiv:1303.4891 [hep-ph].

[27] W. Pauli, Cambridge Monogr. Part. Phys. Nucl. Phys. Cosmol., 14, 122, 2000.

[28] Neutrino Physics, Kai Zuber (2004 by Taylor and Francis Group, LLC)

[29] R. Barate et al. [LEP Working Group for Higgs boson searches and ALEPH and DELPHI
and L3 and OPAL Collaborations], Phys. Lett. B 565, 61 (2003) [hep-ex/0306033].

[30] T. Aaltonen et al. [CDF and D0 Collaborations], Phys. Rev. Lett. 104, 061802 (2010)
[arXiv:1001.4162 [hep-ex]].

113
[31] S. Chatrchyan et al. [CMS Collaboration], JHEP 06, 081 (2013) [arXiv:1303.4571
[hep-ex]].

[32] G. Aad et al. [ATLAS Collaboration], Phys. Lett. B 716, 1 (2012) [arXiv:1207.7214
[hep-ex]].

[33] S. Chatrchyan et al. [CMS Collaboration], Phys. Lett. B 716, 30 (2012) [arXiv:1207.7235
[hep-ex]].

[34] D. V. Forero, M. Tortola and J. W. F. Valle, Phys. Rev. D 86, 073012 (2012)
[arXiv:1205.4018 [hep-ph]].

[35] P. Minkowski, Phys. Lett., B67, 421, 1977.

[36] Kim C W and Pevsner A 1993 Massive Neutrinos in Physics and Astrophysics (Harwood
Academic)

[37] David Griffiths, 1987 Introduction to elementary particles (Jhon Wiley and Sons, Inc.)

[38] E. K. Akhmedov, hep-ph/0001264.

[39] C. Giunti and A. Studenikin, arXiv: 0812.3646v5 (2010)

[40] W. J. Marciano and A. I. Sanda, Phys. Lett. B67, 303, (1977)

[41] R. E. Shrock, Nucl. Phys. B206, 359 (1982)

[42] M. Fukugita and T. Yanagida, Physics of Neutrinos and Application to Astrophysics, M.


Fukugita and T. Yanagida (Springer-Verlag Berlin Heidelberg 2003).

[43] W. Grimus, M. Maltoni, T. Schwetz, M. A. Tortola and J. W. F. Valle, Nucl. Phys. B 648,
376 (2003) [hep-ph/0208132].

[44] R. N. Mohapatra and P. B. Bal, Massive Neutrinos in Physics and Astrophysics (World
Scientific, 1998) Second Edition, Lectures Notes in Physics, Vol 60.

[45] J.W.F. Valle, Gauge theories and the physics of neutrino mass, Prog. Part. Nucl. Phys. 26
(1991) 91, and references therein.

[46] J. Schechter and J. W. F. Valle, Phys. Rev. D24, 7 (1981)

[47] V. B. Semikoz, Nucl. Phys. B 498, 39 (1997) [hep-ph/9611383].

[48] F. Reines, H. S. Gurr, and H. W. Sobel, Phys. Rev. Lett. 37 6, (1976)

114
[49] P. Vogel and J. Engel, Phys. Rev. D 39, 11 (1989)

[50] G.S. Vidyakin, et. al., Pis’ma Zh. Eksp. Teor. Fiz. 55, No. 4, 212-215 (1992).

[51] R. C. Allen, et. al., Phys. Rev. D 47 1, 11-28 (1993)

[52] A. I. Derbin, et.al., Pis’ma Zh. Eksp. Teor. Fiz. 57, No. 12, 755-759 (1993)

[53] L. B. Auerbach, et.al., Phys. Rev. D 63 112001 (2001)

[54] Z. Daraktchieva, et.al., MUNU Collaboration, Phys. Lett. B 615 153-159 (2005)

[55] M. Deniz, et.al.,TEXONO Collaboration, arXiv:0911.1597v2 (2010)

[56] A. G. Beda, et. al., Advances in High Energy Physics. 2012350150 (2012)

[57] F.P. An, et. al., Daya Bay Collaboration, arXiv:1210.6327v2 (2012)

[58] Y. Abe, et. al., Double Chooz Collaboration, arXiv:1301.2948v1 (2013)

[59] J.K. Ahn, et. al., RENO Collaboration, arXiv:1204.0626v2 (2012)

[60] P. Blasi, arXiv:1311.7346 [astro-ph.HE].

