The IceCube Collaboration:
    Contributions to the 31
    st
    International Cosmic Ray Conference*
    Łό
    d
    ź
    , Poland, 7-15 July 2009
    ------------------
    *
    Includes some related papers submitted by individual members of
    the Collaboration

    This page intentionally left blank

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁOD´ Z´ 2009
    1
    IceCube COLLABORATION
    R. Abbasi
    24
    , Y. Abdou
    18
    , M. Ackermann
    36
    , J. Adams
    13
    , J. Aguilar
    24
    , M. Ahlers
    28
    , K. Andeen
    24
    ,
    J. Auffenberg
    35
    , X. Bai
    27
    , M. Baker
    24
    , S. W. Barwick
    20
    , R. Bay
    7
    , J. L. Bazo Alba
    36
    , K. Beattie
    8
    ,
    J. J. Beatty
    15,16
    , S. Bechet
    10
    , J. K. Becker
    17
    , K.-H. Becker
    35
    , M. L. Benabderrahmane
    36
    ,
    J. Berdermann
    36
    , P. Berghaus
    24
    , D. Berley
    14
    , E. Bernardini
    36
    , D. Bertrand
    10
    , D. Z. Besson
    22
    ,
    M. Bissok
    1
    , E. Blaufuss
    14
    , D. J. Boersma
    24
    , C. Bohm
    30
    , J. Bolmont
    36
    , S. Boser¨
    36
    , O. Botner
    33
    ,
    L. Bradley
    32
    , J. Braun
    24
    , D. Breder
    35
    , T. Castermans
    26
    , D. Chirkin
    24
    , B. Christy
    14
    , J. Clem
    27
    ,
    S. Cohen
    21
    , D. F. Cowen
    32,31
    , M. V. D’Agostino
    7
    , M. Danninger
    30
    , C. T. Day
    8
    , C. De Clercq
    11
    ,
    L. Demirors¨
    21
    , O. Depaepe
    11
    , F. Descamps
    18
    , P. Desiati
    24
    , G. de Vries-Uiterweerd
    18
    , T. DeYoung
    32
    ,
    J. C. Diaz-Velez
    24
    , J. Dreyer
    17
    , J. P. Dumm
    24
    , M. R. Duvoort
    34
    , W. R. Edwards
    8
    , R. Ehrlich
    14
    ,
    J. Eisch
    24
    , R. W. Ellsworth
    14
    , O. Engdegard˚
    33
    , S. Euler
    1
    , P. A. Evenson
    27
    , O. Fadiran
    4
    ,
    A. R. Fazely
    6
    , T. Feusels
    18
    , K. Filimonov
    7
    , C. Finley
    24
    , M. M. Foerster
    32
    , B. D. Fox
    32
    ,
    A. Franckowiak
    9
    , R. Franke
    36
    , T. K. Gaisser
    27
    , J. Gallagher
    23
    , R. Ganugapati
    24
    , L. Gerhardt
    8,7
    ,
    L. Gladstone
    24
    , A. Goldschmidt
    8
    , J. A. Goodman
    14
    , R. Gozzini
    25
    , D. Grant
    32
    , T. Griesel
    25
    ,
    A. Groß
    13,19
    , S. Grullon
    24
    , R. M. Gunasingha
    6
    , M. Gurtner
    35
    , C. Ha
    32
    , A. Hallgren
    33
    ,
    F. Halzen
    24
    , K. Han
    13
    , K. Hanson
    24
    , Y. Hasegawa
    12
    , J. Heise
    34
    , K. Helbing
    35
    , P. Herquet
    26
    ,
    S. Hickford
    13
    , G. C. Hill
    24
    , K. D. Hoffman
    14
    , K. Hoshina
    24
    , D. Hubert
    11
    , W. Huelsnitz
    14
    ,
    J.-P. Hulߨ
    1
    , P. O. Hulth
    30
    , K. Hultqvist
    30
    , S. Hussain
    27
    , R. L. Imlay
    6
    , M. Inaba
    12
    , A. Ishihara
    12
    ,
    J. Jacobsen
    24
    , G. S. Japaridze
    4
    , H. Johansson
    30
    , J. M. Joseph
    8
    , K.-H. Kampert
    35
    , A. Kappes
    24,a
    ,
    T. Karg
    35
    , A. Karle
    24
    , J. L. Kelley
    24
    , P. Kenny
    22
    , J. Kiryluk
    8,7
    , F. Kislat
    36
    , S. R. Klein
    8,7
    ,
    S. Klepser
    36
    , S. Knops
    1
    , G. Kohnen
    26
    , H. Kolanoski
    9
    , L. Kopk¨ e
    25
    , M. Kowalski
    9
    , T. Kowarik
    25
    ,
    M. Krasberg
    24
    , K. Kuehn
    15
    , T. Kuwabara
    27
    , M. Labare
    10
    , S. Lafebre
    32
    , K. Laihem
    1
    ,
    H. Landsman
    24
    , R. Lauer
    36
    , H. Leich
    36
    , D. Lennarz
    1
    , A. Lucke
    9
    , J. Lundberg
    33
    , J. Lunemann¨
    25
    ,
    J. Madsen
    29
    , P. Majumdar
    36
    , R. Maruyama
    24
    , K. Mase
    12
    , H. S. Matis
    8
    , C. P. McParland
    8
    ,
    K. Meagher
    14
    , M. Merck
    24
    , P. Mesz´ ar´ os
    31,32
    , E. Middell
    36
    , N. Milke
    17
    , H. Miyamoto
    12
    , A. Mohr
    9
    ,
    T. Montaruli
    24,b
    , R. Morse
    24
    , S. M. Movit
    31
    , K. Munich¨
    17
    , R. Nahnhauer
    36
    , J. W. Nam
    20
    ,
    P. Nießen
    27
    , D. R. Nygren
    8,30
    , S. Odrowski
    19
    , A. Olivas
    14
    , M. Olivo
    33
    , M. Ono
    12
    , S. Panknin
    9
    ,
    S. Patton
    8
    , C. Per´ ez de los Heros
    33
    , J. Petrovic
    10
    , A. Piegsa
    25
    , D. Pieloth
    36
    , A. C. Pohl
    33,c
    ,
    R. Porrata
    7
    , N. Potthoff
    35
    , P. B. Price
    7
    , M. Prikockis
    32
    , G. T. Przybylski
    8
    , K. Rawlins
    3
    , P. Redl
    14
    ,
    E. Resconi
    19
    , W. Rhode
    17
    , M. Ribordy
    21
    , A. Rizzo
    11
    , J. P. Rodrigues
    24
    , P. Roth
    14
    , F. Rothmaier
    25
    ,
    C. Rott
    15
    , C. Roucelle
    19
    , D. Rutledge
    32
    , D. Ryckbosch
    18
    , H.-G. Sander
    25
    , S. Sarkar
    28
    ,
    K. Satalecka
    36
    , S. Schlenstedt
    36
    , T. Schmidt
    14
    , D. Schneider
    24
    , A. Schukraft
    1
    , O. Schulz
    19
    ,
    M. Schunck
    1
    , D. Seckel
    27
    , B. Semburg
    35
    , S. H. Seo
    30
    , Y. Sestayo
    19
    , S. Seunarine
    13
    , A. Silvestri
    20
    ,
    A. Slipak
    32
    , G. M. Spiczak
    29
    , C. Spiering
    36
    , M. Stamatikos
    15
    , T. Stanev
    27
    , G. Stephens
    32
    ,
    T. Stezelberger
    8
    , R. G. Stokstad
    8
    , M. C. Stoufer
    8
    , S. Stoyanov
    27
    , E. A. Strahler
    24
    ,
    T. Straszheim
    14
    , K.-H. Sulanke
    36
    , G. W. Sullivan
    14
    , Q. Swillens
    10
    , I. Taboada
    5
    , O. Tarasova
    36
    ,
    A. Tepe
    35
    , S. Ter-Antonyan
    6
    , C. Terranova
    21
    , S. Tilav
    27
    , M. Tluczykont
    36
    , P. A. Toale
    32
    , D. Tosi
    36
    ,
    D. Turcanˇ
    14
    , N. van Eijndhoven
    34
    , J. Vandenbroucke
    7
    , A. Van Overloop
    18
    , B. Voigt
    36
    ,
    C. Walck
    30
    , T. Waldenmaier
    9
    , M. Walter
    36
    , C. Wendt
    24
    , S. Westerhoff
    24
    , N. Whitehorn
    24
    ,
    C. H. Wiebusch
    1
    , A. Wiedemann
    17
    , G. Wikstrom¨
    30
    , D. R. Williams
    2
    , R. Wischnewski
    36
    ,
    H. Wissing
    1,14
    , K. Woschnagg
    7
    , X. W. Xu
    6
    , G. Yodh
    20
    , S. Yoshida
    12
    1
    III Physikalisches Institut, RWTH Aachen University, D-52056 Aachen, Germany
    2
    Dept. of Physics and Astronomy, University of Alabama, Tuscaloosa, AL 35487, USA
    3
    Dept. of Physics and Astronomy, University of Alaska Anchorage, 3211 Providence Dr., Anchorage, AK 99508,
    USA
    4
    CTSPS, Clark-Atlanta University, Atlanta, GA 30314, USA
    5
    School of Physics and Center for Relativistic Astrophysics, Georgia Institute of Technology, Atlanta, GA 30332.
    USA
    6
    Dept. of Physics, Southern University, Baton Rouge, LA 70813, USA
    7
    Dept. of Physics, University of California, Berkeley, CA 94720, USA
    8
    Lawrence Berkeley National Laboratory, Berkeley, CA 94720, USA
    9
    Institut fur¨ Physik, Humboldt-Universitat¨ zu Berlin, D-12489 Berlin, Germany
    10
    Universite´ Libre de Bruxelles, Science Faculty CP230, B-1050 Brussels, Belgium
    11
    Vrije Universiteit Brussel, Dienst ELEM, B-1050 Brussels, Belgium
    12
    Dept. of Physics, Chiba University, Chiba 263-8522, Japan

    2
    PROCEEDINGS OF THE 31
    st
    ICRC, ŁOD´ Z´ 2009
    13
    Dept. of Physics and Astronomy, University of Canterbury, Private Bag 4800, Christchurch, New Zealand
    14
    Dept. of Physics, University of Maryland, College Park, MD 20742, USA
    15
    Dept. of Physics and Center for Cosmology and Astro-Particle Physics, Ohio State University, Columbus, OH
    43210, USA
    16
    Dept. of Astronomy, Ohio State University, Columbus, OH 43210, USA
    17
    Dept. of Physics, TU Dortmund University, D-44221 Dortmund, Germany
    18
    Dept. of Subatomic and Radiation Physics, University of Gent, B-9000 Gent, Belgium
    19
    Max-Planck-Institut fur¨ Kernphysik, D-69177 Heidelberg, Germany
    20
    Dept. of Physics and Astronomy, University of California, Irvine, CA 92697, USA
    21
    Laboratory for High Energy Physics, Ecole´
    Polytechnique Fed´ er´ ale, CH-1015 Lausanne, Switzerland
    22
    Dept. of Physics and Astronomy, University of Kansas, Lawrence, KS 66045, USA
    23
    Dept. of Astronomy, University of Wisconsin, Madison, WI 53706, USA
    24
    Dept. of Physics, University of Wisconsin, Madison, WI 53706, USA
    25
    Institute of Physics, University of Mainz, Staudinger Weg 7, D-55099 Mainz, Germany
    26
    University of Mons-Hainaut, 7000 Mons, Belgium
    27
    Bartol Research Institute and Department of Physics and Astronomy, University of Delaware, Newark, DE
    19716, USA
    28
    Dept. of Physics, University of Oxford, 1 Keble Road, Oxford OX1 3NP, UK
    29
    Dept. of Physics, University of Wisconsin, River Falls, WI 54022, USA
    30
    Oskar Klein Centre and Dept. of Physics, Stockholm University, SE-10691 Stockholm, Sweden
    31
    Dept. of Astronomy and Astrophysics, Pennsylvania State University, University Park, PA 16802, USA
    32
    Dept. of Physics, Pennsylvania State University, University Park, PA 16802, USA
    33
    Dept. of Physics and Astronomy, Uppsala University, Box 516, S-75120 Uppsala, Sweden
    34
    Dept. of Physics and Astronomy, Utrecht University/SRON, NL-3584 CC Utrecht, The Netherlands
    35
    Dept. of Physics, University of Wuppertal, D-42119 Wuppertal, Germany
    36
    DESY, D-15735 Zeuthen, Germany
    a
    affiliated with Universitat¨ Erlangen-Nurnber¨
    g, Physikalisches Institut, D-91058, Erlangen, Germany
    b
    on leave of absence from Universita` di Bari and Sezione INFN, Dipartimento di Fisica, I-70126, Bari, Italy
    c
    affiliated with School of Pure and Applied Natural Sciences, Kalmar University, S-39182 Kalmar, Sweden
    Acknowledgments
    We acknowledge the support from the following agencies: U.S. National Science Foundation-Office of Polar
    Program, U.S. National Science Foundation-Physics Division, University of Wisconsin Alumni Research Foun-
    dation, U.S. Department of Energy, and National Energy Research Scientific Computing Center, the Louisiana
    Optical Network Initiative (LONI) grid computing resources; Swedish Research Council, Swedish Polar Research
    Secretariat, and Knut and Alice Wallenberg Foundation, Sweden; German Ministry for Education and Research
    (BMBF), Deutsche Forschungsgemeinschaft (DFG), Germany; Fund for Scientific Research (FNRS-FWO), Flanders
    Institute to encourage scientific and technological research in industry (IWT), Belgian Federal Science Policy Office
    (Belspo); the Netherlands Organisation for Scientific Research (NWO); M. Ribordy acknowledges the support of
    the SNF (Switzerland); A. Kappes and A. Groß acknowledge support by the EU Marie Curie OIF Program;
    J. P. Rodrigues acknowledge support by the Capes Foundation, Ministry of Education of Brazil.

    1339: IceCube
    Albrecht Karle, for the IceCube Collaboration (Highlight paper)
    0653: All-Sky Point-Source Search with 40 Strings of IceCube
    Jon Dumm, Juan A. Aguilar, Mike Baker, Chad Finley, Teresa Montaruli, for the
    IceCube Collaboration
    0812: IceCube Time-Dependent Point Source Analysis Using Multiwavelength
    Information
    M. Baker, J. A. Aguilar, J. Braun, J. Dumm, C. Finley, T. Montaruli, S. Odrowski, E.
    Resconi for the IceCube Collaboration
    0960: Search for neutrino flares from point sources with IceCube
    (0908.4209)
    J. L. Bazo Alba, E. Bernardini, R. Lauer, for the IceCube Collaboration
    0987: Neutrino triggered high-energy gamma-ray follow-up with IceCube
    Robert Franke, Elisa Bernardini for the IceCube collaboration
    1173: Moon Shadow Observation by IceCube
    D.J. Boersma, L. Gladstone and A. Karle for the IceCube Collaboration
    1289: IceCube/AMANDA combined analyses for the search of neutrino sources at
    low energies
    Cécile Roucelle, Andreas Gross, Sirin Odrowski, Elisa Resconi, Yolanda Sestayo
    1127: AMANDA 7-Year Multipole Analysis
    (0906.3942)
    Anne Schukraft, Jan-Patrick Hülß for the IceCube Collaboration
    1418: Measurement of the atmospheric neutrino energy spectrum with IceCube
    Dmitry Chirkin for the IceCube collaboration
    0785: Atmospheric Neutrino Oscillation Measurements with IceCube
    Carsten Rott for the IceCube Collaboration
    1565: Direct Measurement of the Atmospheric Muon Energy Spectrum with
    IceCube
    (0909.0679)
    Patrick Berghaus for the IceCube Collaboration
    1400: Search for Diffuse High Energy Neutrinos with IceCube
    Kotoyo Hoshina for the IceCube collaboration
    1311:
    A Search For Atmospheric Neutrino-Induced Cascades with IceCube
    (0910.0215) Michelangelo D’Agostino for the IceCube Collaboration
    0882: First search for extraterrestrial neutrino-induced cascades with IceCube
    (0909.0989) Joanna Kiryluk for the IceCube Collaboration

    0708: Improved Reconstruction of Cascade-like Events in IceCube
    Eike Middell, Joseph McCartin and Michelangelo D’Agostino for the IceCube
    Collaboration
    1221: Searches for neutrinos from GRBs with the IceCube 22-string detector and
    sensitivity estimates for the full detector
    A. Kappes, P. Roth, E. Strahler, for the IceCube Collaboration
    0515: Search for neutrinos from GRBs with IceCube
    K. Meagher, P. Roth, I. Taboada, K. Hoffman, for the IceCube Collaboration
    0393: Search for GRB neutrinos via a (stacked) time profile analysis
    Martijn Duvoort and Nick van Eijndhoven for the IceCube collaboration
    0764: Optical follow-up of high-energy neutrinos detected by IceCube
    (0909.0631)
    Anna Franckowiak, Carl Akerlof, D. F. Cowen, Marek Kowalski, Ringo Lehmann,
    Torsten Schmidt and Fang Yuan for the IceCube Collaboration and for the ROTSE
    Collaboration
    0505: Results and Prospects of Indirect Searches for Dark Matter with IceCube
    Carsten Rott and Gustav Wikström for the IceCube collaboration
    1356: Search for the Kaluza-Klein Dark Matter with the AMANDA/IceCube
    Detectors
    (0906.3969), Matthias Danninger & Kahae Han for the IceCube Collaboration
    0834: Searches for WIMP Dark Matter from the Sun with AMANDA
    (0906.1615)
    James Braun and Daan Hubert for the IceCube Collaboration
    0861: The extremely high energy neutrino search with IceCube
    Keiichi Mase, Aya Ishihara and Shigeru Yoshida for the IceCube Collaboration
    0913: Study of very bright cosmic-ray induced muon bundle signatures measured
    by the IceCube detector
    Aya Ishihara for the IceCube Collaboration
    1198: Search for High Energetic Neutrinos from Supernova Explosions with
    AMANDA
    (0907.4621)
    Dirk Lennarz and Christopher Wiebusch for the IceCube Collaboration
    0549: Search for Ultra High Energy Neutrinos with AMANDA
    Andrea Silvestri for the IceCube Collaboration
    1372: Selection of High Energy Tau Neutrinos in IceCube
    Seon-Hee Seo and P. A. Toale for the IceCube Collaboration
    0484: Search for quantum gravity with IceCube and high energy atmospheric
    neutrinos,
    Warren Huelsnitz & John Kelley for the IceCube Collaboration

