You are on page 1of 49

Neutrino Oscillations

Rohit R.
S. R. 11-01-00-10-91-11-1-09106

A thesis presented for the degree of Bachelor of Science


(Research)

Undergraduate Programme
Indian Institute of Science
Bengaluru - 560012
India
20 April, 2015

Abstract
Neutrino Oscillations
Rohit R.
Although the field of neutrino oscillations has been experimentally active for many
years now, the sterile neutrino has eluded detection. We attempt to come up with a
proposal to use matter effects on neutrino oscillations to enhance sterile neutrino
oscillations and hence detect the sterile neutrino.

Acknowledgements
First and foremost, I would like to thank my mentors Prof. Sudhir Vempati and Prof. Uma
Shankar for their invaluable help and guidance. I would also like to thank all my friends for their
support and scintillating discussions. This endeavour would definitely have not been possible
without my parents support and faith in me. I would also like to thank IISc for giving me the
opportunity to work in this field. A special thanks to the Kishore Vaigyank Protsahana Yojana
for financial support.

Contents
1 Introduction

2 Neutrino Oscillations
2.1 2-flavour oscillations . . . . . . . . . . . . . . . . . . . . . . . .
2.2 n-flavour Oscillations . . . . . . . . . . . . . . . . . . . . . . . .
2.2.1 Decomposition of the n-neutrino problem . . . . . . . .
2.3 Oscillations in the presence of matter . . . . . . . . . . . . . . .
2.3.1 Uniform Matter Density . . . . . . . . . . . . . . . . . .
2.3.2 Varying matter density: The Adiabatic approximation .
2.3.3 Non-adiabatic effects and the Adiabaticity Parameter .
2.3.4 The big picture: Looking at the Solar Neutrino problem

.
.
.
.
.
.
.
.

6
6
9
10
13
13
16
17
19

3 Experiments in Neutrino Physics: Measuring the Oscillation Parameters


3.1 Detection of Neutrinos . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2 Solar neutrinos . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2.1 The Solar Neutrino Flux . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2.2 Measuring the Solar Neutrino Flux . . . . . . . . . . . . . . . . . . . . .
3.2.3 The Solution of the Solar Neutrino Problem . . . . . . . . . . . . . . . .
3.3 Experiments involving Atmospheric Neutrinos . . . . . . . . . . . . . . . . . . .
3.3.1 Cosmic Rays . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3.2 Production of Atmospheric Neutrinos . . . . . . . . . . . . . . . . . . .
3.3.3 Characteristics of the atmospheric neutrino flux[10] . . . . . . . . . . . .
3.3.4 Zenith angle dependence of the atmospheric neutrino flux . . . . . . . .
3.3.5 Measuring the Atmospheric Neutrino Flux . . . . . . . . . . . . . . . . .
3.4 Accelerator experiments to measure 13 . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.

21
21
22
22
23
27
28
28
29
29
30
30
33

4 Sterile Neutrino Oscillations


4.1 Derivation of matter term for s oscillations .
4.1.1 Certain Issues . . . . . . . . . . . . . . . . .
4.2 Propagation in the Earths crust . . . . . . . . . .
4.2.1 Baseline at Resonance . . . . . . . . . . . .
4.3 Outlook and Summary . . . . . . . . . . . . . . . .

.
.
.
.
.

36
36
37
37
38
39

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

A The vacuum Hamiltonian in a more illuminating form

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.

41

CONTENTS

B Some Mathematica demos


B.1 3-flavour Oscillations in Vacuum . . . . . .
B.2 Solar Neutrinos . . . . . . . . . . . . . . . .
B.2.1 Survival Probability of e . . . . . .
B.2.2 Density profile of the Sun and sin2
B.2.3 The Adiabaticity parameter . . . . .
B.2.4 The Exclusion Plot . . . . . . . . . .

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

42
42
44
44
45
45
45

Chapter 1

Introduction
Neutrinos are very light leptons with masses of the order of meVs. Since they are neutral
particles, they do not take part in electromagnetic interactions. They come in three flavours:
e, , and being produced/annihilated alongside the complementary lepton in weak processes.
Neutrinos produced alongside electrons are of e flavour; and similarly for muons and taons.
Since neutrinos are affected only by the weak and gravitational interactions (which are extremely
weak), they have very high mean free paths even in dense matter and hence may be used to
probe and understand the internal structure of otherwise inaccessible bodies like stars and other
astrophysical structures.
Whether or not neutrinos are massless was an open question until the recent discovery of
the phenomenon of neutrino oscillations, which conclusively establish the existence of massive
neutrinos. This phenomenon is the main topic of discussion of this thesis. This article is
organised as follows. In Chapter II, we attempt to review the theoretical framework behind
the phenomenon of neutrino oscillations. In Chapter III, we shall look at the experimental
advances in the field and the various estimations of the parameters associated with oscillations.
In Chapter IV, we shall take a brief look at a simple model incorporating sterile neutrinos and
consider a stand alone accelerator-based experiment to look for sterile neutrinos.

Chapter 2

Neutrino Oscillations
Neutrino Oscillations refers to the phenomenon whereby the flavour of a neutrino - earlier
believed to be a constant of motion, varies with time without any external influence. The
chief reasons for this phenomenon are the non-trivial relation between the flavour and mass
neutrino eigenstates (meaning that states with a definite mass do not have a definite flavour
and vice versa), and the non-zero mass difference between neutrino mass eigenstates. Hence,
this phenomenon could be used to calculate the mass differences of neutrino states and also to
establish the existence of the massive neutrino.
We would like to note that the derivation outlined here is a highly simplified version which
does not take into account various conceptual issues[1, 2, 3, 4, 5, 6, 7, 8]. However, approaches
that do rectify these issues (field-theoretical, wave-packet based) are more elaborate and give
the same results.

2.1

2-flavour oscillations

We shall first look at the simplified case of oscillations between two flavours because, as shall
be seen later, the general case may be decoupled into many copies of these 2-flavour oscillations.
Now, the left-handed components of the flavour states are unitary linear combinations of the
left-handed components of the mass eigenstates. Denoting the two flavour states by ,
(flavour states shall stand for left-handed states from here on) and the mass states by 1, 2,




(t)
1 (t)
=U
(2.1)
(t)
2 (t)
where U is the unitary mixing matrix. Now the mass (Hamiltonian) eigenstates evolve under
Schrodinger dynamics as follows:

 
 

d
1 (t)
E1 0
1 (t)

=
.
(2.2)
0 E2
2 (t)
dt 2 (t)

i.e.,

1 (t)
2 (t)


=

Call


(t) :=

eE1 t 0
0
eE2 t
eE1 t 0
0
eE2 t



1 (0)
2 (0)


(2.3)


(2.4)

CHAPTER 2. NEUTRINO OSCILLATIONS

Now substituting into equation (1),






(0)
(t)

= (t)U
U
(0)
(t)

(2.5)

Left multiply by U and use unitarity of U to obtain the relation between flavour states at the
point of production (t = 0) and point of detection (at t).




(0)
(t)

(2.6)
= U (t)U
(0)
(t)
Strictly speaking, the indices and may only stand for the active flavour states since
neutrinos are produced and detected via weak processes. Sterile neutrinos, if they exist, must
be detected indirectly via the disappearance of active flavour neutrinos.
It is clear that the amplitudes for oscillation are given by the off-diagonal elements of
U (t)U . Hence there will be no oscillations if this matrix is diagonal. One way this could
happen is if U is the identity matrix (trivial mixing).
Now, the most general unitary matrix in two dimensions is of the form,


cos
e sin
(2.7)
U=
e sin
cos
But since Dirac fermions can absorb phases, some of the phase factors are physically
irrelevant and the mixing matrix maybe subject to the transformation
Ui ei Ui e

(2.8)

without changing the physics. Especially in the 2-dimensional case, it is easy to see that the only
phase factor present maybe eliminated in this manner and the general mixing matrix becomes:


cos sin
(2.9)
U=
sin cos

eE1 t 0
0
eE2 t



cos sin
sin cos

eE1 t cos2 + eE2 t sin2 (eE1 t eE2 t )sincos


(eE1 t eE2 t )sincos
eE1 t sin2 + eE2 t cos2

U U =
=



cos sin
sin
cos

Hence,
(E1 E2 )t
2
Applying the ultra relativistic approximation, with p1 p2 , v c and E1 E2
P( ) = sin2 2 sin2

(2.10)

m2j
Ej pj +
2Ej
2
m m22
m2
E := E1 E2 1
=
2E
2E
And the final result becomes(in natural units):
P( ) = sin2 2 sin2

m2 L
L
= sin2 2 sin2
4E
Losc

(2.11)

CHAPTER 2. NEUTRINO OSCILLATIONS

4E
where L := ct is the distance travelled and Losc = m
2 is defined as the oscillation length. This
is the celebrated neutrino oscillation formula.
Note that m2 = 0 P( ) = 0 and hence the mass differences being non-zero is
the second reason for this effect. This dependence of m2 on an experimentally measurable
quantity (P( )) allows the measurement of m2 and this establishes the existence of
the massive neutrino.
In practical cases, the beam of neutrinos tend to have a range of energies as opposed to the
mono energetic beam we have considered. In such cases, the probability is averaged over the
distribution (say (E)) and this average (denoted hPi) is the observed quantity.
Z
P(E)(E)dE
(2.12)
hPi =
0

Figure 2.1 compares the oscillation probabilities for a mono energetic beam and a beam
with energies normally distributed about some value. Note that for very high energies, the
probability approaches 0.5 as the sine squared dependence on energy averages out to give half.
Hence, it is to be noted that oscillations may be observed only at certain length scales

Figure 2.1: Observed oscillation probabilities in theoretical and realistic scenarios plotted against 1 L/Losc . The lighter curve is the monoenergetic case
and the bold line shows the averaged out version over a Gaussian energy
spectrum of mean E and Std. deviation E/10 (Fig taken from [10]).
(depending upon the mass difference and energy). If oscillation length is given by Losc , it is
clear from the figure that if L  Losc , oscillation probability is negligible; while if L  Losc , it
is a constant. Hence, each experiment on neutrinos is said to be sensitive only upto a certain
m2 (defined as that for which L/Losc ' 1). Table 2.1 lists several classes of experiments and
the m2 they are sensitive to.
Experiments studying neutrino oscillation provide bounds on transition probabilities in particular channels. Now these may be used to mark out regions in parameter space that are
allowed by the measured bounds. To understand this, consider an effective two flavour oscillation channel with a mixing of driven by mass difference m2 . Say the upper and lower
bounds on the transition probability are given by A and B respectively.
B sin2 2. sin2
8

m2 L
A
4E

(2.13)

CHAPTER 2. NEUTRINO OSCILLATIONS

L
m

E
MeV

m2
eV2

Reactor SBL

102

102

Reactor LBL

103

103

Accelerator SBL

103

103

Accelerator LBL

106

103

103

Atmospheric

107

103

104

Solar

1011

1011

Experiment

Table 2.1: The table lists various classes of experiments and their m2 sensitivities (taken from
[10]).
Now for a given baseline and energy, this relation marks out regions in the sin2 2 m2
plane. (Note that any one-one function of onto the required range may be used instead of
sin2 2. Indeed, various functions such as tan2 and sin2 2/cos2 are used.) Such plots are
known as exclusion plots and are used to compute the values of the relevant parameters. Figure
2.2 shows exclusion plots from various accelerator experiments.

Figure 2.2: Figure shows exclusion plots (90% CL) from various accelerator
experiments (taken from [10]).

