1. 3 4 5 6 7 8 9
      2. -9 10
      3. -8 10
      4. -7 10
      5. -6 10
      6. -5 10
      7. -4 10

Measurements of atmospheric muons using AMANDA with
emphasis on the Prompt component
by
Raghunath Ganugapati
A dissertation submitted in partial fulfillment of the
requirements for the degree of
Doctor of Philosophy
(Physics)
at the
University of Wisconsin – Madison
2008

?c
Copyright by Raghunath Ganugapati 2008
All Rights Reserved

Measurements of atmospheric muons using
AMANDA with emphasis on the Prompt
component
Raghunath Ganugapati
Under the supervision of Professor Albrecht Karle
At the University of Wisconsin — Madison
The main aim of AMANDA is to detect diffuse extra-terrestrial neutrinos. While at-
mospheric muons can be easily filtered out atmospheric neutrinos are an irreducible
back-ground for diffuse extra-terrestrial neutrino fluxes. At GeV energies the at-
mospheric neutrino fluxes are dominated by conventional neutrinos. However with
increasing energy, the harder “prompt” neutrinos that arise through semi-leptonic
decays of hadrons containing heavy quarks, most notably charm, become dominant.
Estimates of the magnitude of the prompt atmospheric fluxes differ by almost two
orders of magnitude making the significance of evaluating their intensity very impor-
tant. The main principle in this thesis is that it is possible to overcome the theoretical
uncertainty in the magnitude of the prompt neutrino fluxes by deriving their intensity
from a measurement of the down-going prompt muon flux. An attempt to constrain
this flux using this principle was made and analysis of the down-going muon data was
performed to constrain the RPQM model of prompt muons by a factor of 3.67 under
a strict set of simplifying assumptions.
Albrecht Karle (Adviser)

i
Acknowledgements
I consider myself immensely fortunate to have pursued my PhD. in physics at the
University of Wisconsin-Madison. I feel an immense sense of accomplishment in this
intense effort and extended study to attain a PhD, the highest academic degree anyone
can earn. To see a high school knack for solving physics and mathematics problems
culminate in this great accomplishment is something special. UW-Madison took ex-
ceptionally good care of me, so much so, I felt it was a“home away from home”. Closer
to home, I would like to thank the Wisconsin Icecube collaboration.
First and foremost, I would like to thank my advisor, Professor Albrecht Karle,
for taking me on as a student a few years ago, even though he knew little about
me at the time except for the fact that I was that graduate student in engineering
at UW-Madison. I would like to thank my mentor the late Bruce Koci and my co-
supervisor Professor Bob Morse for help with the IceCube drill modeling work that I
did which has been a huge success for the project. I would also like to thank Professor
Francis Halzen, Principal Investigator for IceCube who is my role model for modesty
and Professor Teresa Montaruli for her stint at advising me. In addition, I would also
like to thank Professor David Brown of the Finance department of UW-Madison for
advising me on my PhD minor course work in Mathematical Finance and on my stint
on Wall St as a “Rocket Scientist”.

ii
I would like to thank all the members of the AMANDA/IceCube collaboration
who have interacted with me during the course of this thesis and all the physicists
I interacted with at conferences. Notable mentions are Gary Hill, for valuable the-
sis advice and suggesting that cooking at home was the only way of getting rich in
graduate school! Paolo Desiati for convincing me that the idea to investigate prompt
muons for a PhD thesis not to be such a horrible one. Mark Krasberg for all the non-
physics talk in low times and being one of my close tennis buddies on campus. My
fellow graduate students John Kelley (doesn’t need a reason to be cheerful!), Jessica
Hodges, Jim Braun and Karen Andeen. I would also like to thank Mike Stamatikos,
Jodi Cooley, Katherine Rawlins, David Steele and Rellen Hardtke from the older class
of graduated students.
With all my heart I thank my parents Lakshmi and Haranath Ganugapati for
their encouragement and instilling a knife-edged competitive spirit to prepare me
and my brother to gain an admission into the Indian Institute of Technologies and
subsequent higher education in the United States. Incidentally my brother is my role
model for mathematics, the physics counterpart.
Thanks to Karle’s lab and each one of you all for making this possible!

iii
Contents
Acknowledgements
i
1 Introduction
1
2 Strategies to optimize the hotwater drilling method for IceCube
4
2.1 Description of the thermal process . . . . . . . . . . . . . . . . . . . . . 7
2.2 Methodology and Assumptions . . . . . . . . . . . . . . . . . . . . . . 11
2.3 Description of the model and assumptions . . . . . . . . . . . . . . . . 12
2.4 Optimal drilling algorithm . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.5 Results. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.6 Robustness of Predictions . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.7 Conclusions and Summary . . . . . . . . . . . . . . . . . . . . . . . . . 22
3 Measuring the Prompt Atmospheric Neutrino Flux with Downgoing
Muons in AMANDA-II
24
3.1 AMANDADetector. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.2 ConventionalandPromptAtmosphericNeutrinos . . . . . . . . . . . . 26
3.3 Constraining the Prompt Neutrino Flux with the Downgoing Muon Flux 28

iv
3.4 Prompt Atmospheric Neutrino Models . . . . . . . . . . . . . . . . . . 29
3.5 CharminCORSIKA . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
4 Data streams and quality cuts on the 2005 Sample
36
4.1 Firstguessreconstructions,livetimeandtriggers . . . . . . . . . . . . . 36
4.1.1 Direct Walk Reconstruction . . . . . . . . . . . . . . . . . . . . 36
4.1.2 JAMS Reconstruction . . . . . . . . . . . . . . . . . . . . . . . 36
4.1.3 Highqualitystream. . . . . . . . . . . . . . . . . . . . . . . . . 37
4.1.4 Minimum biasstream . . . . . . . . . . . . . . . . . . . . . . . 37
4.2 Reconstruction Methods . . . . . . . . . . . . . . . . . . . . . . . . . . 38
4.3 Techniques to Further Improve Background Rejection . . . . . . . . . . 38
4.4 Event Simulation and Reweighting . . . . . . . . . . . . . . . . . . . . 38
4.4.1 Preparation ofSimulated Events. . . . . . . . . . . . . . . . . . 39
5 Response of AMANDA-II to Cosmic Ray Muons
40
5.1 Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
5.2 Results. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
6 Model dependencies and systematic error calculations for a down-
going muon analysis
55
6.1 StatisticalErrors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
6.2 Systematic Uncertainties . . . . . . . . . . . . . . . . . . . . . . . . . . 56
6.2.1 Normalization of Cosmic Ray Flux . . . . . . . . . . . . . . . . 56
6.2.2 Spectral Index of Cosmic Ray Spectrum . . . . . . . . . . . . . 57
6.2.3 Detector Sensitivity . . . . . . . . . . . . . . . . . . . . . . . . . 57

v
6.2.4 InteractionModelUncertainity . . . . . . . . . . . . . . . . . . 57
6.2.5 Ice Model Uncertainty . . . . . . . . . . . . . . . . . . . . . . . 58
6.2.6 Other Source ofErrors . . . . . . . . . . . . . . . . . . . . . . . 58
6.3 Result of Systematics Study . . . . . . . . . . . . . . . . . . . . . . . . 58
7 Hadronic Interaction Models and Extended Air showers
63
7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
7.2 Interaction and Extended Air Shower models . . . . . . . . . . . . . . . 65
7.2.1 Available Codes and Model Comparisons . . . . . . . . . . . . . 65
7.2.2 CrossSections. . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
7.2.3 Particle Production . . . . . . . . . . . . . . . . . . . . . . . . . 67
7.2.4 Impact of shower simulations . . . . . . . . . . . . . . . . . . . 68
7.3 Results. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
7.3.1 InteractionModel . . . . . . . . . . . . . . . . . . . . . . . . . . 69
7.3.2 Extended Air Shower . . . . . . . . . . . . . . . . . . . . . . . . 70
7.3.2.1 lateral distribution function . . . . . . . . . . . . . . . 70
7.3.2.2 Zenith Angle and Energy Spectra . . . . . . . . . . . . 72
8 Results and Conclusions
84
8.1 ShapeAnalysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
8.2 Simulation and Fitting Procedure . . . . . . . . . . . . . . . . . . . . . 85
8.3 FittingProcedure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
8.4 Prompt Atmospheric Neutrino Upper limits . . . . . . . . . . . . . . . 87
8.5 Discussion for Better Analysis in Future . . . . . . . . . . . . . . . . . 87
8.6 Conclusion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88

vi
List of Tables
2.1 The input parameters that go into the old AMANDA drill and the New
IceCube drill are compared. . . . . . . . . . . . . . . . . . . . . . . . . 20
2.2 The outputs for the optimum strategy and consequently the fuel con-
sumption are quoted by varying the thermal conductivity of the hose. . 20
2.3 The changes in the optimum strategy and consequently the fuel con-
sumptionarestudied byvarying thedeployment time . . . . . . . . . . 21
2.4 The changes in the optimum strategy and consequently the fuel con-
sumption are studied by cutting the power available at the surface . . 21
2.5 The ouput parameters for optimum strategy and consequently the fuel
consumption are quoted by varying the desired target diameter. . . . . 22
3.1 Critical energy for different particles. . . . . . . . . . . . . . . . . . . . 28
5.1 Presents the statistics of zenith angle resolution after quality cuts for
various zenith ranges. Values in brackets are before quality cuts for the
QGSJETmodel. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

vii
5.2 Presents the statistics of space angle resolution after quality cuts for
various zenith angle ranges. Values in brackets are before quality cuts
fortheQGSJETmodel. . . . . . . . . . . . . . . . . . . . . . . . . . . 51
6.1 Average simulation uncertainties for different sources of errors. . . . . . 59

viii
List of Figures
2.1 The figure describes the schematic view of the IceCube Enhanced Hot
Water Drill (EHWD) at the surface.. . . . . . . . . . . . . . . . . . . . 8
2.2 DepthdependnceofthetemperatureofSouthPoleIce. . . . . . . . . . 9
2.3 Schematicviewofthehotwaterdrillingmethod. . . . . . . . . . . . . . 10
2.4 The figure illustrates the heat transfer procedure that asymptotically
approaches the far-field ice temperature. . . . . . . . . . . . . . . . . . 14
2.5 The hole diameter as a function of time for a range of depths. The drill
strategy delivers a hole of uniform diameter at a required time of 30
hours after the drilling is completed. . . . . . . . . . . . . . . . . . . . 17
2.6 Thisfiguresummarizestheevolutionoftheholesize. . . . . . . . . . . 18
2.7 This figure summarizes the evolution of the hole diameter as a function
of time. The drill strategy delivers a hole of uniform diameter at a
required time of 30 hours after the drilling is completed. . . . . . . . . 19
3.1 The figure shows the layout of the AMANDA detector. The top view
shows 19 strings that were deployed. AMANDA detector is roughly
200mwideand500mlong . . . . . . . . . . . . . . . . . . . . . . . . . 25

ix
3.2 The figure schemtically shows the interaction of the primary cosmic ray
proton with the atmosphere and the formation of several particles as
the shower evolves. (Image credit: Milagro) . . . . . . . . . . . . . . . 27
3.3 Prompt atmospheric neutrinos are predicted to follow a harder spec-
trum than conventional atmospheric neutrinos. The flux of prompt
atmospheric neutrinos is highly uncertain and predictions range over
several orders of magnitude. Image Credit: Jessica Hodges . . . . . . . 30
3.4 The distribution of lateral separation from shower core for the DPMJET-
II for charm and coventional muons in each event with the first inter-
action and multiple interactions isolated. . . . . . . . . . . . . . . . . . 33
3.5 The total energy distribution for the DPMJET-II for charm and con-
ventional muons for each event with first interaction and multiple in-
teractionsisolated. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
5.1 The angular distribution of atmospheric muons in AMANDA-II at a
depth of 1730m using the MAM ice model with the SYBILL interaction
model and the 2001 experimental data. . . . . . . . . . . . . . . . . . . 45
5.2 The depth-intensity of atmospheric muons in AMANDA-II using the
MAM ice model and the SYBILL interaction model with the 2001 ex-
perimentaldata.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
5.3 The relative difference between MAM SYBILL Monte Carlo and AMANDA-
II 2001 data as a function of depth. . . . . . . . . . . . . . . . . . . . . 47
5.4 The relative difference between MAM SYBILL Monte Carlo and AMANDA-
IIexperimentaldataasafunctionofzenithangle. . . . . . . . . . . . . 47

x
5.5 The angular distribution of atmospheric muons in AMANDA-II at a
depth of 1730m using the Millenium ice model with the SYBILL inter-
action model and the 2005 experimental data. . . . . . . . . . . . . . . 48
5.6 The depth-intensity of atmospheric muons in AMANDA-II using the
Millenium ice model and the SYBILL interaction model with the 2005
experimentaldata. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
5.7 The relative difference between Millenium SYBILL Monte Carlo and
AMANDA-II 2005 experimental data as a function of depth. . . . . . . 49
5.8 The relative difference between Millenium SYBILL Monte Carlo and
AMANDA-II 2005 experimental data as a function of zenith angle. . . 49
5.9 The zenith angle difference between the reconstructed and true zenith
angle known from simulation is plotted on x-axis while normalized
counts are plotted on y-axis. The respective slices in zenith are indi-
cated in the plot. Red is before quality cuts while blue is after quality
cuts. From left to right and top to bottom there are 10 slices shown
that go from0.0-0.5 in increments of 0.05. . . . . . . . . . . . . . . . . 52
5.10 The zenith angle difference between the reconstructed and true zenith
angle known from simulation is plotted on x-axis while normalized
counts are plotted on y-axis. The respective slices in zenith are indi-
cated in the plot. Red is before quality cuts while blue is after quality
cuts. From left to right and top to bottom there are 10 slices shown
that go from0.5-1.0 in increments of 0.05. . . . . . . . . . . . . . . . . 53

xi
5.11 Comparison of CORSIKA vertical muon flux for various interaction
models. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
6.1 The N
ch
variation for the DPMJET-II for signal and background (at the
final level after event selection criteria are implemented) when spectral
index is varied by
±0.02
shown as a ratio. . . . . . . . . . . . . . . . . 60
6.2 The N
ch
distribution for the DPMJET-II for signal and background
(at the final level after event selection criteria are implemented) when
spectralindexisvariedby±0.02shownasaratio.. . . . . . . . . . . . 60
6.3 The N
ch
variation for the DPMJET-II for background (at the final level
after event selection criteria are implemented) when compared with an
equally weighted simulation of SYBILL and DPMJET-II is shown. . . . 61
6.4 The N
ch
distribution for the DPMJET-II for background (at the final
level after event selection criteria are implemented) when compared
with an equally weighted simulation of SYBILL and DPMJET-II is
shown. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
6.5 The N
ch
variation for the DPMJET-II millenium ice model (at the
final level after event selection criteria are implemented) when com-
pared with an equally weighted simulation of DPMJET-II millenium
and DPMJET-II AHA model is shown. . . . . . . . . . . . . . . . . . . 62
6.6 The N
ch
distribution for the DPMJET-II millenium ice model (at the
final level after event selection criteria are implemented) when com-
pared with an equally weighted simulation of DPMJET-II millenium
and DPMJET-II AHA model is shown. . . . . . . . . . . . . . . . . . . 62

xii
7.1 Energy fraction distributions using various models for charmed baryon
and mesons for energies of 10, 10
2
, 10
3
, and 10
4
TeV . . . . . . . . . . 73
7.2 The mean multiplicity and the Z-moments of pions and kaons as a
function of primary energy. The top ensemble of points denote pions
while the bottom ones denote kaons . . . . . . . . . . . . . . . . . . . . 74
7.3 The trasverse momentum, longitudinal momentum and lateral separa-
tion of the secondary particles produced by air showers for a 1 PeV
monoenergetic beam of primary protons at a fixed zenith angle of 65
degrees. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
7.4 Shows the average number of muons produced per event as a function
the lateral separation from the shower core at surface of earth for show-
ers initiated by the full cosmic ray spectrum, full cosmic ray spectrum
for zenith>80 degrees, for primaries in the energy range of 1-1000 PeV
and monoenergetic primary energy of 1 PeV with no showering (only
the first interaction) and after the full shower develops (multiple inter-
actions) with events containing atleast 1 prompt muon (produced from
a charmed particle) tagged as “PROMPTS” and for no prompt muon
involved as “CONV”. All data has been normalized to 1 years worth
lifetime. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75

xiii
7.5 Shows the average number of muons produced per event as a function
the lateral separation from the shower core at surface of earth for show-
ers initiated by the full cosmic ray spectrum, full cosmic ray spectrum
for zenith>80 degrees, for primaries in the energy range of 1-1000 PeV
and monoenergetic primary energy of 1 PeV after the full shower devel-
ops (multiple interactions) with showers produced by protons and iron
identified separately. All data has been normalized to 1 years worth
lifetime. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
7.6 Shows the average number of muons produced per event as a func-
tion the lateral separation from the most energetic muon at surface of
earth for showers initiated by the full cosmic ray spectrum, full cosmic
ray spectrum for zenith>80 degrees, for primaries in the energy range
of 1-1000 PeV and monoenergetic primary energy of 1 PeV with no
showering (only the first interaction) and after the full shower develops
(multiple interactions) withevents containing atleast 1 prompt muon
(produced from a charmed particle) tagged as “PROMPTS” and for no
prompt muon involved as “CONV”. All data is normalized to 1 years
worthlifetime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77

xiv
7.7 Shows the average number of muons produced per event as a function
the lateral separation from the most energetic muon at surface of earth
for showers initiated by the full cosmic ray spectrum, full cosmic ray
spectrum for zenith>80 degrees, for primaries in the energy range of
1-1000 PeV and monoenergetic primary energy of 1 PeV after the full
shower develops (multiple interactions) with showers produced by pro-
tons and iron identified separately. All data is normalized to 1 years
worthlifetime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
7.8 Shows the average number of muons produced per event as a function
the lateral separation from the most energetic muon at detector for
showers initiated by the full cosmic ray spectrum, full cosmic ray spec-
trum for zenith>80 degrees, for primaries in the energy range of 1-1000
PeV and monoenergetic primary energy of 1 PeV with no showering
(only the first interaction) and after the full shower develops (multiple
interactions) with events containing atleast 1 prompt muon (produced
from a charmed particle) tagged as “PROMPTS” and for no prompt
muon involved as “CONV”. All data is normalized to 1 years worth
lifetime. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