    0970: A First All-Particle Cosmic Ray Energy Spectrum From IceTop
    Fabian Kislat, Stefan Klepser, Hermann Kolanoski and Tilo Waldenmaier for the
    IceCube Collaboration
    0518: Reconstruction of IceCube coincident events and study of composition-
    sensitive observables using both the surface and deep detector
    Tom Feusels, Jonathan Eisch and Chen Xu for the IceCube Collaboration
    0737: Small air showers in IceTop
    Bakhtiyar Ruzybayev, Shahid Hussain, Chen Xu and Thomas Gaisser for the IceCube
    Collaboration
    1429: Cosmic Ray Composition using SPASE-2 and AMANDA-II
    K. Andeen and K. Rawlins For the IceCube Collaboration
    0519: Study of High pT Muons in IceCube
    (0909.0055)
    Lisa Gerhardt and Spencer Klein for the IceCube Collaboration
    1340: Large Scale Cosmic Rays Anisotropy With IceCube
    (0907.0498)
    Rasha U Abbasi, Paolo Desiati and Juan Carlos Velez for the IceCube Collaboration
    1398: Atmospheric Variations as observed by IceCube
    Serap Tilav, Paolo Desiati, Takao Kuwabara, Dominick Rocco,
    Florian Rothmaier, Matt Simmons, Henrike Wissing for the IceCube Collaboration
    1251: Supernova Search with the AMANDA / IceCube Detectors
    (0908.0441)
    Thomas Kowarik, Timo Griesel, Alexander Piégsa for the IceCube Collaboration
    1352: Physics Capabilities of the IceCube DeepCore Detector
    (0907.2263)
    Christopher Wiebusch for the IceCube Collaboration
    1336: Fundamental Neutrino Measurements with IceCube DeepCore
    Darren Grant, D. Jason Koskinen, and Carsten Rott for the IceCube collaboration
    1237: Implementation of an active veto against atmospheric muons in IceCube
    DeepCore
    Olaf Schulz, Sebastian Euler and Darren Grant for the IceCube Collaboration
    1293: Acoustic detection of high energy neutrinos in ice: Status and
    results from the South Pole Acoustic Test Setup
    (0908.3251 – revised)
    Freija Descamps for the IceCube Collaboration
    0903: Sensor development and calibration for acoustic neutrino detection in ice
    (0907.3561)
    Timo Karg, Martin Bissok, Karim Laihem, Benjamin Semburg, and Delia Tosi
    for the IceCube collaboration

    PAPERS RELATED TO ICECUBE
    0466: A new method for identifying neutrino events in IceCube data
    Dmitry Chirkin
    0395: Muon Production of Hadronic Particle Showers in Ice and Water
    Sebastian Panknin, Julien Bolmont, Marek Kowalski and Stephan Zimmer
    0642: Muon bundle energy loss in deep underground detector
    Xinhua Bai, Dmitry Chirkin, Thomas Gaisser, Todor Stanev and David Seckel
    0542: Constraints on Neutrino Interactions at energies beyond 100 PeV with
    Neutrino Telescopes
    Shigeru Yoshida
    0006: Constraints on Extragalactic Point Source Flux from Diffuse Neutrino Limits
    Andrea Silvestri and Steven W. Barwick
    0418: Study of electromagnetic backgrounds in the 25-300 MHz frequency band at
    the South Pole
    Jan Auffenberg, Dave Besson
    y
    , Tom Gaisser, Klaus Helbing, Timo Karg, Albrecht
    Karle,and Ilya Kravchenko

    PROCEEDINGS OF 31
    st
    ICRC, ŁÓDŹ 2009
    1
    IceCube
    Albrecht Karle
    *
    , for the IceCube Collaboration
    *
    University of Wisconsin-Madison, 1150 University Avenue, Madison, WI 53706
    Abstract
    . IceCube is a 1 km
    3
    neutrino telescope
    currently under construction at the South Pole.
    The detector will consist of 5160 optical sensors
    deployed at depths between 1450 m and 2450 m in
    clear Antarctic ice evenly distributed over 86
    strings. An air shower array covering a surface
    area of 1 km
    2
    above the in-ice detector will meas-
    ure cosmic ray air showers in the energy range
    from 300 TeV to above 1 EeV. The detector is de-
    signed to detect neutrinos of all flavors: ν
    e
    , ν
    μ
    and
    ν
    τ
    . With 59 strings currently in operation, con-
    struction is 67% complete. Based on data taken
    to date, the observatory meets its design goals.
    Selected results will be presented.
    Keywords:
    neutrinos, cosmic rays, neutrino as-
    tronomy.
    I. INTRODUCTION
    IceCube is a large kilometer scale neutrino tele-
    scope currently under construction at the South Pole.
    With the ability to detect neutrinos of all flavors over
    a wide energy range from about 100 GeV to beyond
    10
    9
    GeV, IceCube is able to address fundamental
    questions in both high energy astrophysics and neu-
    trino physics. One of its main goals is the search for
    sources of high energy astrophysical neutrinos which
    provide important clues for understanding the origin
    of high energy cosmic rays.
    The interactions of ultra high energy cosmic rays
    with radiation fields or matter either at the source or
    in intergalactic space result in a neutrino flux due to
    the decays of the produced secondary particles such
    as pions, kaons and muons. The observed cosmic
    ray flux sets the scale for the neutrino flux and leads
    to the prediction of event rates requiring kilometer
    scale detectors, see for example
    1
    . As primary candi-
    dates for cosmic ray accelerators, AGNs and GRBs
    are thus also the most promising astrophysical point
    source candidates of high energy neutrinos. Galactic
    source candidates include supernova remnants, mi-
    croquasars, and pulsars. Guaranteed sources of neu-
    trinos are the cosmogenic high energy neutrino flux
    from interactions of cosmic rays with the cosmic mi-
    crowave background and the galactic neutrino flux
    resulting from galactic cosmic rays interacting with
    the interstellar medium. Both fluxes are small and
    their measurement constitutes a great challenge.
    Other sources of neutrino radiation include dark mat-
    ter, in the form of supersymmetric or more exotic
    particles and remnants from various phase transitions
    in the early universe.
    The relation between the cosmic ray flux and the
    atmospheric neutrino flux is well understood and is
    based on the standard model of particle physics. The
    observed diffuse neutrino flux in underground labo-
    ratories agrees with Monte Carlo simulations of the
    primary cosmic ray flux interacting with the Earth's
    atmosphere and producing a secondary atmospheric
    neutrino flux
    2
    .
    Although atmospheric neutrinos are the primary
    background in searching for astrophysical neutrinos,
    they are very useful for two reasons. Atmospheric
    neutrino physics can be studied up to PeV energies.
    The measurement of more than 50,000 events per
    year in an energy range from 500 GeV to 500 TeV
    will make IceCube a unique instrument to make pre-
    cise comparisons of atmospheric neutrinos with
    model predictions. At energies beyond 100 TeV a
    harder neutrino spectrum may emerge which would
    be a signature of an extraterrestrial flux. Atmos-
    pheric neutrinos also give the opportunity to cali-
    brate the detector. The absence of such a calibration
    beam at higher energies poses a difficult challenge
    for detectors at energies targeting the cosmogenic
    neutrino flux.
    Fig. 1 Schematic view of IceCube. Fifty-nine of 86 strings are in
    operation since 2009.

    2
    A. Karle et al., IceCube
    II. DETECTOR AND CONSTRUCTION STA-
    TUS
    IceCube is designed to detect muons and cascades
    over a wide energy range. The string spacing was
    chosen in order to reliably detect and reconstruct
    muons in the TeV energy range and to precisely
    calibrate the detector using flashing LEDs and at-
    mospheric muons.
    The optical properties of the
    South Pole Ice have been measured with various
    calibration devices
    3
    and are used for modeling the
    detector response to charged particles. Muon recon-
    struction algorithms
    4
    allow measuring the direction
    and energy of tracks from all directions.
    In its final configuration, the detector will consist
    of 86 strings reaching a depth of 2450 m below the
    surface. There are 60 optical sensors mounted on
    each string equally spaced between 1450m and
    2450m depth with the exception of the six Deep
    Core strings on which the sensors are more closely
    spaced between 1760m and 2450m. In addition there
    will be 320 sensors deployed in 160 IceTop tanks on
    the surface of the ice directly above the strings. Each
    sensor consists of a 25cm photomultiplier tube
    (PMT), connected to a waveform recording data ac-
    quisition circuit capable of resolving pulses with
    nanosecond precision and having a dynamic range of
    at least 250 photoelectrons per 10ns. With the most
    recent construction season ending in February 2008,
    half of the IceCube array has been deployed.
    The detector is constructed by drilling holes in
    the ice, one at a time, using a hot water drill. Drilling
    is immediately followed by deployment of a detector
    string into the water-filled hole. The drilling of a
    hole to a depth of 2450m takes about 30 hours. The
    subsequent deployment of the string typically takes
    less than 10 hours. The holes typically freeze back
    within 1-3 weeks. The time delay between two sub-
    sequent drilling cycles and string deployments was
    in some cases shorter than 50 hours. By the end of
    February 2009, 59 strings and IceTop stations had
    been deployed. We refer to this configuration as
    IC59. Once the strings are completely frozen in the
    commissioning can start. Approximately 99% of the
    deployed DOMs have been successfully commis-
    sioned. The 40-string detector configuration (IC40)
    has been in operation from May 2008 to the end of
    April 2009.
    III. MUONS AND NEUTRINOS
    At the depth of IceCube, the event rate from
    downgoing atmospheric muons is close to 6 orders
    of magnitude higher than the event rate from atmos-
    pheric neutrinos. Fig. 2 shows the observed muon
    rate (IC22) as a function of the zenith angle
    5
    .
    IceCube is effective in detecting downward going
    muons.
    A first measurement of the muon energy
    spectrum is provided in the references
    6
    .
    A good angular resolution of the experiment is the
    basis for the zenith angle distribution and much more
    so for the search of point sources of neutrinos from
    galactice sources, AGNs or GRBs. Figure 3 shows
    the angular resolution of IceCube for several detector
    configurations based on high quality neutrino event
    selections as used in the point source search for
    IC40
    7
    .
    The median angular resolution of IC40
    achieved is already 0.7°, the design parameter for the
    full IceCube.
    The muon flux serves in many ways also as a
    calibration tool. One method to verify the angular
    resolution and absolute pointing of the detector uses
    the Moon shadow of cosmic rays.
    The Moon
    reaches an elevation of about 28° above the horizon
    at the South Pole. Despite the small altitude of the
    Moon, the event rate and angular resolution of
    IceCube are sufficient to measure the cosmic ray
    shadow of the Moon by mapping the muon rate in
    the vicinity of the Moon. The parent air showers
    have an energy of typically 30 TeV, well above the
    energy where magnetic fields would pose a signifi-
    Fig. 2 Muon rate in IceCube as a function of zenith angle
    5
    .The
    data agree with the detector simulation which includes atmos-
    pheric neutrinos, atmospheric muons, and coincident cosmic ray
    muons (two muons erroneously reconstructed as a single track.)
    Fig. 3 The angular resolution function of different IceCube con-
    figurations is shown for two neutrino energy ranges samples from
    an E
    -2
    energy spectrum.

    PROCEEDINGS OF 31
    st
    ICRC, ŁÓDŹ 2009
    3
    cant deviation from the direction of the primary
    particles. Fig. 4 shows a simple declination band
    with bin size optimized for this analysis. A deficit of
    ~900 events (~4.2σ) is observed on a background of
    ~28000 events in 8 months of data taking. The defi-
    cit is in agreement with expectations and confirms
    the assumed angular resolution and absolute point-
    ing.
    The full IceCube will collect of order 50 000 high
    quality atmospheric neutrinos per year in the TeV
    energy range. A detailed understanding of the re-
    sponse function of the detector at analysis level is the
    foundation for any neutrino flux measurement. We
    use the concept of the neutrino effective area to
    describe the response function of the detector with
    respect to neutrino flavor, energy and zenith angle.
    The neutrino effective area is the equivalent area for
    which all neutrinos of a given neutrino flux imping-
    ing on the Earth would be observed. Absorption ef-
    fects of the Earth are considered as part of the detec-
    tor and folded in the effected area.
    Figure 5 provides an overview of effective areas
    for various analyses that are presented at this confer-
    ence. First we note that the effective area increases
    strongly in the range from 100 GeV to about 100
    TeV. This is due to the increase in the neutrino-
    nucleon cross-section and, in case of the muons, the
    workhorse of high energy neutrino astronomy, due to
    the additional increase of the muon range. Above
    about 100 GeV, the increase slows down because of
    radiative energy losses of muons.
    The IC22
    8
    and IC80 as well as IC86 (IC80+6
    Deep core) atmospheric ν
    μ
    area are shown for upgo-
    ing neutrinos. The shaded area (IC22) indicates the
    range from before to after quality cuts. The effective
    area of IC40 point source analysis
    7
    is shown for all
    zenith angles. It combines the upward neutrino sky
    (predominantly energies < 1PeV) with downgoing
    neutrinos (predominantly >1 PeV). Also shown is
    the all sky ν
    μ
    + ν
    τ
    area of IC80.
    The ν
    e
    effective area is shown for the current
    IC22 contained cascade analysis
    9
    as well as the IC22
    extremely high energy (EHE) analysis
    10
    . It is inter-
    esting to see how two entirely different analysis
    techniques match up nicely at the energy transition
    of about 5 PeV.
    The cascade areas are about a factor of 20 smaller
    than the ν
    μ
    areas, primarily because the muon range
    allows the detection of neutrino interactions far out-
    side the detector, increasing the effective detector
    volume by a large factor. However, the excellent
    energy resolution of contained cascades will benefit
    the background rejection of any diffuse analysis, and
    makes cascades a competitive detection channel in
    the detector where the volume grows faster than the
    area with the growing number of strings.
    The figure illustrates why IceCube, and other
    large water/ice neutrino telescopes for that matter,
    can do physics over such a wide energy range. Un-
    like typical air shower cosmic ray or gamma ray de-
    tectors, the effective area increases by about 8 orders
    of magnitude (10
    -4
    m
    2
    to 10
    +4
    m
    2
    ) over an energy
    range of equal change of scale (10 GeV to 10
    9
    GeV).
    The analysis at the vastly different energy scales re-
    Fig. 4 4.2σ deficit of events from direction of Moon in the
    IceCube 40-string detector confirms pointing accuracy.
    Fig. 6 The energy resolution for muons is approximately 0.3 in
    log(energy) over a wide energy range
    Fig. 5 The neutrino effective area is shown for a several IceCube
    configurations (IC22, IC 40, IC86), neutrino flavors, energy
    ranges and analysis levels (trigger, final analysis).

    4
    A. Karle et al., IceCube
    quires very different approaches, which are pre-
    sented in numerous talks in the parallel sessions
    11, 45
    .
    The measurement of atmospheric neutrino flux
    requires a good understanding of the energy re-
    sponse. The energy resolution for muon neutrinos in
    the IC22 configuration is shown in Fig. 6
    8
    . Over a
    wide energy range (1 – 10000 TeV) the energy reso-
    lution is ~0.3 in log(energy). This resolution is
    largely dominated by the fluctuations of the muon
    energy loss over the path length of 1 km or less.
    IV. ATMOSPHERIC NEUTRINOS AND THE
    SEARCH FOR ASTROPHYSICAL NEUTRINOS
    We have discussed the effective areas, as well as
    the angular and energy resolution of the detector.
    Armed with these ingredients we can discuss some
    highlights of neutrino measurements and astrophysi-
    cal neutrino searches.
    Figure 7 shows a preliminary measurement ob-
    tained with the IC22 configuration. An unfolding
    procedure has been applied to extract this neutrino
    flux. Also shown is the atmospheric neutrino flux as
    published previously based on 7 years of
    AMANDA-II data. The gray shaded area indicates
    the range of results obtained when applying the pro-
    cedure to events that occurred primarily in the top or
    bottom of the detector. The collaboration is devoting
    significant efforts to understand and reduce system-
    atic uncertainties as the statistics increases. The data
    sample consists of 4492 high quality events with an
    estimated purity of well above 95%. Several atmos-
    pheric neutrino events are observed above 100 TeV,
    pushing the diffuse astrophysical neutrino search
    gradually towards the PeV energy region and higher
    sensitivity. A look at the neutrino effective areas in
    Fig. 5 shows that the full IceCube with 86 strings
    will detect about one order of magnitude more
    events: ~50000 neutrinos/year.
    The search for astrophysical neutrinos is summa-
    rized in Fig. 8. While the figure focuses on diffuse
    fluxes, it is clear that some of these diffuse fluxes
    may be detected as point sources. Some examples of
    astrophysical flux models that are shown include
    AGN Blazars
    46
    , BL Lacs
    47
    , Pre-cursor GRB models
    and Waxman Bahcall bound
    48
    and Cosmogenic neu-
    Fig. 7 Unfolded muon neutrino spectrum
    8
    averaged over zenith
    angle, is compared to simulation and to the AMANDA result.
    Data are taken with the 22 string configuration.
    Fig. 8 Measured neutrino atmospheric neutrino fluxes from AMANDA and IceCube are shown together with a number of models for
    astrophysical neutrinos and several limits by IceCube and other experiments