2.2

n-flavour Oscillations

The derivation of the oscillation formula in the case of n massive neutrinos runs on similar
lines, and the amplitudes are given by the matrix:
A = U (t)U
9

(2.14)

CHAPTER 2. NEUTRINO OSCILLATIONS

A( ) =

n
X

Ek t
Uk
e
Uk

(2.15)

k=1

Using unitarity

Pn

k=1 Uk Uk

= and simplifying using the ultra-relativistic approximation,


 2


n

2 L
X
m

k1
1
P( ) = +
exp
Uk Uk


2E

(2.16)

k=2

Now for anti-neutrinos, the mixing matrix is simply the complex conjugate of the matrix U
and hence it is easily seen that
(2.17)
P = P
This is a consequence of CPT invariance that is a feature of any local, unitary quantum field
theory. It is also clear from the above equation that the survival probabilities for neutrinos and
anti-neutrinos of the same flavour are identical. But the transition probabilities are in general
different unless the mixing matrix happens to be real. In this case, there is CP invariance
and hence the only physical phase in the PMNS matrix, which is the mixing matrix associated
with the three active flavour neutrinos in the Standard Model (see Section 2.2.1), is known as
the CP-violating phase. It is to be noted that for the same reason, CP violation cannot be
determined through 2-flavour oscillations (since the mixing matrix can be made real as shown
in Eqn. 2.9).

2.2.1

Decomposition of the n-neutrino problem

To show how n-flavour oscillations can reduce to effective two flavour oscillations, consider
the case where where one state is much more massive than others, i.e., mn >> mk k =
1, 2, . . . , n 1. Define M 2 := m2n1 , and m2 := m2k1 . Now choose an appropriate length
scale such that:


M 2 L
&1
(2.18)
2E
|m2 |L
which naturally means that 2E  1. Using this in the n-neutrino oscillation formula means
that all phases with m2 are small and hence their exponentials are close to one. This allows
us to neglect all the terms in the summation except the nth term and hence the oscillation
formula becomes:
P

(n)





 2


M 2 L


( ) + Un Un exp
1
2E

(2.19)

where the superscript n denotes that the P is calculated keeping in mind that is is an nneutrino system. Expanding the absolute value,



M 2 L
2
2
(n)
P
( ) 2 |Un | |Un |
1 cos
(2.20)
2E
Now let us assume that we start out with a neutrino beam of flavour . The survival and
transition probabilities of are respectively given by:



M 2 L
2
2
(n)
(2.21)
P
( ) 1 2 |Un | 1 |Un |
1 cos
2E
10

CHAPTER 2. NEUTRINO OSCILLATIONS

(n)

( ) 2 |Un | |Un |

M 2 L
1 cos
2E


(2.22)

Observe that this is starting to look like a 2-flavour system. Motivated by this, define:
sin2 2eff := 4 |Un |2 (1 |Un |2 )
0 :=

Un
p

1 |Un |2
6=

(2.23)
(2.24)

It is easy to see that sin2 2eff is always less that unity and that the new flavour 0 is normalised.
Now the survival and transition probabilities look like:


M 2 L
1 2
(n)
(2.25)
P
( ) 1 sin 2eff 1 cos
2
2E
P (n) ( 0 ) = |h 0 | i|2

2

P

Un


= 6= p
h | i
2


1 |Un |


 2

P
2L
U
M


n
2E
U
6= p
Un

n 1 e
2


1

|U
|
n


2
L
= 12 sin2 2eff 1 cos M
2E

(2.26)

where unitarity of the mixing matrix U was used to obtain the last line.
Hence, under the one-dominant m2 condition and at the right length scales, we have shown
that the n-neutrino problem reduces to a 2-flavour oscillation between 0 with effective
mixing angle eff . Note that may be any flavour (chosen according to the physical situation)
and 0 is then accordingly defined.
Now according to present data[9],
|m221 |
= 7.5 105 eV 2

(2.27)

|m231 |
= 2.5 103 eV 2

(2.28)

|m221 |
= 0.03
|m231 |

(2.29)

Hence, the physical 3-neutrino system, at small length scales, decomposes into an effective two
neutrino system driven by m231 . This is seen in the oscillation of atmospheric neutrinos. To see
this a little more clearly, consider the parametrisation of the 3-neutrino mixing matrix known
as the Pontercovo-Maki-Nagakawa-Sakata (or PMNS) matrix.

Ue1 Ue2 Ue3


U1 U2 U3
(2.30)
U 1 U 2 U 3

1
0
0
cos13
0
= 0 cos23 sin23
0 sin23 cos23
sin13 eCP

0 sin13 eCP
cos12 sin12 0
sin12 cos12 0
1
0
0
0
1
0
cos13
(2.31)
11

CHAPTER 2. NEUTRINO OSCILLATIONS

s13 eCP
s23 c13
c23 c13
(2.32)
where the s are the various mixing angles, CP the only physical (and hence CP-violating as
discussed earlier) phase. (The obvious redactions have been made for sines and cosines in the
last line.) Now, 0 and the effective mixing angle may be calculated taking e as the starting
flavour:
sin2 2eff = 4 |Ue3 |2 (1 |Ue3 |2 )

2
(2.33)
= 4 sin13 eCP (1 |sin13 eCP |2 )
= sin2 213

c12 c13
= s12 c23 c12 s23 s13 eCP
s12 s23 c12 c23 s13 eCP

0 =

6=e

s12 c13
c12 c23 s12 s23 s13 eCP
c12 s23 s12 c23 s13 eCP

U3

1|Ue3 |2

sin cos13
cos23 cos13
p 23
+ p

2
1 sin 13
1 sin2 13
= sin23 + cos23

(2.34)

Hence, short-range atmospheric neutrino oscillations are e 0 oscillations driven by


m213 with effective mixing angle 13 at length scales: (using a typical energy scale 1 GeV for
atmospheric neutrinos) L & 200 km.
Solar neutrino oscillations, on the other hand, are a manifestation of the energy averaging
discussed earlier. To see this, consider the survival probability of some flavour as computed
from Eqn. 2.15 with the one-dominant m2 approximation as before, but now at length scales
such that


m2 L
'1
(2.35)
2E
|M 2 |L
which means that 2E  1. So,

2




2L
M 2 L

(n)
2
2
2 m
2

4E
4E
P
( ) = |U1 | + (|U2 | + + |U(n1) | )e
+ |Un | e


|
{z
}


X


2 L 2

2
2
m

4E + 2|U
= |U1 | + X e
n |

M 2 m2
M 2 L
|U1 |2 cos
+ X cos
2E
2E

(2.36)

!
+ |Un |4

(2.37)
Now, cos
is a highly oscillating function of energy (in the limit chosen) and cancels out
to zero when averaged over an energy distribution. Hence the middle term completely cancels
out (remember |M 2 |  |m2 |), and we expand the rest of the terms to get:




1 2
m2 L
(n)
2 2
+ |Un |4
(2.38)
P
( ) = 1 |Un |
1 sin 2eff 1 cos
2
2E
M 2 L
2E

with
sin2 eff :=

X
1 |Un |2

12

(2.39)

CHAPTER 2. NEUTRINO OSCILLATIONS

It is easy to see that this reduces to a 2-flavour oscillation formula in the limit |Un |2 0.
Coming back to 3-flavour mixing ( = e ), we observe that:




1 2
m212 L
(3)
2 2
P
(e e ) = 1 |Ue3 |
1 sin 2eff 1 cos
+ |Ue3 |4
(2.40)
2
2E
with
sin2 eff =

|Ue2 |2
= sin2 12
1 |Ue3 |2

(2.41)

And with[9]
|Un |2 = |Ue3 |2 = sin2 13 0.01

(2.42)

this reduces to the 2-flavour solar neutrino oscillations.


Hence, the 3-flavour problem maybe decomposed into two 2-flavour oscillations (atmospheric
and solar) occurring at two different length scales and hence the parameters of the PMNS matrix
associated with these oscillations have been suitably dubbed:
m2 := m221

(2.43)

:= 12

(2.44)

m2atm := m231

(2.45)

atm := 23

(2.46)

Note that the atmospheric mixing angle is 23 instead of 13 . This is because the atmospheric
neutrino oscillations are dominantly (recall we started with e ), and in the limit
cos2 13 0, sin2 2eff = sin2 223 for these oscillations.

2.3

Oscillations in the presence of matter

Neutrino oscillations were shown to be modified significantly due to their interaction with
matter. The dominant process that contributes towards this effect, as shown by Wolfenstein[12]
and by Mikheyev and Smirnov[13] is coherent forward scattering. In this section, we shall study
neutrino oscillations in matter by considering the case of the solar neutrino problem1 .

2.3.1

Uniform Matter Density

We shall first consider the rather simple case of oscillations in uniform matter. The case of
interest shall involve two flavours - e and (taken here to be electron and muon neutrinos but
may be generalised as described later). The dominant process that contributes to this effect is
coherent forward scattering as shown by the Feynmann diagrams in Fig. 2.3.
Note that in the regimes of interest (namely the Sun and the Earth), matter is mostly
composed of protons, electrons and neutrons. Even in the Suns core, temperatures are not
high enough to generate heavier leptons via processes such as pair production. Hence, processes
involving these leptons are not considered. The low energy charged current (CC) Lagrangian is
given by:
4GF
LCC = [
eL (p1 ) eL (p2 )][
eL (p3 ) eL (p4 )]
(2.47)
2
1

The outline of this section shall closely follow [11]

13

CHAPTER 2. NEUTRINO OSCILLATIONS


e (p2 )

x (p4 )

e(p1 )

W+

x (p3 )

e(p4 )

f(p2 )

e (p3 )

f(p1 )

Figure 2.3: These Feynmann diagrams show the dominant (coherent forward
scattering) processes in the interaction between neutrinos and matter. f
stands for protons, neutrons and electrons, and x stands for any active
flavour neutrino.
Now, coherent forward scattering implies that p1 = p4 = pe and p2 = p3 =p. Using this, a
Fierz rearrangement gives:
4GF
LCC = [
eL (p) eL (p)][
eL (pe ) eL (pe )]
2

(2.48)

Averaging over the momentum of electrons , and applying the non-relativistic reduction formulae
h
e 5 ei hspini = 0

(2.49)

h
e~ ei hvelocityi = 0

(2.50)

h
e0 ei = Ne

(2.51)

Leff
eL 0 eL
CC = 2GF Ne

(2.52)

gives the effective CC Lagrangian

Repeating the same procedure for the neutral current (NC) processes:






4GF
1 5
2

LN C =
f (p1 ) I3L
Qsin W f (p2 ) [L (p3 ) L (p4 )]
2
2

(2.53)

Averaging over the background fields f, the contributions from the electrons and protons cancel
out because they have opposite values of I3L and charge.
1
Leff
eL 0 eL + L 0 L )
N C = GF Nn (
2

(2.54)

Both the CC and NC interaction terms add to the Hamiltonian in the form a potential
energy and the effective Hamiltonian is hence given by:
"
#

2GF Ne 12 Nn
0
Heff = Hvac +
(2.55)
0
12 GF Nn
where
Hvac = E +

m21 + m22 M 2
+
4E
2E
14

(2.56)

CHAPTER 2. NEUTRINO OSCILLATIONS

with

M =
2
2

cos2 sin2
sin2 cos2


(2.57)

is the vacuum Hamiltonian in the flavour basis (see Appendix A). Casting this into a more
suggestive form,


A
m21 + m22
1
1
(cos2 A)
sin2
(2.58)
Heff = E +
G F nn +
+
sin2
cos2 A
4E
4E 4E
2
with

A = 2 2GF ne E

gives the modified eigenvalues (the tilde indicating the respective values in matter).
q
2
2
= E + m1 + m2 1 GF nn + A 1
(cos2 A)2 + 2 sin2 2
E
4E
4E 4E
2

(2.59)

(2.60)

and the modified mixing angle


sin2 2 =

2 sin2 2
2 sin 2 + (cos2 A)2
2

(2.61)