xv
7.9 Shows the average number of muons produced per event as a function
the lateral separation from the most energetic muon at detector for
showers initiated by the full cosmic ray spectrum, full cosmic ray spec-
trum for zenith>80 degrees, for primaries in the energy range of 1-1000
PeV and monoenergetic primary energy of 1 PeV after the full shower
develops (multiple interactions) with showers produced by protons and
iron identified separately. All data is normalized to 1 years worth lifetime 80
7.10 Shows the sum total of surface energy of all the muons in an event for
showers initiated by the full cosmic ray spectrum, full cosmic ray spec-
trum for zenith>80 degrees, for primaries in the energy range of 1-1000
PeV and monoenergetic primary energy of 1 PeV with no showering
(only the first interaction) and after the full shower develops (multiple
interactions) with events containing atleast 1 prompt muon (produced
from a charmed particle) tagged as “PROMPTS” and for no prompt
muon involved as “CONV”. All data is normalized to 1 years worth
lifetime. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

xvi
7.11 Shows the sum total of energy of all the muons in an event at the
detector for showers initiated by the full cosmic ray spectrum, full cos-
mic ray spectrum for zenith>80 degrees, for primaries in the energy
range of 1-1000 PeV and monoenergetic primary energy of 1 PeV with
no showering (only the first interaction) after the full shower develops
(multiple interactions) with events containing atleast one prompt muon
(produced from a charmed particle) tagged as “PROMPTS” and for no
prompt muon involved as “CONV”. All data is normalized to 1 years
worthlifetime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
7.12 Shows the zenith angle distribution of showers initiated by the full cos-
mic ray spectrum, full cosmic ray spectrum for zenith>80 degrees, for
primaries in the energy range of 1-1000 PeV and monoenergetic primary
energy of 1 PeV with no showering (only the first interaction) and after
the full shower develops (multiple interactions) with events containing
atleast 1 prompt muon (produced from a charmed particle) tagged as
“PROMPTS” and for no prompt muon involved as “CONV”. All data
is normalized to 1 years worth lifetime . . . . . . . . . . . . . . . . . . 83
8.1 The minimized value of chisquare is shown for different levels of signal. 89
8.2 The elliptical contours of chisquare for the fraction of Millenium and
AHA backgrounds are shown forcing the signal contribution to be zero
while making a fit to the data. . . . . . . . . . . . . . . . . . . . . . . . 89

xvii
8.3 The elliptical contours of chisquare for the fraction of Millenium and
AHA backgrounds are shown for best fit value of signal while making a
fittothedata.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
8.4 The elliptical contours of chisquare for the fraction of Millenium and
AHA backgrounds are shown for the allowed level of signal at 90%
confidence level while making a fit to the data. . . . . . . . . . . . . . . 90
8.5 The signal and background spectra for the AHA and Millenium models
together with the minimum bias experimental data before fitting are
shown. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
8.6 The scaled levels at the best fit values of signal and background spec-
tra for the AHA and Millenium models are shown together with the
minimum bias experimental data. . . . . . . . . . . . . . . . . . . . . . 91

1
Chapter 1
Introduction
The Antarctic Muon and Neutrino Detector Array (AMANDA) is designed to detect
high energy neutrinos using the three kilometer thick ice cap covering the South Pole.
AMANDA in its design consists of a large array of phototubes located under the ice.
This array of phototubes embedded in the icecap at depths of 1500 to 2000m captures
the Cherenkov radiation from the ultra relativistic charged leptons that are produced
when neutrinos undergo charged current interactions with nucleons in the ice.
The main aim of AMANDA is to detect extra-terrestrial neutrinos. The back-
ground to the observation of these neutrinos is the flux of atmospheric muons and
neutrinos produced in cosmic ray showers in the atmosphere. Atmospheric muons can
reach the detector only from above, because the range of muons in earth is only a
few kilometers. Atmospheric muons are therefore only downgoing and these can be
easily filtered out by using the earth as a filter and looking at upward-moving neutri-
nos produced in the northern hemisphere. Atmospheric neutrinos can instead reach
the detector from all directions. Hence they are an irreducible background for diffuse
astrophysical neutrino fluxes. It is very important to evaluate their intensity with
reasonable accuracy.

2
At GeV energies the atmospheric fluxes are dominated by the decays of relatively
long-lived particles such as π
±
and K
±
mesons. With increasing energy, the probability
increases that such particles interact in the atmosphere before decaying. This implies
that even a small fraction of short-lived charmed particles can give the dominant
contribution to high energy muon and neutrino fluxes. These “prompt” muons and
neutrinos arise through semi-leptonic decays of hadrons containing heavy quarks, most
notably charm. Estimates of the magnitude of the prompt atmospheric fluxes differ by
almost two orders of magnitude making the significance of evaluating their intensity
very important.
The main principle in this thesis is that it is possible to overcome the theoretical
uncertainty in the magnitude of the prompt neutrino fluxes by deriving their intensity
from a measurement of the down-going prompt muon flux. The suggestion is based
on the observation that due to the charmed particle decay kinematics for the semi-
leptonic decays into neutrino and muon fluxes, the prompt muon and neutrino flux
are essentially the same at sea level. Importantly, this result is independent of the
charm production model. It should be stressed that down-going prompt muons and
not up-going neutrino induced muons are used to get limits on the prompt neutrino
flux. Prompt muons are easy to detect and there are ways of separating them from
the conventional muons using different zenith angle and energy spectral shapes.
This analysis of the atmospheric charm component is challenging for the fact that
there are no robust simulations for producing atmospheric prompt muons. Further,
the limited angular and energy resolutions of AMANDA combined with the lack of
availabilty of a model makes life very hard for a researcher.

3
Studies were also done on the introduction of theoretical models used to calculate
the heat transfer and refreezing rates in boreholes in cold ice at the south pole and
a comparison of these results with experimental data from the AMANDA holes. The
calculations are based on models derived for phase change with a moving boundary
layer in cylindrical coordinates. This work improved estimates of fuel consumption
and contributed to a better efficiency in the drilling of holes for project IceCube.

4
Chapter 2
Strategies to optimize the hotwater
drilling method for IceCube
During the 1990s 23 holes were drilled to depths ranging from 1000 to 2450 meters
at the South Pole to build the AMANDA neutrino telescope. A large hot water drill
was used to drill the holes. This technique was chosen because it was the only one
conceivable to meet the requirements to produce 60cm diameter holes filled with water.
Pioneering efforts to 1000 m depth were successful and allowed the installation of the
first optical sensors in polar ice at the South Pole. The drill grew in size as depth
and hole diameter requirements increased until it reached a maximum thermal power
of 2.2 MW. It became clear that this drill would not be adequate to drill to depths
of 2450 meters, required for IceCube. At depths below 2000 meters drilling became
slow and inefficient. The next generation IceCube detector would require drilling 80
holes in a five year period. A new enhanced hot water drill (EHWD) with a power
of about 5 MW would need to be designed to drill two holes per week. The following
study was initiated to optimize the drill design and drilling strategy and confirm
relatively rough estimates on fuel consumption. We developed a model that allowed

5
us to calculate a drilling procedure that will result in a hole diameter large enough
to ensure successful deployment of detector strings with acceptable fuel requirements.
The model suggests drilling and reaming rates based on heat input that will provide
an optimum hole diameter profile. The result is a barrel shaped hole that compensates
for different freezing rates that are a function of ice temperature and time of exposure
to heat. The calculations and predictions are verified from simulations with drill data
from the AMANDA holes. The robustness of the calculation is checked by applying
perturbations to critical system parameters.
IceCube is a one-cubic-kilometer international high-energy neutrino observatory
being installed in the clear deep ice at the South Pole. It will open unexplored bands
for astronomy, including the PeV energy region, where the Universe is opaque to
high-energy gamma rays originating from beyond the edge of our own galaxy, and
where cosmic rays do not carry directional information because of their deflection
by magnetic fields. The detector will consist of 80 strings of optical modules placed
between 1450 and 2450 meters depth. This depth range takes advantage of the clear
ice below the bubbly ice region and avoids the shear layer between the bottom of
the ice and the detector. The holes will have a diameter of approximately 60 cm to
support the installation of the optical module strings, which are 43 cm in diameter.
The additional size is required to compensate for freezing that takes place on the hole
walls after drilling and during deployment. Hot water drills operate by pumping water
that has been heated under high pressure to a drill head where the hot water jet is used
to melt the ice. In impermeable ice, the hole is filled with water, which is recycled to
the surface by a submersible pump. The water is then reheated and pumped through

6
the drill head to melt more ice. Approximately 8% of the ice volume must be replaced
by water from the surface to make up for the volume lost in the phase change. Before
the AMANDA project these drills were generally portable and limited to less than 1500
meters in depth and to smaller diameters. The cold
?
50
C ice at Sthe South Pole,
depth requirement of 2400m and hole diameter of 60 cm required a much larger drill.
The drill used for AMANDA evolved over the life of the project and grew from 1.6
to 2.2 MW maximum heat input. The drill was pressure limited to 1000 psi operating
pressure, which limited flow as lengths of hose were added. At depths beyond 2000
meters the drill became increasingly inefficient. Fuel consumption per hole was over
10000 gallons and the time required to drill a 2400 meter hole was well over 100 hours,
in one case more than 150 hours. Neither of these figures was acceptable for the 80
holes required for IceCube. A new drill design was proposed that would provide a
constant heat input of about 5MW over the entire drilling depth. A larger hose was
needed to accommodate the higher flow and the hose was to be housed on a single
large reel. Single point power generation, with heat scavenging, have replaced ancient
generators and diesel driven pumps to improve efficiency. In addition the entire drill
heating and pumping plant have been placed in mobile drilling structures that reduce
set up and build down time. The goal is to drill up to 18 holes per year, completing
2 holes per week. The 40 hour drilling time per hole drives the heat input. Since the
drills have all been equipped with calipers and navigation packages, it is possible to
create a map of the hole including a diameter vs depth curve. These measurements
and curves are required to assure the hole diameter remains large enough to permit
the deployment of the optical modules over a 30-hour period after drilling ceases.

7
Crude models were constructed during the AMANDA project to predict the amount
of freezing that would occur during this period. This paper is an introduction to the
theoretical models used to calculate the heat transfer and refreezing rates in boreholes
in cold ice and to compare the results with experimental data from the AMANDA
holes. The calculations are based on models derived for phase change with a moving
boundary layer in cylindrical coordinates.
2.1 Description of the thermal process
A large hot water drilling system consists typically of the following main com-
ponents.
High-pressure pumps to pump the water to the drill head.
Heaters to heat the water to near boiling.
Drill hose to deliver the water to the drill.
Drill head and nozzle to direct water to the front to warm and melt ice.
A return submersible pump to recycle cold water from the bore-hole.
The surface components consist of high-pressure pumps and a heating plant. Heat
losses from this part of the system are low compared of the total heat budget in large
drill systems because the hoses are well insulated. The hose hangs vertically in the
hole, which is filled with water below a depth of 50m. A parcel of water moving
through the hose loses heat to the surrounding water through conduction across the
hose wall and convection to the surrounding water, which is moving slowly up the
hole. The advective term is ignored in these calculations. It is desirable to move the

8
14
13
15
17
16
7
8
7
1
2
2
5
6
10
10
10
10
3
4
9
0
9a
Ic e Cub e
Drill Site La yo ut
Hea ting
S tri ng Deployment
Drill Control
Tanks
Ge ne rators
Pre-Hea t
Tanks
Pumps
Water
Well
Parts storage
work area
7/1 3/0 0
S witching
gear
12
11
Hot-Water Drilling
Figure 2.1: The figure describes the schematic view of the IceCube En-
hanced Hot Water Drill (EHWD) at the surface.
water through the hose as rapidly as possible to keep the residence time in the hose
to a minimum. The amount of heat available at the nozzle drops exponentially with a
decay length λ of the hose where the heat available has fallen to half of that available
at the surface. The decay length depends on the conductivity of the hose material, the
hose wall thickness and the velocity of fluid through the hose. Heat lost through the
hose helps keep the hole from refreezing. Later we will discuss how the decay length
influences drilling and reaming efficiency.
The hot water drill head consists of a massive steel pipe to keep the drill plumb,
and a housing for the electronics and navigation package. A nozzle is designed to
accelerate the speed of the water while keeping the flow intact to create turbulence
ahead of the drill. In cold ice as at South Pole the ice must first be warmed before

9
Figure 2.2: Depth dependnce of the temperature of South Pole Ice.
it can be melted. This can limit drilling speed if the narrow portion of the drill is
short or the velocity at the nozzle is low. A temperature profile of the ice at the South
Pole is shown in figure 2.2. The heat provided by the injected hot water of about 800
liters/min is not dissipated immediately. Energy remains in the form of hot water that
continues to warm and melt the surrounding ice. Some of the energy is conducted into
the surrounding ice. The result is a long plume of warm water that gradually melts
the hole wall to a larger diameter with time. With large hot water drills such as the
AMANDA or IceCube drill this plume can extend more than 100 meters behind the
drill.
As the return flow drifts back up, the cold water column is slightly heated by
losses from the hose. These losses slow the refreezing rate. The effects will be discussed
in a later section. Water at the top of the hole having a temperature of 2
C is recycled
using a submersible pump.

10
Figure 2.3: Schematic view of the hotwater drilling method.
At the surface water is pumped through heaters into the high-pressure hose,
which transports the energy into the borehole where it is used for melting the ice. The
heat input to the borehole is determined from the flux and temperature of pumped
water. The hose at the surface is usually well insulated and the water typically loses
only 2-3
C from the point on the surface where it is pumped until it reaches the top
of the borehole.
A hot water packet takes several minutes travel down the hose to the drill tip,
so an imperfectly insulated hose conducts heat to the surrounding borehole which is
typically filled with cold water. The hose and cable tension is monitored to ensure that
the drill tip does not touch the bottom of the hole. A parcel of fluid traveling down the
hose can be said to lose heat only through the walls of the hose, since the temperature
gradients along the direction of flow are negligible as compared to the radial gradients.

11
These hose losses slow down the process of refreezing that occurs at the top of the hole
and cannot be regarded as a waste. Because of the exponential decay of temperature
and consequently the power available, it is important to maximize the flow within
the pressure capability of the hose. For the EHWD the thermal conductivity of the
hose is 0.4W/m-K [2] and the flow rate is 200 gallons/minute and thus λ is around
8000m. There would also be an advective term, which comes from the conductivity of
the water that fills the hole. In the problem we shall neglect advective heat transfer,
which would fluctuate as the hole size changes. In the past, water was heated to
90
C, losing heat as it travels from the heating plant to the surface of the hole. The
calculations that follow were done with a surface temperature of 88
C, which can be
obtained by suitably insulating the hose from the heaters to the hole opening without
much heat lost into the surrounding atmosphere.
2.2 Methodology and Assumptions
Symmetry allows for an assumption of azimuthal independence. We will simplify
the problem to the one along the radial direction alone because the longitudinal gradi-
ents are negligible. As the heat is exchanged across the ice-water boundary refreezing
will occur. The moving boundary of the phase change interface presents special dif-
ficulties for numerical procedures, since the position of the boundary is dependent
upon a varying temperature field in ice. The boundary condition would be that the
temperature of the ice water interface is always at 0
C. At each time step the change
in radius is calculated from the amount of heat that enters the system in the form of
hose losses and the amount of left over heat energy that goes from the hot water that
is left after the initial melting has occurred. The heat left decays exponentially and

12
for purposes of our calculation, we assume that this heat continues to melt the ice for
a period of four hours after which it goes down to about 2C when it becomes cold
enough that it can no longer melt ice. However, in this section we will briefly look at
the underlying equations. The problem may be formulated in cylindrical co-ordinates
with the z-axis directed downward from the surface along the axis of the hole. The
system is azimuthally symmetric. We will simplify the problem to one along the radial
direction alone because the longitudinal gradients are negligible.
2.3 Description of the model and assumptions
Here we describe a list of the model assumptions in the calculation:
a) The method used to solve the heat transfer differential equation is a numerical
finite difference method. For this purpose, we modeled the problem with a space mesh
where the grid element was 0.15cm wide (1% of the initial radius). We used 1300
elements to span a total area of 3.6m
2
at each depth. The number of elements in the
grid is chosen in such a way that the solution to the problem converges at each depth
in a reasonable time and with sufficient precision. The number of grid points chosen
was increased until the result didn’t change significantly with increasing resolution.
b) The time resolution for each time step was taken as 0.0001 times the charac-
teristic time of the system, which is the total borehole closure time (100 hours). Each
time the size of the hole changed, the number of grid points inside the boundary were
suitably adjusted, allowing us to locate the point on the ice-water interface.
c) Total heatflow is 200 gal/min of water with a temperature of 88
C at the top
of the borehole. This corresponds to a total power of 4.65MW at the top of the hole.
d) The circulated water loses around 40% of the heat before it travels from the