[61] V. F. Hess, Phys. Z. 13 (1912) 1084

[62] P. Auger, R. Maze, T. Grivet-Meyer, Académic des Sciences 206 (1938) 228.

[63] K. Greisen, Phys.Rev.Lett. 16, 748 (1966); G. T. Zatsepin and V. A. Kuzmin, JETP Lett. 4
78 (1966).

[64] [HiRes Collaboration], R. Abbasi, Phys. Rev. Lett. 100 101101 (2008).

[65] http://www.cosmic-ray.org/

[66] J. Abraham [Pierre Auger Collaboration], Phys. Rev. Lett. 101, 061101 (2008)

[67] F. Salamida [Pierre Auger Collaboration], [arXiv: 1107.4809]

[68] M. A. Markov, Procs. Int. Conf. on High Energy Phys., p. 183, Univ. of Rochester (1960).

[69] F. Halzen and S. R. Klein, Rev. Sci. Istrum. 81 081101 (2010).

[70] M. Ageron et al. [ANTARES Coll.], NIM A 656 (2011) 11-38

[71] Klebesadel R. W., Strong I. B. and Olson R. A., 1973, Astrophys. J. Lett., 182, L85.

115
[72] D. Eichler, et. al., Nature, 340, 126 (1989)

[73] R. Narayan, B. Pacznski and T. Piran ApJ, 395, L83 (1992)

[74] C. Fryer and S. E. Woosley, ApJ, 502, L9 (1998)

[75] S. E. Woosley, ApJ, 405, 273 (1993)

[76] E. Waxman and J. Bahcall, Phys. Rev. Lett. 78, 12 (1997)

[77] E. Waxman and J. Bahcall, Ap. J. 541, 707 (2000)

[78] E. Waxman, arXiv:1312.0558 [astro-ph.HE].

[79] W. Rhode, et al., Astroparticle Physics 4 (1996) 217.

[80] M. Ambrosio, et al., Astroparticle Physics 19 (2003) 1.

[81] M. Achterberg, et al., Physical Review D 76 (2007) 042008

[82] A. V. Avrovin, et al., Astronomy Letters 35 (2009) 651.

[83] ANTARES Collaboration, Nuclear Instruments and Methods in Physics Research A 692
(2012) 41-45.

[84] R. Abbasi et al. [IceCube Collaboration], Nature 484, 351 (2012) [arXiv:1204.4219
[astro-ph.HE]].

[85] F. Halzen, arXiv:1311.6350 [hep-ph].

[86] J. Barranco, O.G. Miranda, C.A. Moura and A. Parada, Phys. Lett. B 718, 26 (2012).

[87] A. Cisneros, Astrophysics and Space Science 10 (1971) 87-92.

[88] E.K. Akhmedov, hep-ph/9705451

[89] H. Athar, J.T. Peltoniemi, A. Y. Smirnov, Phys. Rev D 51 (1995) 6647.

[90] P. Mehta and W. Winter, arXiv: 1101.2673v2 [hep-ph] (2011)

[91] A. M. Hillas, Ann. Rev. Astron. Astrophys. 1984. 22: 425-44

[92] J. Schechter and J. W. F. Valle, Phys. Rev. D 24, 1883 (1981) [Erratum-ibid. D 25, 283
(1982)].

116
[93] M. Tortola, J. W. F. Valle and D. Vanegas, arXiv:1205.4018 [hep-ph]; T. Schwetz,
M. Tortola and J. W. F. Valle, New J. Phys. 13, 063004 (2011) [arXiv:1103.0734 [hep-ph]].

[94] G. Domokos and S. Kovesi-Domokos, Phys. Lett. B 410, 57 (1997) [hep-ph/9703265].

[95] J. Alvarez-Muniz and P. Meszaros, Phys. Rev. D 70, 123001 (2004) [astro-ph/0409034].

[96] K. Mannheim, R. J. Protheroe and J. P. Rachen, Phys. Rev. D 63, 023003 (2000)
[astro-ph/9812398].

[97] F. Halzen and E. Zas, Astrophys. J. 488, 669 (1997) [astro-ph/9702193].