It is easily seen that nn = 0 and ne = 0 give back the vacuum energy eigenvalues and
mixing angles. Note that the
sine-squared of the mixing angle is of the form of a resonance
curve with peak at A = 2 2GF ne E = cos2 and width = sin2. Hence, mixing is
strongly suppressed for high matter densities and tends to the vacuum angle for low densities,
with a sharp peak at the resonance density :
cos2
nR :=
2 2GF E

(2.62)

Note that the oscillation length is also enhanced during resonance and is given by:
L(R)
osc =

4E
m2 sin2

(2.63)

This shows that even oscillations with a small mixing angle may be dramatically enhanced in
the presence of matter. This is an idea that will be exploited later on.
Note that although we have limited ourselves to electron and muon neutrinos, e may stand
for any flavour with a CC interaction with the medium in question and may stand for any
flavour with only NC interactions with the medium. Indeed for the solar neutrino problem, we
shall assign e to electron neutrinos and to a linear combination of muon- and taon-neutrinos
(Since neither flavour has CC interactions with the solar medium, nor does their combination).
It is also clear that if both flavours merely interact with matter via NC interactions, the matter
term merely adds a phase to both mass eigenstates and the oscillations proceed practically as
in vacuum.
For antiparticles, the same analysis follows through except for the fact all number densities
reverse their sign (since the expectation of the number operator for fermions must now be
replaced with that of anti-fermions which have the same value but opposite sign). Hence,
depending on the sign of m2 , matter can amplify only one of or oscillations
but not both. This means that even for 2-flavour oscillations, P( ) 6= P(
).
Matter oscillations as in the case we have considered are neither CP nor CPT invariant because
the medium is composed of matter and not anti-matter.
15

CHAPTER 2. NEUTRINO OSCILLATIONS

2.3.2

Varying matter density: The Adiabatic approximation

In the event that matter density varies along the path of propagation, we go back to the
Schrodinger equation in natural units (Eqn. 2.2).

2
d (f ) M
=
(f )
dx
2E

(upto scalar matrices, which only contribute an overall phase) with




1 cos2 + 2A sin2
2

M =
sin2
cos2
2

(2.64)

(2.65)

(p)
and f denoting the flavour basis (let p denote the mass basis). Now using (f ) = U

dU
d (p) = 1 M 2 U
(p)
(p) + U
dx
dx
2E


1 2
dU
d (p)

U M U U
(p)
=
dx
2E
dx
 !




1
m1 2
sin cos d
0
cos sin

(p)
cos sin dx
0
m2 2
sin cos
2E
"
#
m
21
d
d (p)

2E
dx
(p)
=
m
22
d
dx

dx

(2.66)
(2.67)
(2.68)

(2.69)

2E

where Eqn. 2.68 is correct only upto the addition of scalar matrices.
Now, we assume that propagation is slow enough that state, at each point of propagation
collapses to a Hamiltonian (at that point) eigenstate. Hence, states propagate continuously as
Hamiltonian eigenstates. This is known as the adiabatic approximation.
Calculating the survival probability of e in this limit,
P(ad)
(x) = |he (x)|e (0)|2
e e
inserting complete mass bases , ,

2


X



= he (x)| (x)ih (x)| (0)ih (0)|e (0)
,

2



R
X


(x0 )
0x dx0 E

h (0)|e (0)
= he (x)| (x)i e

,
R 0
2
R
0



= e dx E1 cos0 cos + e dx E2 sin0 sin

Z x

1
0

=
1 + cos20 cos2 + sin20 sin2cos
dx (E2 E1 )
2
0

(2.70)

(2.71)

(2.72)

(2.73)
(2.74)

The subscript 0 denotes that the parameter value is calculated at the point of origin of the
neutrino. It is to be noted that the matrix element
h | i = e
16

Rx
0

(x0 )
dx0 E

(2.75)

CHAPTER 2. NEUTRINO OSCILLATIONS

because we have made the adiabatic approximation. The sine term vanishes on averaging over
beam energy[11] provided m2  1010 eV2 and we get
P(ad)
(x) =
e e


1
1 + cos20 cos2
2

(2.76)

We close this section by observing that adiabatic effects alone can cause a dramatic reduction
in the survival probability[11]. Consider an electron neutrino produced deep in the Sun where
' 2 and hence e ' 2 . This beam propagates adiabatically and reaches the surface of the
Sun where = and 2 = e cos + sin. The survival in this case is given by:
P(ad)
= sin2
e e

(2.77)

This is a drastic reduction if happens to be small.

2.3.3

Non-adiabatic effects and the Adiabaticity Parameter

Now, we try to quantify the notion of adiabatcity so as the identify the regimes where
this approximation may be applied. Examining equation (2.69), the adiabaticity condition is
equivalent to saying that the off-diagonal terms of the Hamiltonian are much smaller than the
diagonal terms (non-uniformity of medium comes into play via the off-diagonal terms). But
since any scalar matrix may be freely subtracted from the Hamiltonian without changing the
physics, it only makes sense to talk of the difference of the diagonal terms rather than their
magnitudes. Hence, the adiabaticity condition translates to the inequality:


2
d
m
1 m
22

.
(2.78)

dx
2E
Substituting for from Eqn. 2.61,



d
dne
sin2



.
= 2GF E.
dx
(cos2 A)2 + 2 sin2 2 dx

(2.79)

And hence the adiabaticity condition translates to



3


dne
A)2 + 2 sin2 2 2
 (cos2

dx
2 2GF E 2 sin2

(2.80)

(x)  1

(2.81)

or
with

(/E)2 sin2 2 1
(x) :=
.
. e
(2.82)

2 2GF sin3 2 dn
dx
dn
e

(x) depends on position via the density gradient dx . Note that if matter density is very
high so that /2, then sin2 0 and (x) . On the other end, if matter density
approaches 0 so that , (x) is once again large (assuming a small mixing angle). In Fig.
2.4, (x) is plotted against x for the solar problem to demonstrate that non-adiabatic effects
are not important in the Sun except near resonance points.
To study the behaviour of neutrinos at the resonance regions, define X to be the probability
17

CHAPTER 2. NEUTRINO OSCILLATIONS

Figure 2.4: The figure shows a plot of (x) against distance from the solar
core at various values of /E. The solid lines represent (x) and the dashed
The vacuum mixing angle is taken as 0.1.
lines sin2 2.
of discrete, non-adiabatic jumps between Hamiltonian eigenstates when the neutrino beam
passes through a resonance region. Now since non-adiabaticity is important only near resonance
points, the expression for X shall depend only on the value of at resonance (say R ). This
means that if R  1, there is full adiabaticity and discrete jumps do not come into play at all.
One of the ways this happens is if E . Note that this means that high energy neutrinos
are less prone to non-adiabatic effects.
Now an expression for X may be derived using the Landau-Zener method as described in
[11].
X = exp (F R )
(2.83)
with F = 4 assuming a linear density variation near resonance.
Although this shall suffice for out purposes, it is a semi classical expression which gives only
the leading term in the limit of large values of the exponent. Also, the Suns density variation
is more exponential than linear[14]. The exact solution, obtained for an exponentially varying
matter density[15] is of the form:


R F
exp(R F ) exp sin
2


X=
(2.84)
R F
1 exp sin
2
where F is a quantity that depends on the variation of ne near resonance.
Considering a beam of electron neutrinos originating in the Suns core and travelling towards
its surface, the survival probability (accounting for resonance jumps) is given by:
(ad)

Pee = (1 X)P
 ee
=

(1X)
2
1
2 1+

(ad)

+ XPe 
1 + cos20 cos2 +

X
2


(1 2X)cos20 cos2
18

1 cos20 cos2

(2.85)

CHAPTER 2. NEUTRINO OSCILLATIONS

Now, if the beam is produced somewhere outside the core so that it passes the resonance region
twice on its way out,
Pee =
=

 (ad)
(ad)
2
2
(1
 X) + X Pee + 2X(1 X)Pe
1
2

2 1 + (1 2X) cos20 cos2

(2.86)

Hence, the general expression for survival probability is given by:


(n)
Pee
=


1
1 + (1 2X)n cos20 cos2
2

(2.87)

where the superscript n stands for the number of jumps and can take values 1 or 2.

2.3.4

The big picture: Looking at the Solar Neutrino problem

Armed with the formalism developed so far, we shall take a preliminary look at the solar
neutrino problem (see Section 3.2). The following simplifications shall be made:
1. All neutrinos are produced in the core of the Sun (defined as the region with r . 0.07R )
so that N0 = 98.8 Navo (where Navo := 6.023 1023 /cc).

d
1
ln ne R = 10R
irrespective of where resonance takes place inside the Sun. This is
2. dx
10r

justified because the density profile is very well approximated by ne 200 Navo e R
everywhere in the Sun except in the inner 15%[11]. So, this is not too bad an assumption
unless resonance occurs deep within the Sun.
The density profile of the Sun as in the model considered is an exponentially decreasing
profile and hence looking at the expression for mixing angle in matter, it is clear that (x)
drops down continuously and monotonically from 0 to the vacuum mixing angle, which is less
that 4 (assuming that ne (x) decreases as well). Now, since resonance is assumed to not occur
in the inner 15% of the Sun and production occurs in the inner 7% of the Sun, it is clear that
neutrinos pass resonance only once, if at all. Hence, at the point of production, if 0 4 , it
must pass 4 and undergo resonance once and exactly once. So the condition for resonance is
given by:

cos2 < 2 2GF N0 = 1.5 1011


(2.88)
E
Non-adiabatic effects, as maybe seen from the expression for Pee are unimportant as long
as X  0.5. Let us say that they are unimportant if X 0.05. Now this occurs for (using
the simplified expression from Eqn. 2.83) R < 3.8. Using the given model for density, this
translates to:
sin2 2
.
< 1014 eV
(2.89)
E cos2
using r = 7 1010 cm. Fig. 2.5 shows a plot of survival probability against
E for various
values of vacuum mixing angle. The dotted line marks the onset of non-adiabatic effects, which
become important to the left. At the right end,
E is too large for either the non-adiabaticity or
the resonance condition to be satisfied. Hence, the survival probability reverts back to the average value in vacuum. As
E decreases, there is a sudden plunge due to the onset of resonance.

The range of E over which resonance takes place is determined by the width of the resonance.
Clearly, the width increases with increasing as expected.
19

CHAPTER 2. NEUTRINO OSCILLATIONS

Figure 2.5: The survival probability is plotted as a function of


E . Nonadiabatic effects become important to the left of the vertical dashed line.
In the bottom-most
figure, the dashed line has been computed with F =


2
4 1 tan (recall Eqn. 2.83).
But the probability rises from the resonance basin as non-adiabatic effects become more and
more important. The
E for which these effects become significant decreases with increasing .
For = 0.01, non-adiabaticity effects take over even before the onset of resonance and hence
the basin is never reached.
We shall also look at the exclusion plots for at constant energies. There are three branches

Figure 2.6: Exclusion plot for the problem discussed. The dashed lines stand
for 0.19 < Pee < 0.35 at E=5 MeV.The dotted lines stand for 0.18 < Pee <
0.55 at E=10 MeV. The shaded region is allowed by both conditions
of possible solutions: a horizontal, vertical and diagonal branch (the horizontal branch is easily
eliminated since the exclusion plots do not overlap there for different energies). The simplified picture discussed here is very close to the more elaborate calculations done for solar
neutrinos[11]. In fact, the dashed and dotted lines have been drawn with probabilities close to
those observed at the Chlorine and Kamiokande experiments a while ago[11].

20

Chapter 3

Experiments in Neutrino Physics:


Measuring the Oscillation
Parameters
The are two prominent natural sources of neutrinos on earth besides radioactivity. These are
neutrinos generated by the effect of cosmic rays (atmospheric neutrinos) and those generated
in the sun (solar neutrinos). It was through these sources that the existence of oscillations
was experimentally established. This chapter involves a brief discussion of experiments that
measured the relevant parameters using these sources. Our discussion in this chapter shall
closely follow [9].
But first a brief note on some ways used for the detection of neutrinos.