13
nozzle to the widest point of the drill hole (15m above the drill head). This heat is
available for initial melting and for constructing a sufficient diameter hole for the drill
to fit. This process takes 1-2 minutes and is assumed to be non-heat conducting with
regards to the surrounding ice. The temperature to which the water cools during this
instantaneous conduction is given by 54exp(-z/λ), where z is the depth and λ is the
decay length of the hose.
e) Once the initial melting described above is complete, the hot water remains
inside the hole and continues to increase the diameter of the hole, as some of the heat
is lost into its surroundings. This process has been modeled as a typical steady-state
heat conduction problem with a 4-hour duration and an exponential decay function,
where half of the heat decays in the first hour.
f) We modeled the down-hole path as segments of length 100m at constant
temperature. The conditions inside each slice are assumed to be non-varying and
the heat transfer equations are solved inside the mesh surrounding it. Once all the
calculations are done for one slice we move onto the next slice and so on.
g) A final target diameter at 45cm in each of the 100m slices at the end of
operation was set.
h) The thermal properties of ice (specific heat and freezing point) are assumed
to be constant throughout the hole.
i) We obtained the hose losses by dividing the total heat available at any depth
by the λ of the hose. We neglected the conductivity of the water, so all thermal
conductivity is assumed to come from the hose alone. By neglecting the advective
heat transfer, we may have overestimated the hose losses and underestimated the heat

14
Figure 2.4: The figure illustrates the heat transfer procedure that asymp-
totically approaches the far-field ice temperature.
available at the bottom of the hole but all these corrections can be neglected to first
order.
j) We model the reams as an instantaneous process in which no heat is conducted
into the surrounding ice and it is entirely used for melting the ice on the borehole walls.
2.4 Optimal drilling algorithm
The drilling strategy is optimized to obtain a hole of constant diameter at some
specified time of typically 35 hours after drilling is completed. The minimum amount
of time that we spend inside the hole subjected to constraints like the target uniform
diameter after deployment being 45 cm and the drill fitting through can be accom-
plished by a simplex minimizer however it can be observed that the solutions to the
problem exhibit monotonic nature and thus we can find the optimum just by moving
in one direction. The final time step to obtain optimum solution is in turn dependent

15
on the solutions in each of the slices so there is a need for a first guess solution. We
start out at a high drill speed of 200m/hr and a ream speed of 450m/hr respectively
inside the slice. We calculate the initial diameter that we obtain from the cooling
of water from the nozzle to the drill head. If we don’t get sufficient size of the hole
(initial diameter) for the drill to fit through then we reduce the drill speed by 5m/hr
until the condition of the drill passing through is met. There is a maximum of the
drill speed that we use for our optimization. Once this condition is met we calculate
the position of the ice-water boundary at a time t, the initial guess on total time spent
inside the hole plus the estimated deployment time. If the diameter of the hole is less
than the target diameter of 45cm then we reduce the ream speed in steps of 10m/hr
and check if we reach the target diameter at which point we stop and move onto the
next slice. Once the minimum limit on ream speed (180m/hr) is reached without the
target diameter getting to 45 cm, we reduce the drill speed from the maximum limit
obtained for the drill to fit through in steps of 5m/hr till we get the required target
diameter. Once the required target diameter is established we record the values of the
drill speed and the ream speed used for accomplishing the required target diameter
and calculate the time spent by the drill head in this slice during the drilling and the
reaming operations. Then we move on to the next 100metre slice, the drilling oper-
ation in this slice is delayed by a factor of time that we spent in the previous slices
drilling, in other words it takes time for the drill to get here so we subtract the time
we spent till we get to this slice from our initial guess on total time. This analysis is
repeated in slices of 100 metres until we get to the bottom of the hole. By summing
up the times spent drilling and reaming in each of these slices we get the total time

16
spent inside the hole. This should for iterative purposes be close to our initial guess
of 50 hours. If this is lesser (which is usually the case, considering that we start out
with a larger value than what we think our solution is going to be) then we reduce the
initial guess (50 hours) in steps of 1 hour and repeat the entire procedure described
in this section till the iteration is established. Thus the drill speed, ream speed and
other parameters at each depth that satisfy all these conditions are the solutions to
our optimization problem. At these parameter values we get a uniform hole of 45cm in
diameter at all depths and this also minimizes the total time spent by the drill inside
the hole.
2.5 Results
A baseline model assumes the most likely values for the parameters encountered
in the actual field operation. The results of the optimization process are shown in the
following figures. Figure 2.5 illustrates the optimal refreeze process. While the hole is
drilled at different times and the refreeze rates differ with depth they reach an identical
diameter at a predefined target time. This is the time theoretically available for the
deployment team to deploy the string. After this time the hole would be too small
and the string would get stuck and freeze in prematurely. Obviously some contingency
time needs to be taken into account to ensure a safe deployment process.

17
0
10
20
30
40
50
60
70
80
0
10
20
30
40
50
60
70
TIME(hrs)
DIAMETER(cm) ,..
0m
600m
1200m
1800m
2300m
Figure 2.5: The hole diameter as a function of time for a range of depths.
The drill strategy delivers a hole of uniform diameter at a required time
of 30 hours after the drilling is completed.

18
0
10
20
30
40
50
60
70
80
0
500
1000
1500
2000
2500
Depth(m)
10 hour(amidst drilling)
20 hour
30hour(partly reamed)
40 hour
50 hour
60hour
63 hour(operation end)
Diameter (cm)
Figure 2.6: This figure summarizes the evolution of the hole size.

19
-60
-50
-40
-30
-20
-10
0
0
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
2
Distance From Centre (m)
10 hours
20 hours
30 hours
40 hours
50 hours
60 hours
Temperatture (Celsius)
Figure 2.7: This figure summarizes the evolution of the hole diameter as
a function of time. The drill strategy delivers a hole of uniform diameter
at a required time of 30 hours after the drilling is completed.
2.6 Robustness of Predictions
The above analysis assumes that the inputs used are known accurately. However,
the inputs could generally vary due to fluctuations during the drilling operation or
simply because they haven’t been accurately determined. So we performed another
analysis in which we allowed a perturbation on one input at a time and checked the
results against our original overall drilling strategy, paying special attention to the
total time we spend inside the hole and our fuel estimates. The decay length, power
available at surface, deployment time and desired target diameter are varied and a
range of outputs corresponding to optimal strategy are quoted in tabular forms.

20
λ = 5000m
λ = 8000m
Heat (Surface to 800m) (MW)
5.0
2.2
Heat In (at 2000m) (MW)
5.0
1.9
Heat Out (at 2000m) (MW)
4.0
1.6
Drill Rate (at 2000m) (meters/minute)
1.5
?
2.0
0.5
Flow Rate (at 2000m) (gallons/minute)
200
85
Fuel Consumption (gallons/hour)
200
85
Weight (pounds)
400000
250000
Set-upTime(days)
18? 25
35? 42
Fuel (gallons/hole)
6000? 7000 10000? 12000
Table 2.1: The input parameters that go into the old AMANDA drill (λ
= 5000m) and the new IceCube drill (λ = 8000m) are compared.
λ =
λ =
λ =
5000m 8000m 11000m
Drill Time (hrs)
19.28 19.06 21.17
Ream Time (hrs)
10.67 12.36 13.00
Total Time (hrs)
30.00 31.42 32.55
Total Energy Deposited (GJ)
510
538
549
Energy lost to Surroundings (GJ) 358
386
397
Energy Ratio
0.948 1.00
1.02
Drill Fuel (gal)
3856 3812 4234
Total Fuel (gal)
6000 6284 6550
Table 2.2: The outputs for the optimum strategy and consequently the
fuel consumption are quoted by varying the thermal conductivity of the
hose λ

21
20hr deploy 25hr deploy 35hr deploy 40hr deploy
Drill
17.96
17.96
20.52
23.53
Time (hrs)
Ream
9.50
11.29
13.16
13.14
Time (hrs)
Total
27.47
29.25
33.68
36.68
Time (hrs)
Total Energy
465.3
495.6
570.48
621
Deposited (GJ)
Energy lost
313.3
343.6
418.48
469
to Surroundings (GJ)
Energy
0.864
0.921
1.06
1.154
Ratio
Drill
3592
3592
4104
4706
Fuel (gallons)
Total
5494
5850
6736
7336
Fuel (gallons)
Table 2.3: The output parameters corresponding to the optimum strategy
and consequently the fuel consumption are quoted by varying the deploy-
ment time.

22
5%
10 %
15 %
20 %
powercut powercut powercut powercut
Drill Time (hrs)
20.93
23.6
26.9
31.25
Ream Time (hrs)
12.90
13.18
13.40
13.63
Total Time (hrs)
33.83
36.78
40.36
44.89
Total Energy Deposited (GJ)
544.3
560
580.6
608.2
Energy lost to Surroundings (GJ) 392.3
408
428.6
456.2
Energy Ratio
1.01
1.04
1.077
1.128
Drill Fuel (gallons)
3977
4248
4573
5000
Total Fuel (gallons)
6437.2
6620
6860
7182.4
Table 2.4: The output parameters corresponding to the optimum strategy
and consequently the fuel consumption are quoted by cutting down the
power available.
Diameter=45cm Diameter=50cm
Drill Time (hrs)
19.06
22.24
Ream Time (hrs)
12.36
13.29
Total Time (hrs)
31.42
35.53
Total Energy Deposited (GJ)
538
601.8
Energy lost to Surroundings (GJ)
386
424.3
Energy Ratio
1.01
1.12
Drill Fuel (gallons)
3812
4448
Total Fuel (gallons)
6284
7106
Table 2.5: The output parameters corresponding to the optimum strategy
and consequently the fuel consumption are quoted by varying the desired
target diameter.

23
2.7 Conclusions and Summary
A fundamental analysis of the available drill data and a heat transfer simulation
was performed. Further a thermodynamic analysis of the process of drilling and re-
freeze and compared our results with existing data. This provided a better prediction
of the refreeze rates and an optimal strategy for efficiently drilling uniform holes. We
also infer that it is always better strategy to put energy into the ice as late as possible
to prevent it from refreezing; in other words drill fast and ream slow. This is because
when we put in heat energy late we are fighting less steep temperature gradients in
the surrounding ice. In order to make full use of the analysis we propose a more
regulated system called the smart drill. In this system the computer would regulate
the drill speed in such a way that the borehole diameter is uniform and of the size as
predicted by an optimized freeze back prediction. Studying the system perturbations
on a one at a time basis gave us valuable insights into the development of such a
system. A significant improvement in hole quality and fuel consumption will be to the
benefit of the proposed IceCube project. The modifications suggested in this analysis
contributed to a significant reduction of the fuel consumption for the IceCube holes
that have been drilled to date (40 holes as of 2008).

24
Chapter 3
Measuring the Prompt Atmospheric
Neutrino Flux with Downgoing Muons in
AMANDA-II
3.1 AMANDA Detector
The AMANDA-II detector, Antarctic Muon and Neutrino Detesctor Array, is
located at the South Pole. It consists of a total of 677 optical modules. Each module
comprises of photomultiplier tube and the hardware inside a glass pressure sphere.
These optical modules are attached to 19 strings frozen into the ice, these sensors
are deployed across a range of depths from 1500m to 2000m in a cylinder of 100 m
radius. These modules detect Cherenkov light from secondary charged particles that
are produced from the interaction of a neutrino with ice. AMANDA is integrated
into IceCube detector which is still under construction. IceCube will consist of 70-80
detector strings, each with 60 optical modules. Currently 40 strings are deployed and
are taking data.

25
Figure 3.1: The figure shows the layout of the AMANDA detector. The
top view shows 19 strings that were deployed. AMANDA detector is
roughly 200m wide and 500m long

26
Detectors like AMANDA-II are sensitive to an energy region in which contribu-
tions from prompt charm decays in cosmic ray showers cannot be neglected and may
constitute an interesting signal as well as a significant background depending on the
nature of the analysis. In searches for diffuse fluxes of astrophysical neutrinos the sig-
nal must be separated at high energies from the background of atmospheric neutrinos.
Atmospheric muons can reach the detector only from above (downgoing through the
earth) because the range of muons in earth is only a few kilometers. Atmospheric
muons are therefore only downgoing. Their flux is typically so high that the region
of sky accessible to even very deep neutrino telescopes is only the hemisphere below
the horizon. Atmospheric neutrinos can instead reach the detector from all directions.
Hence they are an irreducible background for diffuse astrophysical neutrino fluxes. It
is therefore, very important to evaluate their intensity with reasonable accuracy.
3.2 Conventional and Prompt Atmospheric Neutrinos
When cosmic rays interact with the nuclei in the atmosphere there are two types
of particles that are produced that could decay subsequently to give the muon and
neutrino fluxes observed. Once these particles are produced in the atmosphere there
is a competition between interaction and decay. The critical energy at which the
interaction and decay lengths become equal is defined as
ǫ
crit
=
mc
2
h
o
(3.1)
where mc
2
is the particle’s rest energy, τ the mean life time and the scale constant
h
o
that comes from the assumption of an isothermal atmosphere [36].

27
Figure 3.2: The figure schemtically shows the interaction of the primary
cosmic ray proton with the atmosphere and the formation of several par-
ticles as the shower evolves. (Image credit: Milagro)

28
Particle ǫ
crit
(GeV)
µ
±
1.0
π
±
115
K
±
850
D
±
3.8×10
7
D
0
,D¯
0
9.6
×10
7
D
±
s
8.5
×10
7
Λ
+
c
2.4
×10
8
Table 3.1: Critical energy for different particles.
Table 3.1 lists the critical energy of parent particles of muons and neutrinos.
These particles could decay and contribute to the atmospheric muon and neutrino
fluxes. As can be seen, this critical energy is very high for charmed particles so
charmed particles decay readily and muons from them are called “prompt” muons.
Above this energy the parent particle is likely to interact or be slowed down before
decaying into a neutrino and muon. Since prompt muons are produced readily they
follow a Φ
E
?
2.7
spectrum resembling the primary cosmic ray spectrum. The π
±
and K
±
mesons decay into conventional neutrinos only if they dont interact in the
atmosphere. If interaction takes place they disappear in the atmosphere and hence
the spectra of conventional neutrinos follows a Φ
E
?
3.7
spectrum.
3.3 Constraining the Prompt Neutrino Flux with the Down-
going Muon Flux
With increasing energy prompt neutrinos become the biggest source of uncer-
tainty in predictions of the atmospheric neutrino flux. The DPMJET-II.55 is the only
model available for simulating prompt muons and it uses a Naumov RQPM model

29
[38]. Details of the model are discussed in the next section. As can be seen in figure
3.3 the crossover between conventional neutrinos and the Naumov RQPM model of
charm is between 40 and 200 TeV. The level of prompt neutrinos is a potential problem
which would limit the search for diffuse astrophysical neutrinos at energies above a few
tens of TeV. The suggestion in this thesis is based on the observation that due to the
charmed particle decay kinematics for semi-leptonic decays into muon and neutrino
fluxes these fluxes are essentially the same at sea level. This result is independent of
the charm production model and hence a constraint on a prompt muon flux is equiv-
alent to a constraint on the prompt neutrino flux [31]. There are ways of separating
the prompt muons from the conventional muons in underwater or under-ice detectors,
such as the different zenith angle dependence, the different depth dependence at a
given zenith angle, and the different spectral shape at a given depth and zenith angle
[31].
3.4 Prompt Atmospheric Neutrino Models
The prompt atmospheric neutrino fluxes are uncertain by more than 2 orders
of magnitude. This is because ground-based particle accelerators cannot reach the
energies at which particles are produced in the atmosphere. The uncertainty stems
from the need to extrapolate accelerator data to the high energies probed and the
uncertainities in parameters that go into each model: the primary spectral index, the
critical energy for decay and the Interaction and decay lengths can also play a role.
The Naumov RQPM (Recombination Quark Parton Model) [38, 39] tested in
this thesis is a phenomenological non-perturbative model that takes into account the
contribution of intrinsic charm to the production process. It is assumed that c¯c pairs

30
(GeV)]
ν
[E
10
log
3
4
5
6
7
8
9
]
-1
sr
-1
s
-2
dN/dE [GeV cm
2
E
-9
10
-8
10
-7
10
-6
10
-5
10
-4
10
ν
Conventional atmospheric
ν
Martin GBW prompt atmospheric
ν
Naumov RQPM prompt atmospheric.
ν
CharmC prompt atmospheric
ν
CharmD prompt atmospheric
Figure 3.3: Prompt atmospheric neutrinos are predicted to follow a harder
spectrum than conventional atmospheric neutrinos. The flux of prompt
atmospheric neutrinos is highly uncertain and predictions range over sev-
eral orders of magnitude. Image Credit: Jessica Hodges

31
are coupled to a number of constituents in the projectile hadron [40]. The production
of mesons such as π
±
and K
±
in the fragmentation region of proton collisions pro-
ceeds via quark recombination. The evidence for the recombination mechanism comes
from the observation that the longitudinal momentum distribution of the pions in the
fragmentation region of an incident proton is very similar to the distribution of the
valence quarks that they share in the proton, as is revealed in deep inelastic scattering
experiments [44].
3.5 Charm in CORSIKA
CORSIKA is a simulation software package to simulate air showers induced by
primary cosmic rays. The program produces large amounts of data of all the secondary
particles (electromagnetic particles, muons and hadrons). An interaction model is used
to describe the physics of the interactions between cosmic rays and the atmosphere.
The older DPMJET interaction model of CORSIKA produced charmed particles but
they were never allowed to decay so prompt muons could not be simulated. The energy
spectra for prompt muons had to be taken at the surface of the earth from empirical
parametrizations [33] and muons of multiplicity one were simulated because of lack of
prior knowledge of prompt muon multiplicities. In the DPMJET-II-55 framework [20]
charmed particles were treated and hence prompt muons produced subsequently were
tagged through a generation counter based on their parent [16]. For demonstration
purposes the case of first interaction alone was isolated from multiple interactions
in the atmosphere. The plots at the surface of the earth for the energy spectra and
lateral separation from shower axis are shown in figures 3.4 and 3.5 respectively. Figure
3.4 shows that there is not much discrimination power between prompt muons and

32
conventional muons when lateral distributions are compared using the DPMJET-II.55
model. This contradicts the hypothesis that prompt muons are single muons as was
previously hypothesized when we didn’t have a prompt muon simulation. DPMJET-
II.55 provided us with a reference model for prompt muons and proved to us that
any strategy to separate prompt muons from conventional muons has to be focused
on using the flatter energy and zenith dependence of prompt muons; if one were to
identify these signal events on top of the background events from conventional muons.
An important part of being able to do a prompt muon analysis using the DPMJET-
II.55 interaction model of CORSIKA is being able to identify them. If one is interested
in muons which come from a decay of ordinary mesons generated in the first interac-
tion, CORSIKA is run and the particle file is scanned for those muons which come
from the first interaction by looking for the generation counter. In CORSIKA the
generation counter is available that tracks the parents of the particles produced. For
instance, for the decays from pions the counter is augmented by 51, for charmed parti-
cles it is 31 to get a discrimination against all other muons. The tagging could also be
used to identify the muons from first interaction using the number on the generation
counter during the production. If an event contains one or more muons that have
their parent as a charmed particle we identify them as prompt muons. A need to tag
the prompt muon events was accomplished this way.
An important part of being able to do a prompt muon analysis using the DPMJET-
II.55 interaction model of CORSIKA is being able to identify them. If one is intrested
in muons which come from a decay of ordinary mesons generated in the first interac-
tion, CORSIKA is run and the particle file is scanned for those muons which come

33
Lateral Separation Plot(all tracks)
FIRST INT ONLY(PROMPTS)
FIRST INT ONLY(CONV)
AFTER MULTIPLE INTER(PROMPTS)
AFTER MULTIPLE INTER(CONV)
Lateral Separation(metres) at Surface Of Earth
Percentage of Events in Bin
10
-4
10
-3
10
-2
10
-1
1
0
25
50
75
100
125
150
175
200
225
250
Figure 3.4: The distribution of lateral separation from shower core for the
DPMJET-II for charm and coventional muons in each event with the first
interaction and multiple interactions isolated.