    PROCEEDINGS OF 31
    st
    ICRC, ŁÓDŹ 2009
    5
    trinos
    49
    .
    The following limits are shown for AMANDA
    and IceCube:
    • AMANDA-II, 2000-2006, atmospheric muon
    neutrino flux
    50
    • IceCube-22 string, atmospheric neutrinos,
    (preliminary)
    8
    • AMANDA-II, 2000-2003, diffuse E
    -2
    muon
    neutrino flux limit
    51
    • AMANDA-II, 2000-2002, all flavors, not con-
    tained events, PeV to EeV, E
    -2
    flux limit
    52
    • AMANDA-II, 2000-2004, cascades, contained
    events, E
    -2
    flux limit
    53
    • IceCube-40, muon neutrinos, throughgoing
    events, preliminary sensitivity
    29
    • IceCube-22, all flavor, throughgoing, downgo-
    ing, extremely high energies (10 PeV to
    EeV)
    10
    Also shown are a few experimental limits from
    other experiments, including Lake Baikal
    54
    (diffuse,
    not contained), and at higher energies some differen-
    tial limits by RICE, Auger and at yet higher energies
    energies from ANITA.
    The skymap in Fig. 9 shows the probability for a
    point source of high-energy neutrinos. The map was
    obtained from 6 months of data taken with the 40
    string configuration of IceCube. This is the first re-
    sult obtained with half of IceCube instrumented.
    The “hottest spot” in the map represents an excess of
    7 events, which has a post-trial significance of 10
    -4.4
    .
    After taking into account trial factors, the probability
    for this event to happen anywhere in the sky map is
    not significant. The background consists of 6796
    neutrinos in the Northern hemisphere and 10,981
    down-going muons rejected to the 10
    -5
    level in the
    Southern hemisphere. The energy threshold for the
    Southern hemisphere increases with increasing ele-
    vation to reject the cosmic ray the muon background
    by up to a factor of ~10
    -5
    . The energy of accepted
    downgoing muons is typically above 100 TeV.
    This unbinned analysis takes the angular resolu-
    tion and energy information on an event-by-event
    basis into account in the significance calculation.
    The obtained sensitivity and discovery potential is
    shown for all zenith angles in the figure.
    V. SEARCH FOR DARK MATTER
    IceCube performs also searches for neutrinos pro-
    duced by the annihilation of dark matter particles
    gravitationally trapped at the center of the Sun and
    the Earth. In searching for generic weakly interacting
    massive dark matter particles (WIMPs) with spin-
    independent interactions with ordinary matter,
    IceCube is only competitive with direct detection
    experiments if the WIMP mass is sufficiently large.
    On the other hand, for WIMPs with mostly spin-
    dependent interactions, IceCube has improved on the
    previous best limits obtained by the SuperK experi-
    ment using the same method. It improves on the best
    limits from direct detection experiments by two or-
    ders of magnitude. The IceCube limit as well as a
    limit obtained with 7 years of AMANDA are shown
    in the figure. It rules out supersymmetric WIMP
    models not excluded by other experiments. The in-
    stallation of the Deep Core of 6 strings as shown in
    Fig. 1 will greatly enhance the sensitivity of IceCube
    for dark matter. The projected sensitivity in the
    range from 50 GeV to TeV energies is shown in Fig.
    11. The Deep Core is an integral part of IceCube
    and relies on the more closely spaced nearby strings
    for the detection of low energy events as well as on a
    highly efficient veto capability against cosmic ray
    muon backgrounds using the surrounding IceCube
    strings.
    Fig, 9: The map shows the probability for a point source of high-
    energy neutrinos on the atmospheric neutrino background. The
    map was obtained by operating IceCube with 40 strings for half a
    year
    7
    . The “hottest spot” in the map represents an excess of 7
    events. After taking into account trial factors, the probability for
    this event to happen anywhere in the sky map is not significant.
    The background consists of 6796 neutrinos in the Northern hemi-
    sphere and 10,981 down-going muons rejected to the 10
    -5
    level in
    the Southern hemisphere.
    Fig. 10 Upper limits to E
    -2
    -type astrophysical muon neutrino spec-
    tra are shown for the newest result of ½ year of IC40 and a num-
    ber of earlier results obtained by IceCube and other experiments.

    6
    A. Karle et al., IceCube
    VI. COSMIC RAY MUONS AND HIGH ENERGY
    COSMIC RAYS
    IceCube is a huge cosmic-ray muon detector and
    the first sizeable detector covering the Southern
    hemisphere. We are using samples of several billion
    downward-going muons to study the enigmatic large
    and small scale anisotropies recently identified in the
    cosmic ray spectrum by Northern detectors, namely
    the Tibet array
    55
    and the Milagro array
    56
    . Fig. 12
    shows the relative deviations of up to 0.001 from the
    average of the Southern muon sky observed with the
    22-string array
    11
    . A total of 4.3 billion events with a
    median energy of 14 TeV were used. IceCube data
    shows that these anisotropies persist at energies in
    excess of 100 TeV ruling out the sun as their origin.
    Having extended the measurement to the Southern
    hemisphere should help to decipher the origin of
    these unanticipated phenomena.
    IceCube can detect events with energies ranging
    from 0.1 TeV to beyond 1 EeV, neutrinos and cos-
    mic ray muons.
    The surface detector IceTop consists of ice Cher-
    enkov tank pairs. Each IceTop station is associated
    with an IceCube string. With a station spacing of
    125 m, it is efficient for air showers above energies
    of 1 PeV. Figure 13 shows an event display of a
    very high-energy (~EeV) air shower event. Hits are
    recorded in all surface detector stations and a large
    number of DOMs in the deep ice. Based on a pre-
    liminary analysis some 2000 high-energy muons
    would have reached the deep detector in this event if
    the primary was a proton and more if it was a nu-
    cleus. With 1 km
    2
    surface area, IceTop will acquire
    a sufficient number of events in coincidence with the
    in-ice detector to allow for cosmic ray measurements
    up to 1 EeV. The directional and calorimetric meas-
    urement of the high energy muon component with
    the in-ice detector and the simultaneous measure-
    ment of the electromagnetic particles at the surface
    with IceTop will enable the investigation of the en-
    ergy spectrum and the mass composition of cosmic
    rays.
    Events with energies above one PeV can deposit
    an enormous amount of light in the detector. Figure
    14 shows an event that was generated by flasher
    pulse produced by an array of 12 UV LEDs that are
    mounted on every IceCube sensor. The event pro-
    duces an amount of light that is comparable with that
    of an electron cascade on the order of 1 PeV. Pho-
    tons were recorded on strings at distances up to 600
    m from the flasher. The events are somewhat
    brighter than previously expected because the deep
    ice below a depth of 2100m is exceptionally clear.
    Fig. 13 A very high energy cosmic ray air shower ob-
    served both with the surface detector IceTop and the in-
    ice detector string array.
    Fig. 11 The red boxes show the upper limits at 90% confi-
    dence level on the spin-dependent interaction of dark matter
    particles with ordinary matter
    18, 20
    . The two lines represent the
    extreme cases where the neutrinos originate mostly from
    heavy quarks (top line) and weak bosons (bottom line) pro-
    duced in the annihilation of the dark matter particles. Also
    shown is the reach of the complete IceCube and its DeepCore
    extension after 5 years of observation of the sun. The shaded
    area represents supersymmetric models not disfavored by di-
    rect searches for dark matter. Also shown are previous limits
    from direct experiments and from the Superkamiokande ex-
    periment.
    Fig.12 The plot shows the skymap of the relative intensity in
    the arrival directions of 4.3 billion muons produced by cosmic
    ray interactions with the atmosphere with a median energy of
    14 TeV; these events were reconstructed with an average angu-
    lar resolution of 3 degrees. The skymap is displayed in equa-
    torial coordinates.

    PROCEEDINGS OF 31
    st
    ICRC, ŁÓDŹ 2009
    7
    The scattering length is substantially larger than in
    average ice at the depth of AMANDA.
    Extremely high energy (EHE) events, above about
    1 PeV, are observed near and above the horizon. At
    these energies, the Earth becomes opaque to neutri-
    nos and one needs to change the search strategy. In
    an optimized analysis, the neutrino effective area
    reaches about 4000m
    2
    for IC80 at 1 EeV. IC80 can
    therefore test optimistic models of the cosmogenic
    neutrino flux. IceCube is already accumulating an
    exposure with the current data that makes detection
    of a cosmogenic neutrino event possible.
    IceCube construction is on schedule to completion
    in February 2011. The operation of the detector sta-
    ble and data analysis of recent data allows a rapid in-
    crease of the sensitivity and the discovery potential
    of IceCube.
    VII. ACKNOWLEDGEMENTS
    We acknowledge the support from the following
    agencies: U.S. National Science Foundation-Office
    of Polar Program, U.S. National Science Foundation-
    Physics Division, University of Wisconsin Alumni
    Research Foundation, U.S. Department of Energy,
    and National Energy Research Scientific Computing
    Center, the Louisiana Optical Network Initiative
    (LONI) grid computing resources; Swedish Research
    Council, Swedish Polar Research Secretariat, and
    Knut and Alice Wallenberg Foundation, Sweden;
    German Ministry for Education and Research
    (BMBF), Deutsche Forschungsgemeinschaft (DFG),
    Germany; Fund for Scientific Research (FNRS-
    FWO), Flanders Institute to encourage scientific and
    technological research in industry (IWT), Belgian
    Federal Science Policy Office (Belspo); the Nether-
    lands Organisation for Scientific Research (NWO);
    M. Ribordy acknowledges the support of the SNF
    (Switzerland); A. Kappes and A. Groß acknowledge
    support by the EU Marie Curie OIF Program.
    REFERENCES
    [1] E. Waxman and J. Bahcall, Phys. Rev. D 59 (1998)
    023002
    [2] T. Gaisser,
    Cosmic Rays and Particle Physics
    Cambridge University Press 1991
    [3] M.Ackermann et al., J. Geophys. Res.
    111
    , D13203
    (2006)
    [4] J.Ahrens et al., Nucl. Inst. Meth. A
    524
    , 169 (2004)
    [5] P.
Berghaus
for
the
IceCube

Collaboration,

proc.

    Proc.
for
the
XV
International
Symposium
on
Very

    High
Energy
Cosmic
Ray
Interactions
(ISVHECRI

    2008),
Paris,
France,
arXiv:0902.0021
    [6] P. Berghaus
    et al.
    , Direct Atmospheric Muon Energy
    Spectrum Measurement with IceCube (IceCube col-
    laboration), Proc. of the 31
    st
    ICRC HE1.5; astro-
    ph.HE/arXiv: 0909.0679
    [7] J. Dumm
    et al.
    , Likelihood Point-Source Search with
    IceCube (IceCube collaboration), Proc. of the 31
    st
    ICRC OG2.5D.
    [8] D. Chirkin
    et al.
    , Measurement of the Atmospheric
    Neutrino Energy Spectrum with IceCube (IceCube
    collaboration), Proc. of the 31
    st
    ICRC HE2.2
    [9] J. Kiryluk
    et al.
    , First Search for Extraterrestrial Neu-
    trino-Induced Cascades with IceCube (IceCube col-
    laboration), Proc. of the 31
    st
    ICRC OG2.5; astro-
    ph.HE/arXiv: 0909.0989
    [10] K. Mase
    et al.
    , The Extremely High-Energy Neutrino
    Search with IceCube (IceCube collaboration), Proc. of
    the 31
    st
    ICRC HE1.4
    [11] R. Abbasi and P. Desiati, JC Díaz Vélez
    et al.
    , Large-
    Scale Cosmic-Ray Anisotropy with IceCube (IceCube
    collaboration), Proc. of the 31
    st
    ICRC SH3.2; astro-
    ph.HE/0907.0498
    [12] K. Andeen
    et al.
    , Composition in the Knee Region
    with SPASE-AMANDA (IceCube collaboration),
    Proc. of the 31
    st
    ICRC HE1.2
    [13] C. Rott et al., Atmospheric Neutrino Oscilaton
    Measurement with IceCube (IceCube collaboration),
    Proc. of the 31
    st
    ICRC HE2.3
    [14] M. Baker
    et al.
    , IceCube Time-Dependent Point-
    Source Analysis Using Multiwavelength Information
    (IceCube collaboration), Proc. of the 31
    st
    ICRC OG2.5
    [15] J. Bazo Alba
    et al.
    , Search for Neutrino Flares from
    Point Sources with IceCube (IceCube collaboration,
    to
    appear in
    Proc. of the 31
    st
    ICRC OG2.5
    [16] M. Bissok
    et al.,
    Sensor Development and Calibration
    for Acoustic Neutrino Detection in Ice (IceCube col-
    laboration), Proc. of the 31
    st
    International Cosmic Ray
    Conference, Lodz, Poland (2009) HE2.4; astro-
    ph.IM/09073561
    [17] D. Boersma
    et al.
    , Moon-Shadow Observation by
    IceCube (IceCube collaboration), Proc. of the 31
    st
    ICRC OG2.5
    Figure 14: A flasher event in IceCube. Such events,
    produced by LEDs built in the DOMs, can be used for
    calibration purposes.

    8
    A. Karle et al., IceCube
    [18] J. Braun and D. Hubert
    et al.
    , Searches for WIMP
    Dark Matter from the Sun with AMANDA (IceCube
    collaboration), Proc. of the 31
    st
    ICRC HE2.3; astro-
    ph.HE/09061615
    [19] M. D’Agostino
    et al.
    , Search for Atmospheric Neu-
    trino-Induced Cascades (IceCube collaboration), Proc.
    of the 31
    st
    ICRC HE2.2; ; astro-
    ph.HE/arXiv:0910.0215
    [20] M. Danninger and K. Han
    et al.
    , Search for the Ka-
    luza-Klein Dark Matter with the AMANDA/IceCube
    Detectors (IceCube collaboration), Proc. of the 31
    st
    ICRC HE2.3; astro-ph.HE/0906.3969
    [21] F. Descamps
    et al.
    , Acoustic Detection of High-
    Energy Neutrinos in Ice (IceCube collaboration), Proc.
    of the 31
    st
    ICRC HE2.4
    [22] M. Duvoort
    et al.
    , Search for GRB Neutrinos via a
    (Stacked) Time Profile Analysis (IceCube collabor-
    ation), Proc. of the 31
    st
    ICRC OG2.4
    [23] S. Euler
    et al.
    , Implementation of an Active Veto
    against Atmospheric Muons in IceCube DeepCore
    (IceCube collaboration), Proc. of the 31
    st
    ICRC OG2.5
    [24] T. Feusels
    et al.
    , Reconstruction of IceCube Coinci-
    dent Events and Study of Composition-Sensitive Ob-
    servables Using Both the Surface and Deep Detector
    (IceCube collaboration), Proc. of the 31
    st
    ICRC HE1.3
    [25] A. Francowiak
    et al.
    , Optical Follow-Up of High-
    Energy Neutrinos Detected by IceCube (IceCube and
    ROTSE collaborations), Proc. of the 31
    st
    ICRC OG2.5
    [26] R. Franke
    et al.
    , Neutrino Triggered High-Energy
    Gamma-Ray Follow-Up with IceCube (IceCube col-
    laboration), Proc. of the 31
    st
    ICRC OG2.5
    [27] L. Gerhardt
    et al.
    , Study of High p_T Muons in
    IceCube (IceCube collaboration), Proc. of the 31
    st
    ICRC HE1.5; astro-ph.HE/arXiv 0909.0055
    [28] D. Grant
    et al.
    , Fundamental Neutrino Measurements
    with IceCube DeepCore (IceCube collaboration),
    Proc. of the 31
    st
    ICRC HE2.2
    [29] K. Hoshina
    et al.
    , Search for Diffuse High-Energy
    Neutrinos with IceCube (IceCube collaboration), Proc.
    of the 31
    st
    ICRC OG2.5
    [30] W. Huelsnitz
    et al.,
    Search for Quantum Gravity with
    IceCube and High-Energy Atmospheric Neutrinos
    (IceCube collaboration), Proc. of the 31
    st
    ICRC HE2.3
    [31] A. Ishihara
    et al.
    , Energy-Scale Calibration Using
    Cosmic-Ray Induced Muon Bundles Measured by the
    IceCube Detector with IceTop Coincident Signals
    (IceCube collaboration), Proc. of the 31
    st
    ICRC HE1.5
    [32] A. Kappes
    et al.
    , Searches for Neutrinos from GRBs
    with the IceCube 22-String Detector and Sensitivity
    Estimates for the Full Detector (IceCube collabor-
    ation), Proc. of the 31
    st
    ICRC OG2.4
    [33] F. Kislat
    et al.
    , A First All-Particle Cosmic-Ray En-
    ergy Spectrum from IceTop (IceCube collaboration),
    Proc. of the 31
    st
    ICRC HE1.2
    [34] M. Kowarik
    et al.
    , Supernova Search with the
    AMANDA / IceCube Neutrino Telescopes (IceCube
    collaboration), Proc. of the 31
    st
    ICRC OG2.2
    [35] D. Lennarz
    et al.
    , Search for High-Energetic Neutrinos
    from Supernova Explosions with AMANDA (IceCube
    collaboration), Proc. of the 31
    st
    ICRC OG2.5; astro-
    ph.HE/09074621
    [36] K. Meagher
    et al.
    , Search for Neutrinos from GRBs
    with IceCube (IceCube collaboration), Proc. of the 31
    st
    ICRC OG2.4
    [37] E. Middell
    et al.
    , Improved Reconstruction of Cas-
    cade-Like Events (IceCube collaboration), Proc. of the
    31
    st
    ICRC OG2.5
    [38] C. Portello-Roucelle
    et al.
    , IceCube / AMANDA
    Combined Analyses for the Search of Neutrino Sour-
    ces at Low Energies (IceCube collaboration), Proc. of
    the 31
    st
    ICRC OG2.5
    [39] C. Rott
    et al.
    , Results and Prospects of Indirect
    Searches fo Dark Matter with IceCube (IceCube col-
    laboration), Proc. of the 31
    st
    ICRC HE2.3
    [40] B. Ruzybayev
    et al.
    , Small Air Showers in IceCube
    (IceCube collaboration), Proc. of the 31
    st
    ICRC HE1.1
    [41] A. Schukraft and J. Hülß
    et al.
    , AMANDA 7-Year
    Multipole Analysis (IceCube collaboration), Proc. of
    the 31
    st
    ICRC OG2.5; astro-ph.HE/09063942
    [42] S. Seo
    et al.
    , Search for High-Energy Tau Neutrinos in
    IceCube (IceCube collaboration), Proc. of the 31
    st
    ICRC OG2.5
    [43]
    43
    A. Silvestri
    et al.
    , Search for Ultra–High-Energy
    Neutrinos with AMANDA (IceCube collaboration),
    Proc. of the 31
    st
    ICRC OG2.5
    [44] S. Tilav
    et al.
    , Atmospheric Variations as Observed by
    IceCube (IceCube collaboration), Proc. of the 31
    st
    ICRC HE1.1
    [45] C. Wiebusch
    et al.
    , Physics Capabilities of the
    IceCube DeepCore Detector (IceCube collaboration),
    Proc. of the 31
    st
    ICRC OG2.5; astro-ph.IM/09072263
    [46] F. W. Stecker. Phys. Rev. D, 72(10):107301, 2005
    [47] A. Mücke et al. Astropart. Phys., 18:593, 2003
    [48] S. Razzaque, P. Meszaros, and E. Waxman. Phys.
    Rev. D, 68(8):083001, 2003
    [49] R. Engel, D. Seckel, and T. Stanev. Phys. Rev. D,
    64(9):093010, 2001
    [50] A. Achterberg
    et al
    . (IceCube collaboration),
    Phys.Rev.D
    79
    102005, 2009
    [51] A. Achterberg
    et al
    . (IceCube collaboration),, Phys.
    Rev. D
    76
    042008 (2007);
    [52] Ackerman
    et al
    . (IceCube collaboration), Astrophys.
    Jour.
    675
    2 1014-1024 (2008); astro-ph/07113022
    [53] A. Achterberg,
    et al
    ., Astrophys. Jour.
    664
    397-410
    (2007); astro-ph/0702265v2
    [54] A.V.Avrorin
    et al.
    , Astronomy Letters, Vol.35, No.10,
    pp.651-662, 2009; arXiv:0909.5562
    [55] M. Amenomori et al. [The Tibet AS-Gamma
    Collaboration], Astrophys. J. 633, 1005 (2005)
    arXiv:astroph/0502039
    [56] A.A. Abdo et al., Phys. Rev. Lett. 101: 221101
    (2008), arXiv:0801.3827