3.1

Detection of Neutrinos

We shall briefly discuss three of the principal methods of detecting neutrinos and their
advantages/disadvantages:
1. Radiochemical detectors: Radiochemical detectors use the inverse beta decay reaction:
e + X e + Y

(3.1)

to detect electron neutrinos. The detector is exposed to the flux for a while and then the
number of Y atoms are counted to obtain the flux. The Y atoms are usually radioactive
and hence the detector cannot be exposed for arbitrary periods of time.
The principal disadvantage of this technique is that it is blind to important properties of
the flux such as its energy spectrum and arrival rate. Radiochemical detectors measure
just the integrated flux over a period of time. However, this being said, radiochemical
detectors have extremely low detection thresholds and hence can be used to gauge low
energy neutrinos such as those from the pp and pep flux of solar neutrinos (see Section 3.2).
Given in Table 3.1 are some materials used as detectors and their thresholds along with
other parameters. Experiments that used radiochemical detectors include (see Section3.2)
Homestake, SAGE, GALLEX and GNO.
2. Water Cherenkov detectors: These use the scattering reactions:
x + e x + e
21

CHAPTER 3. EXPERIMENTS IN NEUTRINO PHYSICS: MEASURING THE


OSCILLATION PARAMETERS
Initial Nucleus (X)

Final Nucleus (Y)

Threshold
MeV

Half-Life of Y

Capture Rate
SNU

37 Cl

37 Ar

0.814

35 days

7.9 2.6

71 Ga

71 Ge

0.233

11.4 days

132+20
17

7 Li

7 Be

0.862

53.4 days

51.8 16

127 I

127 Xe

0.789

36 days

80

81 Br

81 Kr

0.470

2 105 years

27.8+17
11

98 Mo

98 Tc

1.68

4 106 years

17.4+18.5
11

205 Tl

205 Pb

0.062

107 years

263

Table 3.1: The table lists various elements used as radiochemical detectors and their relevant
parameters (taken from [11]). The solar neutrino unit (SNU) is defined as the neutrino flux
which produces 1036 captures per target atom per second.
x + p x + n

(3.2)

and the electron (or lepton in case of inelastic scattering) hence scattered has a strong
forward energy peak in the direction of the incident neutrino. The accelerated lepton
produces radiation via the Cherenkov effect in water and hence provides a way for detecting neutrinos. The advantage of such detectors is that they can detect neutrinos in
real-time, and can also distinguish between electron and muon neutrinos (muons produce
clear sharp rings while electron rings tend to be fuzzy). They can also measure the direction and energy of the incoming neutrinos. The catch is that the reactions (Eqn. 3.2)
have higher thresholds (the threshold for taon neutrinos is so high that they can rarely
be detected; see Section 3.2.2) than the radiochemical ones (discussed above) and hence
cannot be used to measure low energy neutrinos.
Famous experiments that used water Cherenkov detectors include Super-Kamiokande and
SNO (used heavy water instead; see Section 3.2.2).
3. Liquid Scintillation Detectors There are also the numerous experiments like Borexino
(see Section 3.2.2), Daya Bay and RENO (see Section 3.4), which use a liquid scintillation
detector. They rely on certain materials that develop excitations on passage of neutrinos
through them. These excitations then lead to the emission of light which is then observed.

3.2
3.2.1

Solar neutrinos
The Solar Neutrino Flux

Solar neutrinos are produced as by-products during nuclear fusion processes in the sun. The
net reaction (called the pp chain) may be written as:
4p

He + 2e+ + 2e

(3.3)

and produces roughly 25 MeV of energy[11] along with two electron neutrinos. However, this
reaction occurs via a large number of sub-processes (listed in 3.2) and the electron neutrinos
22

CHAPTER 3. EXPERIMENTS IN NEUTRINO PHYSICS: MEASURING THE


OSCILLATION PARAMETERS

produced hence have their own characteristic energy spectrum (shown in Fig. 3.1)). There is
also another chain of reactions known as the CNO cycle (also listed in 3.2) that contribute to
the neutrino flux.

Abbr.

Flux
cm2 s1

2p d + e+ + e

pp

5.97(1 0.006) 1010

p + e + p d + e

pep

1.41(1 0.011) 108

hep

7.90(1 0.15) 103

7 Be

5.07(1 0.06) 109

8B

5.94(1 0.11) 106

13 N

2.88(1 0.15) 108

+ e+ + e

15 O

8
2.15(1+0.17
0.16 ) 10

+ e+ + e

17 F

6
5.82(1+0.19
0.17 ) 10

Reaction

3 He

+ p

7 Be

+ e

8B

13 N

15 O

17 F

8 Be
13 C
15 N
17 O

4 He

+ e+ + e

7 Li

+ e

+ e+ + e
+ e+ + e

Table 3.2: The table lists important neutrino producing processes in the Sun and the SSM
predictions of their contributions to the solar neutrino flux (taken from [9]). The first five
reactions form the pp chain, while the last three are called the CNO cycle.
The energy created via these reactions is predominantly in the form of photons, but these
photons take a long time to come to the surface of the Sun. Neutrinos, on the other hand, have
very long mean free paths even in such high matter densities and hence reach the Earth in a much
shorter time frame. Hence, studying the solar neutrino flux, together with the information on
solar radiation showed that the Sun has been in thermal equilibrium for a long time.
Solar
neutrino fluxes are calculated using Standard Solar Models (or SSMs). These are complied
calculations on the composition, temperature, density, etc of the Sun based on observations.
The most elaborate SSMs are believed to have been developed by the late John Bahcall and
collaborators. SSMs labelled BS05(OP)[24], BS06(GS) and BSB06(AGS)[25], BPS08(GS) and
BPS08(AGS)[26] represent some their recent calculations. The energy spectrum of the solar
neutrino flux, as predicted by BS05(OP) is given in Fig. 3.1.

3.2.2

Measuring the Solar Neutrino Flux

Some of the major collaborations (types of detectors used given in parenthesis; see Section 3.1) that have studied the solar neutrino flux till date are Homestake (Chlorine), SAGE,
GALLEX, GNO (Gallium) Kamiokande, Super-Kamiokande (water Cherenkov), SNO (heavy
water) and Borexino (liquid scintillation). These experiments and their obsevations shall be
discussed next.
Homestake
Homestake was the first experiment to study solar neutrinos. It was started by Davis et al.
at Homestake in the late 1960s, and used the 37 Cl 37 Ar method proposed by Pontecorvo[27]
23

CHAPTER 3. EXPERIMENTS IN NEUTRINO PHYSICS: MEASURING THE


OSCILLATION PARAMETERS

Figure 3.1: Figure shows the flux predictions of the SSM BS05(OP)[24]
(taken from [23]).
for detection.
37

Cl + e

37

Ar + e

(3.4)

This experiment had a threshold of around 814 keV and was sensitive to(according to SSMs)
pep, 13 N, and 15 O neutrinos. Right from the very first observations, it was clear that
there was a deficit in the solar neutrino flux as predicted by the SSMs and this was the origin
of the solar neutrino problem.
8 B, 7 Be,

The Gallium Experiments


Gallium radiochemical detectors employ the reaction
71

Ga + e

71

Ge + e

(3.5)

to detect neutrinos beyond a threshold energy of about 233 keV. Hence, they are sensitive the
most dominant component of the solar neutrino flux - the pp neutrinos. SSMs predict that
more than 80% of the flux measured by Gallium detectors is due to pp and 7 Be neutrinos.
The SAGE collaboration at Baskan in Russia published its first results in 1991, followed
soon after by the GALLEX experiment (at Gran Sasso, Italy) in 1992. GALLEX finished
observations in early 1997 and since April 1998, a newly defined collaboration: GNO continued
measurements till April 2003. The results from all these experiments along with Homestake are
summarised in Table 3.3.

24

CHAPTER 3. EXPERIMENTS IN NEUTRINO PHYSICS: MEASURING THE


OSCILLATION PARAMETERS

Experiment

Flux from

37 Cl

(SNU)

Flux from

71 Ga

(SNU)

Homestake[28]

2.56 0.16 0.16

GALLEX[29]

77.5 6.2+4.3
4.7

GNO [30]

73.4+5.5
5.3 2.5

GNO + GALLEX[31]

+3.2
67.6+4.0
4.0 3.2

SAGE[32]

2.6
65.4+3.1
3.0 2.8

8.46+0.87
0.88

127.9+8.1
8.2

SSM (BPS08[GS][26])

Table 3.3: The table summarises results by the discussed radiochemical detector-based experiments and compares them with predictions of the SSM (taken from [9]).
Super-Kamiokande
The Kamioka Neutrino Detection Experiment (Kamiokande), located in the Mozumi mine
of the Kamioka Mining Company in the Gifu prefecture in Japan, observed real-time scattering
of solar neutrinos in 1987. Using a water Cherenokov detector, it was sensitive to the direction
of the neutrinos and hence provided the first direct evidence of the fact that neutrinos came
from the direction of the Sun[21]. It was later upgraded to the Super Kamiokande (Super-K)
experiment, which started observations on April 1, 1996. The Super K detector is a 50-kton
cylindrical water Cherenkov detector 41.4 m high and 39.3 m in diameter.[18].
The inner detector is lined with 11,146 photo-multiplier tubes (PMTs) facing a fiducial 22.5
kton volume of ultra-pure water. The inner volume is surrounded by a roughly 2 meter thick
outer detector equipped with 1885 out-facing PMTs.
Events were classified into two categories - fully contained (FC) and partially contained (PC)
based on whether the Cherenkov light was deposited entirely or partly in the inner detector.
FC events were again separated into those with a single Cherenkov ring and those with multiple
Cherenkov rings, and only single-ring events were used in the analysis of FC events. The FC
events were also classified as sub-GeV (Evis < 1330 M eV ) and multi-GeV (Evis > 1330 M eV ),
where (Evis was defined as the energy of the electron that would produced the observed amount
of Cherenkov radiation.
The Super-K experiment has thresholds of around 5 MeV [18] and measures the 8 B flux.
A summary of their results is given in Table 3.4.
The SNO Experiment
The SNO (Sudbury Neutrino Observatory) experiment in Canada used a heavy water
Cherenkov detector that used both CC and NC reactions (Eqn.3.6, 3.7) to detect neutrinos.
Located 6800 ft underground, it used 1000 tonnes of heavy water contained in a 12 meter diameter acrylic vessel and the radiated Chrenkov light was picked up by 9600 photomultiplier
tubes mounted on a support structure.[16]
e + d e + p + p [CC]

(3.6)

x + d x + p + n [N C]

(3.7)

25

CHAPTER 3. EXPERIMENTS IN NEUTRINO PHYSICS: MEASURING THE


OSCILLATION PARAMETERS

Experiment

Reaction

8B

Flux (106 cm2 s1 )

Kamiokande[33]

2.80 0.19 0.33

Super-K[34]

2.32 0.04 0.05

SNO (Ph. III) [17]

CC

+0.07
1.67+0.05
0.04 0.08

NC

+0.36
5.54+0.33
0.31 0.34

Borexino[35]

2.4 0.4 0.1

SSM (BPS08[GS][26])

5.94(1 0.11)

SSM (SHP11[GS][37])

5.58(1 0.14)

Table 3.4: The table summarises results by the discussed real-time observation experiments and
compares them with predictions of the SSM (taken from [9]).
Part of
spectrum

Measured Flux
(cm2 s1 )

BPS08[GS][26]
(cm2 s1 )

SHP11[GS][37]
(cm2 s1 )