[98] F. W. Stecker and M. H. Salamon, Space Sci. Rev. 75, 341 (1996) [astro-ph/9501064].

[99] F. W. Stecker, C. Done, M. H. Salamon and P. Sommers, Phys. Rev. Lett. 66, 2697 (1991)
[Erratum-ibid. 69, 2738 (1992)].

[100] M. Ahlers, M. C. Gonzalez-Garcia and F. Halzen, Astropart. Phys. 35, 87 (2011)


[arXiv:1103.3421 [astro-ph.HE]].

[101] K. Murase, Phys. Rev. D 76, 123001 (2007) [arXiv:0707.1140 [astro-ph]].

[102] J. P. Rachen and P. Meszaros, AIP Conf. Proc. 428, 776 (1997) [astro-ph/9811266].

[103] R. Beck, Astrophys. Space Sci. Trans., 5, 43-47, 2009

[104] F. Tavecchio, G. Ghisellini, L. Foschini, G. Bonnoli, G. Ghirlanda and P. Coppi, Mon.


Not. Roy. Astron. Soc. 406, L70 (2010) [arXiv:1004.1329 [astro-ph.CO]].

[105] P. J. E. Peebles, Principles of Physical Cosmology, (1976)

[106] D. W. Hogg, arXiv:astro-ph/9905116v4 (2000)

[107] WMAP Collaboration, E. Komatsu, et. al., Astrophys. J. Suppl. 192, 18 (2011)

[108] S. Juneau, et. al., arXiv:1211.6436v1 [astro-ph.CO] (2012)

[109] T. Kaluza, Zum Unittsproblem in der Physik. Sitzungsber. Preuss. Akad. Wiss. Berlin.
(Math. Phys.) 96: 6972. (1921).

[110] O. Klein, Quantentheorie und fnfdimensionale Relativittstheorie . Zeitschrift fr Physik


A 37 (12): 895906. Bibcode 1926ZPhy...37..895K. doi:10.1007/BF0139748 (1926).

[111] G.C. McLaughlin, J.N. Ng, Phys. Lett. B 470 (1999) 157-162

117
[112] K. R. S. Balaji, A. S. Dighe and R. N. Mohapatra, hep-ph/0202267.

[113] G. Bertone, D. Hooper and J. Silk, Phys. Rept. 405, 279 (2005) [hep-ph/0404175].

[114] J. A. Tyson, G. P. Kochanski and I. P. Dell’Antonio, Astrophys. J. 498, L107 (1998)


[astro-ph/9801193].

[115] Particle Dark Matter, Giafranco Bertone (Cambridge University Press 2010).

[116] M. Archidiacono, S. Hannestad, A. Mirizzi, G. Raffelt and Y. Y. Y. Wong, JCAP 1310,


020 (2013) [arXiv:1307.0615 [astro-ph.CO]].

[117] S. Dodelson and L. M. Widrow, Phys. Rev. Lett. 72, 17 (1994) [hep-ph/9303287].

[118] M. Drewes, Int. J. Mod. Phys. E 22, 1330019 (2013) [arXiv:1303.6912 [hep-ph]].

[119] S. Schael et al. [ALEPH and DELPHI and L3 and OPAL and SLD and LEP
Electroweak Working Group and SLD Electroweak Group and SLD Heavy Flavour Group
Collaborations], Phys. Rept. 427, 257 (2006) [hep-ex/0509008].

[120] A. A. Penzias, and R. W. Wilson, 1965, ApJ, 142, 419

[121] P. A. R. Ade et al. [Planck Collaboration], arXiv:1303.5076 [astro-ph.CO].

[122] J. Lesgourgues, J. Mangano, G. Miele, and S. Pastor, 2013, Neutrino Cosmology


(Cambridge: Cambridge Univ. Press).

[123] A. Aguilar-Arevalo et al. [LSND Collaboration], Phys. Rev. D 64, 112007 (2001)
[hep-ex/0104049].

[124] A. A. Aguilar-Arevalo et al. [MiniBooNE Collaboration], Phys. Rev. Lett. 110, 161801
(2013) [arXiv:1207.4809 [hep-ex], arXiv:1303.2588 [hep-ex]].