34
Muon Energy Plot
FIRST INT ONLY(PROMPTS)
FIRST INT ONLY(CONV)
AFTER MULTIPLE INTER(PROMPTS)
AFTER MULTIPLE INTER(CONV)
log10(Total Muon Energy)GeV at Surface Of Earth
Percentage of Events in Bin
10
-4
10
-3
10
-2
10
-1
1
3
3.5
4
4.5
5
5.5
6
Figure 3.5: The total energy distribution for the DPMJET-II for charm
and conventional muons for each event with first interaction and multiple
interactions isolated.

35
from the first interaction by looking for the generation counter. In CORSIKA the
generation counter is available that tracks the parents of the particles produced. For
instance, for the decays from pions the counter is augmented by 51, for charmed parti-
cles it is 31 to get a discrimination against all other muons. The tagging could also be
used to identify the muons from first interaction using the number on the generation
counter during the production. If an event contains one or more muons that have
their parent as a charmed particle we identify them as prompt muons. A need to tag
the Prompt muon events was accomplished this way.

36
Chapter 4
Data streams and quality cuts on the 2005
Sample
4.1 First guess reconstructions, livetime and triggers
Likelihood reconstructions need a first guess for the muon track direction to be
able to iterate through and get the final direction. The criteria for designing a first
guess algorithm should be to have fast computing time while approximating the track
direction. Two popular first guess algorithms are described below.
4.1.1 Direct Walk Reconstruction
Direct Walk is a pattern recognition algorithm. The basic algorithm reconstructs
the track direction using only direct hits(hits unscattered in ice).
4.1.2 JAMS Reconstruction
JAMS (Just Another Muon Search) is a pattern recognition algorithm. The basic
algorithm for JAMS creates hit selections and stores an event. A fast algorithm to
find first guess candidate tracks is implemented using hit clusters and these are stored.

37
The first guess candidates are narrowed down into a few viable track solutions. One or
two of these best track fits are stored in the event. The quality of the fit is measured
by variable σ
ψ
and in this analysis is set to be less than 0.05. It measures the deviation
of hit clusters from a single track hypothesis.
During the 2005 data run, 1.85 billion events were recorded by AMANDA-II. The
livetime for the 2005 filtered data is 199.25 days. The 2005 data filtering is different
from earlier years in two aspects [19].
1. Modified cross talk is applied.
2. JAMS and Direct Walk at level 1 and level 2 cuts are swapped
The discussion in this chapter is confined to the analysis of downgoing muons.
The output streams relevant to our analysis are the high quality stream and the
minimum bias stream.
4.1.3 High quality stream
Downgoing muon events near the horizon with a high quality in JAMS recon-
struction were selected. Every event with a zenith angle greater then 55 degrees in
JAMS reconstruction fit with a low number for the spread σ
ψ
< 0.05 of its JAMS fit
solutions was identified in this stream.
4.1.4 Minimum bias stream
Every 100th event regardless of how it was triggered was included in this stream.

38
4.2 Reconstruction Methods
An arrival timing based likelihood approach as described by the Pandel function
[13] was used to reconstruct the 2005 experimental data and simulation. This function
was subsequently modified for PMT jitter and and a convoluted pandel probability
distribution function was used for this analysis.
4.3 Techniques to Further Improve Background Rejection
Several other techniques were used to improve background rejection. Methods
were also employed to remove electronic crosstalk and other fake events. The quality
of the JAMS fit was checked to ensure high quality events
4.4 Event Simulation and Reweighting
This analysis relied on simulated data sets of background and signal events. The
downgoing conventional atmospheric muons were simulated using the DPMJET-II.55
and SYBILL interaction models. For this work, a preliminary version of CORSIKA
in which the charmed particle decays are enabled in DPMJET-II.55 [20] was used for
signal simulation. Further, we know that the threshold primary energy for muons at
large zenith angles to make it to the detector is high so we do the simulation in two
stages to get enhanced statistics at large zenith angles.
1. Primary threshold energy of 800 GeV of primary cosmic ray energy for muons
between 0 to 70 degrees zenith angle.
2. Primary threshold energy of 10 TeV for primary cosmic ray between 70 to 90
degrees zenith angle.

39
The events were simulated with a Φ
E
?
1.7
primary energy spectrum, one power
harder than the spectrum present in nature. The advantage of this approach is that
it reduces the simulation time. The generated events must then be reweighted to the
original cosmic ray spectrum with appropriate normalization factor applied [17].
4.4.1 Preparation of Simulated Events
It would be a huge demand on computer time to simulate the same number of
days livetime worth of simulation as actual experimental data. Simulation is thus
generated with a flatter spectra and event weights are used to reweight to the spectra
we wish to simulate [18]. The simulation events were scaled to match the livetime of
the data during a particular period or year. The simulated events underwent the same
reconstruction procedures as the data and had to satisfy the same zenith angle and
energy requirements. The reconstructed muons deviate from their actual direction due
to mis-reconstruction and we use quality criteria to improve the angular resolution.

40
Chapter 5
Response of AMANDA-II to Cosmic Ray
Muons
The flux of downgoing muons detected by the AMANDA-II neutrino telescope is
used as a test beam to check the experimental systematic error and to improve the
knowledge of its response. This work shows the outcome of the effort for a better
understanding of AMANDA-II performance, an improved data filter and event recon-
struction. The simulated predictions for preliminary experimental downgoing muon
angular and depth intensities are compared with experimental results and theoretical
calculations. This report encompasses large statistics of simulated data generated at
large zenith angles using the QGSJET interaction model with a live time of 30 days
and compared to a 30 day span of experimental data for the year 2001. This was
aimed to minimize the effect of statistical and systematic errors on the angular and
depth intensities (particularly close to the horizon). The use of a new Convoluted
Pandel likelihood function ensures improved event reconstruction and stability. This
also ensures good agreement between simulated and experimental data. Further, anal-
ysis was also done using the 2005 experimental data and 2005 simulation using the

41
SYBILL interaction model. CORSIKA air shower generator was used to simulate the
interaction of the cosmic rays with the atmosphere.
5.1 Analysis
In order to measure the atmospheric muon angular distribution it is necessary
to evaluate the event trigger and reconstruction efficiencies as a function of the zenith
angle. This requires a Monte Carlo with the complete simulation chain from the pri-
mary interaction in the atmosphere to the detector response based on best knowledge
of the physical processes involved. The event generation is done using CORSIKA
v6.020 with the QGSJET01 interaction model. The model incorporates the curvature
of the earth and the south pole average atmospheric profile. A multi-component pri-
mary cosmic ray energy spectrum [5] is used to get the composition. The generated
muons are propagated to the earth’s surface and then through the ice, considering all
pertinent energy losses. The muons passing through AMANDA-II or near it are folded
into the detector trigger simulation. At this stage the detector response is completely
simulated in order to reproduce the experimentally detected events. This is based on
our overall understanding of the physics and the detector.
The event reconstruction chain for simulation is identical to the one used for
the experimental data. A cleaning procedure then removes the optical modules that
are dead or have odd transient behavior. A time calibration which also accounts
for the signal propagation time through the cables is then employed. A Convoluted
Pandel probability distribution function (a time likelihood based reconstruction) is
used. This accounts for the finite photo-multiplier tube timing resolution of the pulse
obtained when a photon emitted by the muon passes through it. Due to the limited

42
angular resolution of the reconstruction, additional cuts are used to improve the event
sample quality for both the experimental and simulated data. These cuts improve the
resolution of the zenith angle and the space angle. .
Tables 5.1 and 5.2 summarize the mean, median, RMS and the extreme tail
of the zenith angle and the space angle resolution. We misreconstruct events closer
to the horizon and this becomes more prominent as we get closer to the horizon.
This can partially be explained by the fact that vertical muons would have larger
track length and larger number of direct hits (hits that are minimally scattered in the
ice before being detected by a optical module) and hence their direction determined
more accurately than a horizontal event. With this resolution we can derive the
experimental angular distribution at AMANDA-II depth by merely calculating the
detector acceptance at each zenith angle bin using the simulated detector response
to unfold the measured data, neglecting the inter-bin correlations. When we neglect
interbin correlations, for each bin in cosine of the zenith angle, the ratio of the true
events generated (that trigger AMANDA) to the reconstructed events (accounting for
various efficiencies during reconstruction) for both the Monte Carlo simulation (known
ratio) and the corresponding experimental data (unknown ratio) can be equated. With
interbin correlations, one needs to account for leakages across bins due to finite angular
resolution and event quality. Neglecting inter-bin correlations can be justified by
demonstrating that these correlations are minimal. We accomplish this by making our
best event quality selections on a sample of 30 days of simulation data. The quality
cuts chosen for this purpose are chi square for the reconstructed track less than 7.3,
track length greater than 120 meters(-15ns to 75ns direct hits only), absolute value of

43
the the difference between the zenith angle reconstructed using odd hits only and using
even hits only from hits ordered. Further these inter-bin overlaps can be minimized
by increasing or decreasing the bin size in cosine zenith. The bin-size for the actual
intensity distribution is defined at twice the RMS value of statistics shown in tables
5.1 and 5.2 so that most events are self contained and overlaps are minimal. This
implies we need to use larger bin sizes to account for poorer angular resolution closer
to the horizon. Experimental data spanning 30 days from the year 2001 is used.
5.2 Results
Ice properties(scattering and interaction lengths) are an important uncertainty
that affects the count rates of muons in the Monte carlo simulation when results are
shown. In this chapter in particular and the thesis in general we use three models.
MAM (Muon Absorption Model) is obtained by increasing the absorption such that
the time residuals match between the data and the Monte Carlo. It includes the effect
of OM sensitivity as well as ice model and is based on the model of layered ice. The
millennium ice model incorporates actual AMANDA ice measurements and is believed
to be more accurate than the MAM model. A new and a better ice model called the
AHA incorporates ”stretched” layer structure, i.e., dust peaks are higher and valleys
lower and corrects for the systematic smearing of the layers with the measurement
techniques used in AMANDA [10].
Figure 5.1 shows the plot of the angular distribution of downgoing muons in the
AMANDA-II detector (the flux of atmospheric muons versus the cosine of the zenith
angle) using the 2001 experimental data. The triangles represent the AMANDA-II
detector data and the boxes represent the simulated data. The plot shows simulated

44
data using the MAM ice model. Figure 5.2 shows the atmospheric muon flux as
a function of slant depth. The slant depth is a function of the zenith angle and
represents the distance the muon travels to the AMANDA depth. Imagine looking
at the surface of the Earth from the AMANDA depth at different angles. Figure 5.3
and figure 5.4 shows the comparison between the simulation and experimental data.
It can be seen that the experimental data and the simulation differ by about 25% for
vertical muons and is as high as 40-45% for horizontal muons.
Likewise, figures 5.5, 5.6, 5.7, 5.8 represent similar plots for the 2005 experimen-
tal data and SYBILL millenium model. Better angular resolution was also ensured
using tighter quality cuts compared to the earlier MAM ice model. Figure 5.7 and
5.8 show the comparison between the simulation and experimental data and is a mea-
sure of performance of our simulations to replicate experimental data and helps us
understand the systematic error. It can be seen that the experimental data and the
simulation differ by only 10-20% till 80 degrees in zenith angle when the Millenium
SYBILL Monte Carlo is used.
Figure 5.11 shows the comparison of the down going muon intensity as mea-
sured by the L3+C collaboration. It can be noticed that hadronic interaction model
QGSJET01 has the maximum disagreement with the observed data while SYBILL
has the minimum difference.

45
cosΘ
Φ
μ
(cm
-2
sec
-1
sr
-1
)
10
-15
10
-14
10
-13
10
-12
10
-11
10
-10
10
-9
10
-8
10
-7
10
-6
10
-5
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
Figure 5.1: The angular distribution of atmospheric muons in AMANDA-
II at a depth of 1730m using the MAM ice model with the SYBILL inter-
action model and the 2001 experimental data.

46
Depth (Kmwe)
Φ
μ
(cm
-2
sec
-1
sr
-1
)
10
-15
10
-14
10
-13
10
-12
10
-11
10
-10
10
-9
10
-8
10
-7
1
10
Figure 5.2: The depth-intensity of atmospheric muons in AMANDA-II
using the MAM ice model and the SYBILL interaction model with the
2001 experimental data.

47
depth(km)
(MC-Data)/Data
-0.5
-0.4
-0.3
-0.2
-0.1
0
0.1
0.2
1
10
Figure 5.3: The relative difference between MAM SYBILL Monte Carlo
and AMANDA-II 2001 data as a function of depth.
cosine(zenith)
(MC-Data)/Data
-0.5
-0.4
-0.3
-0.2
-0.1
0
0.1
0.2
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
Figure 5.4: The relative difference between MAM SYBILL Monte Carlo
and AMANDA-II experimental data as a function of zenith angle.

48
Cos(Zenith)
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
-1
sr
-1
sec
-2
Flux(cm
-14
10
-13
10
-12
10
-11
10
-10
10
-9
10
-8
10
-7
10
-6
10
Angular Distribution
Exp Data (2005)
MMC(SYBILL) Data (2005)
Figure 5.5: The angular distribution of atmospheric muons in AMANDA-
II at a depth of 1730m using the Millenium ice model with the SYBILL
interaction model and the 2005 experimental data.
Depth (Kmwe)
10
)
-1
sr
-1
sec
-2
Flux(cm
-15
10
-14
10
-13
10
-12
10
-11
10
-10
10
-9
10
-8
10
-7
10
-6
10
Depth Dependency Plot
Exp Data (2005)
MMC(SYBILL) Data (2005)
Figure 5.6: The depth-intensity of atmospheric muons in AMANDA-II
using the Millenium ice model and the SYBILL interaction model with
the 2005 experimental data.

49
Depth (Kmwe)
10
(MC-Data)/Data
-1.5
-1
-0.5
0
0.5
1
1.5
Data Agreement Plot
Depth (Kmwe)
10
(MC-Data)/Data
-1.5
-1
-0.5
0
0.5
1
1.5
Figure 5.7: The relative difference between Millenium SYBILL Monte
Carlo and AMANDA-II 2005 experimental data as a function of depth.
cos(zenith)
0
0.2
0.4
0.6
0.8
1
(MC-Data)/Data
-1.5
-1
-0.5
0
0.5
1
1.5
Data Agreement Plot
Figure 5.8: The relative difference between Millenium SYBILL Monte
Carlo and AMANDA-II 2005 experimental data as a function of zenith
angle.

50
cos
Mean
Median
RMS 90% quantile 90% quantile
(zenith)
(degrees) (degrees) (degrees) right(degrees) left(degrees)
0.0-0.05 14.23(21.22) 11.75(22) 7.11(5.55)
25.5(28.0)
7.0(+13.0)
0.05-0.1 8.02(19.10) 6.75(19) 5.09(6.55)
16(27.0)
3.0(+10.0)
0.1-0.15 5.27(15.49) 4.5(15.5) 4.14(7.01)
11.75(25.0)
1.0(+6.0)
0.15-0.20 3.22(11.46) 2.75(11) 3.34(6.96)
7.5(21.0)
-0.2(+2.5)
0.20-0.25 2.36(8.04)
2.0(7.0) 2.93(6.34)
6.0(18.0)
-0.75(+0.5)
0.25-0.30
1.77(5.6)
1.6(4.5) 2.71(5.51)
5.0(14.0)
-1.25(-0.5)
0.30-0.35 1.43(3.86)
1.2(3.0) 2.57(4.76)
4.0(11.0)
-1.75(-1.5)
0.35-0.40
1.17(2.6)
1.0(2.0) 2.45(4.13)
4.25(8.5)
-1.75(-2)
0.40-0.45 0.97(1.55)
0.7(1.5) 2.34(3.64)
4.0(6.5)
-1.9(-2.75)
0.45-0.50 0.69(1.55) 0.6(0.75) 2.13(3.26)
3.5(4.5)
-2.0(-5.5)
0.50-0.55 0.84(1.99) 0.25(1.25) 2.26(4.96)
4(8)
-2(-3)
0.55-0.60 0.53(1.63)
0.5(1) 2.24(4.88)
3.5(7.5)
-2(-3)
0.60-0.65 0.53(1.41)
0.5(0.5) 2.25(4.86)
3.75(7)
-2(-3.5)
0.65-0.70 0.51(0.99) 0.12(0.75) 2.06(4.57)
3.0(6)
-2(-3.5)
0.70-0.75 0.30(0.74) 0.12(0.5) 1.98(4.39)
2.5(5.5)
-2(-3.5)
0.75-0.80 0.29(0.49) 0.12(0.5) 1.79(4.30)
2.5(5)
-2(-3.5)
0.80-0.85 0.21(0.30) 0.12(0.25) 1.75(4.21)
2.0(4.5)
-2(-3.5)
0.85-0.90 0.17(0.05) 0.1(0.25) 1.72(4.09)
2.0(4.5)
-2(-4)
0.90-0.95 0.13(-0.30)
0.1(0.0) 1.77(3.98)
2.25(3.8)
-2(-4.5)
0.95-0.1.0 -0.14(-1.05) 0.1(-0.5) 1.94(4.12)
2.0(2.75)
-2.5(-5.75)
Table 5.1: Presents the statistics of zenith angle resolution after quality
cuts for various zenith ranges. Values in brackets are before quality cuts
for the QGSJET model.