8B

(2.4 0.4 0.1) 106


(3.10 0.15) 109
(1.0 0.2) 108
(6.6 0.7) 1010

5.94(1 0.11) 106


5.07(1 0.06) 109
1.41(1 0.011) 108
5.98(1 0.006) 1010

5.58(1 0.14) 106


5.00(1 0.07) 106
1.44(1 0.012) 109
5.97(1 0.006) 1010

7 Be

pep
pp

Table 3.5: The table summarises the measurements by the Borexino experiment (taken from
[9, 71, 72]).
Note that the CC reaction is only senstitive to e while the NC reaction is undergone by all
active neutrino flavours. Hence, SNO had the great advantage of being able to measure both
the electron- and the total (active) neutrino flux simultaneously and hence demonstrate that
the missing electron-neutrinos do indeed manifest as other flavour neutrinos.
The CC and NC events were distinguished on the basis that while the momentum of the
electron produced in the NC reaction had a strong forward peak (with respect to the Earth-Sun
axis), the same for the CC reaction had an approximate angular distribution of 1 13 cos.
The SNO experiment was run in three different phases with different mechanisms to detect the neutrons from the NC reaction. In the first phase, the neutrons were captured on
deuterium nucleii in D2 O. In the second phase, 2 tons of NaCl were added and the neutrons
were captured on the 35 Cl to raise detection efficiency. The third phase relied on an array
of proportional counters (the Neutral Current Detection or NCD array) deployed in the heavy
water and neutrons were detected via the reaction[17]:
3

He + n

H + p

(3.8)

The SNO experiment could accurately determine the shape of the 8 B spectrum owing to
the fact that the energy of the electrons produced via the CC reactions had a strong correlation
with that of the original neutrinos. A combined analysis of the data from all phases of SNO
could be found in [17]. The SNO detector is currently being upgraded for the planned SNO+
26

CHAPTER 3. EXPERIMENTS IN NEUTRINO PHYSICS: MEASURING THE


OSCILLATION PARAMETERS

experiment[16].
The Borexino experiment
The Borexino experiment also uses a liquid scintillation detector to study solar neutrinos.
It came online in 2007, and was highly sensitive to all solar neutrino components, particularly
those below 2 MeV. Borexino recently succeeded in directly measuring the pp neutrino flux[71]
- something that had not been possible over the past 30 years due to the inability to suppress
the high background noise at this energy range.
Borexino was also the first experiment to directly observe 7 Be neutrinos[72].
Borexino has also measured the pep and 8 B part of the solar neutrino spectrum (see Table
3.5 for a summary of results and comparison with SSM predictions), and set an upper limit for
the CNO flux, assuming the LMA solution (see Section 3.2.3).

3.2.3

The Solution of the Solar Neutrino Problem

Until recently, the solution to the solar neutrino problem had only been narrowed down to
one of three regions in parameter space[22][10] (90% CL):
The small mixing angle (SMA) MSW solution:
4 106 eV2 . m2 . 1.2 105 eV2
3 103 . sin2 2 . 1.1 102
The large mixing angle (LMA) MSW solution:
8 106 eV2 . m2 . 3.0 105 eV2
0.42 . sin2 2 . 0.74
The vacuum oscillation solution:
6 1011 eV2 . m2 . 1.1 1010 eV2
0.70 . sin2 2 . 1
In 2001, the SNO experiment and Super-K[38] provided evidence for flavour conversion in
solar neutrino flux[39]. The NC flux measured by SNO showed a good agreement with the
8 B flux of the solar model[24]. The results were analysed and found to be consistent with the
LMA solution (with MSW effects) to the solar neutrino problem[13, 12]. However, SNO alone
could not rule out other possibilities with sufficient significance, and this job was completed b
KamLAND.
KamLAND
KamLAND is a long-baseline experiment with a liquid scintillation detector. It observed
anti-neutrinos from reactors and was sensitive down to m2 105 . Hence, KamLAND should
not observe a substantial e disappearance should SMA (mixing angle is too small) or vacuum
oscillations (oscillations happen over a much bigger range) happened to be the right solution.

27

CHAPTER 3. EXPERIMENTS IN NEUTRINO PHYSICS: MEASURING THE


OSCILLATION PARAMETERS

KamLAND reported its first results[40] in December 2002, with a clear observation of e
disappearance.
Nobs NBG
= 0.611 0.085 0.041
(3.9)
R :=
NN oOsc
with Nobs being the oberved number of events, NBG the background, and NN oOsc the expected
number of events in case of no oscillations. A combined analysis with other data showed that
LMA was the unique solution to the solar neutrino problem at more than 5 CL [41, 42, 43,
44, 45].
Later on, KamLAND also observed the periodic variation of e survival probability,

Figure 3.2: The e survival probability as observed by KamLAND, as a L0 /E


where L0 = 180 km. The histograms show the expected distributions from
a 2 and 3-flavour analysis using the best-fit values of the parameters (taken
from [9], originally from [46]).
that is expected with oscillations, for the first time (see Fig. 3.2). Including recent data on 13
measurements, a 3-flavour oscillation analysis of the solar neutrino and KamLAND gives the
current values of the solar oscillation parameters.
m221 = (7.53 0.18) 105 eV2
2

sin 13 = 0.023 0.002


2

tan 12 =

3.3
3.3.1

0.436+0.029
0.025

(3.10)
(3.11)
(3.12)

Experiments involving Atmospheric Neutrinos


Cosmic Rays

Cosmic rays are highly energetic beams of particles that consist of about 90% protons, 9%
alpha particles and 1% heavier nucleii. The absolute flux of cosmic rays is of the order of 1000
28

CHAPTER 3. EXPERIMENTS IN NEUTRINO PHYSICS: MEASURING THE


OSCILLATION PARAMETERS

m2 s1 sr1 for energies of a few GeV.


Above a few GeVs the energy spectrum for protons falls off, to a good approximation, as
E 2.7 [57]. The fluxes are less well known at higher energies. It is to eliminate the uncertainity
in cosmic ray fluxes that experiments rely on flux ratios rather than the absolute values. Also,
the H to He ratio is not a constant with energy. At around 100 MeV, this ratio is less than 5.
It rises to about 10 at energies around 1 GeV and then to about 30 at around 100 GeV[57, 58].
Cosmic rays fluxes are modulated via two processes: the solar wind and the geomagnetic
cut-off. The solar wind weakens the galactic cosmic ray flux at lower energies. The magnitude of
this effect on a particle is determined by its momentum to charge ratio (known as rigidity)-the
effect being weaker at higher rigidities. By around 2 GeV, the solar wind influence is nearly
negligible[10].
The Earths magnetic field influences the charged particles in the cosmic ray flux to produce
what are known as geomagnetic effects. To present a somewhat simplified picture, consider a
positively charged particle in a uniform magnetic field. Now, this particle moves in a helical
path around the field lines and in the direction of the field line (or against, depending on its
initial momentum). This is mechanism behind geomagnetic effects.
It is easy to see that the trajectory of the particle depends on the ratio of the momentum
of the particle to the charge (known as the rigidity). Now, particles with a high rigidity
move in a helix with a large radius and hence appear to be uninfluenced by the field. This
is (roughly speaking) how the high energy part of the cosmic ray spectrum remains untouched
by geomagnetic effects. Meanwhile, particles with a low enough rigidity never make it to the
Earths surface (some of them follow the field lines to the Earths magnetic poles) - hence
creating a cut-off rigidity for cosmic ray particles.
The cosmic rays that do make it past the barriers discussed above react with atmospheric
nucleii to form neutrinos that constitute the so-called atmospheric neutrino flux.

3.3.2

Production of Atmospheric Neutrinos

The production of atmospheric neutrinos proceeds essentially in three steps. First, the
primary cosmic rays react with nuceii in the atmosphere to give pions and kaons. The pions
then decay into muons which then give muon neutrinos (branching ratios given in square braces).
[9]
+ ( )
[99.99%]

e + ( ) + e (e ) [100.0%]
The kaons either directly give muons or decay into pions. [9]
K
K
K
K

3.3.3

+ ( )
+ 0
+ +
0 + e + e (e )

[63.6%]
[20.7%]
[5.59%]
[5.07%]

Characteristics of the atmospheric neutrino flux[10]

1. The muon and electron neutrino fluxes follow the power laws [ ] E 3 and [e ] E 3.5
E
for energies 1 . GeV
. 10 ([e ] denoting the flux of e ).
2. The largest contributions to the neutrino flux at energy E come from cosmic rays of energy
10E.

29

CHAPTER 3. EXPERIMENTS IN NEUTRINO PHYSICS: MEASURING THE


OSCILLATION PARAMETERS

[ +
]

3. From the production mechanisms, it is easy to see that [e +e ] ' 2. But for E & 1 GeV ,
the muons start reaching the Earths surface before decay and hence stop contributing.
Therefore, once again looking at the production mechanisms, [e + e ] falls and the ratio
[ +
]
[e +
e ] rises. Around 100 GeV , muons cease to be the dominant source of e and the KL
and K + reactions take over.
4.

[e ]
[
e ]

> 1 because protons heavily dominate the primary cosmic ray flux and hence during
[ + ]
[+ ]
[e ]
hadronic interactions, the production of + is favoured over . So [ ] ' [ ] ' [
e ] > 1.

5. For the same reason, while ' 1 for E . 1 GeV (since muon decay supplies the
corresponding antiparticles), this ratio increases beyond a few GeV when the muons start
reaching the Earth before decay.
6. The geomagnetic effects are negligible for E & 5 GeV .

3.3.4

Zenith angle dependence of the atmospheric neutrino flux

There are two main effects (besides oscillations) that lead to atmospheric neutrino flux being
a function of angle.
The cosmic ray reactions are dependant of the gradient of the density of air. As zenith
angle approaches 2 , the density variation becomes less steep and hence pion and kaon
decay is enhance as compared to the flux along the vertical. However, it is easy to see
that the angular dependence induced by this mechanism is up-down symmetric, ie, the
dependence is on |cos|.
The geomagnetic effect causes a zenith as well as azimuthal angle dependence for the low
energy part of the spectrum. This effect is up-down asymmetric but does not effect the
high energy neutrinos.
Neutrino oscillations also create a zenith angle dependence in the atmospheric neutrino
flux. This is because of the variation in the path length traversed by neutrinos approaching the detector on the Earths surface at various angles. Up-going neutrinos have to cross
the entire diameter of the Earth (L 13000 km), while down-going neutrinos merely cross
part of the atmosphere (L 20 km). This effect is clearly not up-down symmetric (dependence on cos) and can be decoupled from geomagnetic effects by using the high energy
component of the spectrum. This zenith angle dependence was exploited to obtain the
relevant mass-squared difference and mixing angle for atmospheric neutrino oscillations.

3.3.5

Measuring the Atmospheric Neutrino Flux

Super Kamiokande
The Super-Kamiokande provided evidence of disappearance of atmospheric muon neutrinos
by measuring a [ ]:[e ] ratio of less than 2 for the valid energy range (see Section 3.3.3)[47]. The
data, consistent with oscillations, also drew rival interpretations such as the neutrino
decay[48] and quantum decoherence[49]. However, observation of a dip in the L/E distribution of
the data[20] (see Fig. 3.4), which exhibited a sinusoidal behaviour characteristic of oscillations,
ruled out these theories: neutrino decay being disfavoured at 3.4 and decoherence even more
strongly.
30

CHAPTER 3. EXPERIMENTS IN NEUTRINO PHYSICS: MEASURING THE


OSCILLATION PARAMETERS

Figure 3.3: The figure shows the zenith angle distribution for atmospheric
neutrinos as measured by Super-K. The dotted histograms show the MonteCarlo predictions for the no-oscillation hypothesis and the solid histograms
show the best-fit expectations for oscillations (taken from [9]).