[125] M Antonello, B Baibussinov, P Benetti, E Calligarich, N Canci, S Centro, A Cesana and


K Cieslik et al., Eur. Phys. J. C 73, 2345 (2013) [arXiv:1209.0122 [hep-ex]].

[126] A. Palazzo, Mod. Phys. Lett. A 28, 1330004 (2013) [arXiv:1302.1102 [hep-ph]].

[127] F. Kaether, W. Hampel, G. Heusser, J. Kiko and T. Kirsten, Phys. Lett. B 685, 47 (2010)
[arXiv:1001.2731 [hep-ex]].

[128] J. N. Abdurashitov et al. [SAGE Collaboration], Phys. Rev. C 80, 015807 (2009)
[arXiv:0901.2200 [nucl-ex]].

118
[129] C. Giunti and M. Laveder, Phys. Rev. C 83 (2011) 065504 [arXiv:1006.3244 [hep-ph]].

[130] G. Mention, M. Fechner, T. .Lasserre, T. .A. Mueller, D. Lhuillier, M. Cribier and


A. Letourneau, Phys. Rev. D 83, 073006 (2011) [arXiv:1101.2755 [hep-ex]].

[131] L. Wolfenstein, Phys. Rev. D 17, 2369 (1978). S. P. Mikheev and A. Y. .Smirnov, Nuovo
Cim. C 9, 17 (1986).

[132] O. G. Miranda, C. A. Moura and A. Parada, arXiv:1308.1408 [hep-ph].

[133] Z. G. Berezhiani and R. N. Mohapatra, Phys. Rev. D 52, 6607 (1995) [hep-ph/9505385].

[134] Z. G. Berezhiani, A. D. Dolgov and R. N. Mohapatra, Phys. Lett. B 375, 26 (1996)


[hep-ph/9511221].

[135] Y. Zhang, X. Ji and R. N. Mohapatra, arXiv:1307.6178 [hep-ph].

[136] J. Bramante, Int. J. Mod. Phys. A 28, 1350067 (2013) [arXiv:1110.4871 [hep-ph]].

[137] G. Mangano, A. Melchiorri, P. Serra, A. Cooray and M. Kamionkowski, Phys. Rev. D 74,
043517 (2006) [astro-ph/0606190].

[138] P. J. Fox and E. Poppitz, Phys. Rev. D 79, 083528 (2009) [arXiv:0811.0399 [hep-ph]].

[139] R. Laha, B. Dasgupta and J. F Beacom, arXiv:1304.3460 [hep-ph].

[140] R. Foot and R. R. Volkas, Phys. Rev. Lett. 75, 4350 (1995) [hep-ph/9508275].

[141] R. Foot and R. R. Volkas, Phys. Rev. D 52, 6595 (1995) [hep-ph/9505359].

[142] R. N. Mohapatra and S. Nasri, Phys. Rev. D 71, 053001 (2005) [hep-ph/0407194].

[143] Z. Berezhiani, Int. J. Mod. Phys. A 19, 3775 (2004) [hep-ph/0312335].

[144] R. Foot, Phys. Rev. D 86, 023524 (2012) [arXiv:1203.2387 [hep-ph]].

[145] C. Boehm, T. A. Ensslin and J. Silk, J. Phys. G 30, 279 (2004) [astro-ph/0208458].

[146] C. Boehm, D. Hooper, J. Silk, M. Casse and J. Paul, Phys. Rev. Lett. 92, 101301 (2004)
[astro-ph/0309686].

[147] D. Hooper, F. Ferrer, C. Boehm, J. Silk, J. Paul, N. W. Evans and M. Casse, Phys. Rev.
Lett. 93, 161302 (2004) [astro-ph/0311150].

[148] I. M. Shoemaker, arXiv:1305.1936 [hep-ph].

119
[149] P. Mitropoulos, arXiv:1307.2823 [hep-ph].

[150] S. Kanemura, T. Matsui and H. Sugiyama, arXiv:1305.4521 [hep-ph].

[151] W. Chao and M. J. Ramsey-Musolf, arXiv:1212.5709 [hep-ph].

[152] K. Kainulainen, Phys. Lett. B 244, 191 (1990).