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁOD´ Z´ 2009
    1
    All-Sky Point-Source Search with 40 Strings of IceCube
    Jon Dumm
    , Juan A. Aguilar
    , Mike Baker
    , Chad Finley
    , Teresa Montaruli
    ,
    for the IceCube Collaboration
    Dept. of Physics, University of Wisconsin, Madison, WI, 53706, USA
    See the special section of these proceedings.
    Abstract. During 2008-09, the IceCube Neutrino
    Observatory was operational with 40 strings of
    optical modules deployed in the ice. We describe
    the search for neutrino point sources based on a
    maximum likelihood analysis of the data collected
    in this configuration. This data sample provides the
    best sensitivity to high energy neutrino point sources
    to date. The field of view is extended into the down-
    going region providing sensitivity over the entire
    sky. The 22-string result is discussed, along with
    improvements leading to updated angular resolution,
    effective area, and sensitivity. The improvement in
    the performance as the number of strings is increased
    is also shown.
    Keywords: neutrino astronomy
    I. INTRODUCTION
    The primary goal of the IceCube Neutrino Obser-
    vatory is the detection of high energy astrophysical
    neutrinos. Such an observation could reveal the origins
    of cosmic rays and offer insight into some of the
    most energetic phenomena in the Universe. In order to
    detect these neutrinos, IceCube will instrument a cubic
    kilometer of the clear Antarctic ice sheet underneath the
    geographic South Pole with an array of 5,160 Digital
    Optical Modules (DOMs) deployed on 86 strings from
    1.5–2.5 km deep. This includes six strings with a smaller
    DOM spacing and higher quantum efficiency compris-
    ing DeepCore, increasing the sensitivity to low energy
    neutrinos <∼ 100 GeV. IceCube also includes a surface
    array (IceTop) for observing extensive air showers of
    cosmic rays. Construction began in the austral sum-
    mer 2004–05, and is planned to finish in 2011. Each
    DOM consists of a 25 cm diameter Hamamatsu photo-
    multiplier tube, electronics for waveform digitization,
    and a spherical, pressure-resistant glass housing. The
    DOMs detect Cherenkov photons induced by relativistic
    charged particles passing through the ice. In particular,
    the directions of muons (either from cosmic ray showers
    above the surface, or neutrino interactions within the ice
    or bedrock) can be well reconstructed from the track-like
    pattern and timing of hit DOMs.
    The 22-string results presented in the discussion are
    from a traditional up-going search. In such a search,
    neutrino telescopes use the Earth as a filter for the large
    background of atmospheric muons, leaving only an irre-
    ducible background of atmospheric neutrinos below the
    horizon. These have a softer spectrum (∼ E
    −3.6
    above
    100 GeV) than astrophysical neutrinos which originate
    from the decays of particles accelerated by the first order
    Fermi mechanism and thus are expected to have an E
    −2
    spectrum. This search extends the field of view above
    the horizon into the large background of atmospheric
    muons. In order to reduce this background, strict cuts on
    the energy of events need to be applied. This makes the
    search above the horizon primarily sensitive to extremely
    high energy (> PeV) sources.
    II. METHODOLOGY
    An unbinned maximum likelihood analysis, account-
    ing for individual reconstructed event uncertainties and
    energy estimators, is used in IceCube point source anal-
    yses. A full description can be found in Braun et al. [1].
    This method improves the sensitivity to astrophysical
    sources over directional clustering alone by leveraging
    the event energies in order to separate hard spectrum
    signals from the softer spectrum of the atmospheric
    neutrino or muon background. For each tested direction
    in the sky, the best fit is found for the number of signal
    events n
    s
    over background and the spectral index of
    a power law γ of the excess events. The likelihood
    ratio of the best-fit hypothesis to the null hypothesis
    (n
    s
    = 0) forms the test statistic. The significance of
    the result is evaluated by performing the analysis on
    scrambled data sets, randomizing the events in right
    ascension but keeping all other event properties fixed.
    Uniform exposure in right ascension is ensured as the
    detector rotates completely each day, and the location
    at 90
    south latitude gives a uniform background for
    each declination band. Events that are nearly vertical
    (declination < −85
    or > 85
    ) are left out of the
    analysis, since scrambling in right ascension does not
    work in the polar regions.
    Two point-source searches are performed. The first is
    an all-sky search where the maximum likelihood ratio
    is evaluated for each direction in the sky on a grid,
    much finer than the angular resolution. The significance
    of any point on the grid is determined by the fraction
    of scrambled data sets containing at least one grid point
    with a log likelihood ratio higher than the one observed
    in the data. This fraction is the post-trial p-value for
    the all-sky search. Because the all-sky search includes
    a large number of effective trials, the second search is
    restricted to the directions of a priori selected sources
    of interest. The post-trial p-value for this search is again

    2
    J. DUMM et al. ICECUBE POINT SOURCE ANALYSIS
    calculated by performing the same analysis on scrambled
    data sets.
    III. EVENT SELECTION
    Forty strings of IceCube were operational from April
    2008 to May 2009 with ∼ 90% duty cycle after a good
    run selection based on detector stability. The ∼ 3× 10
    10
    triggered events per year are first reduced to ∼ 1 × 10
    9
    events using low-level likelihood reconstructions and
    energy estimators as part of an online filtering system
    on site. These filtered events are sent over satellite to a
    data center in the North for further processing, including
    higher-level likelihood reconstructions for better angular
    resolution. Applying the analysis-level cuts (described
    below) that optimize the sensitivity to point sources
    finally yields a sample of ∼ 3 × 10
    4
    events. Due to
    offline filtering constraints, 144 days of livetime were
    used to design the analysis strategy and finalize event
    selection, keeping the time and right ascension of the
    events blinded. This represents about one-half of the
    final 40-string data sample. Because the northern sky
    and southern sky present very different challenges, two
    techniques are used to reduce the background due to
    cosmic ray muons.
    For the northern sky, the Earth filters out atmospheric
    muons. Only neutrinos can penetrate all the way through
    the Earth and interact near the detector to create up-
    going muons. However, since down-going atmospheric
    muons trigger the detector at ∼ 1 kHz, even a small
    fraction of mis-reconstructed events contaminates the
    northern sky search. Events may be mis-reconstructed
    due to random noise or light from muons from indepen-
    dent cosmic ray showers coincident in the same readout
    window of ± 10 µs. Therefore, strict event selection
    is still required to reject mis-reconstructed down-going
    events. This selection is based on track-like quality
    parameters (the reduced likelihood of the track fit and
    the directional width of the likelihood space around
    the best track fit [2]), a likelihood ratio between the
    best up-going and down-going track solution, and a
    requirement that the event’s set of hits can be split
    into two parts which both reconstruct as nearly-upgoing.
    Although the track-like quality parameters have very
    little declination dependence, these last two parameters
    only work for selecting up-going neutrino candidates
    and remove down-going events. This event selection
    provides an optimal sensitivity to sources of neutrinos
    in the TeV–PeV energy range.
    In the southern sky, energy estimators were used to
    separate the large number of atmospheric muons from a
    hypothetical source of neutrinos with a harder spectrum.
    After track-quality selections, similar but tighter than
    for the up-going sample, a cut based on an energy
    estimator is made until a fixed number of events per
    steradian is achieved. Because only the highest energy
    events pass the selection, sensitivity is primarily to
    neutrino sources at PeV energies and above. Unlike
    for the northern sky, which is a ∼ 90% pure sample
    ν
    / GeV )
    10
    (E
    log
    2
    3
    4
    5
    6
    7
    8
    9
    (E/GeV)]
    10
    dP/d[log
    0
    0.1
    0.2
    0.3
    0.4
    0.5
    0.6
    0.7
    0.8
    Atmospheric (up-going)
    -2
    (up-going)
    E
    -2
    (down-going)
    E
    -1.5
    (down-going)
    E
    Fig. 1: Probability density (P) of neutrino energies at
    final cut level for atmospheric and an E
    −2
    spectrum of
    neutrinos averaged over the northern sky and E
    −1.5
    in
    the southern sky.
    of neutrino-induced muons, the event sample in the
    southern sky is almost entirely well-reconstructed high
    energy atmospheric muons and muon bundles.
    IV. PERFORMANCE
    The performance of the detector and the analysis
    is characterized using a simulation of ν
    µ
    and ν¯
    µ
    . At-
    mospheric muon background is simulated using COR-
    SIKA [3]. Muon propagation through the Earth and
    ice are done using MMC [4]. A detailed simulation of
    the ice [5] propagates the Cherenkov photon signal to
    each DOM. Finally, a simulation of the DOM, including
    angular acceptance and electronics, yields an output
    treated identically to data. For an E
    −2
    spectrum of
    neutrinos the median angular difference between the
    neutrino and the reconstructed direction of the muon in
    the northern (southern) sky is 0.8
    (0.6
    ). The different
    energy distributions in each hemisphere shown in Fig. 1
    cause this effect, since the reconstruction performs better
    at higher energies. The cumulative point spread functions
    for the 22-, 40-, and 80-string configurations of IceCube
    are shown in Fig. 2 for two different ranges of energy.
    Fig. 3 shows the effective area to an equal-ratio flux
    of ν
    µ
    + ν¯
    µ
    . Fig. 4 shows the 40-string sensitivity to
    an E
    −2
    spectrum of neutrinos for 330 days of livetime
    and compared to the 22-string configuration of IceCube,
    as well as ANTARES sensitivity, primarily relevant for
    the southern sky. The 80-string result uses the same
    methodology and event selection for the up-going region
    as this work.
    V. DISCUSSION
    The previous season of IceCube data recorded with
    the 22-string configuration has already been the subject
    of point source searches [7]. The analysis included
    5114 atmospheric neutrino events including a contam-
    ination of about 5% of atmospheric muons during a
    livetime of about 276 days. No evidence was found
    for a signal, and the largest significance is located at

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁOD´ Z´ 2009
    3
    Fig. 2: The point spread function of the 22-, 40-, and 80-
    string IceCube configurations in two energy bins. This
    is the cumulative distribution of the angular difference
    between the neutrino and recostructed muon track using
    simulated neutrinos. The large improvement between the
    22- and 40-string point spread function at high energies
    is due to an improvement in the reconstruction, which
    now uses charge information.
    ν
    / GeV )
    10
    ( E
    log
    2
    3
    4
    5
    6
    7
    8
    9
    ]
    2
    Effective Area [m
    μ
    ν
    +
    μ
    ν
    -3
    10
    -2
    10
    -1
    10
    1
    10
    2
    10
    3
    10
    4
    10
    zenith range (0
    °
    , 30
    °
    )
    zenith range (30
    °
    , 60
    °
    )
    zenith range (60
    °
    , 90
    °
    )
    zenith range (90
    °
    , 120
    °
    )
    zenith range (120
    °
    , 150
    °
    )
    zenith range (150
    °
    , 180
    °
    )
    Fig. 3: The IceCube 40-string solid-angle-averaged ef-
    fective area to an equal-ratio flux of ν
    µ
    and ν¯
    µ
    , recon-
    structed within 2
    of the true direction. The different
    shapes of each zenith band are due to a combination of
    event selection and how much of the Earth the neutrinos
    must travel through. Since the chance of a neutrino
    interacting increases with its energy, in the very up-going
    region high energy neutrinos are absorbed in the Earth.
    Only near the horizon do muons from > PeV neutrinos
    often reach IceCube. Above the horizon, low energy
    events are removed by cuts, and in the very down-going
    region effective area for high energies is lost due to
    insufficient target material.
    153.4
    r.a., 11.4
    dec. Accounting for all trial factors,
    this is consistent with the null hypothesis at the 2.2 σ
    level. The events in the most significant location did
    not show a clear time dependent pattern, and these
    coordinates have been included in the catalogue of
    sources for the 40-string analysis.
    Declination [degrees]
    -80 -60 -40 -20
    0
    20
    40
    60
    80
    ]
    -1
    s
    -2
    dN/dE [TeV cm
    2
    E
    -12
    10
    -11
    10
    -10
    10
    -9
    10
    22 strings 275.7 d
    40 strings prel. sens. 330 d
    IceCube prelim 365 d
    ANTARES prel. 365 d
    Fig. 4: 40-string IceCube sensitivity for 330 days as
    a function of declination to a point source with dif-
    ferential flux
    dE
    = Φ
    0
    (E/TeV)
    −2
    . Specifically, Φ
    0
    is
    the minimum source flux normalization (assuming E
    −2
    spectrum) such that 90% of simulated trials result in a
    log likelihood ratio log λ greater than the median log
    likelihood ratio in background-only trials (log λ = 0).
    Comparison are also shown for the 22-string and the
    expected performance of the 80-string configuration, as
    well as the ANTARES [6] sensitivity.
    Since the 22-string analysis, a number of improve-
    ments have been achieved. An additional analysis of
    the 22-string data optimized for E
    −2
    and harder spectra
    was performed down to −50
    declination with a binned
    search [8]. These analyses are now unified into one
    all-sky search which uses the energy of the events and
    extends to −85
    declination. Secondly, a new recon-
    struction that uses the charge observed in each DOM
    performs better, especially on high energy events. Third,
    an improved energy estimator, based on the photon
    density along the muon track, has a better muon energy
    resolution.
    With construction more than half-complete, IceCube
    is already beginning to demonstrate its potential as an
    extraterrestrial neutrino observatory. The latest science
    run with 40 strings was the first detector configuration
    with one axis the same length as that of the final array.
    Horizontal muon tracks reconstructed along this axis
    provide the first class of events of the same quality as
    those in the finished 80-string detector.
    There are now 59 strings of IceCube deployed and
    taking data. Further development of reconstruction and
    analysis techniques, through a better understanding of
    the detector and the depth-dependent properties of the
    ice, have continued to lead to improvements in physics
    results. New techniques in the southern sky may include
    separating muon bundles of cosmic ray showers from
    single muons induced by high energy neutrinos. At lower
    energies, the identification of starting muon tracks from
    neutrinos interacting inside the detector will be helped
    with the addition of DeepCore [9].

    4
    J. DUMM et al. ICECUBE POINT SOURCE ANALYSIS
    REFERENCES
    [1] J. Braun et al. Methods for point source analysis in high energy
    neutrino telescopes. Astropart. Phys. 29, 155, 2006.
    [2] Neunhoffer, T. Astropart. Phys., 25, 220. 2006.
    [3] D. Heck et al. CORSIKA: A Monte Carlo code to simulate
    extensive air showers, FZKA, Tech. Rep., 1998.
    [4] D. Chirkin and W. Rhode. Preprint hep-ph/0407075, 2004.
    [5] J. Lundberg et al. Nucl. Inst. Meth., vol. A581, p. 619, 2007.
    [6] J. A. Aguilar et al. Expected discovery potential and sensitivity to
    neutrino point-like sources of the ANTARES neutrino telescope.
    in proceedings 30
    th
    ICRC, Merida. 2007.
    [7] R. Abbasi et al. (IceCube Collaboration) First Neutrino Point-
    Source Results from IceCube in the 22-String Configuration.
    submitted, 2009. astro-ph/09052253
    [8] R. Lauer, E. Bernardini. Extended Search for Point Sources of
    Neutrinos Below and Above the Horizon. in proceedings of 2
    nd
    Heidelberg Workshop, 2009, astro-ph/09035434.
    [9] C. Wiebusch et al. (IceCube Collaboration), these proceedings.