Figure 3.4: The figure shows the results of the analysis of SK-I data. The
solid line shows the best-fit with 2-flavour oscillations. The dashed
and dotted lines show the predictions by the neutrino decay and quantum
decoherence hypotheses (taken from [9], originally from [20]

31

CHAPTER 3. EXPERIMENTS IN NEUTRINO PHYSICS: MEASURING THE


OSCILLATION PARAMETERS

The final step was the demonstration of appearance of tau neutrinos. Now because of
the high threshold of the corresponding reaction, (the event rate is extremely small) and the
short lifetime of the tau lepton, Super-K cannot detect taon neutrinos as easily as it does the
others. Hence techniques such as neural network analysis were employed in carefully selecting
the occasional tau events from the large collection of data [50][51] and appearance was
established at the 3.8 level in 2013.
Although the first measurements of m2atm and atm were made from atmospheric neutrinos,
the uncertainties involved prevented accurate measurement of these parameters. Accelerator
based experiments had more merits on this front since the variables involved like the length of
the baseline and the energy spectrum of the neutrinos were known more precisely. These shall
be discussed next.
The K2K experiment
The KEK-to-Kamioka (K2K) experiment was a long-baseline experiment with a baseline of
250 km. It used a muon neutrino beam from a 12 GeV proton synchrotron (KEK-PS) with an
average beam energy of about 1.3 GeV. The energy profile of the source was determined by a
near detector located 300 m downstream. Super-K served as the far detector.
The K2K experiment started taking data in 1999 and finished its observations in 2004. The
number of FC events observed at Super-K was 112 as opposed to the no-oscillation prediction
of 158.1+9.2
8.6 . The measured energy spectrum also showed the necessary distortion required by
neutrino oscillations. The probability that these results are due to a statistical fluctuation were
estimated to be 0.0015% or 4.3 [52].
The MINOS Experiment
MINOS uses neutrinos produced at the NuMI (Neutrinos at the Main Injector) facility using
120 GeV protons from the Fermilab Main Injector. It has a baseline of 735 km and most of its
data employed a neutrino beam with an energy spread of 1-5 GeV (peaked at 3 GeV).
MINOS data (Fig. 3.5) came down in favour of neutrino oscillations, disfavouring decay
and decoherence at 7 and 9 respectively[53]. The result (at 86% CL) disfavoured maximal
mixing. Recently, MINOS has also published[54] a combined analysis of disappearance and
nue appearance using both accelerator and atmospheric neutrino data. The results now put the
parameter values at |m2atm | = (2.28 2.46) 103 eV2 (68% CL) and sin2 atm = 0.35 0.65
(90% CL) for normal hierarchy, and |m2atm | = (2.322.53)103 eV2 (68% CL) and sin2 atm =
0.34 0.67 (90% CL) for inverted hierarchy.
The T2K Experiment
The T2K experiment is the first off-axis long-baseline experiment with a baseline of 295 km
from J-PARC in Tokai, Japan to Super-K. It uses a narrow-band beam with a peak energy
of 0.6 GeV directed 2.5o off-axis to Super-K.
T2K started its first run in 2010. Recent measurements[55] report 58 1-ring -like events
against an expectation of 205 17 for no oscillations. T2K results[56] currently put down the
relevant parameter values to be |m2atm | = (2.51 0.10) 103 eV2 (68% CL) and sin2 atm =
2
2
3
0.514+0.055
0.056 (68% CL) for normal hierarchy, and |matm | = (2.48 0.10) 10 eV (68% CL)
2
and sin atm = 0.511 0.055 (68% CL) for inverted hierarchy. Note that T2K results are
consistent with maximal mixing.

32

CHAPTER 3. EXPERIMENTS IN NEUTRINO PHYSICS: MEASURING THE


OSCILLATION PARAMETERS

Figure 3.5: The top part of the figure shows the energy spectrum of the fully
reconstructed CC events from the MINOS far detector. The bottom panel
shows a plot of R (see Eqn. 3.9) against reconstructed neutrino energy. The
best-fit to oscillations, decay and decoherence is also shown (taken from [9]).
OPERA
The OPERA experiment has been been designed to directly detect tau neutrinos via CC
interactions. It uses an emulsion technique which is able to identify the short-lived leptons
on an event-by-event basis. It has a baseline of 730 km length with the source at CERN and
the detector at Gran Sasso in Italy. The average beam energy is 17 GeV.
The detector is a combination of the Emulsion Cloud Chamber and magnetised spectrometer. and uses neutrinos produced by high-energy protons from the CERN-SPS. OPERA
collected data during 2008 and 2012, and analysis has not yet been completed[9].

3.4

Accelerator experiments to measure 13

13 measurements were carried out by long-baseline experiments (Daya Bay, RENO, Double
CHOOZ) using e oscillations which were driven by m2atm . The accelerator driven experiments
discussed earlier (MINOS, T2K, etc) studying oscillations are also sensitive to 13 . The results

33

CHAPTER 3. EXPERIMENTS IN NEUTRINO PHYSICS: MEASURING THE


OSCILLATION PARAMETERS

Figure 3.6: The figure shows the 90% CL allowed regions in the atmospheric
neutrino parameter space as estimated by T2K (2013[55], 2011[68]) SuperK[69] and MINOS[70] (taken from [9]).
from these experiments shall be discussed next.
CHOOZ and Double CHOOZ
CHOOZ[59] was a long baseline experiment located at the CHOOZ nuclear power station
in France. With an average value of L/E 300 eV2 (L 1 km and hEi 3 MeV), it studied
the disappearance of reactor e driven by m2atm and 13 (recall Section 2.2.1). CHOOZ used
a liquid scintillation detector and neutrinos were detected via the reaction :
e + p e+ + n

(3.13)

CHOOZ initially found no evidence for e oscillations[59]. However, Double CHOOZ has
recently reported a non-zero value for 13 , measuring sin2 213 = 0.109 0.030 0.025[60].
Daya Bay and RENO
The Daya Bay experiment uses the Daya Bay nuclear power complex in China as source, and
a gadolinium-doped liquid scintillation detector to observe reactor e . RENO studies reactor
e s from the Yonggwang nuclear power plant in Korea with two identical liquid scintillation
detectors located 294 m and 1383 m away from the center of the reactor array.
Daya Bay observed a non-zero deficit in the e flux, with the ratio of observed to expected
events being R = 0.944 0.007 0.003, and a 3-flavour oscillation analysis yielded sin2 213 =
0.089 0.010 0.005 . This result excluded the no-oscillation hypothesis at 7.7[61]. Later
on, the bounds were improved to sin2 213 = 0.090+0.008
0.009 [62]. RENO initially measured R =
0.920 0.009 0.014, and sin2 213 = 0.113 0.013 0.019 from a rate-only analysis, hence
34

CHAPTER 3. EXPERIMENTS IN NEUTRINO PHYSICS: MEASURING THE


OSCILLATION PARAMETERS

Figure 3.7: The figure shows the regions in parameter space allowed/excluded from various neutrino experiments (taken from [74]).
excluding the no-oscillation hypothesis at the 4.9 . Later on, an improved data analysis[64]
gave sin2 213 = 0.113 0.013 0.019.
T2K and MINOS
In 2011, T2K, while studying e oscillations reported a non-zero 13 value with a signifi- 
+0.045
cance of 2.5[65]. Recently, this was improved to 7.3 [66] and sin2 213 = 0.140+0.038
0.032 0.1700.037
2
was reported for normal (inverted) hierarchies, assuming CP = 0, sin 23 = 0.5 and |m231 | =
2.4 103 eV2 . Comparison with Daya Bays results[62] immediately leads to 6= 0 or
sin2 23 6= 0.5.
MINOS has also recently reported its estimates
for 13 . The best-fit value at 90 % CL being

+0.054
+0.038
2
2
2 sin 23 sin 213 = 0.0510.030 0.0930.049 [70] for normal (inverted) hierarchy.

35

Chapter 4

Sterile Neutrino Oscillations


Having taken a brief glance at the recent developments in the field and the current status
of the oscillation parameters, we now attempt to come up with a proposal to look for sterile
neutrinos with the aid of the MSW effect.

4.1

Derivation of matter term for s oscillations

The fact that sterile neutrinos have eluded detection in the experiments discussed so far
places strong constraints on the mixing between active and sterile states[19]. Hence, we shall
recall that even small mixing angles maybe drastically enhanced by means of the MSW effect
(recall Section 2.3.1) and use this to derive the baseline for a stand-alone accelerator-driven
experiment that shall look for sterile neutrinos. The derivation in this section shall closely
follow that in Section 2.3.1.
Consider an effective two neutrino system with and s , with a small mixing angle and
driven by a m2 = 0.01eV2 , propagating through a uniform matter density. Assuming that
the dominant process is once again coherent forward scattering, the Lagrangian is given by:
(assuming that the thermal energy of the particles is much less than the mass of Z)






4GF
1 5

2
L=
f (p1 ) I3L
Qsin W f (p2 ) {
L (p3 ) L (p4 )}
2
2
where f 0 stands for the proton, electron and neutron. The conditions for coherent forward
scattering are given by:
p1 = p2 = p
p3 = p4 = p
The effective Lagrangian for the process is obtained by integrating the Lagrangian over p,
assuming a thermal distribution for the matter particles and using the formulae:
hf 5 f i
hspini
= 0 [unpolarised medium]
hfj f i hvelocityi = 0
[stationary medium]

hf 0 f i =
nf
[hnumber operatori]
Leff


X  f

nf I3L 2sin2 W Qf [
L 0 L ]
= 2GF
f

36

CHAPTER 4. STERILE NEUTRINO OSCILLATIONS

Now, the contribution due to the protons and electrons cancel out because they have opposite
values of I3L and charge. The only term left is from the neutron density nn .
1
Leff = GF nn L 0 L
2
This contributes in the form of a potential for (with no modification for s ) and the modified
Hamiltonian becomes:
#

 "
2 + m2
1 GF nn 0

cos2
sin2
2
=E+ 1
2
H
+
+
sin2 cos2
4E
4E
0
0
|
{z
}
Hvac

= 1
H
4E

cos2 + 2A sin2
sin2
cos2


+ I2

where I2 is the identity matrix (addition of the scalar matrix merely adds an overall phase and
doesnt affect the physics), and A is given by:

A = 2GF nn E
This gives the effective mixing angle and mass difference in vacuum (tilde indicating that the
parameter value is in the presence of matter effects.)
sin2 2 =
=

4.1.1

2 sin2 2
(cos2 A)2 + 2 sin2 2

q
(cos2 A)2 + 2 sin2 2

Certain Issues

The usual calculation of resonance neutron density yields (assuming mixing angle to be 0.01
rad):
(R)

nn

= cos2
2G E
F

0.01eV 1.000
= 21.1710
5 GeV 2 1GeV

= 8.017 1025 cc1

We recall the discussion in Section 2.3.1 (regarding only one of x y and y x being
amplified by matter effects) and conclude that this may be fixed and that a resonance may be
achieved by using antiparticles.

4.2

Propagation in the Earths crust

Firstly, we calculate the density of neutrons in the Earths crust. The contribution of the
ith constituent element toward this is given by:
n(i)
n =

Pi
Yi NA
Mi

where is the average density of the Earths crust (3.2 gcc1 ), NA is the Avogadro number, Pi
is the percentage composition of the element (by weight), Mi its molar mass, and Yi its neutron
number. The total neutron density is then obtained by summing these.
37

CHAPTER 4. STERILE NEUTRINO OSCILLATIONS

Element

% by wt.