[153] K. Kainulainen, J. Maalampi and J. T. Peltoniemi, Nucl. Phys. B 358, 435 (1991).

[154] K. Enqvist, K. Kainulainen and J. Maalampi, Nucl. Phys. B 349, 754 (1991).

[155] R. Barbieri and A. Dolgov, Nucl. Phys. B 349, 743 (1991).

[156] K. Enqvist, K. Kainulainen and M. J. Thomson, Nucl. Phys. B 373, 498 (1992).

[157] A. D. Dolgov, S. H. Hansen, G. Raffelt and D. V. Semikoz, Nucl. Phys. B 590, 562 (2000)
[hep-ph/0008138].

[158] M. -R. Wu, T. Fischer, G. Martnez-Pinedo and Y. -Z. Qian, arXiv:1305.2382


[astro-ph.HE].

[159] P. Chen, arXiv:1302.5319 [astro-ph.HE].

[160] G. Sigl, arXiv:1202.0466 [astro-ph.HE].

[161] P. Meszaros, K. Asano and P. Veres, arXiv:1209.2436 [astro-ph.HE].

[162] P. Abreu et al. [The Pierre Auger Collaboration], arXiv:1107.4805 [astro-ph.HE].

[163] S. Biagi, Nucl. Phys. Proc. Suppl. 212-213, 109 (2011) [arXiv:1101.3670 [astro-ph.HE]].

[164] M. G. Aartsen et al. [IceCube Collaboration], arXiv:1304.5356 [astro-ph.HE].

[165] K. Murase and S. Nagataki, Phys. Rev. D 73, 063002 (2006); H. -N. He, R. -Y. Liu,
X. -Y. Wang, S. Nagataki, K. Murase and Z. -G. Dai, Astrophys. J. 752, 29 (2012);

[166] J. Barranco, O. G. Miranda, C. A. Moura, T. I. Rashba and F. Rossi-Torres, JCAP 1110,


007 (2011) [arXiv:1012.2476 [astro-ph.CO]].

[167] J. Barranco, O. Miranda, C. Moura and A. Parada, Phys.Lett. B718, 26 (2012).

[168] A. Esmaili and Y. Farzan, JCAP 1212, 014 (2012).

[169] A. S. Joshipura, S. Mohanty and S. Pakvasa, arXiv:1307.5712 [hep-ph].

120
[170] S. Pakvasa, A. Joshipura and S. Mohanty, 1209.5630.

[171] A. Esmaili, Phys. Rev. D 81, 013006 (2010) [arXiv:0909.5410 [hep-ph]].

[172] J. F. Beacom, N. F. Bell, D. Hooper, J. G. Learned, S. Pakvasa and T. J. Weiler, Phys. Rev.
Lett. 92, 011101 (2004) [hep-ph/0307151].

[173] P. Keranen, J. Maalampi, M. Myyrylainen and J. Riittinen, Phys. Lett. B 574, 162 (2003)
[hep-ph/0307041].

[174] R. M. Crocker, F. Melia and R. R. Volkas, Astrophys. J. Suppl. 141, 147 (2002)
[astro-ph/0106090].

[175] M. A. de Avillez, A. Asgekar, D. Breitschwerdt and E. Spitoni, arXiv:1204.1511


[astro-ph.GA].

[176] J. F. Navarro, C. S. Frenk and S. D. M. White, Astrophys. J. 490, 493 (1997)


[astro-ph/9611107].

[177] J. F. Navarro, C. S. Frenk and S. D. M. White, Astrophys. J. 462, 563 (1996)


[astro-ph/9508025].

[178] N. Whitehorn, C. Kopper, and N. K. Neilson, talk given at the IceCube Particle
Astrophysics Symposium (IPA) 2013, Madison, WI, USA.

[179] J. Diemand and B. Moore, Adv. Sci. Lett. 4, 297 (2011) [arXiv:0906.4340 [astro-ph.CO]].

[180] R. A. Chevalier and Z. Y. Li, Astrophys. J. 520, L29 (1999) [astro-ph/9904417].

[181] X. Hernandez and W. H. Lee, arXiv:1002.0553 [astro-ph.CO].

[182] D. Hooper, Phys. Dark Univ. 1, 1 (2012) [arXiv:1201.1303 [astro-ph.CO]].

121

You might also like