51
cos
Mean
Median
RMS 90% quantile
(zenith)
(degrees) (degrees) (degrees)
(degrees)
0.0-0.05 16.25(24.34) 9.13(24.5) 2.72(7.30)
26(34)
0.05-0.1 9.88(21.21)
6.63(21) 2.29(7.91)
19(32)
0.1-0.15 6.85(17.6)
4.39(17) 2.73(8.4)
13(29)
0.15-0.20 5.34(13.95) 3.79(12.75)
8.3
10(25.2)
0.20-0.25 4.63(10.95) 3.23(9.2)
7.7
8.9(21.2)
0.25-0.30 4.31(9.05)
3.0(7.2)
6.9
8.0(18.2)
0.30-0.35
4.10(7.75 2.9(6.05)
6.14
7.9(16)
0.35-0.40 4.00(6.88) 2.76(5.5)
5.55
7.8(13.2)
0.40-0.45 3.84(6.22) 2.62(5.0)
5.11
7.0(12.0)
0.45-0.50 3.63(5.65) 2.47(4.5)
4.81
6.8(11.0)
0.50-0.55 3.87(6.63) 2.72(5.0)
6.04
7.0(13.5)
0.55-0.60
3.16(6.3) 2.29(4.75)
5.68
6.0(13.5)
0.60-0.65 3.47(6.17)
2.7(4.5)
5.62
6.3(13.0)
0.65-0.70 3.20(5.91) 2.17(4.5)
5.33
6.0(12.5)
0.70-0.75 2.98(5.58) 2.04(4.2)
5.16
5.75(12.0)
0.75-0.80 2.73(5.31) 1.84(4.0)
4.99
5.1(11.5)
0.80-0.85 2.65(5.13) 1.92(3.75)
5.01
5.0(11.0)
0.85-0.90 2.56(5.09) 1.88(3.75)
5.00
5.0(11.0)
0.90-0.95 2.61(5.13) 2.0(3.75)
5.2
5.0(11.0)
0.95-0.1.0 2.89(5.14) 2.3(3.75)
5.18
5.7(11.0)
Table 5.2: Presents the statistics of space angle resolution after quality
cuts for various zenith angle ranges. Values in brackets are before quality
cuts for the QGSJET model.

52
RecoZenith-TrueZenith(In Degrees)
-30
-20
-10
0
10
20
30
normalized counts
-3
10
-2
10
-1
10
1
zenith range 0.0-0.05
RecoZenith-TrueZenith(In Degrees)
-30
-20
-10
0
10
20
30
normalized counts
-3
10
-2
10
-1
10
1
zenith range 0.05-0.1
RecoZenith-TrueZenith(In Degrees)
-30
-20
-10
0
10
20
30
normalized counts
-3
10
-2
10
-1
10
1
zenith range 0.1-0.15
RecoZenith-TrueZenith(In Degrees)
-30
-20
-10
0
10
20
30
normalized counts
-3
10
-2
10
-1
10
1
zenith range 0.15-0.20
RecoZenith-TrueZenith(In Degrees)
-30
-20
-10
0
10
20
30
normalized counts
-3
10
-2
10
-1
10
1
zenith range 0.20-0.25
RecoZenith-TrueZenith(In Degrees)
-30
-20
-10
0
10
20
30
normalized counts
-3
10
-2
10
-1
10
1
zenith range 0.25-0.30
RecoZenith-TrueZenith(In Degrees)
-30
-20
-10
0
10
20
30
normalized counts
-3
10
-2
10
-1
10
1
zenith range 0.30-0.35
RecoZenith-TrueZenith(In Degrees)
-30
-20
-10
0
10
20
30
normalized counts
-3
10
-2
10
-1
10
1
zenith range 0.35-0.40
RecoZenith-TrueZenith(In Degrees)
-30
-20
-10
0
10
20
30
normalized counts
-3
10
-2
10
-1
10
1
zenith range 0.40-0.45
RecoZenith-TrueZenith(In Degrees)
-30
-20
-10
0
10
20
30
normalized counts
-3
10
-2
10
-1
10
1
zenith range 0.45-0.50
Figure 5.9: The zenith angle difference between the reconstructed and true
zenith angle known from simulation is plotted on x-axis while normalized
counts are plotted on y-axis. The respective slices in zenith are indicated
in the plot. Red is before quality cuts while blue is after quality cuts.
From left to right and top to bottom there are 10 slices shown that go
from 0.0-0.5 in increments of 0.05.

53
RecoZenith-TrueZenith(In Degrees)
-30
-20
-10
0
10
20
30
normalized counts
-3
10
-2
10
-1
10
1
zenith range 0.50-0.55
RecoZenith-TrueZenith(In Degrees)
-30
-20
-10
0
10
20
30
normalized counts
-3
10
-2
10
-1
10
1
zenith range 0.55-0.60
RecoZenith-TrueZenith(In Degrees)
-30
-20
-10
0
10
20
30
normalized counts
-3
10
-2
10
-1
10
1
zenith range 0.60-0.65
RecoZenith-TrueZenith(In Degrees)
-30
-20
-10
0
10
20
30
normalized counts
-3
10
-2
10
-1
10
1
zenith range 0.65-0.70
RecoZenith-TrueZenith(In Degrees)
-30
-20
-10
0
10
20
30
normalized counts
-3
10
-2
10
-1
10
1
zenith range 0.70-0.75
RecoZenith-TrueZenith(In Degrees)
-30
-20
-10
0
10
20
30
normalized counts
-3
10
-2
10
-1
10
1
zenith range 0.75-0.80
RecoZenith-TrueZenith(In Degrees)
-30
-20
-10
0
10
20
30
normalized counts
-3
10
-2
10
-1
10
1
zenith range 0.80-0.85
RecoZenith-TrueZenith(In Degrees)
-30
-20
-10
0
10
20
30
normalized counts
-3
10
-2
10
-1
10
1
zenith range 0.85-0.90
RecoZenith-TrueZenith(In Degrees)
-30
-20
-10
0
10
20
30
normalized counts
-3
10
-2
10
-1
10
1
zenith range 0.90-0.95
RecoZenith-TrueZenith(In Degrees)
-30
-20
-10
0
10
20
30
normalized counts
-3
10
-2
10
-1
10
1
zenith range 0.95-1.0
Figure 5.10: The zenith angle difference between the reconstructed and
true zenith angle known from simulation is plotted on x-axis while nor-
malized counts are plotted on y-axis. The respective slices in zenith are
indicated in the plot. Red is before quality cuts while blue is after quality
cuts. From left to right and top to bottom there are 10 slices shown that
go from 0.5-1.0 in increments of 0.05.

54
Figure 5.11: Comparison of CORSIKA vertical muon flux for various in-
teraction models.

55
Chapter 6
Model dependencies and systematic error
calculations for a down-going muon
analysis
Traditional cut and count analysis place a limit or make a discovery based on excess
of events over the predicted background in the experimental data. Due to uncertain-
ties in the simulation, the number of signal and background events predicted may not
accurately reflect the true signal and background numbers. The cosmic ray spectrum
is uncertain both in normalization and spectral index. There are also detector-related
uncertainties due to the uncertain sensitivity of optical modules and modeling of light
propagation. Different interaction models produce different number of muons and
there is a wide variety of choices of models but by far the biggest uncertainty affect-
ing this analysis is the ice. Our incomplete understanding of the dust layers in the
ice makes ice a nuisance parameter that affects the sensitivity to the prompt muon
signal. For purposes of computing final results we use two models, the AHA and the
Millenium ice models without being biased to either but allow the fit to decide the

56
right proportions of these models to explain the experimental data. The systematic
uncertainties are summed in quadrature separately for background and signal. These
uncertainities are a useful exercise to determine the mismatch between experimental
data and simulation that could be used for any future downgoing muon analysis.
Due to the nature of uncertainties and difficulties in quantifying them we deviate
and take another approach that is based on shape based fitting for background and
signal to experimental data for calculating the final limit.
6.1 Statistical Errors
Due to the computational requirements, background and signal simulation statis-
tics are somewhat limited. However, the optimized background simulations used in
this analysis have large variation in event weights (we use a reweighted MC simula-
tion). The statistical errors are kept track by ROOT and these were subsequently
used for the construction of the limit.
6.2 Systematic Uncertainties
6.2.1 Normalization of Cosmic Ray Flux
The different absolute normalizations between the experiments are caused by
uncertainities in the energy calibration. The average energy of cosmic ray particles
is 4.4*10
7
GeV, which is considerably above the knee in the all-particle cosmic ray
spectrum. Numerous experiments have measured a large spread in the absolute nor-
malization of the flux of cosmic rays at this energy [9]. Estimates of the uncertainty
in the normalization of the Horandel cosmic ray flux are 20% [5]. This uncertainty

57
translates to a 20% variation in the number of background and signal events.
6.2.2 Spectral Index of Cosmic Ray Spectrum
The best fit values for the spectral index of the cosmic ray data is -2.71±0.02 in
which the errors specify the statistical uncertainties [5]. Varying the spectral index by
0.02 in the DPMJET-II interaction model (model being tested) produces an average
variation of 35% in the number of background and signal events in the High-Energy
region (N
ch
400) as is shown in figure 6.1. The corresponding plot with the actual
number of event counts in one year of data is shown in figure 6.2.
6.2.3 Detector Sensitivity
The properties of the refrozen ice around each OM, the absolute sensitivity of
individual OMs, the obscuration of OMs by nearby power cables can effect the detector
sensitivity. The analysis uses the values obtained in [6] where reasonable variations of
these parameters in the simulation were found to cause a 15% variation in the signal
and background passing rate.
6.2.4 Interaction Model Uncertainity
For this analysis, two interaction models SYBILL and DPMJET-II are consid-
ered equally likely options for the background simulation. Each of the these models
were renormalized to match the number of data events observed in the low N
ch
region,
where the signal was expected to be insignificant compared to the background. By
rescaling the simulation to the number of observed data events, the uncertainty of the
background simulation was reduced to the uncertainty in the spectral shape. We have
only one model of signal simulation (DPMJET-II) hence no uncertainty was assumed

58
on it. Varying the models produces an average variation of 80% on the background in
the high energy region (N
ch
400) as is shown in figure 6.3. The corresponding plot
with the actual number of event counts in one year of data is shown in figure 6.4.
6.2.5 Ice Model Uncertainty
Based on results from ice properties systematics studies, the millennium ice
model has been modified and a new ice model (called the aha model) has been con-
structed. Two types of modifications were made to the millennium model: a) the ice
model was corrected for a systematic smearing of the dust layer structure introduced
by the analysis methods used in AMANDA, and b) the extrapolation of the optical
properties to larger depths was redone with new ice core data on dust concentration
to produce cleaner ice below the big “dust peak”. The details are explained in [10].
Varying the models produces an average variation of 40% on the background in the
high energy region (N
ch
400) as is shown in figure 6.5. The corresponding plot with
the actual number of event counts in one year of data is shown in figure 6.6.
6.2.6 Other Source of Errors
The systematic errors due to the rock density (below the detector), and muon
energy loss do not contribute significantly to this analysis.
6.3 Result of Systematics Study
The systematic uncertainties are summed in quadrature separately for back-
ground and signal and the total systematic error numbers for background and signal
are calculated to be 70% and 60% respectively.

59
Source Background
Signal
Name simulation simulation
Cosmic Ray Normalization
20
20
Cosmic Ray Spectral Index
35.0
35.0
Detector Sensitivity
15.0
15.0
Ice Properties
40.0
40.0
Interaction Model Uncertainty
40.0
X
Total Error
70.0
60.0
Table 6.1: Average simulation uncertainties for different sources of errors.

60
Number Of Optical Module Hit
0
100
200
300
400
500
Ratio(systematic variation)
-1
-0.8
-0.6
-0.4
-0.2
-0
0.2
0.4
0.6
0.8
1
DPMJET (2005)(charm) After Quality cut Index=-2.68
DPMJET (2005)(conventional) After Quality Cut Index=-2.68
DPMJET (2005)(charm) After Quality Cuts Index=-2.72
DPMJET (2005)(conventional) After Quality Cuts Index=-2.72
Figure 6.1: The N
ch
variation for the DPMJET-II for signal and back-
ground (at the final level after event selection criteria are implemented)
when spectral index is varied by
±0.02
shown as a ratio.
Number Of Optical Module Hit
0
100
200
300
400
500
Counts
1
10
2
10
3
10
4
10
5
10
6
10
7
10
8
10
9
10
DPMJET (2005)(charm) After Quality cut Index=-2.68
DPMJET (2005)(conventional) After Quality Cut Index=-2.68
DPMJET (2005)(charm) After Quality Cuts Index=-2.72
DPMJET (2005)(conventional) After Quality Cuts Index=-2.72
DPMJET (2005)(charm) After Quality Cuts Index=-2.7
DPMJET (2005)(conventional) After Quality Cuts Index=-2.7
Figure 6.2: The N
ch
distribution for the DPMJET-II for signal and back-
ground (at the final level after event selection criteria are implemented)
when spectral index is varied by
±0.02
shown as a ratio.

61
Number of Optical Modules Hit
0
100
200
300
400
500
Variation
-1
-0.5
0
0.5
1
1.5
NChannel After Quality Cut Plot
Compared to an Equally Weighted Sample (2005) After Quality Cuts
Figure 6.3: The N
ch
variation for the DPMJET-II for background (at the
final level after event selection criteria are implemented) when compared
with an equally weighted simulation of SYBILL and DPMJET-II is shown.
Number of Optical Modules Hit
0
100
200
300
400
500
Counts
1
10
2
10
3
10
4
10
5
10
6
10
7
10
8
10
9
10
NChannel After Quality Cut Plot
SYBILL simulation (2005) After Quality Cuts
DPMJET (2005)(conventional) After Quality Cuts
Equally Weighted sample (2005) of DPMJET and SYBILL after Quality Cuts
Figure 6.4: The N
ch
distribution for the DPMJET-II for background (at
the final level after event selection criteria are implemented) when com-
pared with an equally weighted simulation of SYBILL and DPMJET-II is
shown.

62
Number Of Optical Module Hit
0
100
200
300
400
500
Variation
-1
10
1
10
Figure 6.5: The N
ch
variation for the DPMJET-II millenium ice model
(at the final level after event selection criteria are implemented) when
compared with an equally weighted simulation of DPMJET-II millenium
and DPMJET-II AHA model is shown.
Number Of Optical Module Hit
0
100
200
300
400
500
Counts
1
10
2
10
3
10
4
10
5
10
6
10
7
10
8
10
9
10
SYBILL simulation (2005) After Quality Cuts(Photonics)
SYBILL simulation (2005) After Quality Cuts (Photonics AHA)
Equally Weighted Photonics AHA+Photonics simulation (2005) After Quality Cuts
Figure 6.6: The N
ch
distribution for the DPMJET-II millenium ice model
(at the final level after event selection criteria are implemented) when
compared with an equally weighted simulation of DPMJET-II millenium
and DPMJET-II AHA model is shown.