Atomic

Atomic Mass

Neutron

N. Density

Number

(g/mol)

Number

(1023 /cc)

47.2

16.00

8.00

4.49

Si

28.8

14

28.00

14

2.67

Al

7.96

13

26.98

13.98

0.81

Fe

4.32

26

55.84

29.84

0.82

Ca

3.85

20

40.08

20.08

0.35

Na

2.36

11

22.96

11.96

0.28

2.14

19

39.10

20.10

0.26

Mg

2.20

12

24.30

12.30

0.15

Total

9.60

Table 4.1: The table shows the computation of the neutron density of continental crust. Elements below 1% by weight not considered since the oscillation length estimated is important
only to the order of magnitude. Data regarding chemical compositions of Earths crust taken
from [73].
Using this to calculate A,

A = 2GF Enn

9.521023 cc1
2(1.17 105 GeV 2 )(1GeV ) (51000eV
1 /cm)3
4
2
= 1.19 10 eV
This gives a modifies mass difference:
p
2
2
2
=

p(cos2 A) + sin 2
=
(0.01 eV 2 1.000 + 1.19 104 eV 2 )2 + (0.01 eV 2 0.020)2
= 0.988 102 eV 2
And the baseline for the oscillation:
L'

4.2.1

4E
4 1 GeV

= 80 km

0.01 eV 2

Baseline at Resonance

4E
4 1 GeV
4E
=

= 2 1013 eV 1 4000 km
2 0.02

sin2
0.01
eV

However, attaining resonance requires a neutron density about 100 times that of the Earths
crust.
L(R) '

38

CHAPTER 4. STERILE NEUTRINO OSCILLATIONS

An alternative would be to increase the energy instead, so that the resonance neutron
density falls. But since a beam of 100 GeV energy does not seem pragmatic, we attempt to
optimise-tweaking beam energy so that a sizeable amplification is obtained without having to
attain resonance. The catch is that increasing energy increases the baseline length. As is

Figure 4.1: Plot of sin2 2 against Energy(in GeV) for propagation through
the Earth (generated in Mathematica)

Figure 4.2: Plot of oscillation length (in km) against Energy (in GeV) for
propagation through the Earth (generated in Mathematica)
evident from the figure, this leaves a very narrow window of high energies around 84 GeV to
obtain a reasonable degree of amplification. This is owing to the fact that the width of this
peak is proportional to sin2 2, which in this case is a very small quantity.
The increase in energy also sends up the oscillation length. The resonance baselines at
various values of m2 , E and are given in Tables 4.2.
Although our attempt via the MSW effect did not immediately yield a practical set of
parameters, the idea of using matter effects to enhance oscillations is by no means exhausted.
For instance, the parametric resonance effect[67], which is different from MSW, is believed to
play a role in oscillations in the presence of step-function like matter densities like the Earth
(the core, mantle and crust densities may be approximated by a step function). This hints at
future directions in which this work could be continued.

4.3

Outlook and Summary

The field of neutrino physics is at present a very active and developing one. There are
new experiments (like SNO+, OPERA and many others) being planned that promise exciting
39

CHAPTER 4. STERILE NEUTRINO OSCILLATIONS

aa

aa

m2 /E aaa (rad)
aa
(eV2 /GeV)
a

103
102
101
100
101
102
103

101

102

103

104

3980
398.0
39.8
3.98
0.398
0.0398
0.00398

39500.0
3950.0
395.0
39.5
3.95
0.395
0.0395

395000
39500.0
3950.0
395.0
39.5
3.95
0.395

3950000.0
395000.0
39500.0
3950.0
395.0
39.5
3.95

Table 4.2: The resonance baselines (km, rounded off to 3 s.f.) for various values of m2 /E
(eV2 /GeV) and mixing angle (rad) are tabulated at E=1 GeV
results in the years to come. Although the oscillation parameters have been measured, attempts
are still on to improve the accuracy of estimates. In addition, there are several open problems
like determining the sign of the atmospheric mass-squared difference, the magnitude of the CPviolation phase (these two are closely linked) and the question of whether neutrinos are Dirac
or Majorana particles. There is also the question of the existance of the sterile neutrino, (which
we attempted to tackle via the MSW effect) and the number of sterile neutrinos. Tackling these
questions will be the aim of the community in the immediate future.

40

Appendix A

The vacuum Hamiltonian in a more


illuminating form
In this Appendix, we shall take a look at the vacuum Hamiltonian for 2-flavour oscillations
and rewrite it in the form used in Eqn. 2.56. We start from Eqn.2.2,

 
 

d
E1 0
1 (t)
1 (t)
(A.1)
=
.

0 E2
2 (t)
dt 2 (t)
and shift to the flavour basis (the mixing matrix is
of the differential operator).
 

E1
(t)
d
.U
=
0
dt (t)

E1
Hvac = U
0

now a constant and can be taken to the left

0
E2

0
E2

(t)
(t)

Substituting for the mixing matrix (Eqn. 2.9),






E1 0
cos sin
cos sin
Hvac =
0 E2
sin cos
sin cos
Using the ultra-relativistic approximation and simplifying,





1
m221
1 0
cos2 sin2
Hvac =
(E1 + E2 )

0 1
sin2 cos2
2
2E

(A.2)
(A.3)

(A.4)

(A.5)

where m221 := m22 m21 . Suppressing the identity, we write (employing the notation in Section
2.3.1)


m21 + m22 m221
cos2 sin2
+
(A.6)
Hvac = E +
sin2 cos2
4E
4E
which is the required form.

41

Appendix B

Some Mathematica demos


In this appendix, we shall present the code for some Mathematica demos that shall illustrate
the concepts discussed about neutrino oscillations.

B.1

3-flavour Oscillations in Vacuum

(* This notebook demonstrates 3 flavour neutrino oscillations in \


vacuum *)
ClearAll["Global*"]
p = 3.1415;
e = 2.71828182845904523536028747135266249775;
X[\[Theta]23_] := {{1, 0, 0}, {0, Cos[\[Theta]23],
Sin[\[Theta]23]}, {0, -Sin[\[Theta]23], Cos[\[Theta]23]}}
Y[\[Theta]13_, \[Delta]_] := {{Cos[\[Theta]13], 0,
Sin[\[Theta]13]*Power[e, -I*\[Delta]]}, {0, 1,
0}, {-Sin[\[Theta]13]*Power[e, I*\[Delta]], 0, Cos[\[Theta]13]}}
Z[\[Theta]12_] := {{Cos[\[Theta]12], Sin[\[Theta]12],
0}, {-Sin[\[Theta]12], Cos[\[Theta]12], 0}, {0, 0, 1}}
U[\[Theta]23_, \[Theta]13_, \[Theta]12_, \[Delta]_] :=
X[\[Theta]23].Y[\[Theta]13, \[Delta]].Z[\[Theta]12]
V[\[Theta]23_, \[Theta]13_, \[Theta]12_, \[Delta]_] :=
Inverse[U[\[Theta]23, \[Theta]13, \[Theta]12, \[Delta]]]
(* Amplitude matrix *)
Q[\[Alpha]_, \[Beta]_,
L_, \[Theta]23_, \[Theta]13_, \[Theta]12_, \[Delta]_, m21_,
m31_, \[Epsilon]_] :=
U[\[Theta]23, \[Theta]13, \[Theta]12, \[Delta]][[\[Alpha],
1]]\[Conjugate]*
U[\[Theta]23, \[Theta]13, \[Theta]12, \[Delta]][[\[Beta], 1]]*
Power[e, -I*2.53*0.*Power[10, L - \[Epsilon]]] +

42

APPENDIX B. SOME MATHEMATICA DEMOS

U[\[Theta]23, \[Theta]13, \[Theta]12, \[Delta]][[\[Alpha],


2]]\[Conjugate]*
U[\[Theta]23, \[Theta]13, \[Theta]12, \[Delta]][[\[Beta], 2]]*
Power[e, -I*2.53*Power[10, m21 + L - \[Epsilon]]] +
U[\[Theta]23, \[Theta]13, \[Theta]12, \[Delta]][[\[Alpha],
3]]\[Conjugate]*
U[\[Theta]23, \[Theta]13, \[Theta]12, \[Delta]][[\[Beta], 3]]*
Power[e, -I*2.53*Power[10, m31 + L - \[Epsilon]]];
(* Energy in GeV *)
\[Epsilon]0 = 2;
(* Plot of survival and transition probability for electron neutrinos \
with distance (log scale in km) for vaious values of oscillation \
parameters *)
Manipulate[
Plot[{Abs[
Q[1, 1, L, \[Theta]23, \[Theta]13, \[Theta]12, \[Delta], m21,
m31, 2]]^2,
Abs[Q[1, 2, L, \[Theta]23, \[Theta]13, \[Theta]12, \[Delta], m21,
m31, 2]]^2,
Abs[Q[1, 3, L, \[Theta]23, \[Theta]13, \[Theta]12, \[Delta], m21,
m31, 2]]^2, (Abs[
Q[1, 1, L, \[Theta]23, \[Theta]13, \[Theta]12, \[Delta], m21,
m31, 2]]^2 +
Abs[Q[1, 2, L, \[Theta]23, \[Theta]13, \[Theta]12, \[Delta], m21,
m31, 2]]^2 +
Abs[Q[1, 3, L, \[Theta]23, \[Theta]13, \[Theta]12, \[Delta], m21,
m31, 2]]^2)}, {L, 0,
7}], {m21, -7, -1}, {m31, -7, -1}, {\[Theta]23,
0, \[Pi]/2}, {\[Theta]13, 0, \[Pi]/2}, {\[Theta]12,
0, \[Pi]/2}, {\[Delta], 0, 2 \[Pi]}]
(* Plot of survival and transition probability for electron neutrinos \
into mixtures of \[Mu] and \[Tau] *)
Manipulate[Plot[{
Abs[Q[1, 1, L, \[Theta]23, \[Theta]13, \[Theta]12, \[Delta], m21,
m31, 2]]^2,
Abs[Cos[\[Theta]]*
Q[1, 2, L, \[Theta]23, \[Theta]13, \[Theta]12, \[Delta], m21,
m31, 2] +
Sin[\[Theta]]*
Q[1, 3, L, \[Theta]23, \[Theta]13, \[Theta]12, \[Delta], m21,
m31, 2]]^2,
Abs[-Sin[\[Theta]]*
Q[1, 1, L, \[Theta]23, \[Theta]13, \[Theta]12, \[Delta], m21,
m31, 2] +
Cos[\[Theta]]*
43

APPENDIX B. SOME MATHEMATICA DEMOS

Q[1, 3, L, \[Theta]23, \[Theta]13, \[Theta]12, \[Delta], m21,


m31, 2]]^2, (Abs[
Q[1, 1, L, \[Theta]23, \[Theta]13, \[Theta]12, \[Delta], m21,
m31, 2]]^2 +
Abs[Cos[\[Theta]]*
Q[1, 2, L, \[Theta]23, \[Theta]13, \[Theta]12, \[Delta], m21,
m31, 2] +
Sin[\[Theta]]*
Q[1, 3, L, \[Theta]23, \[Theta]13, \[Theta]12, \[Delta], m21,
m31, 2]]^2 +
Abs[-Sin[\[Theta]]*
Q[1, 2, L, \[Theta]23, \[Theta]13, \[Theta]12, \[Delta], m21,
m31, 2] +
Cos[\[Theta]]*
Q[1, 3, L, \[Theta]23, \[Theta]13, \[Theta]12, \[Delta], m21,
m31, 2]]^2)
}, {L, 0, 3*10^5}], {m21, -7, -1}, {m31, -7, -1}, {\[Theta]23,
0, \[Pi]/2}, {\[Theta]13, 0, \[Pi]/2}, {\[Theta]12,
0, \[Pi]/2}, {\[Delta], 0, 2 \[Pi]}, {\[Theta], 0, \[Pi]/2}]

B.2

Solar Neutrinos

In this section, we shall present code that pertains to the toy version of the solar neutrino
problem we considered in Section 2.3.4.