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    1
    IceCube Time-Dependent Point Source Analysis Using
    Multiwavelength Information
    M. Baker
    , J. A. Aguilar
    , J. Braun
    , J. Dumm
    , C. Finley
    , T. Montaruli
    , S. Odrowski
    , E. Resconi
    for the IceCube Collaboration
    Dept. of Physics, University of Wisconsin, Madison, WI 53706, USA
    Max-Planck-Institut fur Kernphysik, D-69177 Heidelberg, Germany
    see special section of these proceedings
    Abstract
    . In order to enhance the IceCube’s sen-
    sitivity to astrophysical objects, we have developed
    a dedicated search for neutrinos in coincidence
    with flares detected in various photon wavebands
    from blazars and high-energy binary systems. The
    analysis is based on a maximum likelihood method
    including the reconstructed position, the estimated
    energy and arrival time of IceCube events. After a
    short summary of the phenomenological arguments
    motivating this approach, we present results from
    data collected with 22 IceCube strings in 2007-2008.
    First results for the 40-string IceCube configuration
    during 2008-2009 will be presented at the conference.
    We also report on plans to use long light curves and
    extract from them a time variable probability density
    function.
    Keywords
    : Neutrino astronomy, Multiwavelength
    astronomy
    I. INTRODUCTION
    IceCube is a high-energy neutrino observatory cur-
    rently under construction at the geographic South Pole.
    The full detector will be composed of 86 strings of
    60 Digital Optical Modules (DOMs) each, deployed
    between 1500 and 2500m below the glacier surface. A
    six string Deep Core with higher quantum efficiency
    photomultipliers and closer DOM spacing in the lower
    detector will enhance sensitivity to low energy neutrinos.
    Muons passing through the detector emit Cˇ erenkov light
    allowing reconstruction with
    ? 1
    angular resolution
    in the full detector and about
    1.5
    (median) in the
    22 string configuration. In this paper we describe the
    introduction of a time dependent term to the standard
    search for steady emission of neutrinos presented in Ref.
    [3]. We apply it in a search for periodic emission of
    neutrinos from seven high-energy binary systems and
    for a neutrino emission coincident with a catalogue of
    flares occurring when IceCube was taking data in its 22
    string configuration. We also describe an extension of the
    method that uses multi-wavelength (MWL) lightcurves
    to characterize neutrino emission.
    II. TIME DEPENDENT POINT SOURCE SEARCH
    An unbinned maximum likelihood ratio method, using
    a test statistic that compares a signal plus background
    hypothesis to a background-only one, has been used for
    the search for point sources of neutrinos in IceCube [1].
    We use the angular and energy distribution of events
    as information to characterize the signal with respect to
    the background. In the analysis of the 22-string data
    we use the number of hit DOMs in an event as an
    energy estimator, while for the 40-string configuration
    we use a more sophisticated energy estimator based on
    the photon density along the muon track. The analysis
    method returns a best-fit number of signal events and
    spectral index (though with a large error that depends
    on the number of events near the celestial coordinate
    being tested).
    We use the IceCube 22-string upward-going neutrino
    event data sample of 5114 events collected in 275
    days of livetime between May 31, 2007 and April
    5, 2008 (which includes misreconstructed atmospheric
    muon contamination of about 5%). Selection cuts are
    based on the quality of the reconstruction, on the angular
    uncertainty of the track reconstruction (
    σ < 3
    ) and on
    other variables such as the number of DOMs hit by the
    direct Cˇ erenkov light produced by muons. Fig 1 shows
    that the time distribution of these atmospheric neutrino
    events is consistent with a flat distribution.
    Neutrinos from a point source are expected to cluster
    around the direction of the source and to have a spectrum
    dN
    dE
    ∝ E
    γ
    with spectral index
    γ ∼  2
    as predicted
    by
    1
    st
    order Fermi acceleration mechanisms. On the
    other hand, the background of atmospheric neutrinos
    is distributed uniformly in right ascension and has an
    energy spectrum with
    γ ∼  3.6
    above 100 GeV. We
    construct a signal probability distribution function (pdf):
    S
    i
    =
    1
    2 πσ
    2
    i
    e
    |
    ?
    xi  ?
    xs
    |
    2
    2 σ
    2
    i
    ∗ E (E
    i
    |γ ) ∗ T
    i
    ,
    (1)
    where
    σ
    i
    is the reconstructed angular error of the event
    [2],
    ?x
    i
     ?x
    s
    the angular separation between the recon-
    structed event and the source, E is the energy pdf with
    spectral index
    γ
    , and
    T
    i
    is the time pdf of the event. The
    background pdf is given by:
    B
    i
    = B (?x
    i
    ) ∗ E
    bkg
    (E
    i
    ) ∗
    1
    L
    (2)
    where
    B (?x
    i
    )
    is the background event density (a func-
    tion of the declination of the event),
    E
    bkg
    the energy

    2
    BAKER
    et al.
    ICECUBE TIME DEPENDENT POINT SOURCE ANALYSIS
    Fig. 1. Time distribution of the 22-string neutrino events.
    distribution of the background, and L the livetime.
    The background pdf is determined using the data, and
    the final p-values for these analyses are obtained by
    comparing scrambled equivalent experiments to data.
    Scrambled times are drawn from the distribution of
    measured atmospheric muon event times, taking one
    event per minute to obtain a constant rate.
    The analysis method gives more weight to events
    which are clustered in space and at energies higher than
    expected from the atmospheric background. In this work
    we present the results which include for the first time
    a time dependent term in the pdf. Results are given
    in terms of p-values, or the fraction of the scrambled
    samples with a higher test statistic than found for the
    data.
    III. BINARY SYSTEM PERIODICITY SEARCH
    One class of high-energy binary systems, micro-
    quasars, includes a compact object with an accretion
    disk emitting relativistic jets of matter. Jets are assumed
    to accelerate protons, hence
    pp
    and
    interactions are
    possible. The two microquasars LS 5039 (which is out
    of the IceCube field of view) and LSI 61 +303 [4] have
    been observed to emit TeV gamma-rays modulated with
    the orbital phase of the systems. H.E.S.S. detects the
    minimum of the photon emission for LS 5039 during
    the superior conjunction, where the compact object is
    behind the massive star [5]. The gamma ray modulation
    can be interpreted as an indication of absorption of
    gammas emitted from the compact object. Nonetheless,
    the modulation could be very different in neutrinos,
    where neutrino production depends on how much matter
    is crossed by the proton beam on which interactions and
    decays depend. Since we assume that the modulation is
    related to the relative position of the accelerator with
    respect to the observer, we also include in our search
    objects for which no TeV modulation has yet been
    observed, using the period obtained from spectroscopic
    observations of the visible binary partner. We then leave
    the phase as a free parameter to be fit. Due to low
    statistics, a Gaussian will be adequate to describe the
    Fig. 2. Comparison of discovery potential at 5σ and 50% probability
    between the time-integrated and time-dependent methods for LSI +61
    303.
    time modulation. Hence our time-dependent pdf is:
    T
    i
    =
    1
    2 πσ
    w
    e
    |
    φi  φ
    0
    |
    2
    2 σ
    2
    w
    ,
    (3)
    where
    σ
    w
    is the width of the Gaussian in the period,
    φ
    i
    is the phase of the event and
    φ
    0
    is the phase of peak
    emission. The phase takes a value between 0 to 1.
    We find that this time-dependent method has a better
    discovery potential than the time-integrated analysis if
    the sigma of the emission is less than about 20% of
    the total period (Fig. 2). Since there are more degrees
    of freedom, the time-dependent analysis will perform
    worse if neutrinos are emitted over a large fraction of
    the period.
    We examined seven binary systems, listed in Tab. I,
    covering a range of declinations and periods. There was
    no evidence of periodicity seen for any of the sources
    tested. The most significant result for this search has
    a pre-trial p-value of 6%, we expect to see this level
    of significance from one of our seven trials in 35%
    of scrambled samples, hence we find no evidence for
    periodicity.
    Object
    RA (deg)
    Dec (deg)
    Period (d)
    p-value
    LSI +61 303
    40.1
    +61.2
    26.5
    0.51
    Cygnus X-1
    299.6
    +35.2
    5.6
    0.63
    Cygnus X-3
    308.1
    +40.9
    0.2
    0.09
    XTE J1118+480
    169.5
    +48.0
    0.2
    0.11
    GRS1915
    288.8
    +10.9
    30.8
    0.61
    SS 433
    287.9
    +5.0
    13.1
    0.06
    GRO 0422+32
    65.4
    +32.9
    0.2
    0.39
    TABLE I
    SYSTEM NAME, EQUATORIAL COORDINATES, PERIOD AND
    PRE-TRIAL P-VALUE.
    IV. MULTIWAVELENGTH FLARES ANALYSIS
    In high-energy environments,
    and
    pp
    interactions
    produce pions and kaons that decay into photons and
    neutrinos. Thus, we expect a correlation between TeV
    γ
    and
    ν
    µ
    fluxes. Blazars and binary systems exhibit

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    3
    variability, with flares often observed to correlate in
    several photon wavebands. Hence, if TeV information is
    not available, we can use X-ray and optical data as well.
    We use this expected time correlation between photons
    and neutrinos to suppress the background of atmospheric
    neutrinos, which have a random distribution in time, by
    looking for neutrino emission in time windows selected
    based on MWL information. By restricting our search
    we need fewer events to achieve a
    signal than with
    the time-integrated search.
    We use MWL observations to create a catalogue of
    flares from blazars and binary systems which have states
    of heightened non-thermal emission. We determine the
    time window of our search based on the MWL data to
    characterize the time and duration of peak brightness.
    A. Selection of Flares
    To collect a list of interesting flares we monitored
    alerts such as Astronomer’s Telegram or GCN for
    sources observed undergoing a change of state which
    may produce heightened neutrino emission. The selected
    catalogue is presented in Tab. II and illustrated here:
    3C 454.3
    flares were measured by AGILE GRID
    during July 24-30, 2007 [8] and again during Nov.
    12-22, 2007 [9].
    1ES 1959+650
    was seen by INTEGRAL in a hard
    flux state (Nov. 25-28 2007 [6]). Later Whipple
    obtained a few measurements around December 2-7
    [7] which we also selected for investigation.
    Cygnus X-1
    had a ”giant outburst” seen by Konus-
    Wind and Suzaku-WAM [10]. These giant outbursts
    have been modeled in [11].
    S5 0716+71
    was seen flaring in GeV, optical and
    radio bands during two periods, September 7-13,
    2007 and Oct 19-29, 2007 [12].
    B. Method and Results
    We tested two methods to search for neutrino flares:
    the first case (hereafter the ”box method”), uses a pdf
    which counts only events which fall inside the selected
    time window:
    T
    i
    =
    H (t
    max
     t
    i
    )H (t
    i
     t
    min
    )
    t
    max
    t
    min
    ,
    (4)
    where H is the Heaviside step function, and
    t
    min
    and
    t
    max
    are fixed from MWL data. The second case is to
    find a best-fit Gaussian to describe the neutrino emission,
    fitting the mean of the flare and its duration inside the
    selected time window. The time factor in the source term
    will be:
    T
    i
    =
    1
    2 πσ
    t
    e
    |
    ti  t
    0
    |
    2
    2 σ
    2
    t
    (5)
    where
    t
    0
    is the peak emission and
    σ
    t
    is the width. The
    Gaussian search method yields more information about
    the flare, such as width and time of the peak of the
    emission, and also can use events outside of the time
    window. To focus the search on correlation with photon
    emission instead of an all-year search, we confined the
    Fig. 3.
    Comparison of the box and Gaussian method for the flare
    search. The mean number of events needed for a
    5  σ
    detection is
    plotted against the width of neutrino emission.
    mean to the time window, and the sigma can not be
    longer than the time window. The Gaussian introduces
    two additional parameters to fit, while the box method
    has no additional parameters over the time-integrated
    search.
    To compare the two methods, we generated signal
    events with Gaussian time distributions of different
    widths to add to scrambled data. Our figure of merit
    is the minimum flux required for 50% probability of
    5
    σ
    discovery. We find the box method outperforms the
    Gaussian unless the FWHM of the signal function is
    less than 10% or greater than 110% of the width of the
    time window. We show the discovery potential curves for
    time windows of 3 and 10 days in Fig. 3. We also tested
    the possibility that the time window we chose based on
    MWL information is not centered on a neutrino flare
    by injecting events with an offset in the window, still
    finding a region where the box requires fewer events
    for discovery. Hence the box method, which performs
    better than the Gaussian method in a broad part of the
    signal parameter space was selected for providing the
    final p-values.
    We found that 5 of 7 flares we examined were best
    fit by 0 source events, while S5 0716+71 and 1ES
    1959+650 each showed one contributing event during
    a flare. Considering that we looked at 7 flares, the post
    trials p-value is 14% for the most significant result, the
    10 day flare of S5 0716+71. This value is compatible
    with background fluctuations.
    Source
    Alert Ref.
    Time Window
    p-value
    1ES 1959+650
    [6]
    MJD 54428-54433
    1
    1ES 1959+650
    [7]
    MJD 54435.5-54440.5
    0.08
    3C 454
    [8]
    MJD 54305-54311
    1
    3C 454
    [9]
    MJD 54416-54426
    1
    Cyg X-1
    [10]
    MJD 54319.5-54320.5
    1
    S5 0716+71
    [12]
    MJD 54350-54356
    1
    S5 0716+71
    [12]
    MJD 54392-54402
    0.02
    TABLE II
    FLARE LIST: SOURCE NAME, REFERENCES FOR THE ALERT,
    INTERVAL IN MODIFIED JULIAN DAY, PRE-TRIAL P-VALUE.

    4
    BAKER
    et al.
    ICECUBE TIME DEPENDENT POINT SOURCE ANALYSIS
    V. FUTURE DEVELOPMENTS: ANALYSIS BASED ON
    LONG LIGHT CURVES
    With the advent of Fermi, long and regularly sampled
    high energy
    γ
    -ray light curves will be available soon.
    The Fermi public data [13] already provide a first
    glimpse of the variable behavior of bright sources and
    the quality of the data. We plan to analyze Fermi light
    curves using the method described in [14]. Following
    this approach, the analysis of long light curves will
    provide:
    A systematic selection of flaring periods: until now
    the selection of flaring periods is biased because
    detections are often triggered by alerts. The moni-
    toring of the sky provided by Fermi will eliminate
    this.
    A systematic criterion to define the threshold for a
    flare: once enough data will be accumulated, the
    flare statistics will provide a characteristic level
    and a standard deviation. With a safe 3
    σ
    threshold,
    flaring periods cannot be confused with intrinsic
    fluctuations of the detector and can be selected
    uniformly across the entire period considered.
    The possibility to select more than one flare in the
    same light curve, to estimate the frequency of the
    high states.
    A non-parametric time dependent signal pdf.
    Our analysis of long Fermi light curves is still in de-
    velopment and for the moment limited by the relatively
    short duration of the Fermi data taking. We illustrate the
    method using the light curve collected by RXTE-ASM
    for Mkn 421 (Fig. 4). About 10 years of RXTE-ASM
    data are analyzed in order to extract a characteristic
    level of the source and determine flaring periods, as in
    [14]. For example here the threshold for flaring has been
    fixed at the 3
    σ
    level that corresponds to 1.7 RXTE/ASM
    count/sec. Interpreting periods selected above this level
    with the Maximum Likelihood Block algorithm provides
    the time dependent pdf (see Fig. 5).
    While the rate of events observed in IceCube is
    approximately stable over timescales of a few days, the
    variability of the background has to be considered if
    longer periods are tested. The main source of variations
    of the observed event rates are changes in the detector
    uptime. These will be implemented in the description of
    the background.
    VI. CONCLUSIONS
    We have presented the results of a time dependent
    analysis of the IceCube 22 string data sample. We
    searched for a periodic time structure of neutrinos from
    binary systems, and neutrinos in coincidence with high
    Fig. 4. Subperiod of Mkn 421 light curve collected by ASM/RXTE
    for illustration of the method.
    Fig. 5. The time pdf resulting from application of the 3σ threshold
    described in the text to the Mkn 421 light curve.
    flux states from sources for which other experiments
    issued alerts. Results in all cases were consistent with
    background fluctuations. We also provide insight on how
    MWL information may in the future be directly used to
    create a time pdf to analyze correlations of photon and
    neutrino emission.
    REFERENCES
    [1] J. Braun
    et al., Astropart. Phys.
    29
    , 299 (2008)
    [2] T. Neunho¨ffer, Astropart. Phys.
    25
    , 220 (2006)
    [3] J. Dumm
    et al., for the IceCube Collaboration, these proceedings.
    [4] J. Albert
    et al., 2009, ApJ
    693
    , 303
    [5] F. Aharonian
    et al., 2006 A&A
    460
    , 743
    [6] E. Bottacini
    et al., http://www.astronomerstelegram.org/?read=
    1315
    [7] VERITAS
    collab.
    http://veritas.sao.arizona.edu/documents/
    summary1es1959.table
    [8] S. Vercellone
    et al., 2008, ApJ
    676
    , L13
    [9] S. Vercellone
    et al., 2009, ApJ
    690
    , 1018
    [10] S. Golenetskii,
    et al., GCN 6745, Aug 10 2007
    [11] G.E. Romero, M.M. Kaufman Bernado and I.F. Mirabel, 2002,
    A&A
    393
    , L61
    [12] M. Villata
    et al., 2008, A&A
    481
    , L79
    [13] http://fermi.gsfc.nasa.gov/ssc/data/access/lat/msl
    lc/
    [14] E. Resconi
    et al., arXiv:0904.1371.