63
Chapter 7
Hadronic Interaction Models and
Extended Air showers
7.1 Introduction
The development of a shower is influenced by the properties of the hadronic
interactions and the mechanisms of the transport of secondary particles through the
atmosphere in addition to the primary cosmic ray mass and energy. The hadronic
and nuclear interactions impose large uncertainties since they are poorly known in
the energy and kinematic ranges of interest. In addition, a detector with its limited
acceptance and efficiency gives a distorted picture of the secondary particles. The
challenge of experimental physics is to understand the shower development and the
detector performance well enough using the experimental data obtained and to gain
an understanding of hadronic and nuclear interaction models.
In the shower development process major uncertainities arise from the hadronic
interactions which are described by phenomenological models. These models are tuned
to fit the available data from p
?
p¯ and heavy ion accelerators but the experimental

64
data are not well suited for CR interactions. Collider experiments do not register
the most energetic particles emitted in the extreme forward direction which are key
to atmospheric shower development and accelerators by far do not reach the energies
encountered in cosmic rays. This extrapolation to higher energies relies solely on
theoretical guidelines and the uncertainities play a huge role.
The muon rate as measured by the AMANDA-II detector is higher by about 30%
than simulations using the QGSJET model with the Wiebel-Sooth parametrization
for the cosmic ray spectrum [25]. Comparisons like these indicate that more bench-
marks with data and improvements of the hadronic interaction model are necessary.
A comparison of interaction models in CORSIKA [26] when used in its interaction
test mode for beams of monoenergetic protons on nitrogen nuclei (the most abundant
component of air) is presented. In this mode only the first interaction of a shower
calculation is performed.
All secondaries from CORSIKA including the spectator nucleons from projectile
and target, are stored in the particle stack, and further shower calculations are omitted.
In this mode, many interactions can be generated, and all information about the
particles can be stored. In the released CORSIKA version, charmed hadrons cannot
be handled properly. In this work we have a preliminary version of CORSIKA in
which the charmed particle decays are enabled in the framework of DPMJET-II.55
[20] hadronic interaction model. The energy fractions, multiplicities and Z-moments
of these particles are compared to the FLUKA+DPMJET-III [27] hadronic interaction
and transport code. For model comparison we have used both diffractive and non-
diffractive events in a mixture as given by respective models. Diffractive events are

65
visible as peaks at large energy fractions, as events typically have a forward-going
meson with a direction slightly different from that of the original proton.
It is suggested in [30] that prompt muons become dominant at large distances
from shower core due to their larger transverse momentum. DPMJET-II model (with
charmed particles allowed to decay) is used to test this hypothesis and lateral muon
density distributions characterizing the region of high energy and large zenith angle
are shown and these distributions are isolated for the first interaction (showering off)
and multiple interaction (showering on) for the following cases:
a) Cosmic ray spectrum
b) Cosmic ray spectrum with zenith angle greater than 80 degrees
c) Showers with 1-1000PeV primary energy
d) Monoenergetic primary energy of 1PeV.
Events involving prompt muon (atleast one or more prompt muons) and conven-
tional muon ( no prompt muon) are distinguished. Characterization of the hardness
of the energy and zenith angle spectra of prompt muons compared to conventional
muons are shown in the framework of DPMJET-II.55 hadronic interaction model.
7.2 Interaction and Extended Air Shower models
7.2.1 Available Codes and Model Comparisons
CORSIKA [26] is a multi-purpose shower simulation program of air shower de-
velopment. The hadronic interaction models DPMJET, QGSJET-I [28] and QGSJET-
II [29], SYBILL in CORSIKA are studied and compared with the predictions from
the FLUKA model. QGSJET and DPMJET are based on the Gribov-Regee the-

66
ory of multi-Pomeron exchange, which has been used successfully for over a decade
to describe elastic and inelastic scattering of hadrons. In particular nucleus-nucleus
collision and diffraction are treated in great detail in these models. SIBYLL [22] is a
minijet model that describes the rise of the cross-section with energy by increasing the
pairwise minijet production and also applies the Glauber theory for hadron-nucleus
collisions and treats projectile nuclei as a superposition of free nucleons. PYTHIA
[32] models hadronic interactions with high momentum transfer according to QCD,
and takes into account resonance formation as well as gluon radiation from quarks
and contains the fragmentation of colour strings into colour neutral hadrons. It also
contains the soft processes which are important for air showers, but cannot handle
primary mesons or nucleus-nucleus collisions on the basis of classical string theory.
For treatment of nucleus-nucleus collisions FRITIOF adopts superposition principle.
Hence a combination of PYTHIA and FRITIOF should be used to simulate EAS. The
major systematic uncertainities in EAS analysis arise from the lack of knowledge of
the total cross-sections and the details of particle production for nuclear and hadronic
reactions at high energies with small momentum transfer.
7.2.2 Cross Sections
The first quantities compared are the inelastic p-air cross-sections. All models
except for SYBILL calculate cross-sections from the experimental data assuming a
distribution of the nucleons in the air nuclei. Therefore, all these models agree reason-
ably with each other (and collider data) and start to diverge only at energies where
no measurements exists anymore. SYBILL adopts a parametrization which exhibits
the flattest rise with values clearly below the experimental results at lower energies

67
and steeply rises and surpasses all other models at higher energies. In DPMJET-
II.55 this has been corrected downwards and agrees nicely with the cross sections Of
QGSJET. The spread between the models amounts to about 35% as shown in figure
4.12. Since the inelastic cross-section determines the mean free path of a particle in
the atmosphere, it influences directly the longitudinal shower development. A larger
cross-section causes shorter showers and consequently, fewer particles at ground level.
The differences in cross-sections for comparable assumptions originate partially from
different applications of the Glauber theory and from varying assumptions regarding
the form of the target nuclei. The discrepancies between the models are rather big,
taking into consideration that all authors use basically the same approach to calcu-
late cross-sections. By agreement on the best method of calculation a big part of the
discrepancies should vanish.
7.2.3 Particle Production
The production of secondaries in hadronic interactions also differs between mod-
els. A variety of quantities need to be examined.The quantity with the largest impact
on air shower development is the inelasticity ,i.e. the fraction of the energy of a par-
ticle that is used for production of secondary particles. Again, a variation in this
quantity directly implies a modification of the longitudinal shower development. The
effects of inelasticity and cross-sections are basically independent and may cancel out
or add up. For DPMJET with the largest inelasticity and the largest cross-sections
the showers are very short and this leads to differences in the muon multiplicity at the
ground level and depth of detector.

68
7.2.4 Impact of shower simulations
In interpretation of shower measurements,it is vitally important to make a com-
mon effort towards a reference simulation program that contains the best and the most
detailed treatment of all physical processes relevant to shower development that are
used and tested by each experiment in a different way. Such a reference should also
serve to estimate the performance of special purpose programs that are optimized for
particular aspects of CR physics such as calculations of highest energy showers, TeV
muons, Cherenkov light production, and so on. Air shower analyses are based on the
comparison of experimental data with MC simulations and so to be able to perform
such a comparison, a spectral form, an energy dependent mass composition and pa-
rameters of the high-energy interactions have to be assumed. Therefore, a discrepancy
between MC and data can have many sources and on the other hand an agreement
does not necessarily mean that all assumptions are right especially when registering
only one observable (i.e. number of muons); several parameter settings may exist
that can reproduce the observation. Fluctuations in the observable are then directly
projected onto uncertainties in the primary energy or mass. When measuring several
quantities it is possible to recognize fluctuations. A big part of the shower fluctuations
originates from the first hadronic interaction and the secondary particles produced are
studied. It has been noted earlier that the SYBILL model predicts fewer muons for
high-energy air showers as compared to other models. Several hadronic observables
have been investigated and compared to various hadronic interaction models. Corre-
spondingly, they are characterized by a restricted number of adjustable parameters,
which can be fitted with the available data. However the microscopic content of any

69
model is restricted by only a number of possible physics mechanisms. Thus, one can
not exclude the possibility that something important is missing, especially, concerning
the very high energy range. This explains the need for alternative model approaches
and for continuing tests of model validity, using both accelerator and cosmic ray data.
7.3 Results
7.3.1 Interaction Model
A comparison of interaction models in CORSIKA when used in its interaction
test model for beams of monoenergetic protons on nitrogen nuclei is done. In this
model only the first interaction of a shower calculation is performed and high-statistics
proton beams in the energy range from 1TeV to 100PeV were used. All secondaries,
including the spectator nucleons from projectile and target, are stored in the particle
stack, and further shower calculations are omitted. In this mode, many interactions
can be generated, and all information about the particles can be stored. In the released
CORSIKA version, charmed hadrons cannot be handled properly. For this work we
have used a preliminary version of CORSIKA in which the charmed particle decays are
enabled in DPMJET-II.55 [20]. The energy fractions, multiplicities and Z-moments
of these particles are compared to the FLUKA+DPMJET-III interaction and trans-
port code. For model comparison we have used both diffractive and non-diffractive
events in a mixture as given by respective models. Diffractive events are visible as
peaks at large energy fractions, as events typically have a forward-going meson with
a direction slightly different from that of the original proton. From figure 7.1 we see
that SIBYLL and FLUKA + DPMJET-III are in very good agreement with each

70
other and in reasonable agreement with DPMJET-II for conventional mesons (pions
and kaons). However, QGSJET-01 [28] and QGSJET-II [29] predict a lower energy
fraction in the diffractive region where secondaries take a very large fraction of the pri-
mary energy. This could explain the disagreement in the AMANDA-II muon intensity
distribution , since the depth of the detector selects higher energy secondaries. The
fact that AMANDA data are about 30% higher than simulations [25] indicates that
models like SIBYLL, FLUKA+DPMJET-III and DPMJET II.55, as well as a harder
proton primary spectrum parametrization could better account for the experimental
observations. For charmed hadrons, this implementation of DPMJET-II in CORSIKA
underestimates diffractive events. This is particularly evident for charmed baryons.
In figure 7.2, the Z-moments and the multiplicities are shown for all energies and
models. Z-moments show a similar trend to what is described for the energy fractions
with a weight that takes into account the slope of the cosmic ray spectrum. It is also
noticeable that the spread between models is much larger for kaons than for pions.
7.3.2 Extended Air Shower
7.3.2.1 lateral distribution function
Measurements of the lateral distribution of the penetrating muon component of
extended air showers (EAS) underground are sensitive to the chemical composition
of primaries, their energy and interaction characteristics. One critical component of
these models is the forward production of pions and kaons in high-energy hadronic
interactions. Most of the pions and kaons are produced at low transverse momentum.
The validity of approaches suggested in [30] in which it is hypothesized that the high

71
transverse momentum muons can be used to infer their production rate from heavy
quarks is tested. For purposes of illustration the first interaction alone(no shower)
is separated from the case of full shower formation for prompts and conventional
muons, further proton showers are separated from iron showers to see the affect of
heavier primaries. Lateral distribution function gives the average number of muons per
metre square of area in the radial direction away from the point of reference. Lateral
distribution functions are plotted for the different cases of full cosmic ray spectrum,
full cosmic ray spectrum with zenith angle greater than 80 degrees, primaries in the
energy range of 1-1000 PeV and primaries of fixed energy 1 PeV at zenith angle of 65
degrees. Most of the muons are produced along the direction of the shower core and
decrease as we go away from the shower core as is reflected by the falling slope of these
plots. Usually the number of muons from first interaction are lower than multiple
interactions considering that multiple interaction encompasses first interaction. All
these plots taken together prove a single important point that the strategy to separate
prompt muons from conventional muons using lateral distribution functions does not
look promising. One possible explanation is that the transverse and longitudinal
momenta are on average larger for charmed secondaries as can be seen in figure 6.3
for the case of 1PeV fixed primaries at a zenith angle of 65 degrees therefore, the
lateral distribution of muons at the surface measured from the shower axis is not much
different for prompt muons than conventional ones neither after the first interaction
nor after full shower development. The various trends can be read from figures section
bearing in mind that at very large lateral distances generation statistics also play a
role. These plots are shown in figures 7.4, 7.5, 7.6, 7.7, 7.8, 7.9.

72
7.3.2.2 Zenith Angle and Energy Spectra
Detailed studies of zenith angle and energy spectra is not the primary objective
of this chapter. Testing of the DPMJET-II.55 model is done and distributions of
surface energy, energy at detector and zenith angle are shown. As is known from [31]
prompts exhibit a harder energy and zenith angle spectra and can be seen in figures
7.10, 7.11 and 7.12. It is hypothesized during my earlier analysis that the production
of charmed particles in the forward direction would produce muons of multiplicity one
and here we test that hypothesis using our simulation and see that is not the case.

73
primary
E/E
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
)
p
dN/d(E/E
-5
10
-4
10
-3
10
-2
10
-1
10
1
10
2
10
3
10
)
avg
Model (N
DPMJET-II (18.6)
QGSJET-01c (21.6)
SIBYLL 2.1 (20.1)
QGSJET-II-03 (20.0)
FLUKA / DPMJET-III (21.6)
GeV
4
at 10
±
π
) for
p
dN/dx vs. x (E/E
primary
E/E
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
)
p
dN/d(E/E
-5
10
-4
10
-3
10
-2
10
-1
10
1
10
2
10
3
10
)
avg
Model (N
DPMJET-II (29.8)
QGSJET-01c (33.2)
SIBYLL 2.1 (32.0)
QGSJET-II-03 (30.1)
FLUKA / DPMJET-III (34.9)
GeV
5
at 10
±
π
) for
p
dN/dx vs. x (E/E
primary
E/E
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
)
p
dN/d(E/E
-5
10
-4
10
-3
10
-2
10
-1
10
1
10
2
10
3
10
)
avg
Model (N
DPMJET-II (48.6)
QGSJET-01c (53.5)
SIBYLL 2.1 (51.9)
QGSJET-II-03 (47.4)
FLUKA / DPMJET-III (55.5)
GeV
6
at 10
±
π
) for
p
dN/dx vs. x (E/E
primary
E/E
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
)
p
dN/d(E/E
-5
10
-4
10
-3
10
-2
10
-1
10
1
10
2
10
3
10
)
avg
Model (N
DPMJET-II (74.2)
QGSJET-01c (88.2)
SIBYLL 2.1 (84.4)
QGSJET-II-03 (81.0)
GeV
7
at 10
±
π
) for
p
dN/dx vs. x (E/E
primary
E/E
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
)
p
dN/d(E/E
-5
10
-4
10
-3
10
-2
10
-1
10
1
10
2
10
)
avg
Model (N
DPMJET-II (6.4)
QGSJET-01c (4.5)
SIBYLL 2.1 (5.4)
QGSJET-II-03 (3.3)
FLUKA / DPMJET-III (4.6)
GeV
4
) for Kaons at 10
p
dN/dx vs. x (E/E
primary
E/E
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
)
p
dN/d(E/E
-6
10
-5
10
-4
10
-3
10
-2
10
-1
10
1
10
2
10
)
avg
Model (N
DPMJET-II (10.9)
QGSJET-01c (8.1)
SIBYLL 2.1 (9.4)
QGSJET-II-03 (5.3)
FLUKA / DPMJET-III (7.8)
GeV
5
) for Kaons at 10
p
dN/dx vs. x (E/E
primary
E/E
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
)
p
dN/d(E/E
-6
10
-5
10
-4
10
-3
10
-2
10
-1
10
1
10
2
10
)
avg
Model (N
DPMJET-II (18.4)
QGSJET-01c (14.1)
SIBYLL 2.1 (16.0)
QGSJET-II-03 (8.6)
FLUKA / DPMJET-III (12.8)
GeV
6
) for Kaons at 10
p
dN/dx vs. x (E/E
primary
E/E
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
)
p
dN/d(E/E
-6
10
-5
10
-4
10
-3
10
-2
10
-1
10
1
10
2
10
3
10
)
avg
Model (N
DPMJET-II (28.5)
QGSJET-01c (24.3)
SIBYLL 2.1 (26.9)
QGSJET-II-03 (14.6)
GeV
7
) for Kaons at 10
p
dN/dx vs. x (E/E
primary
E/E
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
)
p
dN/d(E/E
-5
10
-4
10
-3
10
-2
10
-1
10
1
10
Model (N
avg
)
DPMJET-II (0.3)
FLUKA / DPMJET-III (0.006)
GeV
4
) for Charm at 10
p
dN/dx vs. x (E/E
primary
E/E
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
)
p
dN/d(E/E
-5
10
-4
10
-3
10
-2
10
-1
10
1
10
)
avg
Model (N
DPMJET-II (0.6)
FLUKA / DPMJET-III (0.04)
GeV
5
) for Charm at 10
p
dN/dx vs. x (E/E
primary
E/E
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
)
p
dN/d(E/E
-5
10
-4
10
-3
10
-2
10
-1
10
1
10
2
10
Model (N
avg
)
DPMJET-II (0.9)
FLUKA / DPMJET-III (0.1)
GeV
6
) for Charm at 10
p
dN/dx vs. x (E/E
primary
E/E
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
)
p
dN/d(E/E
-5
10
-4
10
-3
10
-2
10
-1
10
1
10
2
10
Model (N
avg
)
DPMJET-II (1.3)
GeV
7
) for Charm at 10
p
dN/dx vs. x (E/E
Figure 7.1: Energy fraction distributions using various models for charmed
baryon and mesons for energies of 10, 10
2
, 10
3
, and 10
4
TeV

74
Z-Moment as a Function Of Energy
DPMJET-II
QGSJET-01c
SIBYLL 2.1
QGSJET-II-03
FLUKA/DPMJET-III
log
10
E
primary
(GeV)
Z-moment
0
0.05
0.1
0.15
0.2
0.25
0.3
0.35
2
3
4
5
6
7
8
9
Mean Multiplicity as a Function Of Energy
DPMJET-II
QGSJET-01c
SIBYLL 2.1
QGSJET-II-03
FLUKA/DPMJET-III
log
10
E
primary
(GeV)
Mean Multiplicity
10
-1
1
10
10
2
2
3
4
5
6
7
8
9
Figure 7.2: The mean multiplicity and the Z-moments of pions and kaons
as a function of primary energy. The top ensemble of points denote pions
while the bottom ones denote kaons
FIRST INT ONLY(PROMPTS)
FIRST INT ONLY(CONV)
AFT MULTIPLE INTER(PROMPTS)
AFT MULTIPLE INTER(CONV)
All Muon Tracks
log10(Muon Energy)GeV at Surface Of Earth
Percentage of Events in Bin
10
-4
10
-3
10
-2
10
-1
1
3
3.5
4
4.5
5
5.5
6
Lateral Separation Plot(all tracks)
FIRST INT ONLY(PROMPTS)
FIRST INT ONLY(CONV)
MULTIPLE INTER(PROMPTS)
MULTIPLE INTER(CONV)
Tracks that make it to detector only
Maximum Lateral Separation(meters) at Surface Of Earth
Percentage of Events in Bin
10
-4
10
-3
10
-2
10
-1
1
0
25
50
75
100
125
150
175
200
225
250
Figure 7.3: The trasverse momentum, longitudinal momentum and lat-
eral separation of the secondary particles produced by air showers for a 1
PeV monoenergetic beam of primary protons at a fixed zenith angle of 65
degrees.