B.2.1

Survival Probability of e

ClearAll["Global*"];
(* Survival Probability *)
GF = 1.17*10^(-5 - 18); (* eV^-2 *)
n0 = 98.8 * 6.023*10^(23) / 51000^3 ; (* eV^3 *)
R0 = 7*10^(10) *51000 ; (* eV^-1 *)
X[x_, \[Theta]_] :=
Exp[-\[Pi]/4*x*Sin[2 \[Theta]]^2/Cos[2 \[Theta]]*R0/10];
C0[x_, \[Theta]_] := (x*Cos[2 \[Theta]] - 2*1.414*GF*n0 )/
Sqrt[(x*Cos[2 \[Theta]] - 2*1.414*GF*n0 )^2 + (x*
Sin[2 \[Theta]])^2];
P[x_, \[Theta]_] :=
0.5*(1 + (1 - 2 X[x, \[Theta]]) C0[x, \[Theta]]*Cos[2 \[Theta]])
Manipulate[
Plot[P[Power[10, x], \[Theta]], {x, (-15), (-9)},
44

APPENDIX B. SOME MATHEMATICA DEMOS

PlotRange -> {0, 1}], {\[Theta], 0.0001, \[Pi]/4 - 0.0001}]

B.2.2

Density profile of the Sun and sin2

n[z_] := n0/98.8*48.8* Exp[-11.1*z^2 / (z + 0.2)]


S[\[Theta]_, x_, z_] :=
x^2*Sin[2 \[Theta]]^2/(x^2*
Sin[2 \[Theta]]^2 + ( x*Cos[2 \[Theta]] - 2*1.414*GF*n[z] )^2)
Plot[n[z]/n0, {z, 0, 1}, PlotRange -> {0, 1}]
Manipulate[
Plot[S[0.1, Power[10, x], z], {z, 0, 1},
PlotRange -> {0, 1}], {x, -16, 2}]

B.2.3

The Adiabaticity parameter

K[z_] := 1/R0 * 48.8/98.8 *


n0 * ( -2*z*11.1 / (z + 0.2) + z^2 *11.1 / (z + 0.2)^2 )*
Exp[-11.1*z^2 / (z + 0.2)];
\[Gamma][x_, z_, \[Theta]_] :=
x^2/(2*1.414*GF)*
Sin[2 \[Theta]]^2/(Sqrt[S[\[Theta], x, z]]^(3)) /Abs[K[z]]
Plot[K[z], {z, 0, 1}]
Plot[\[Gamma][10^(-14), z, 0.1], {z, 0, 1}, PlotRange -> {0, 30}]

B.2.4

The Exclusion Plot

f[\[Zeta]_] := 0.5*ArcCos[( -\[Zeta] + Sqrt[\[Zeta]^2 + 4] )/2 ];


PP[x_, \[Zeta]_] :=
0.5*(1 + (1 - 2 X[x, f[\[Zeta]]]) C0[x, f[\[Zeta]]]*Cos[2*f[\[Zeta]]])
Manipulate[
Plot[PP[Power[10, x], Power[10, \[Zeta]]], {x, -18, -9},
PlotRange -> {0, 1}], {\[Zeta], -4, 7}]
ContourPlot[
PP[Power[10, x]/(5*10^6), Power[10, \[Zeta]]], {\[Zeta], -4,
1}, {x, -8, -3}, Contours -> {0.19, 0.35}]

45

Bibliography
[1] S.M. Bilenky and B. Pontecorvo, Phys. Rep. 41, 225 (1978).
[2] B. Kayser, Phys. Rev. D 24, 110 (1981).
[3] C. Giunti, C.W. Kim and U.W. Lee, Phys. Rev. D 44, 3635 (1991).
[4] C. Giunti, C.W. Kim and U.W. Lee, Phys. Rev. D 45, 2414 (1992).
[5] C. Giunti, C.W. Kim, J.A. Lee and U.W. Lee, Phys. Rev. D 48, 4310
[6] J. Rich, Phys. Rev. D 48, 4318 (1993).
[7] K. Kiers, S. Nussinov and N. Weiss, Phys. Rev. D 53, 537 (1996).
[8] W. Grimus and P. Stockinger, Phys. Rev. D 54, 3414 (1996).
[9] K.A. Olive et al. (Particle Data Group), Chin. Phys. C, 38, 090001 (2014)
[10] S.M. Bilenky, C. Giunti and W. Grimus, Prog. Part. Nucl. Phys. 43:1-86 (1999), arXiv:hepph/9812360v4
[11] Palash B. Pal, Int. J. Mod. Phys. A, 07, 5387 (1992)
[12] L. Wolfenstein, Phys. Rev. D 17, 2369 (1978)
[13] S. P. Mikheyev, A. Yu. Smirnov, Nuovo Cimento, C9 (1986)J.N. Bahcall, A.M. Serenelli,
and S. Basu, Astrophys. J. 621, L85 (2005)
[14] V. Barger, N. Deshpande, P. B. Pal, R. J. N. Phillips, and K. Whisnant, Phys. Rev. D 43,
R1759(R) (1991)
[15] S.T. Petcov, Phys. Lett. B200, 373 (1988)
[16] The SNO website (http://www.sno.phy.queensu.ca/)
[17] B. Aharmim et al., [SNO Collab.], Phys. Rev. C88, 025501 (2013)
[18] The official Super K website (http://www-sk.icrr.u-tokyo.ac.jp/sk/index-e.html)
[19] K. Abe, et al., [Super-Kamiokande Collab.], Phys. Rev. D 91, 052019 (2015)
[20] Y. Ashie, et al., [Super-Kamiokande Collab.] Phys. Rev. Lett. 93:101801 (2004)
[21] K.S. Hirata et al., [Kamiokande Collab.], Phys. Rev. Lett. 63,16 (1989)
[22] G.L. Fogli, E. Lisi, D. Montanino, Astropart. Phys. 9:119-130 (1998)
46

BIBLIOGRAPHY

[23] John Bahcalls home-page (http://www.sns.ias.edu/ jnb/)


[24] J.N. Bahcall, A.M. Serenelli, and S. Basu, Astrophys. J. 621, L85 (2005)
[25] J.N. Bahcall, A.M. Serenelli, and S. Basu, Astrophys. J. Supp. 165, 400 (2006)
[26] C. Pena-Garay and A.M. Serenelli, arXiv:0811.2424 (2008)
[27] B. Pontecorvo, Chalk River Lab. report PD-205 (1946)
[28] B.T. Cleveland et al., Astrophys. J. 496, 505 (1988)
[29] W. Hampel et al., [GALLEX Collab.], Phys. Lett. B447, 127 (1999)
[30] M. Altmann et al., [GNO Collab.], Phys. Lett. B616, 174 (2005)
[31] F. Kaether et al., Phys. Lett. B685, 47 (2010)
[32] J.N. Abdurashitov et al., [SAGE Collab.], Phys. Rev. C80, 015807 (2009)
[33] Y. Fukuda et al., [Kamiokande Collab.], Phys. Rev. Lett. 77, 1683 (1996)
[34] K. Abe et al., [Super-Kamiokande Collab.], Phys. Rev. D83, 052010 (2011)
[35] G. Bellini et al., [Borexino Collab.], Phys. Rev. D82, 033006 (2010)
[36] B. Aharmim et al., [SNO Collab.], Phys. Rev. Lett. 101, 111301 (2008); Phys. Rev. C87,
015502 (2013)
[37] A.M. Serenelli, W.C. Haxton, and C. Pena-Garay, Astrophys. J. 743, 24 (2011)
[38] Y. Fukuda et al., [Super-Kamiokande Collab.], Phys. Rev. Lett. 86, 5651 (2001)
[39] Q.R. Ahmad et al., [SNO Collab.], Phys. Rev. Lett. 87, 071301 (2001)
[40] K. Eguchi et al., [KamLAND Collab.], Phys. Rev. Lett. 90, 021802 (2003)
[41] G. L. Fogli et al., Phys. Rev. D67, 073002 (2003)
[42] M. Maltoni, T. Schwetz, and J.W. Valle, Phys. Rev. D67, 093003 (2003)
[43] A. Bandyopadhyay et al., Phys. Lett. B559, 121 (2003)
[44] J.N.Bahcall, M.C. Gonzalez-Garcia, and C. Pena-Garay, JHEP 0302, 009 (2003)
[45] P.C. de Holanda and A.Y. Smirnov, JCAP 0302, 001 (2003)
[46] A. Gando et al., [KamLAND Collab.], Phys. Rev. D83, 052002 (2011)
[47] Y. Fukuda et al. [Super-Kamiokande Collab.], Phys. Rev. Lett. 81, 1562 (1998)
[48] V. Barger et al., Phys. Rev. Lett. 82, 2640 (1999)
[49] E. Lisi et al., Phys. Rev. Lett. 85, 1166 (2000)
[50] K. Abe et. al. [Super-Kamiokande Collab.], Phys. Rev. Lett. 97:171801 (2006)
[51] K. Abe et. al. [Super-Kamiokande Collab.], Phys. Rev. Lett. 110, 181802 (2013)
47

BIBLIOGRAPHY

[52] M. H. Ahn, et. al. [K2K Collab.], Phys. Rev. D74: 072003 (2006)
[53] P. Adamson et. al. [MINOS Collab.], Phys. Rev. Lett. 106:181801 (2011)
[54] P. Adamson et. al. [MINOS Collab.], Phys. Rev. Lett. 112, 191801 (2014)
[55] K. Abe et. al. [T2K Collab.], Phys. Rev. Lett. 111, 211803 (2013)
[56] K. Abe et. al. [T2K Collab.], Phys. Rev. Lett. 112, 181801 (2014)
[57] W.R. Webber and J.A. Lezniak, Astrophys. Space Sci. 30, 361 (1974).
[58] E.-S. Seo et al., Proc. of the 22nd International Cosmic Ray Conference, Dublin, Ireland,
edited by M. Cawley et. al., Vol. 2, p. 627, Reprint Ltd., Dublin, Ireland (1991)
[59] M. Apollonio et al., [Chooz Collab.], Phys. Lett. B466, 415 (1999)
[60] Y. Abe et al., [Double Chooz Collab.], Phys. Rev. D86, 052008 (2012)
[61] F.P. An et al., [Daya Bay Collab.], Chinese Phys. C37, 011001 (2013)
[62] F.P. An et al., [Daya Bay Collab.], Phys. Rev. Lett. 112, 061801 (2014)
[63] J.K. Ahn et al., [RENO Collab.], Phys. Rev. Lett. 108, 191802 (2012)
[64] S.-H. Seo [for the RENO Collab.], talk at the TAUP2013 International Workshop, September 9-13, 2013, Asilomar, California, USA
[65] K. Abe et al., [T2K Collab.], Phys. Rev. Lett. 107, 041801 (2011)
[66] K. Abe et al., [T2K Collab.], Phys. Rev. Lett. 112, 061802 (2014)
[67] M. V. Chizhov and S. T. Petcov, Phys. Rev. Lett. 83, 1096 (1999)
[68] K. Abe et al., [T2K Collab.], Phys. Rev. D85, 031103 (R) (2012)
[69] Y. Itow, Nucl. Phys. (Proc. Supp.) B235-236, 79 (2013)
[70] P. Adamson et al., [MINOS Collab.], Phys. Rev. Lett. 110, 171801 (2013)
[71] [Borexino Collab.], Nature 512, 383386 (2014)
[72] G. Bellini et al., [Borexino Collab.], Phys. Rev. Lett. 107, 141302 (2011).
[73] Wedepohl, K. H. The composition of the continental crust. Geochimica et cosmochimica
Acta, 59(7), 1217-1232 (1995)
[74] http://hitoshi.berkeley.edu/neutrino

48

You might also like