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁOD´ Z´ 2009
    1
    Search for neutrino flares from point sources with IceCube
    J. L. Bazo Alba
    , E. Bernardini
    , R. Lauer
    , for the IceCube Collaboration
    DESY, D-15738 Zeuthen, Germany.
    see special section of these proceedings
    Abstract. A time-dependent search for neutrino
    flares from pre-defined directions in the whole sky
    is presented. The analysis uses a time clustering
    algorithm combined with an unbinned likelihood
    method. This algorithm provides a search for sig-
    nificant neutrino flares over time-scales that are
    not fixed a-priori and that are not triggered by
    multiwavelength observations. The event selection
    is optimized to maximize the discovery potential,
    taking into account different time-scales of source
    activity and background rates. Results for the 22-
    string IceCube data from a pre-defined list of bright
    and variable astrophysical sources will be reported
    at the conference.
    Keywords: IceCube, Neutrino Flares, Clustering.
    I. INTRODUCTION
    Several astrophysical sources are known to have a
    variable photon flux at different wavelengths, showing
    flares that last between several minutes to several days.
    Hadronic models of Active Galactic Nuclei (AGNs) pre-
    dict [1][2] neutrino emission associated with these multi-
    wavelength (MWL) emissions. Time integrated analyses
    are less sensitive in this flaring scenario because they
    contain a higher background of atmospheric neutrinos
    and atmospheric muons. Therefore a time dependent
    analysis is more sensitive because it reduces the back-
    ground by searching smaller time scales around the flare.
    A direct approach that looks for this correlation using
    specific MWL observations is reported in [3].
    In order to make the flare search more general, and
    since MWL observations are scarce and not available for
    all sources, we take an approach not triggered by MWL
    observations. We apply a time-clustering algorithm
    (see [4]) to pre-defined source directions looking
    for the most significant accumulation in time (flare)
    of neutrino events over background, considering all
    possible combinations of event times. One disadvantage
    of this analysis is the increased number of trials,
    which reduces the significance. Nevertheless, for flares
    sufficiently shorter than the total observation period,
    the time clustering algorithm is more sensitive than a
    time integrated analysis. The predicted time scales are
    well below this threshold.
    II. FLARE SEARCH ALGORITHM
    The time clustering algorithm chooses the most
    promising flare time windows based on the times of the
    most signal-like events from the analyzed data. Each
    combination of these event times defines a search time
    window (∆t
    i
    ). For each ∆t
    i
    a significance parameter
    λ
    i
    is calculated. The algorithm returns the best λ
    max
    corresponding to the most significant cluster. The signif-
    icance can be obtained using two approaches: a binned
    method, as in the previous implementation [4], and
    an improved unbinned maximum likelihood method [5]
    which enhances the performance.
    The unbinned maximum likelihood method defines
    the significance parameter by:
    λ = −2log
    .
    L(?x
    s
    , n
    s
    = 0)
    L(?x
    s
    , nˆ
    s
    , γˆ
    s
    )
    ¸
    ,
    (1)
    where ?x
    s
    is the source location, nˆ
    s
    and γˆ
    s
    are the best
    estimates of the number of signal events and source
    spectral index, respectively, which are found by max-
    imizing the likelihood, (L):
    L =
    n
    Y
    tot
    i=1
    μ
    n
    s
    n
    tot
    S
    i
    +
    μ
    1 −
    n
    s
    n
    tot
    B
    i
    (2)
    The background probability density function (pdf),
    B
    i
    , calculated purely from data distributions, is given
    by:
    B
    i
    = P
    space
    i
    i
    , φ
    i
    )P
    energy
    i
    (E
    i
    , θ
    i
    )P
    time
    i
    i
    ),
    (3)
    where P
    space
    describes the distribution of events in a
    given area (a zenith band of 8
    is used for convenience).
    In a simple case this probability would be flat because of
    random distribution of background events. However, due
    to applied cuts, Earth absorption properties and detector
    geometry, this probability is dependent on zenith, θ
    i
    , and
    azimuth, φ
    i
    . The irregular azimuthal distribution caused
    by the detector geometry is shown in Fig. 1. For time
    integrated analyses covering one year the dependence
    on the azimuth is negligible because the exposure for
    all right ascension directions is integrated. However, an
    azimuth correction becomes important for time scales
    shorter than 1 day, reaching up to 40% difference, thus
    it should be included in time dependent analyses. P
    space
    has value unity when integrated over solid angle inside
    the test region (i.e. zenith band).
    The energy probability P
    energy
    i
    is determined from
    the energy estimator distribution and depends on the
    zenith coordinate. In the southern sky an energy sensitive
    event selection is the most efficient way to reduce the at-
    mospheric muon background. This energy cut decreases
    with zenith angle, thus creating a zenith dependence

    2
    J. L. BAZO ALBA et al. NEUTRINO FLARES WITH ICECUBE
    of the energy. Therefore a zenith dependent energy
    probability, shown in Fig. 2, is needed. Note that for
    the northern sky this correction is small.
    0
    ]
    Azimuth[
    0
    50
    100
    150
    200
    250
    300
    350
    Normalized Events
    0.2
    0.4
    0.6
    0.8
    1
    Fig. 1. Normalized azimuth distribution of the data sample reported
    in [9].
    cos(
    θ
    )
    -1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1
    (E)
    10
    log
    0
    1
    2
    3
    4
    5
    6
    7
    8
    9
    10
    -3
    10
    10
    -2
    10
    -1
    1
    Fig. 2. Background energy pdf from data as a function of the energy
    estimator and zenith angle. The ultra high energy sample [10] is used.
    The southern sky corresponds to cos(θ) > 0.
    Given the low statistics at final sample level, estimat-
    ing the background by counting events inside a time
    window would introduce significant errors for short time
    scales. Therefore another approach is used, namely, to
    fit the event rates in the entire observed period as a
    function of time. Two regions of the sky (South and
    North) are distinguished because they have different
    properties. The northern sky sample consists mostly of
    atmospheric neutrinos which do not show a significant
    seasonal variation, therefore a constant fit is used. For
    the southern sky, a sinusoidal fit is used because it is
    dominated by a background of high energy atmospheric
    muons which have seasonal variation. These fits are
    shown in Fig. 3 and include the necessary correction
    for the uptime
    1
    of the detector. It has been verified that
    the time modulations for different zenith bands within a
    1
    The uptime takes into account the inefficiency periods and data
    gaps after data quality selection.
    half hemisphere are the same, thus allowing us to use
    all events inside the half hemisphere for the fit of the
    rates.
    Time MJD
    54300 54350 54400 54450 54500 54550
    Rate (Hz)
    20
    25
    30
    35
    40
    45
    50
    -6
    ×
    10
    Time MJD
    54300 54350 54400 54450 54500 54550
    Rate (Hz)
    0.16
    0.165
    0.17
    0.175
    0.18
    0.185
    0.19
    0.195
    -3
    ×
    10
    Fig. 3. Uptime corrected rates and their fits for the southern (left)
    and northern (right) skies.
    The signal pdf, S
    i
    , is given by:
    S
    i
    = P
    space
    i
    (| ?x
    i
    − ?x
    s
    |, σ
    i
    )P
    energy
    i
    (E
    i
    , θ
    i
    , γ
    s
    ), (4)
    where, the spatial probability, P
    space
    i
    is a Gaussian func-
    tion of | ?x
    i
    − ?x
    s
    |, the space angular difference between
    the source location, ?x
    s
    , and each event’s reconstructed
    direction, ?x
    i
    , and σ
    i
    , the angular error estimation of
    the reconstructed track. The estimator used for σ
    i
    is
    the size of the error ellipse around the maximum value
    of the reconstructed event track likelihood. The energy
    probability, P
    energy
    i
    , constructed from signal simulation,
    is a function of the event energy estimation, E
    i
    , the
    zenith coordinate, θ
    i
    , and the assumed energy spectral
    index of the source, γ
    s
    (E
    −γ
    s
    ). A projection of P
    energy
    i
    for the whole sky is shown in Fig. 4. For a given θ
    i
    and γ
    s
    the energy pdf is normalized to unity over E
    i
    .
    For the energy a dedicated estimator of the number of
    photons per track length is used. No flare time structure
    is assumed (i.e. taken to be flat in time). Therefore there
    is no need to include a time dependent term in the signal
    pdf.
    (E)
    10
    log
    0 12 34 5 6 7 8 910
    Energy spectral index
    1
    1.5
    2
    2.5
    3
    3.5
    4
    10
    -4
    -3
    10
    10
    -2
    10
    -1
    Fig. 4. Projection for the whole sky of the energy component of the
    signal pdf as a function of the energy estimator and energy spectral
    index. The ultra high energy sample [10] is used.

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁOD´ Z´ 2009
    3
    TABLE I
    LIST OF VARIABLE ASTROPHYSICAL SOURCES AND THEIR DETECTION PROBABILITY (TIME VARIABLE AND TIME INTEGRATED) FOR A
    SIMULATED FLARE OF 7 DAYS WITH AN E
    −2
    ENERGY SPECTRUM AND POISSON MEAN OF 5 INJECTED EVENTS. THE EQUIVALENT STEADY
    FLUX CORRESPONDS TO 5 EVENTS INJECTED AT ANY TIME IN THE FULL ICECUBE DATA-TAKING PERIOD (276 DAYS).
    Source
    Type
    dec [
    ]
    ra [
    ]
    Detection Probability (3σ) (%)
    Eq. flux ×10
    −11
    Time Variable
    Time Integrated
    TeVcm
    −2s−1
    GEV J0540-4359
    LBL
    -44.1
    84.7
    46.8
    24.5
    57.5
    GEV J1626-2502
    FSRQ
    -25.5
    246.4
    85.6
    80.8
    30.9
    GEV J1832-2128
    FSRQ
    -21.1
    278.4
    77.1
    72.1
    21.2
    GEV J2024-0812
    FSRQ
    -7.6
    306.4
    37.5
    14.4
    3.0
    3C 279
    FSRQ
    -5.8
    194.1
    26.1
    9.8
    2.4
    3C 273
    FSRQ
    2.0
    187.3
    50
    12.4
    1.2
    CTA 102
    FSRQ
    11.7
    338.1
    36.2
    13.6
    1.0
    GEV J0530+1340
    FSRQ
    13.5
    82.7
    31.4
    10.1
    1.1
    3C 454.3
    FSRQ
    16.1
    343.5
    70.1
    12.2
    1.2
    GEV J0237+1648
    LBL
    16.6
    39.7
    69
    11
    1.2
    We use a binned method implementation of the time
    clustering algorithm as a crosscheck of our new un-
    binned analysis. In the case of the binned method, a
    circular angular search bin (2.5
    radius) around the
    source direction is used. The times of the events that
    define the search time windows (∆t
    i
    ) are given by all the
    events inside this angular bin. The significance parame-
    ter is obtained from Poisson statistics, given the number
    of expected background events inside the bin and the
    observed events in each cluster with multiplicity
    2
    m.
    The expected number of background events is calculated
    by integrating, in the given time window, the fit to the
    rates, as described above. This calculation takes into
    account the zenith dependence of the background, in
    zenith bands with the size of the bin, the corresponding
    uptime factor and the azimuth correction.
    The best significance obtained for a cluster is cor-
    rected for trial factors by running several Monte Carlo
    background-only simulations. The simulation is done
    by creating distributions from data of zenith, azimuth,
    reconstruction error and energy estimator. The event
    characteristics are randomly taken from these distri-
    butions while considering the correlations between the
    different parameters. In order to study the performance
    of the algorithm, we calculate the neutrino flare detection
    probability as a function of the signal strength and
    duration of the flare by simulating signal events on top
    of background events
    3
    . The properties of signal events
    are taken from a dedicated signal simulation and depend
    on the assumed energy spectral index. The Point Spread
    Function (PSF) is used to smear the events around
    the source location, thus simulating the effect of the
    direction reconstruction. For each simulation, a random
    time is chosen around which signal events are randomly
    injected inside the time window defined by the flare
    duration. The flare duration is investigated in the range
    from 1 day to 15 days, though the algorithm finds the
    best time window, which could be larger. We constrain
    2
    The integral of the Poisson distribution of the background events
    starts at (m-1) since the beginning and end of the time period are fixed
    from the data itself.
    3
    The number of injected background and signal events is Poisson
    distributed.
    the largest flare duration in the algorithm to be less than
    30 days, which is sensible from γ-ray observations.
    III. SOURCE SELECTION
    Since searching for all directions in the sky would de-
    crease the significance, we consider only a few promis-
    ing sources, thus reducing the number of trials. We
    select variable bright astrophysical sources in the whole
    sky. The selected blazars, including Flat Spectrum Radio
    Quasars (FSRQs) and Low-frequency peaked BL Lacs
    (LBLs), are taken from the confirmed Active Galactic
    Nuclei (AGN) in the third EGRET catalogue (3EG) [6].
    We also require that they are present in the current latest
    Fermi catalogue (0FGL) [7]. The criteria for selecting
    variable and bright source is based on the following
    parameters thresholds:
    • Variability index (3EG) > 1
    • Maximum 3EG flux (E > 100 MeV) > 40 [10
    −8
    ph cm
    −2
    s
    −1
    ]
    • Average 3EG flux (E > 100 MeV) > 15 [10
    −8
    ph
    cm
    −2
    s
    −1
    ]
    • Inside visibility region of IceCube.
    The selected source list consists of 10 directions
    (Table I) that are going to be tested with the time
    clustering algorithm. Models like [2] favor fluxes of
    higher energy neutrinos from FSRQ sources. Given the
    absorption of neutrinos at different energies in the Earth
    and the event cut strategy, southern sky FSRQs are more
    favored by these models because of their higher energy
    range of sensitivity.
    IV. DATA SAMPLES
    IceCube[8] 22-string data from 2007-08 is used. It
    spans 310 days with an overall effective detector uptime
    of 88.9% (i.e. 276 days). The whole sky (declination
    range from -50
    to 85
    ) is scanned. Different selection
    criteria are applied for the northern and southern skies.
    Previously obtained reconstructed datasets are used: the
    standard point source sample for the northern sky [9]
    (5114 events, declination from -5
    to 85
    , 1.4
    sky-
    averaged median angular resolution) and the dedicated
    ultra high energy sample for the southern sky [10] (1877
    events in the whole sky, declination from -50
    to 85
    ,

    4
    J. L. BAZO ALBA et al. NEUTRINO FLARES WITH ICECUBE
    -2
    s
    -1
    )
    -11
    TeV cm
    Equivalent steady flux (x10
    1
    2
    3
    4
    5
    6
    ] (%)
    σ
    Detection Probability [3
    0
    10
    20
    30
    40
    50
    60
    70
    80
    90
    100
    1 day flare
    7 day flare
    15 day flare
    T integrated
    (a) Southern sky at (dec=-7.6, ra=306.4)
    -2
    s
    -1
    )
    -11
    TeV cm
    Equivalent steady flux (x10
    0.5
    1
    1.5
    2
    2.5
    ] (%)
    σ
    Detection Probability [3
    0
    10
    20
    30
    40
    50
    60
    70
    80
    90
    100
    1 day flare
    7 day flare
    15 day flare
    T integrated
    (b) Northern sky at (dec=16.1, ra=343.5)
    Fig. 5. Detection probability (3σ) for two source directions. The curves correspond to different time duration of the flares as function of the
    injected flux with a E
    −2
    energy spectrum, using an unbinned time variable method (dashed), compared to a time integrated method (solid).
    The same mean number of events are injected into the time-windows (1, 7, 15, and 276 days) at each point on the x-axis, which is labeled
    with the equivalent flux corresponding to the full 276 day period.
    1.3
    sky-averaged median angular resolution). The first
    sample is optimized, within an unbinned method, for
    the optimal sensitivity to both hard and soft spectrum
    sources. The second sample was optimized for a binned
    method at ultra high energies. Therefore it should be
    noted that the binned method results are much better in
    the southern sky than in the northern sky. Nevertheless,
    the unbinned method, for an E
    −2
    energy spectrum still
    performs better in the southern sky.
    The energy containment in these two regions is dif-
    ferent, with ranges from TeV to PeV and from PeV
    to EeV, in the northern and southern sky respectively.
    Event tracks are obtained with a multi-photoelectron
    4
    (MPE) [11] reconstruction which improves the angular
    resolution for high energies.
    V. RESULTS
    The probability of a 3σ flare detection using this time
    variable analysis (time clustering algorithm) for a given
    number of injected signal events (i.e. Poisson mean of
    5 events) with a E
    −2
    energy spectrum inside a seven-
    day window is shown for all sources in Table I. For
    comparison purposes, time-integrated detection proba-
    bilities integrated over the whole 22-string IceCube data
    period (276 days) are also given. In the northern sky,
    the same simulated signal was on average four times
    more likely to be detected at 3σ with the unbinned time
    variable search than with the time integrated search, and
    in the southern sky, on average about twice as likely with
    the time variable search. The gain is not as substantial
    as in the northern sky because the discovery potential
    without time properties is already greater since for the
    same number of injected signal events the background
    is relatively smaller. A more detailed example for two
    sources, at the southern and northern skies, for different
    time scales and signal fluxes is presented in Fig. 5. For
    4
    The MPE reconstruction takes the arrival time distribution of the
    first of N photons using the cumulative distribution of the single photon
    pdf.
    shorter flare durations the detection probability increases
    and is well above a time integrated search. It can be seen
    that there is a different behaviour for each part of the sky.
    This is caused by the different type of backgrounds (high
    energy atmospheric muons in the south and atmospheric
    neutrinos in the north) and the difference in number of
    final events in each sample (less events in the southern
    sky) due to the different selection cuts.
    VI. SUMMARY
    We have presented the sensitivity of the time cluster-
    ing algorithm using an unbinned maximum likelihood
    method. This is an improvement over the previous
    performances using a binned method and time integrated
    analyses. The search window for variable sources has
    been extended to the southern sky. IceCube 22-string
    data will be analyzed using this method looking for
    neutrinos flares with no a priori assumption on the time
    structure of the signal.
    REFERENCES
    [1] D. F. Torres and F. Halzen, Astropart. Phys. 27 (2007) 500.
    [2] A. Atoyan and C. D. Dermer, New Astron. Rev. 48 (2004) 381.
    [3] M. Baker et al. for the IceCube Collab.: Contributions to this
    conference.
    [4] K. Satalecka et al. for the IceCube Collab.: Contributions to the
    ICRC, Merida, Mexico, July 3-11, pages 115-118, (2007).
    [5] J. Braun et al., Astropart. Phys. 29 (2008) 299.
    [6] P. L. Nolan et al., The Astrophysical Journal. Vol. 597, No 1,
    (2003) pp. 615-627.
    [7] A. A. Abdo et al., Fermi LAT Collaboration, (2009).
    arXiv:0902.1340
    [8] A. Karle et al., for the IceCube Collab.: ARENA Proceedings
    (2008).
    [9] J. L. Bazo Alba et al. for the IceCube Collab.: NOW 2008
    Proceedings. Nucl. Phys. B (Proc. Suppl.) Vol. 188 (2009) pp.
    267-269. arXiv:0811.4110
    [10] R. Lauer et al. for the IceCube Collab.: Proceedings 2nd
    Heidelberg Workshop ”HE γ-rays and ν’s from Extra-Galactic
    Sources”, (2009). arXiv:0903.5434
    [11] J. Ahrens et al. for the AMANDA Collab., Nucl. Instrum. Meth.
    A 524 (2004) 169.