75
LDF at Earth Surface(CR spectrum)
FIRST INT ONLY(PROMPTS)
FIRST INT ONLY(CONV)
MULTIPLE INTER(PROMPTS)
MULTIPLE INTER(CONV)
Lateral Sep(meters) From Shower Core Surface(Track by Track)
average muons per m
2 per event
10
-15
10
-14
10
-13
10
-12
10
-11
10
-10
10
-9
10
-8
10
-7
10
-6
10
-5
10
-4
10
-3
10
-2
10
-1
1
0
50
100
150
200
250
300
350
400
450
500
LDF Of Muons AMANDA depth(CR Spectrum Zenith80)
FIRST INT ONLY(PROMPTS)
FIRST INT ONLY(CONV)
MULTIPLE INTER(PROMPTS)
MULTIPLE INTER(CONV)
Lateral Sep(meters) From Shower Core Surface(Track by Track)
average muons per m
2 per event
10
-15
10
-14
10
-13
10
-12
10
-11
10
-10
10
-9
10
-8
10
-7
10
-6
10
-5
10
-4
10
-3
10
-2
10
-1
1
0
50
100
150
200
250
300
350
400
450
500
LDF at Surface Of Earth(1-1000PeV)
FIRST INT ONLY(PROMPTS)
FIRST INT ONLY(CONV)
MULTIPLE INTER(PROMPTS)
MULTIPLE INTER(CONV)
Lateral Sep(meters) From Shower Core Surface(Track by Track)
average muons per m
2 per event
10
-15
10
-14
10
-13
10
-12
10
-11
10
-10
10
-9
10
-8
10
-7
10
-6
10
-5
10
-4
10
-3
10
-2
10
-1
1
0
50
100
150
200
250
300
350
400
450
500
LDF at Surface Of Earth(1PeV primaries only)
FIRST INT ONLY(PROMPTS)
FIRST INT ONLY(CONV)
MULTIPLE INTER(PROMPTS)
MULTIPLE INTER(CONV)
Lateral Sep(meters) From Shower Core Surface(Track by Track)
average muons per m
2 per event
10
-15
10
-14
10
-13
10
-12
10
-11
10
-10
10
-9
10
-8
10
-7
10
-6
10
-5
10
-4
10
-3
10
-2
10
-1
1
0
50
100
150
200
250
300
350
400
450
500
Figure 7.4: Shows the average number of muons produced per event as a
function the lateral separation from the shower core at surface of earth for
showers initiated by the full cosmic ray spectrum, full cosmic ray spectrum
for zenith>80 degrees, for primaries in the energy range of 1-1000 PeV
and monoenergetic primary energy of 1 PeV with no showering (only the
first interaction) and after the full shower develops (multiple interactions)
with events containing atleast 1 prompt muon (produced from a charmed
particle) tagged as “PROMPTS” and for no prompt muon involved as
“CONV”. All data has been normalized to 1 years worth lifetime

76
LDF at Surface Of Earth(CR spectrum)
PROTON SHOWERS(PROMPTS)
PROTON SHOWERS(CONV)
IRON SHOWERS(PROMPTS)
IRON SHOWERS(CONV)
Lateral Sep(meters) From Shower Core Surface(Track by Track)
average muons per m
2 per event
10
-15
10
-14
10
-13
10
-12
10
-11
10
-10
10
-9
10
-8
10
-7
10
-6
10
-5
10
-4
10
-3
10
-2
10
-1
1
0
50
100
150
200
250
300
350
400
450
500
LDF at Surface Of Earth(zenith greater than 80)
PROTON SHOWERS(PROMPTS)
PROTON SHOWERS(CONV)
IRON SHOWERS(PROMPTS)
IRON SHOWERS(CONV)
Lateral Sep(meters) From Shower Core Surface(Track by Track)
average muons per m
2 per event
10
-15
10
-14
10
-13
10
-12
10
-11
10
-10
10
-9
10
-8
10
-7
10
-6
10
-5
10
-4
10
-3
10
-2
10
-1
1
0
50
100
150
200
250
300
350
400
450
500
LDF at Surface Of Earth(1-1000PeV)
PROTON SHOWERS(PROMPTS)
PROTON SHOWERS(CONV)
IRON SHOWERS(PROMPTS)
IRON SHOWERS(CONV)
Lateral Sep(meters) From Shower Core Surface(Track by Track)
average muons per m
2 per event
10
-15
10
-14
10
-13
10
-12
10
-11
10
-10
10
-9
10
-8
10
-7
10
-6
10
-5
10
-4
10
-3
10
-2
10
-1
1
0
50
100
150
200
250
300
350
400
450
500
LDF at Surface Of Earth(1PeV primaries only )
PROTON SHOWERS(PROMPTS)
PROTON SHOWERS(CONV)
IRON SHOWERS(PROMPTS)
IRON SHOWERS(CONV)
Lateral Sep(meters) From Shower Core Surface(Track by Track)
average muons per m
2 per event
10
-15
10
-14
10
-13
10
-12
10
-11
10
-10
10
-9
10
-8
10
-7
10
-6
10
-5
10
-4
10
-3
10
-2
10
-1
1
0
50
100
150
200
250
300
350
400
450
500
Figure 7.5: Shows the average number of muons produced per event as
a function the lateral separation from the shower core at surface of earth
for showers initiated by the full cosmic ray spectrum, full cosmic ray spec-
trum for zenith>80 degrees, for primaries in the energy range of 1-1000
PeV and monoenergetic primary energy of 1 PeV after the full shower
develops (multiple interactions) with showers produced by protons and
iron identified separately. All data has been normalized to 1 years worth
lifetime

77
LDF Of Muons at Surface(Cosmic Ray Spectrum)
FIRST INT ONLY(PROMPTS)
FIRST INT ONLY(CONV)
MULTIPLE INTER(PROMPTS)
MULTIPLE INTER(CONV)
Lateral Sep(meters) From Most Energetic Muon Surface(Track by Track)
average muons per m
2 per event
10
-15
10
-14
10
-13
10
-12
10
-11
10
-10
10
-9
10
-8
10
-7
10
-6
10
-5
10
-4
10
-3
10
-2
10
-1
1
0
50
100
150
200
250
300
350
400
450
500
LDF Of Muons at Surface(CR Spectrum Zenith80)
FIRST INT ONLY(PROMPTS)
FIRST INT ONLY(CONV)
MULTIPLE INTER(PROMPTS)
MULTIPLE INTER(CONV)
Lateral Sep(meters) From Most Energetic Muon Surface(Track by Track)
average muons per m
2 per event
10
-15
10
-14
10
-13
10
-12
10
-11
10
-10
10
-9
10
-8
10
-7
10
-6
10
-5
10
-4
10
-3
10
-2
10
-1
1
0
50
100
150
200
250
300
350
400
450
500
LDF Of Muons at Surface(1-1000PeV primaries only)
FIRST INT ONLY(PROMPTS)
FIRST INT ONLY(CONV)
MULTIPLE INTER(PROMPTS)
MULTIPLE INTER(CONV)
Lateral Sep(meters) From Most Energetic Muon Surface(Track by Track)
average muons per m
2 per event
10
-15
10
-14
10
-13
10
-12
10
-11
10
-10
10
-9
10
-8
10
-7
10
-6
10
-5
10
-4
10
-3
10
-2
10
-1
1
0
50
100
150
200
250
300
350
400
450
500
LDF Of Muons at Surface(1PeV primaries only)
FIRST INT ONLY(PROMPTS)
FIRST INT ONLY(CONV)
MULTIPLE INTER(PROMPTS)
MULTIPLE INTER(CONV)
Lateral Sep(meters) From Most Energetic Muon Surface(Track by Track)
average muons per m
2 per event
10
-15
10
-14
10
-13
10
-12
10
-11
10
-10
10
-9
10
-8
10
-7
10
-6
10
-5
10
-4
10
-3
10
-2
10
-1
1
0
50
100
150
200
250
300
350
400
450
500
Figure 7.6: Shows the average number of muons produced per event as a
function the lateral separation from the most energetic muon at surface of
earth for showers initiated by the full cosmic ray spectrum, full cosmic ray
spectrum for zenith>80 degrees, for primaries in the energy range of 1-
1000 PeV and monoenergetic primary energy of 1 PeV with no showering
(only the first interaction) and after the full shower develops (multiple
interactions) withevents containing atleast 1 prompt muon (produced from
a charmed particle) tagged as “PROMPTS” and for no prompt muon
involved as “CONV”. All data is normalized to 1 years worth lifetime

78
LDF Of Muons at Surface(Cosmic Ray Spectrum)
PROTON SHOWERS(PROMPTS)
PROTON SHOWERS(CONV)
IRON SHOWERS(PROMPTS)
IRON SHOWERS(CONV)
Lateral Sep(meters) From Most Energetic Muon Surface(Track by Track)
average muons per m
2 per event
10
-15
10
-14
10
-13
10
-12
10
-11
10
-10
10
-9
10
-8
10
-7
10
-6
10
-5
10
-4
10
-3
10
-2
10
-1
1
0
50
100
150
200
250
300
350
400
450
500
LDF Of Muons at Surface(zenith greater than 80)
PROTON SHOWERS(PROMPTS)
PROTON SHOWERS(CONV)
IRON SHOWERS(PROMPTS)
IRON SHOWERS(CONV)
Lateral Sep(meters) From Most Energetic Muon Surface(Track by Track)
average muons per m
2 per event
10
-15
10
-14
10
-13
10
-12
10
-11
10
-10
10
-9
10
-8
10
-7
10
-6
10
-5
10
-4
10
-3
10
-2
10
-1
1
0
50
100
150
200
250
300
350
400
450
500
LDF Of Muons Surface (1-1000PeV primaries)
PROTON SHOWERS(PROMPTS)
PROTON SHOWERS(CONV)
IRON SHOWERS(PROMPTS)
IRON SHOWERS(CONV)
Lateral Sep(meters) From Most Energetic Muon Surface(Track by Track)
average muons per m
2 per event
10
-15
10
-14
10
-13
10
-12
10
-11
10
-10
10
-9
10
-8
10
-7
10
-6
10
-5
10
-4
10
-3
10
-2
10
-1
1
0
50
100
150
200
250
300
350
400
450
500
LDF Of Muons Surface (1PeV primaries only)
PROTON SHOWERS(PROMPTS)
PROTON SHOWERS(CONV)
IRON SHOWERS(PROMPTS)
IRON SHOWERS(CONV)
Lateral Sep(meters) From Most Energetic Muon Surface(Track by Track)
average muons per m
2 per event
10
-15
10
-14
10
-13
10
-12
10
-11
10
-10
10
-9
10
-8
10
-7
10
-6
10
-5
10
-4
10
-3
10
-2
10
-1
1
0
50
100
150
200
250
300
350
400
450
500
Figure 7.7: Shows the average number of muons produced per event as a
function the lateral separation from the most energetic muon at surface
of earth for showers initiated by the full cosmic ray spectrum, full cosmic
ray spectrum for zenith>80 degrees, for primaries in the energy range of
1-1000 PeV and monoenergetic primary energy of 1 PeV after the full
shower develops (multiple interactions) with showers produced by protons
and iron identified separately. All data is normalized to 1 years worth
lifetime

79
LDF Of Muons AMANDA depth(Cosmic Ray Spectrum)
FIRST INT ONLY(PROMPTS)
FIRST INT ONLY(CONV)
MULTIPLE INTER(PROMPTS)
MULTIPLE INTER(CONV)
Lateral Sep(meters) From Most Energetic Muon Detector(Track by Track)
average muons per m
2
per event
10
-15
10
-14
10
-13
10
-12
10
-11
10
-10
10
-9
10
-8
10
-7
10
-6
10
-5
10
-4
10
-3
10
-2
10
-1
1
0
50
100
150
200
250
300
350
400
450
500
LDF Of Muons AMANDA depth(CR Spectrum Zenith80)
FIRST INT ONLY(PROMPTS)
FIRST INT ONLY(CONV)
MULTIPLE INTER(PROMPTS)
MULTIPLE INTER(CONV)
Lateral Sep(meters) From Most Energetic Muon Detector(Track by Track)
average muons per m
2
per event
10
-15
10
-14
10
-13
10
-12
10
-11
10
-10
10
-9
10
-8
10
-7
10
-6
10
-5
10
-4
10
-3
10
-2
10
-1
1
0
50
100
150
200
250
300
350
400
450
500
LDF Of Muons AMANDA depth(1-1000PeV primaries)
PROTON SHOWERS(PROMPTS)
PROTON SHOWERS(CONV)
IRON SHOWERS(PROMPTS)
IRON SHOWERS(CONV)
Lateral Sep(meters) From Most Energetic Muon Detector(Track by Track)
average muons per m
2
per event
10
-15
10
-14
10
-13
10
-12
10
-11
10
-10
10
-9
10
-8
10
-7
10
-6
10
-5
10
-4
10
-3
10
-2
10
-1
1
10
0
50
100
150
200
250
300
350
400
450
500
LDF Of Muons AMANDA depth(1PeV primaries only)
PROTON SHOWERS(PROMPTS)
PROTON SHOWERS(CONV)
IRON SHOWERS(PROMPTS)
IRON SHOWERS(CONV)
Lateral Sep(meters) From Most Energetic Muon Detector(Track by Track)
average muons per m
2
per event
10
-15
10
-14
10
-13
10
-12
10
-11
10
-10
10
-9
10
-8
10
-7
10
-6
10
-5
10
-4
10
-3
10
-2
10
-1
1
10
0
50
100
150
200
250
300
350
400
450
500
Figure 7.8: Shows the average number of muons produced per event as a
function the lateral separation from the most energetic muon at detector
for showers initiated by the full cosmic ray spectrum, full cosmic ray spec-
trum for zenith>80 degrees, for primaries in the energy range of 1-1000
PeV and monoenergetic primary energy of 1 PeV with no showering (only
the first interaction) and after the full shower develops (multiple inter-
actions) with events containing atleast 1 prompt muon (produced from
a charmed particle) tagged as “PROMPTS” and for no prompt muon
involved as “CONV”. All data is normalized to 1 years worth lifetime

80
LDF Of Muons AMANDA depth(Cosmic Ray Spectrum)
PROTON SHOWERS(PROMPTS)
PROTON SHOWERS(CONV)
IRON SHOWERS(PROMPTS)
IRON SHOWERS(CONV)
Lateral Sep(meters) From Most Energetic Muon Detector(Track by Track)
average muons per m
2
per event
10
-15
10
-14
10
-13
10
-12
10
-11
10
-10
10
-9
10
-8
10
-7
10
-6
10
-5
10
-4
10
-3
10
-2
10
-1
1
0
50
100
150
200
250
300
350
400
450
500
LDF Of Muons AMANDA depth(CR Spectrum Zenith80)
PROTON SHOWERS(PROMPTS)
PROTON SHOWERS(CONV)
IRON SHOWERS(PROMPTS)
IRON SHOWERS(CONV)
Lateral Sep(meters) From Most Energetic Muon Detector(Track by Track)
average muons per m
2
per event
10
-15
10
-14
10
-13
10
-12
10
-11
10
-10
10
-9
10
-8
10
-7
10
-6
10
-5
10
-4
10
-3
10
-2
10
-1
1
0
50
100
150
200
250
300
350
400
450
500
LDF Of Muons AMANDA depth(1-1000PeV primaries)
PROTON SHOWERS(PROMPTS)
PROTON SHOWERS(CONV)
IRON SHOWERS(PROMPTS)
IRON SHOWERS(CONV)
Lateral Sep(meters) From Most Energetic Muon Detector(Track by Track)
average muons per m
2
per event
10
-15
10
-14
10
-13
10
-12
10
-11
10
-10
10
-9
10
-8
10
-7
10
-6
10
-5
10
-4
10
-3
10
-2
10
-1
1
10
0
50
100
150
200
250
300
350
400
450
500
LDF Of Muons AMANDA depth(1PeV primaries only)
PROTON SHOWERS(PROMPTS)
PROTON SHOWERS(CONV)
IRON SHOWERS(PROMPTS)
IRON SHOWERS(CONV)
Lateral Sep(meters) From Most Energetic Muon Detector(Track by Track)
average muons per m
2
per event
10
-15
10
-14
10
-13
10
-12
10
-11
10
-10
10
-9
10
-8
10
-7
10
-6
10
-5
10
-4
10
-3
10
-2
10
-1
1
10
0
50
100
150
200
250
300
350
400
450
500
Figure 7.9: Shows the average number of muons produced per event as a
function the lateral separation from the most energetic muon at detector
for showers initiated by the full cosmic ray spectrum, full cosmic ray spec-
trum for zenith>80 degrees, for primaries in the energy range of 1-1000
PeV and monoenergetic primary energy of 1 PeV after the full shower de-
velops (multiple interactions) with showers produced by protons and iron
identified separately. All data is normalized to 1 years worth lifetime

81
Muon Energy Plot(cosmic ray spectrum)
first int(prompts)
first int(conv)
multiple inter(prompts)
multiple inter(conv)
log10(Total Muon Energy)GeV at Surface
Counts in Bin
10
-4
10
-3
10
-2
10
-1
1
10
10
2
10
3
10
4
10
5
10
6
10
7
10
8
10
9
10
10
10
11
1
2
3
4
5
6
7
8
Muon Energy Plot(cossmic ray spectrum for zenith80)
first int(prompts)
first int(conv)
multiple inter(prompts)
multiple inter(conv)
log10(Total Muon Energy)GeV at Surface
Counts in Bin
10
-4
10
-3
10
-2
10
-1
1
10
10
2
10
3
10
4
10
5
10
6
10
7
10
8
10
9
10
10
10
11
1
2
3
4
5
6
7
8
Muon Energy Plot(1-1000PeV)
first int(prompts)
first int(conv)
multiple inter(prompts)
multiple inter(conv)
log10(Total Muon Energy)GeV at Surface
Counts in Bin
10
-4
10
-3
10
-2
10
-1
1
10
10
2
10
3
10
4
10
5
10
6
10
7
10
8
10
9
10
10
10
11
1
2
3
4
5
6
7
8
Muon Energy Plot(1PeV primaries)
first int(prompts)
first int(conv)
multiple inter(prompts)
multiple inter(conv)
log10(Total Muon Energy)GeV at Surface
Counts in Bin
10
-4
10
-3
10
-2
10
-1
1
10
10
2
10
3
10
4
10
5
10
6
10
7
10
8
10
9
10
10
10
11
1
2
3
4
5
6
7
8
Figure 7.10: Shows the sum total of surface energy of all the muons in an
event for showers initiated by the full cosmic ray spectrum, full cosmic ray
spectrum for zenith>80 degrees, for primaries in the energy range of 1-
1000 PeV and monoenergetic primary energy of 1 PeV with no showering
(only the first interaction) and after the full shower develops (multiple
interactions) with events containing atleast 1 prompt muon (produced
from a charmed particle) tagged as “PROMPTS” and for no prompt muon
involved as “CONV”. All data is normalized to 1 years worth lifetime