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    1
    Neutrino triggered high-energy gamma-ray follow-up with IceCube
    Robert Franke
    , Elisa Bernardini
    for the IceCube collaboration
    DESY Zeuthen,D-15738 Zeuthen, Germany
    See the special section of these proceedings.
    Abstract
    . We present the status of a program
    for the generation of online alerts issued by Ice-
    Cube for gamma-ray follow up observations by Air
    Cherenkov telescopes (e.g. MAGIC). To overcome the
    low probability of simultaneous observations of flares
    of objects with gamma-ray and neutrino telescopes
    a neutrino-triggered follow-up scheme is developed.
    This mode of operation aims at increasing the avail-
    ability of simultaneous multi-messenger data which
    can increase the discovery potential and constrain the
    phenomenological interpretation of the high energy
    emission of selected source classes (e.g. blazars).
    This requires a fast and stable online analysis of
    potential neutrino signals. We present the work
    on a significance-based alert scheme for a list of
    phenomenologically selected sources. To minimize the
    rate of false alerts due to detector instabilities a fast
    online monitoring scheme based on IceCube trigger
    and filter rates was implemented.
    Keywords
    : IceCube neutrino gamma-ray follow-up
    I. INTRODUCTION
    A Neutrino Triggered Target of Opportunity pro-
    gram (NToO) was developed already in 2006 using the
    AMANDA array to initiate quasi-simultaneous gamma-
    ray follow-up observations by MAGIC. The aim of such
    an approach is to increase the chance to discover cosmic
    neutrinos by on-line searches for correlations with estab-
    lished signals (e.g. flares in high-energy gamma-rays)
    triggered by neutrino observations. For sources which
    manifest large time variations in the emitted radiation,
    the signal-to-noise ratio can be increased by limiting
    the neutrino exposures to most favorable periods. The
    chance of discovery can then be enhanced (the so
    called ”multi-messenger approach”) by ensuring a good
    coverage of simultaneous data at a monitoring waveband
    (e.g. gamma-rays). The first realization of such an ap-
    proach led to two months of follow-up observations of
    AMANDA triggers by MAGIC, focused on a selected
    sample of Blazars as target sources [1]. An extension
    of this program to IceCube and also to optical follow-
    up observations has been later realized with the ROTSE
    network of optical telescopes, addressing possible cor-
    relations between neutrino multiplets and either GRBs
    or Supernovae [2].
    Multi-messenger studies can be accomplished off-
    line, searching for correlations between the measured
    intensity curves in the electromagnetic spectrum and the
    time of the detected neutrinos. The major limitations
    encountered so far were due to the scarce availability
    of information on the electromagnetic emission of the
    objects of interest, which typically are not observed
    continuously. Whenever data is available, such an a-
    posteriori approach is however very powerful, and it is
    part of the research plans of the IceCube Collaboration.
    We emphasize that a neutrino telescope at the South
    Pole is continuously and simultaneously sensitive to
    all objects located in the northern hemisphere. The
    investigation of the correlation between the observed
    properties of the electromagnetic emission and the de-
    tected neutrinos is therefore at any time feasible once
    the relevant electro-magnetic information is available.
    In other words, on-line and off-line approaches have to
    be seen as complementary and not mutually exclusive.
    In case of variable objects like Blazars, FSRQs as
    well as Galactic systems like microquasars and magne-
    tars, hadronic models describing the very high energy
    gamma-rays emission also predict simultaneous high
    energy neutrinos. Absorption processes might attenuate
    the gamma-ray luminosity when the objects are brightest
    in neutrinos, so that an anti-correlation or time-lag might
    be predicted as well. In all cases, the availability of
    simultaneous data on high energy gamma-ray emission
    and (possibly) neutrinos is mandatory to test different
    scenarios and shed light on the emission mechanisms
    (e.g. extract information on the optical depth and on
    other astrophysical source parameters).
    II. SELECTION OF TARGET SOURCES
    The most interesting objects as a target for gamma-ray
    follow-up observations of IceCube events are promising
    sources of TeV neutrinos, which are either known to
    exhibit a bright GeV flux in gamma-rays and show
    extrapolated fluxes detectable by Imaging Air Cherenkov
    Telescopes, or are already detected by IACTs and
    are variable. Candidates currently being considered are
    AGNs (HBL, LBL, FSRQs), Microquasars and Magne-
    tars (SGRs). A preliminary source list based on observa-
    tions with the FERMI [6] and EGRET [3] experiments
    is based on the following criteria:
    Source is present in both the third EGRET(3EG)
    and Fermi catalogues;
    Source is classified as variable in the Fermi cata-
    logue;
    Variability Index
    > 1
    in the 3EG catalog (taken
    from [5]);
    Maximum 3EG flux
    > 40 · 10
     8
    ph cm
     2
    s
     1
    , E >
    100
    MeV;

    2
    R. FRANKE
    et al.
    NTOO WITH ICECUBE
    cos(
    θ
    )
    -1
    -0.9
    -0.8
    -0.7
    -0.6
    -0.5
    -0.4
    -0.3
    -0.2
    -0.1
    0
    Rate [Hz]
    0
    0.02
    0.04
    0.06
    0.08
    0.1
    0.12
    0.14
    0.16
    0.18
    0.2
    0.22
    -3
    ×
    10
    Fig. 1. Predicted rate of atmospheric neutrinos based on Monte-Carlo
    for IceCube in its 2009/2010 configuration with
    59
    deployed strings.
    Average 3EG flux
    > 15 · 10
     8
    phcm
     2
    s
     1
    , E >
    100
    MeV;
    Difference between the maximum 3EG flux and the
    minimum 3EG flux
    > 30 · 10
     8
    ph cm
     2
    s
     1
    , E >
    100
    MeV.
    The sources that were selected according to these criteria
    can be found in Table I.
    III. EVENT SELECTION
    The basis for the event selection is an on-line filter
    that searches for up-going muon tracks. The rate of this
    filter is about
    24
    Hz for IceCube in its 2009/2010 con-
    figuration with
    59
    deployed strings. As the computing
    resources at the South Pole are limited one can not
    run more elaborate reconstructions at this rate, so a
    further event selection has to be done. This so called
    Level-2 filter searches events that were reconstructed
    with a zenith angle
    θ > 80
    (
    θ = 0
    equals vertically
    down-going tracks) with a likelihood reconstruction. By
    requiring a good reconstruction quality the background
    of misreconstructed atmospheric muons is further re-
    duced. The parameters used to assess the track quality
    are the likelihood of the track reconstruction and the
    number of unscattered photons with a small time residual
    w.r.t. the Cherenkov cone. The reduced event rate of
    approximately
    2.9
    Hz can then be reconstructed with
    more time intensive reconstructions, like a likelihood
    fit seeded with ten different tracks (iterative fit). The fit
    with the best likelihood is used for further cuts. Based on
    this reconstruction the final event sample is selected by
    employing a zenith angle cut of
    θ > 90
    for the iterative
    reconstruction and further event quality cuts based on
    this reconstruction. In addition to the already mentioned
    parameters we also employ a cut on the longest distance
    between hits with a small time residual compared to
    their expected arrival time calculated from the track
    geometry when projected on the reconstructed track. The
    resulting rate of atmospheric neutrinos as predicted by
    Monte Carlo as a function of zenith angle can be seen
    in Figure 1.
    IV. THE TIME-CLUSTERING ALGORITHM
    The timescale of a neutrino flare is not fixed a-priori
    and thus a simple rolling time window approach is not
    adequate to detect flares. The time clustering approach
    that was developed for an unbiased neutrino flare search
    [7] looks for any time frame with a significant deviation
    of the number of detected neutrinos from the expected
    background. The simplest implementation uses a binned
    approach where neutrino candidates within a fixed bin
    around a source are regarded as possible signal events.
    To exploit the information that can be extracted from
    the estimated reconstruction error and other event prop-
    erties like the energy an unbinned maximum-likelihood
    method is under development.
    If a neutrino candidate is detected at time
    t
    i
    around a
    source candidate the expected background
    N
    i,j
    bck
    is calcu-
    lated for all other neutrino candidates
    j
    with
    t
    j
    < t
    i
    from
    that source candidate. To calculate
    N
    i,j
    bck
    the detector
    efficiency as a function of the azimuth angle and the
    uptime has to be taken into account. The probability to
    observe the multiplet
    (i, j)
    by chance is then calculated
    according to
    ?
    k=N
    i,j
    obs
     1
    (N
    i,j
    bck
    )
    k
    k!
    e
     N
    i,j
    bck
    (1)
    where
    N
    obs
    is the number of detected on-source neutrinos
    between
    t
    j
    and
    t
    i
    . It has to be reduced by
    1
    to take
    into account the bias that one only does this calculation
    when a signal candidate is detected. As typical flares
    in high energy gamma-rays have a maximal duration of
    several days we constrain our search for time clusters of
    neutrinos to three weeks.
    If the cluster with the highest significance exceeds a
    certain threshold (e.g. corresponding to
    5 σ
    ) the detector
    stability will be checked and an alert will be send to an
    Cherenkov telescope to initiate a follow-up observation.
    V. DATA QUALITY
    Data quality is very important for any online alert
    program to minimize the rate of false alerts due to
    detector or DAQ instabilities. IceCube has a very ex-
    tensive monitoring of the DAQ and South Pole on-line
    processing. However, most of the information is only
    available with a certain delay after data-taking and thus
    not useful for follow-up program which requires fast
    alerts. To ensure that alerts are only sent for neutrino
    multiplets that where detected during stable running
    conditions a simple but powerful stability monitoring
    scheme has been developed. It is based on a continuous
    measurement of the relevant trigger and filter rates and
    their respective ratios in time bins of
    10
    minutes. These
    values are then compared to a running average of these
    rates over approximately four days to detect significant
    deviations. The running average is necessary as slow
    seasonal changes in the atmosphere and faster weather
    changes influence the rate of atmospheric muons which
    dominate the Level-2 rate. An example of this behaviour

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    3
    TABLE I
    PRELIMINARY CANDIDATE SOURCE LIST FOR NEUTRINO TRIGGERED FOLLOW-UP OBSERVATIONS. THE TYPE OF AGN HAS BEEN TAKEN
    FROM [4]
    Source name
    Blazar type
    Dec.
    [
    ]
    RA
    [
    ]
    max. 3EG flux
    [10
     8
    cm
     2
    s
     1
    ]
    min. 3EG flux
    [10
     8
    cm
     2
    s
     1
    ]
    avg. 3EG flux
    [10
     8
    cm
     2
    s
     1
    ]
    3C 273
    FSRQ
    2.0
    187.3
    48.3
    8.5
    15.4
    CTA 102
    FSRQ
    11.7
    338.1
    51.6
    12.1
    19.2
    GEV J0530+1340
    FSRQ
    13.5
    82.7
    351.4
    32.4
    93.5
    3C 454.3
    FSRQ
    16.1
    343.5
    116.1
    24.6
    53.7
    GEV J0237+1648
    LBL
    16.6
    39.7
    65.1
    11.6
    25.9
    15/12
    20/12
    25/12
    30/12
    05/01
    10/01
    15/01
    20/01
    25/01
    30/01
    05/02
    10/02
    15/02
    20/02
    25/02
    05/03
    10/03
    15/03
    20/03
    25/03
    30/03
    05/04
    10/04
    15/04
    20/04
    Date
    0.0
    0.5
    1.0
    1.5
    2.0
    2.5
    3.0
    3.5
    4.0
    Rate [Hz]
    Fig. 2. Rate of the Level-2 online filter for four months (December
    2008 till April 2009) in 10-minute bins with IceCube in its 2008/2009
    configuration with 40 deployed strings. The Level-2 filter is an online
    filter that is used for different follow-up observation programs. The
    slow change in the rate is due to seasonal variations in the atmospheric
    muon background rate caused by pressure changes in the atmosphere.
    can be seen in Figures 2 and 3. This system was
    tested off-line on data from IceCube in its 40-string
    configuration and proved to correlate very well with the
    extensive off-line detector monitoring. The fraction of
    data that has to be discarded due to detector or software
    problems was about
    6 %
    , which includes all periods
    in Figures 2 and 3 that significantly deviate from the
    average. This method will be implemented online for
    IceCube in its 2009/2010 configuration with
    59
    deployed
    strings.
    VI. SIGNIFICANCE CALCULATION
    Under the hypothesis that all the neutrinos are of
    atmospheric origin, the probability of observing at least
    N
    obs
    multiplets above the significance threshold and
    detecting at least
    N
    coinc
    coincident gamma-ray flares is
    given by:
    ?
    +∞
    m=N
    obs
    (N
    bck
    )
    m
    m!
    e
     N
    bck
    ?
    m
    j=N
    coinc
    m!
    j!(m  j)!
    (p
    gam
    )
    j
    (1 p
    gam
    )
    m j
    (2)
    where the first term describes the Poisson probability
    of observing at least
    N
    obs
    neutrino multiplets with
    N
    bck
    background expected, and the second term describes the
    probability of observing at least
    N
    coinc
    out of
    m
    – the
    running number of observed multiplets, larger or equal
    0123456
    Level 2 online filter rate [Hz]
    0
    500
    1000
    1500
    2000
    2500
    Number of 10 min bins
    Fig. 3. Histogram of the rates of the online Level-2 filter of IceCube
    in its 2008/2009 configuration with 40 deployed strings for the four
    months shown in figure 2. The bin at a value of
    5.5
    Hz contains all
    entries bigger than
    5.5
    Hz.
    to
    N
    obs
    – each with a probability of
    p
    gam
    . We note that this
    probability can be calculated anytime a-posteriori, once
    a realistic knowledge of the probability
    p
    gam
    to detect
    a gamma-ray flare in a time window
    ∆t
    is available.
    In order to avoid statistical biases it is mandatory,
    however, that the statistical test is defined a-priori, i.e.
    that the conditions to accept an observation and defining
    a coincidence are previously fixed. Methods on how the
    to reliably estimate the probability
    p
    gam
    of detecting a
    gamma-ray flare in a time window
    ∆t
    , which is influ-
    enced by the source elevation and weather conditions,
    from the frequency of the observed gamma-ray flares are
    under development. The significance calculated above
    also does not account for the trial factor correction due
    to the selection of three or more objects, which can
    however be calculated as the product of the individual
    terms corresponding to each source. The probability of
    having at least one coincidence in any of the proposed
    sources is, for example:
    P =1
    N
    ?
    Sources
    i=1
    P
    0
    i
    (3)
    where
    P
    0
    i
    is the probability of having zero coincidences
    at the source
    i
    .
    VII. THE GAMMA-RAY FOLLOW-UP OBSERVATION
    SCHEME
    We propose an observation scheme as follows:

    4
    R. FRANKE
    et al.
    NTOO WITH ICECUBE
    0
    10
    20
    30
    40
    50
    60
    70
    80
    Declination [degree]
    10
    -2
    10
    -1
    10
    0
    10
    1
    Alert rate [1/year]
    3 sigma
    5 sigma
    Fig. 4. Preliminiary alert rate from atmospheric neutrino background
    for IceCube in its 2009/2010 configuration with
    59
    deployed strings
    for an alert threshold for the multiplet significance corresponding to
    (upper points) and
    (lower points) and a bin size of
    2
    .
    Up to 1 day after receiving an IceCube alert from
    one of the pre-defined directions, the source is
    scheduled to be observed as soon as visible and
    observation conditions allow.
    If the gamma-ray observation is possible, it will
    continue for one hour.
    The results of the on-line analysis will be checked
    and, if there is a positive hint (above 3
    σ
    ) the
    gamma-ray observations may be extended. In case
    of a positive observation (i.e. a gamma-ray flux
    trespassing the pre-defined threshold defining a
    flare), the opportunity to trigger multi-wavelength
    observations should then be considered. Due to
    the irreducible background of atmospheric neutri-
    nos (Figure 1) one can estimate the alert rate for
    different zenith regions (Figure 4) for thresholds
    corresponding to
    and
    . The on-source bin has
    been preliminarily chosen to have a radius of
    2
    .
    Based on a simple Monte Carlo simulation that does not
    take into account detector features like the azimuth de-
    pendent efficiency we calculated the discovery probabil-
    ities for different numbers of injected on-source events
    (see Figure 5) at a declination of
    26
    . The discovery
    probability is defined here as the probability to detect a
    5 σ
    deviation with the time clustering method.
    VIII. STATUS
    The event selection and software to calculate the
    significance of a neutrino cluster are implemented and
    ready to be deployed at the South Pole. As IceCube in
    its 2009/2010 configuration with
    59
    deployed strings is
    considerably bigger than the previous detector configu-
    ration the stability monitoring needs to be checked with
    the first weeks of physics data. Pending the approval
    of the follow-up program by a Cherenkov telescope
    collaboration we then aim for a timely implementation
    of this program.
    1234567
    Number of injected signal neutrino events
    0.0
    0.1
    0.2
    0.3
    0.4
    0.5
    0.6
    0.7
    0.8
    Probability to trigger on neutrino flare
    1 day
    2 days
    4 days
    8 days
    Fig. 5.
    Preliminary discovery probability for a certain number of
    injected on-source neutrino events for different flare durations for a
    source at a declination of
    26
    . The discovery probability is defined
    here as the probability to detect a
    5 σ
    deviation with the time-clustering
    method. This does not contain the probability of the gamma-ray
    observation.
    IX. OUTLOOK
    Besides enhancing the chance to discover point
    sources of neutrinos, the gamma-ray follow-up approach
    here discussed can increase the chance of detecting
    unusual gamma-ray emission of the selected objects.
    It also can provide an important contribution to the
    understanding of the flaring behavior of a few emitters of
    high energy gamma-rays in a way complementary to X-
    ray observations. Most relevant, it can provide a series of
    coincidences and therefore represent an important input
    to dedicated multi-wavelength follow-up observations,
    which will assess in more details the phenomenology
    of the potential sources. In fact – thanks to the existing
    communication infrastructures of multi-wavelength cam-
    paigns – the observation of gamma-ray flares can start
    a monitoring of the objects at other wavelengths (e.g.
    X-ray) that would further complement the information
    that are discussed here.
    REFERENCES
    [1] M. Ackermann
    et al.
    , Proc. 29th ICRC, arXiv:astro-ph/0509330.
    [2] Kowalski, M. and Mohr, A., Astropart. Phys., Vol. 27, (2007)
    pp. 533-538, arXiv:astro-ph/0701618.
    [3] Hartman
    et al.
    , The Astrophysical Journal Supplement Series,
    Vol. 123, No. 1. (1999), pp. 79-202.
    [4] Nandikotkur G., Jahoda K.M., Hartman R.C., 2007, ApJ, 657,
    706.
    [5] Nolan P.L.
    et al.
    , The Astrophysical Journal. Vol 597, No 1,
    (2003) pp. 615-627.
    [6] Fermi Collab. Submitted to Astrophysical Journal Supplement
    (2009).
    [7] K. Satelecka
    et al.
    for the IceCube collaboration, Proc. 30th
    ICRC, pp. 115-118.