82
Muon Energy Plot(cosmic ray spectrum)
first int(prompts)
first int(conv)
multiple inter(prompts)
multiple inter(conv)
log10(Total Muon Energy)GeV at Detector
Counts in Bin
10
-4
10
-3
10
-2
10
-1
1
10
10
2
10
3
10
4
10
5
10
6
10
7
10
8
10
9
10
10
10
11
1
2
3
4
5
6
7
8
Muon Energy Plot(cosmic ray spectrum for zenith80)
first int only(prompts)
first int only(conv)
after multiple inter(prompts)
after multiple inter(conv)
log10(Total Muon Energy)GeV at Detector
Counts in Bin
10
-4
10
-3
10
-2
10
-1
1
10
10
2
10
3
10
4
10
5
10
6
10
7
10
8
10
9
10
10
10
11
1
2
3
4
5
6
7
8
Muon Energy Plot(1-1000PeV)
first int(prompts)
first int(conv)
multiple inter(prompts)
multiple inter(conv)
log10(Total Muon Energy)GeV at Detector
Counts in Bin
10
-4
10
-3
10
-2
10
-1
1
10
10
2
10
3
10
4
10
5
10
6
10
7
10
8
10
9
10
10
10
11
1
2
3
4
5
6
7
8
Muon Energy Plot(1PeV primary)
first int(prompts)
first int(conv)
multiple inter(prompts)
multiple inter(conv)
log10(Total Muon Energy)GeV at Detector
Counts in Bin
10
-4
10
-3
10
-2
10
-1
1
10
10
2
10
3
10
4
10
5
10
6
10
7
10
8
10
9
10
10
10
11
1
2
3
4
5
6
7
8
Figure 7.11: Shows the sum total of energy of all the muons in an event
at the detector for showers initiated by the full cosmic ray spectrum, full
cosmic ray spectrum for zenith>80 degrees, for primaries in the energy
range of 1-1000 PeV and monoenergetic primary energy of 1 PeV with
no showering (only the first interaction) after the full shower develops
(multiple interactions) with events containing atleast one prompt muon
(produced from a charmed particle) tagged as “PROMPTS” and for no
prompt muon involved as “CONV”. All data is normalized to 1 years
worth lifetime

83
Zenith Angle Distribution(cosmic ray spectrum)
FIRST INT ONLY(PROMPTS)
FIRST INT ONLY(CONV)
MULTIPLE INTER(PROMPTS)
MULTIPLE INTER(CONV)
Cosine(TrueZenith Angle)
Counts Per Year
10
-3
10
-2
10
-1
1
10
10
2
10
3
10
4
10
5
10
6
10
7
10
8
10
9
10
10
10
11
10
12
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
Zenith Angle Distribution(1PEV SHOWERS)
FIRST INT ONLY(PROMPTS)
FIRST INT ONLY(CONV)
MULTIPLE INTER(PROMPTS)
MULTIPLE INTER(CONV)
Cosine(TrueZenith Angle)
Counts Per Year
10
-3
10
-2
10
-1
1
10
10
2
10
3
10
4
10
5
10
6
10
7
10
8
10
9
10
10
10
11
10
12
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
Zenith Angle Distribution(1-1000PeV)
FIRST INT ONLY(PROMPTS)
FIRST INT ONLY(CONV)
MULTIPLE INTER(PROMPTS)
MULTIPLE INTER(CONV)
Cosine(TrueZenith Angle)
Counts Per Year
10
-3
10
-2
10
-1
1
10
10
2
10
3
10
4
10
5
10
6
10
7
10
8
10
9
10
10
10
11
10
12
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
Zenith Angle Distribution(1PeV primaries)
FIRST INT(PROMPTS)
FIRST INT(CONV)
MULTIPLE INT(PROMPTS)
MULTIPLE INT(CONV)
Cosine(TrueZenith Angle)
Counts Per Year
10
-3
10
-2
10
-1
1
10
10
2
10
3
10
4
10
5
10
6
10
7
10
8
10
9
10
10
10
11
10
12
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
Figure 7.12: Shows the zenith angle distribution of showers initiated by
the full cosmic ray spectrum, full cosmic ray spectrum for zenith>80 de-
grees, for primaries in the energy range of 1-1000 PeV and monoenergetic
primary energy of 1 PeV with no showering (only the first interaction) and
after the full shower develops (multiple interactions) with events contain-
ing atleast 1 prompt muon (produced from a charmed particle) tagged as
“PROMPTS” and for no prompt muon involved as “CONV”. All data is
normalized to 1 years worth lifetime

84
Chapter 8
Results and Conclusions
8.1 Shape Analysis
In a usual AMANDA analysis, confidence intervals are constructed based on the
number of events in the final data sample after taking into account the predicted back-
ground and signal events. Statistical and systematic uncertainities are incorporated
into the confidence interval based on the work of Feldman and Cousin [7]. These cut
and count methods do not consider shape information about the predicted or observed
spectrum of events. This analysis will compare the shape and normalization of the
observed data to simulations based on the model for background and signal flux. The
number of optical modules fired is a powerful variable used in this analysis. By using
shape and normalization information for optical modules fired it is hoped that the
atmospheric muon background can be better understood and limits could be placed
on prompt atmospheric muon models.

85
8.2 Simulation and Fitting Procedure
The simulation for the signal (prompt muon) is done using the DPMJET model
while the background (conventional muon) was done using SYBILL. Two ice models
AHA and Millenium were simulated. The AHA model is flatter compared with the
Millenium [10]. While the Millenium model is simulated both for the background and
signal, the AHA model is simulated only for the background and the simulation for
AHA signal is obtained by appropriate scaling. It is the tradition of most AMANDA
related analysis to use the hybrid frequentist-Bayesian method based on the work of
Cousins and Highland [8] to construct a confidence belt for the signal. The nature
of downgoing muon analysis presents the challenge of dealing with unclear systematic
uncertainities at high energies so in this analysis we use slope fitting of the background
and signal to the experimental data to determine limits on the assumed signal spectra.
8.3 Fitting Procedure
The number of channel hit spectra at the final stage (after all the cuts) for the
experimental data is fitted using the simulated background and signal (assuming there
is no preference for the AHA or the Millenium model). The scaling parameters for
the simulated background and the signal for the Millenium model are denoted by
f
C
?
MIL
and f
PROMPT
respectively and f
C
?
AHA
to denote AHA background and these
are parameters are fitted using a chisquare based approach. The AHA signal was not
simulated but was scaled using the fitted background ratio of the two models. The
idea of slope fitting would be to minimize the chisquare while fitting these parameters
with the experimental data. For purposes of this minimization only bins numbered

86
6 to 17 (number of optical modules fired greater than 180) are used as it would be
more representative of the high-energy response of the detector. The best fit values
of f
C
?
MIL
, f
PROMPT
, f
C
?
AHA
are obtained by minimizing the chisquare and confidence
bands on the scaling of the signal are constructed. In figure 8.1 the best chisquare is
shown at different allowed levels of signal. Error contours for the fraction of allowed
AHA and Millenium background forcing signal to be zero (demonstrative purposes)
and the best fit value of signal and the 90% level for signal contribution are also shown
in figures 8.2, 8.3 and 8.4 respectively. The spectra before scaling and after scaling for
the best fit values are shown in figures 8.5 and 8.6 respectively. Fitting equations and
scaling parameters fitted for are described by the below equations.
χ
2
=
?
bins
(N
DATA
i
?
N
PRED
i
)
2
σ
2
i
(8.1)
N
PRED
i
= f
C
?
AHA
N
PRED
iC
?
AHA
+ f
C
?
MIL
N
PRED
iC
?
MIL
+ f
PROMPT
?
N
PRED
iP
?
MIL
+
f
C
?
AHA
f
C
?
MIL
N
PRED
iP
?
AHA
?
(8.2)
N
DATA
i
is observed experimental data counts in each bin. N
PRED
i
is predicted
simulation data counts in each bin. σ
2
i
is the variance given by ROOT after reweight-
ing. f
C
?
AHA
is the unknown scale factor for Conventional muon background using
AHA model. f
C
?
MIL
is the unknown scale factor for conventional muon background
using Millenium model. f
PROMPT
is the unknown scale factor for prompt muon signal
using the Millenium model. N
PRED
iC
?
AHA
is the predicted conventional background using
AHA model. N
PRED
iC
?
MIL
is the predicted conventional background using the Millenium

87
model. N
PRED
iP
?
MIL
is the predicted prompt signal using the Millenium model. N
PRED
iP
?
AHA
is the predicted prompt signal using the AHA model.
8.4 Prompt Atmospheric Neutrino Upper limits
Since prompt muons have a harder (less steep) spectrum than the conventional
atmospheric neutrinos, it is possible to search for a prompt neutrino flux by separating
the two event classes in energy. A limit on prompt muons is equivalent to a limit of
prompt neutrinos [31]. The Naumov RQPM model is a non-perturbative model of
prompt atmospheric neutrinos and incorporates data from primary cosmic ray and
hadronic interaction experiments. The upper limit of this model at 90% confidence
level using shape based spectral fitting is 3.67·Φ
RQPM
.
8.5 Discussion for Better Analysis in Future
The biggest problem that makes this analysis tricky is the fact that we don’t
have a good model for prompt muon production and DPMJET-II.55, far from being
accurate is the best we could get to use in conjunction with SYBILL conventional muon
Monte Carlo. Events containing one or more muons whose parent is a charmed particle
were tagged as prompt muons and the experimental data was fitted to incorporate this
component to minimize the chisquare and this was used to derive an upper limit on
the charm cross-section. This approach suffers from the fact that the prompt muon
event production rate the way it is defined is not linear with an increase in charm
cross-section and hence placing upper limits this way is far from the correct way of
doing it. A correct approach would be to change the charm cross-section up and down
and produce muon event rates for the prompt muon signal and construct the chisquare

88
surface for the fit to the experimental data and derive upper limits on the cross section
based on it.
Further another issue in this analysis is that the SYBILL conventional Monte
Carlo doesn’t produce any charmed particles and this is used in combination with a
signal simulation of DPMJET that samples charmed particles from the charm cross-
section, these are not two mutually exclusive sets when added up to compare to the
experimental data and is a rough approximation to derive upper-limits on the signal.
8.6 Conclusion
This analysis placed an upper limit on the prompt neutrino model of RQPM
using prompt muon analysis and set a constraint on the model. This result is the first
of its kind in using a downgoing muon analysis to set an upper limit on the prompt
neutrinos. AMANDA-II has now been integrated into IceCube. The main aim of Ice-
Cube is to detect extra-terrestrial neutrinos and the level of uncertainty on the prompt
neutrino flux is of great interest in calculating the sensitivity of IceCube experiment to
extra-terrastrial neutrinos. Using a downgoing muon analysis we constrain the RQPM
model to a factor of 3.67 at 90% confidence.

89
Relative Signal contribution (DPMJET model)
0
0.5
1
1.5
2
2.5
3
3.5
4
chi square
12
12.5
13
13.5
14
14.5
15
15.5
Chisquare as a function of Signal Contribution(No Smoothing)
Figure 8.1: The minimized value of chisquare is shown for different levels
of signal.
Fitted fraction of AHA background
0
0.02 0.04 0.06 0.08
0.1 0.12 0.14 0.16 0.18
0.2
Fitted fraction of Millenium background
1.2
1.25
1.3
1.35
1.4
1.45
1.5
1.55
1.6
14.5
15
15.5
16
16.5
17
17.5
18
18.5
Parameters of Background fit to Data (bins 6-16)
Figure 8.2: The elliptical contours of chisquare for the fraction of Mille-
nium and AHA backgrounds are shown forcing the signal contribution to
be zero while making a fit to the data.

90
Fitted fraction of AHA background
0
0.02 0.04 0.06 0.08
0.1 0.12 0.14 0.16 0.18
0.2
Fitted fraction of Millenium background
1.2
1.25
1.3
1.35
1.4
1.45
1.5
1.55
1.6
12
12.5
13
13.5
14
14.5
15
15.5
16
16.5
Parameters of Background fit to Data at best Signal fit(bins 6-16)
Figure 8.3: The elliptical contours of chisquare for the fraction of Mille-
nium and AHA backgrounds are shown for best fit value of signal while
making a fit to the data.
Fitted fraction of AHA background
0
0.02 0.04 0.06 0.08
0.1 0.12 0.14 0.16 0.18
0.2
Fitted fraction of Millenium background
1.2
1.25
1.3
1.35
1.4
1.45
1.5
1.55
1.6
15
15.5
16
16.5
17
17.5
18
18.5
19
Parameters of Background fit to Data at 90 percent upper limit Signal fit(bins 6-16)
Figure 8.4: The elliptical contours of chisquare for the fraction of Mille-
nium and AHA backgrounds are shown for the allowed level of signal at
90% confidence level while making a fit to the data.

91
Number of Optical Module Hit
0
100
200
300
400
500
Counts
-2
10
-1
10
1
10
2
10
3
10
4
10
5
10
6
10
7
10
NChannel Before Scaling
SYBILL Millenium bef Scaling (2005)
SYBILL AHA bef Scaling (2005)
Experimental Data(2005)
Signal Mill DPMJET(2005)
Signal AHA DPMJET(2005)
Figure 8.5: The signal and background spectra for the AHA and Millenium
models together with the minimum bias experimental data before fitting
are shown.
Number of Optical Module Hit
0
100
200
300
400
500
Counts
-2
10
-1
10
1
10
2
10
3
10
4
10
5
10
6
10
7
10
NChannel After Scaling(Bins>6)
SYBILL Millenium aft Scaling (2005)
SYBILL AHA aft Scaling (2005)
Experimental Data(2005)
Signal Mill DPMJET Scaled(2005)
Signal AHA DPMJET Scaled(2005)
Figure 8.6: The scaled levels at the best fit values of signal and background
spectra for the AHA and Millenium models are shown together with the
minimum bias experimental data.

92
Bibliography
[1] Jim Green, private communication.
[2] C.G.S. Costa, Astropart. Phys. 16 , 193 (2001).
[3] G. Fiorentini, A. Naumov, and F.L. Villante, Phys. Lett. B 510 , 173 (2001).
[4] G. Gelmini, P. Gondolo and G. Varieschi Measuring the prompt atmospheric
neutrino flux with down-going muons in neutrino telescopes , Jan 2003 Phys.
Rev. D 67, 017301.
[5] E.V. Bugaev et al., Il Nuovo Cimento 12C , No. 1, 41 (1989).
[6] E.V. Bugaev et al., Phys. Rev. D 58 , 054001 (1998).
[7] Ochs, Nucl. Phys. B118, 397 (1977).
[8] E. V. Bugaev, A. Misaki, V. A. Naumov, T. S. Sinegovskaya, S. I. Sinegov-
sky, and N. Takahashi Atmospheric muon flux at sea level,underground and
under-water, Phys. Rev. D 58, 054001 (1998), Issue 5, September 1998.
[9] Athina Meli, private communication.
[10] D. Heck, J. Knapp, J.N. Capdevielle, G. Schatz, and T. Thouw,Tech. Rep.
FZKA 6019, Forschungszentrum Karlsruhe (1998).

93
[11] http://www.icecube.wisc.edu/ boersma/filter2005/webpage/index.html.
[12] D. Pandel, Diploma Thesis, Humboldt-Universit¨at zu Berlin, Berlin, Ger-
many, February 1996.
[13] http://www.icecube.wisc.edu/ jkelley/simulation/dcorsweighting.pdf.
[14] http://www.icecube.wisc.edu/ jkelley/simulation/livetimes.pdf.
[15] J. Horandel, Astropart.Phys., 19(2003) 193.
[16] http://www.icecube.wisc.edu/ jkelley/simulation/aha/aha.html.
[17] H. Kampert, Nuclear Physics B (Proc. Suppl.), 165 (2007).
[18] J. Ahrens et al., Phys. Rev. Lett. 90 , 251101 (2003).
[19] P. Desiati et al. , Response of AMANDA-II to Cosmic Ray Muons,Proc. of
28th ICRC, Hamburg, Germany (2003).
[20] D. Heck, T. Pierog, J. Knapp, CORSIKA: an air shower simulation program:
http://www-ik.fzk.de/corsika/.
[21] http://www.fluka.org; A. Fass`o, A. Ferrari, J. Ranft and P.R. Sala, FLUKA:
a multi-particle transport code,CERN-2005-10 (2005), INFN/T C
05/11,
SLAC-R-773;A. Fasso`o et al. , The physics models of FLUKA: status and
recent developments, Computing in High Energy and Nuclear Physics 2003
Conference (CHEP2003), La Jolla, CA, USA, March 24-28, 2003, (paper
MOMT005), eConf C0303241 (2003), [arXiv:hep-ph/0306267].

94
[22] Spencer Klein, Studying High pT Muons in Cosmic-Ray Air Showers,2 Dec
2006 [arXiv:stro-ph/0612051 v1].
[23] N.N. Kalmykov et al. , Nucl. Phys. B (Proc. Suppl.) 52B (1997) 17.
[24] S. Ostapchenko, Nucl. Phys. B (Proc. Suppl.) 151 (2006) 143 [arXiv:hep-
ph/0412332].
[25] R.S. Fletcher et al. , Phys. Rev. D50 (1994) 5710.
[26] T. Sjostrand A brief introduction to PYTHIA 8.080.
[27] G.J. Feldman and R.D. Cousins Unified approach to the classical statistical
analysis of small signals , Phys. Rev. D 57, 3873-3889 (1998).
[28] R.D. Cousins amd V.L. Highland Incorporating systematic uncertainties into
an upper limit , Nucl. Ins. Meth. Phys. Res. A32 0 331 (1992).

Back to top