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    1
    Moon Shadow Observation by IceCube
    D.J. Boersma
    , L. Gladstone
    and A. Karle
    for the IceCube Collaboration
    RWTH Aachen University, Germany
    Department of Physics, University of Wisconsin, Madison, WI 53706, USA
    see special section of these proceedings
    Abstract
    . In the absence of an astrophysical stan-
    dard candle, IceCube can study the deficit of cosmic
    rays from the direction of the Moon. The observation
    of this “Moon shadow” in the downgoing muon flux is
    an experimental verification of the absolute pointing
    accuracy and the angular resolution of the detector
    with respect to energetic muons passing through.
    The Moon shadow has been observed in the 40-
    string configuration of IceCube. This is the first stage
    of IceCube in which a Moon shadow analysis has
    been successful. Method, results, and some systematic
    error studies will be discussed.
    Keywords
    : IceCube, Moon shadow, pointing capa-
    bility
    I. INTRODUCTION
    IceCube is a kilometer-cube scale Cherenkov detector
    at the geographical South Pole, designed to search
    for muons from high energy neutrino interactions. The
    arrival directions and energy information of these muons
    can be used to search for point sources of astrophysical
    neutrinos, one of the primary goals of IceCube.
    The main component of IceCube is an array of optical
    sensors deployed in the glacial ice at depths between
    1450 m and 2450 m. These Digital Optical Modules
    (DOMs), each containing a 25 cm diameter photo-
    multiplier tube with accompanying electronics within
    a pressure housing, are lowered into the ice along
    “strings.” There are currently 59 strings deployed of 86
    planned; the data analyzed here were taken in a 40 string
    configuration, which was in operation between April
    2008 and April 2009. There are 13 lunar months of data
    within that time. In this analysis we present results from
    8 lunar months of the 40 string configuration.
    For a muon with energy on the order of a TeV,
    IceCube can reconstruct an arrival direction with or-
    der
    1
    accuracy. For down-going directions, the vast
    majority of the detected muons do not originate from
    neutrino interactions, but from high energy cosmic ray
    interactions in the atmosphere. These cosmic ray muons
    are the dominant background in the search for astro-
    physical neutrinos. They can also be used to study the
    performance of our detector. In particular, we can verify
    the pointing capability by studying the shadow of the
    Moon in cosmic ray muons.
    As the Earth travels through the interstellar medium,
    the Moon blocks some cosmic rays from reaching the
    Earth. Thus, when other cosmic rays shower in the
    Earth’s atmosphere and create muons, there is a rela-
    tive deficit of muons from the direction of the Moon.
    IceCube detects these muons, not the primary cosmic
    rays. Since the position and size of the Moon is so well
    known, the resulting deficit can be used for detector cal-
    ibration. The idea of a Moon shadow was first proposed
    in 1957 [1], and has become an established observation
    for a number of astroparticle physics experiments; some
    examples are given in references [2], [3], [4], [5]. Exper-
    iments have used the Moon shadow to calibrate detector
    angular resolution and pointing accuracy [6]. They have
    also observed the shift of the Moon shadow due to the
    Earth’s magnetic field [7]. The analysis described here
    is optimized for a first observation, and does not yet
    include detailed studies such as describing the shape of
    the observed deficit. These will be addressed in future
    studies.
    II. METHOD
    A. Data and online event selection
    Data transfer from the South Pole is limited by the
    bandwidth of two satellites; thus, not all downgoing
    muon events can be immediately transmitted. This anal-
    ysis uses a dedicated online event selection, choosing
    events with a minimum quality and a reconstructed
    direction within a window of acceptance around the
    direction of the Moon. The reconstruction used for the
    online event selection is a single (i.e., not iterated) log-
    likelihood fit.
    The online event selection is defined as follows, where
    δ
    denotes the Moon declination:
    The Moon must be at least
    15
    above the horizon.
    At least 12 DOMs must register each event.
    At least 3 strings must contain hit DOMs.
    The reconstructed direction must be within 10
    of
    the Moon in declination.
    The reconstructed direction must be within
    40
    / cos(δ)
    of the Moon in right ascension; the
    cos(δ)
    factor corrects for projection effects.
    These events are then sent via satellite to the northern
    hemisphere for further processing, including running the
    higher-quality 32-iteration log-likelihood reconstruction
    used in further analysis.
    The Moon reached a maximum altitude of
    27
    above
    the horizon (
    δ =  27
    ) in 2008, when viewed from

    2
    D.J. BOERSMA
    et al.
    ICECUBE MOON SHADOW
    log10(E/GeV)
    3
    4
    5
    6
    7
    8
    rate [Hz]
    -3
    10
    10
    -2
    10
    -1
    1
    10
    10
    2
    Fig. 1. The energy spectrum of (simulated) CR primaries of muons
    (or muon bundles) triggering IceCube. Red: all events; blue: primaries
    with
    δ >  30
    .
    time since 3 September 2008 [days]
    0
    10
    20
    30
    40
    50
    60
    event rate [Hz]
    0
    5
    10
    15
    20
    25
    ]
    o
    -(Moon declination)[
    0
    5
    10
    15
    20
    25
    Fig. 2.
    The rate of events passing the Moon filter (in Hz, lower
    curve) averaged hourly, together with the position of the Moon above
    the horizon at the South Pole (in degrees, upper curve), plotted versus
    time over 3 typical months.
    the IceCube detector. The trigger rate from cosmic ray
    muons is more than 1.2 kHz in the 40 string configu-
    ration, but most of those muons travel nearly vertically,
    and thus they cannot have come from directions near the
    Moon. Only
    ∼ 11%
    of all muons that trigger the detector
    come from angles less than
    30
    above the horizon.
    Furthermore, muons which are closer to horizontal (and
    thus closer to the Moon) must travel farther before
    reaching the detector. They need a minimum energy
    to reach this far (see Fig. 1): the cosmic ray primaries
    which produce them must have energies of at least 2 TeV.
    Three typical months of data are shown in Fig. 2,
    along with the position of the Moon above the horizon.
    The dominant shape is from the strong increase in muon
    flux with increasing angle above the horizon: as the
    Moon rises, so do the event rates near the Moon. This
    can be seen clearly in the correlation between the two
    sets of curves. There is a secondary effect from the
    layout of the 40 strings. One dimension of the detector
    layout has the full width (approximately 1km) of the
    completed detector, while the other is only about half
    as long. When the Moon is aligned with the short axis,
    fewer events pass the filter requirements. This causes the
    12 hour modulation in the rate.
    B. Optimization of offline event selection and search bin
    size
    A simulated data sample of
    10
    5
    downgoing muon
    events was generated using CORSIKA [8].
    A set of cuts was developed using the following esti-
    mated relation between the significance
    S
    , the efficiency
    ψ
    [degrees]
    0
    1
    2
    3
    4
    5
    6
    7
    8
    9
    10
    ’)
    ψ
    ’ PSF(
    ψ
    d
    ψ
    0
    0
    0.2
    0.4
    0.6
    0.8
    1
    Fig. 3. The x-axis shows the angular difference
    ψ
    between the true
    and reconstructed track. The y-axis shows the fraction of events with
    this or lower angular error. The blue curve shows the event sample
    after offline event selection, and the red curve shows the event sample
    after online event selection.
    η
    of events passing the cut, and the resulting median
    angular resolution
    Ψ
    med
    of the sample:
    S (cuts) ∝
    ?
    η(cuts)
    Ψ
    med
    (cuts)
    (1)
    Since the deficit is based on high statistics of events in
    the search bin, this function provides a good estimator
    for optimizing the significance.
    The following cuts were chosen:
    At least 6 DOMs are hit with light that hasn’t been
    scattered in the ice, allowing a -15 nsec to +75 nsec
    window from some minimal scattering.
    Projected onto the reconstructed track, two of those
    hits at least 400 meters apart.
    The
    estimated error ellipse on the reconstructed
    direction has a mean radius less than
    1.3
    .
    The cumulative point spread function of the sample after
    the above quality cuts is shown as the blue line in Fig. 3.
    The size
    Ψ
    search
    of the search bin is optimized for a
    maximally significant observation using a similar
    N
    -
    error based argument and the resulting relation, which
    follows. Using the cumulative point spread function of
    the sample after quality cuts, we have:
    S (Ψ
    search
    ) ∝
    ?
    Ψ
    search
    0
    P SF (ψ
    )dψ
    Ψ
    search
    (2)
    Maximizing this significance estimator gives an optimal
    search bin radius of
    0.7
    . This analysis uses square bins
    with an area equal to that of the optimized round bin,
    with side length
    1.25
    .
    C. Calculating significance
    To show that the data are stable in right ascension
    α
    ,
    we show, in Fig. 4, the number of events in the central
    declination band. The errors shown are
    N
    . The average
    of all bins excluding the Moon bin is 27747, which is
    plotted as a line to guide the eye. The Moon bin has 852
    events below this simple null estimate. This represents
    a
    5.2σ
    deficit using
    N
    errors.

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    3
    (
    α
    evt
    -
    α
    Moon
    )cos(
    δ
    Moon
    ) [degrees]
    -20
    -10
    0
    10
    20
    events
    27000
    27500
    28000
    IceCube
    IceCube preliminarypreliminary
    Fig. 4. Number of events per
    1.25
    square bin, relative to the position of the Moon. The declination of the reconstructed track is within
    0.625
    bin from the declination of the Moon. The average of all bins except the Moon bin is shown as a redline to guide the eye.
    -0.07 -0.39 -0.96 -0.05 -0.34 1.62 -0.72 -0.11 0.98 1.69 1.66 -0.14 -0.4 -1.5 0.63 1.17 0.75 -0.22 0.72 -0.04 0.44 -0.54 -0.88
    -1.05 0.52 -0.69 1.94 0.41 1.3 -0.05 -0.06 -1.17 0.83 -0.43 -0.61 -1.59 -0.2 -1.26 0.38 0.36 1.23 -0.91 -0.41 -0.05 0.72 1.13
    -1.27 0.67 -0.09 0.15 -2.16 0.2 1.13 0.45 0.55 0.63 -1.23 -1.14 -1.69 0.21 0.97 -0.75 0.33 -0.98 -0.74 -0.6 0.49 0.93 -0.27
    -0.4 0.05 -0.07 0.02 -0.29 0.98 2.16 0.05 -0.74 -0.86 -0.37 -5.04 -1.89 -0.6 -0.64 0.9 -0.43 2.95 -1.06 -0.77 3.04 0.69 0.8
    1.86 0.34 0.91 -0.92 0.79 -1.27 0.98 -0.56 0.07 -0.5 -1.17 -1.42 -1.52 -0.06 1.14 -0.64 0.59 -0.15 0.62 -0.75 0.89 1.02 0.65
    0.81 1.47 -1.7 -0.28 -1.35 -1.31 -0.74 -1.4 -0.86 1.13 0.5 -0.82 2.84 1.68 -1.3 1.99 0.58 -2.03 -0.02 -0.22 -1.05 0.47 -2.31
    0.02 -0.33 -0.26 0.98 0.29 1.06 -0.81 0.01 -1.22 -1.58 0.47 0.34 -0.46 -0.8 -1.22 -1.98 1.73 3.43 0.03 -0.31 0.86 -0.47 0.49
    (
    α
    event
    -
    α
    Moon
    )cos(
    δ
    Moon
    ) [degrees]
    -10
    -5
    0
    5
    10
    [degrees]
    event
    δ
    -
    Moon
    δ
    -4
    -3
    -2
    -1
    0
    1
    2
    3
    4
    -6
    -4
    -2
    0
    2
    4
    IceCube preliminary
    6
    Fig. 5. The significance of deviations in a region centered on the Moon.
    Significance observed
    -4
    -2
    0
    2
    4
    # of bins
    0
    5
    10
    15
    20
    25
    30
    χ
    2
    / ndf
    9.715 / 9
    Constant
    30.09
    ±
    3.18
    Mean
    -0.02742
    ±
    0.08082
    Sigma
    0.9472
    ±
    0.0633
    Fig. 6. Each of the deviations shown in Fig. 5 is plotted here. The
    deviations of the central 9 bins are shown in red. The surrounding bins
    are shown with a black line histogram, and fit with a Gaussian curve.
    Although this shows that the data are stable, this
    error system is vulnerable to variations in small data
    samples. Although we don’t see such variations here, we
    considered it prudent to consider an error system which
    takes into account the size of the background sample.
    We used a standard formula from Li and Ma [9] for
    calculating the significance of a point source:
    S =
    N
    on
     αN
    off
    ?
    α(N
    on
    + N
    off
    )
    .
    (3)
    where
    N
    on
    is the number of events in the signal sample,
    N
    off
    is the number of events in the off-source region, and
    α
    is the ratio between observing times on- to off-source.
    We take
    α
    instead as the ratio of on- to off-source areas
    observed, since the times are equal.
    The above significance formula is applied to the Moon
    data sample in the following way. The data are first
    plotted in the standard Moon-centered equatorial coor-
    dinates, correcting for projection effects with a factor of
    cos(δ)
    . The plot is binned using the
    1.25
    × 1.25
    bin
    size optimized in the simulation study. Each bin suc-
    cessively is considered as an on-source region. There is
    a very strong declination dependence in the downgoing
    muon flux, so variations of the order of the Moon deficit
    are only detectable in right ascension. Thus, off-source
    regions are selected within the same zenith band as the
    on-source region. Twenty off-source bins are used for
    each calculation: ten to the right and ten to the left of
    the on-source region, starting at the third bin out from
    the on-source bin (i.e., skipping two bins in between).
    III. RESULTS
    For a region of 7 bins or
    8.75
    in declination
    δ
    and 23
    bins or
    28.75
    in right ascension
    α
    around the Moon, the
    significance of the deviation of the count rate in each bin
    with respect to its off-source region was calculated, as
    described in section II-C. The result is plotted in Fig. 5.
    The Moon can be seen as the
    5.0σ
    deficit in the central
    bin, at
    (0, 0)
    .
    To test the hypothesis that the fluctuations in the back-
    ground away from the Moon are distributed randomly
    around 0, we plot them in Fig. 6. The central 9 bins,
    including the Moon bin, are not included in the Gaussian

    4
    D.J. BOERSMA
    et al.
    ICECUBE MOON SHADOW
    fit, but are plotted as the lower, shaded histogram. The
    width of the Gaussian fit is consistent with 1; therefore,
    the background is consistent with random fluctuations.
    IV. CONCLUSIONS AND FUTURE PLANS
    IceCube has observed the shadow of the Moon as
    a
    5.0σ
    deviation from event counts in nearby regions,
    using data from 8 of the total 13 lunar months in the data
    taking period with the 40-string detector setup. From
    this, we can conclude that IceCube has no systematic
    pointing error larger than the search bin,
    1.25
    .
    In the future, this analysis will be extended in many
    ways. First, we will include all data from the 40 string
    detector configuration. We hope to repeat this analysis
    using unbinned likelihood methods, and to describe the
    size, shape, and any offset of the Moon Shadow. We will
    then use the results of these studies to comment in more
    detail on the angular resolution of various reconstruction
    algorithms within IceCube. This analysis is one of the
    only end-to-end checks of IceCube systematics based
    only on experimental data.
    LG acknowledges the support of a National Defense
    Science and Engineering Graduate Fellowship from the
    American Society for Engineering Education.
    REFERENCES
    [1] G.W. Clark,
    Arrival Directions of Cosmic-Ray Air Showers from
    the Northern Sky
    October 15, 1957, Physical Review Vol 108,
    no 2.
    [2] A. Karle for the HEGRA collaboration,
    The Angular Resolution
    of the Hegra Scintillation Counter Array at La Palma
    , Ann Arbor
    1990 Proceedings, High Energy Gamma-Ray Astronomy 127-
    131.
    [3] Giglietto, N.
    Performance of the MACRO detector at Gran Sasso:
    Moon shadow and seasonal variations
    , 1997, Nuclear Physics B
    Proceedings Supplements, Volume 61, Issue 3, p. 180-184.
    [4] M.O. Wasco for the Milagro collaboration,
    Study of the Shadow
    of the Moon and the Sun with VHE Cosmic Rays
    , 1999
    arXiv:astro-ph/9906.388v1
    [5] The Soudan 2 collaboration,
    Observation of the Moon Shadow in
    Deep Underground Muon Flux
    , 1999 arXiv:hep-ex/9905.044v1
    [6] The Tibet AS Gamma Collaboration, M. Amenomori,
    et al.
    ,
    Multi-TeV Gamma-Ray Observation from the Crab Nebula Using
    the Tibet-III Air Shower Array Finely Tuned by the Cosmic-Ray
    Moon’s Shadow
    , arXiv:astro-ph/0810.3757v1
    [7] L3 Collaboration, P. Achard
    et al.
    ,
    Measurement of the Shad-
    owing of High-Energy Cosmic Rays by the Moon: A Search
    for TeV-Energy Antiprotons
    Astropart.Phys.23:411-434,2005,
    arXiv:astro-ph/0503472v1
    [8] D. Heck, J. Knapp, J.N. Capdevielle, G. Schatz, T. Thouw,
    CORSIKA: A Monte Carlo Code to Simulate Extensive Air
    Showers
    , FZKA 6019 (1998)
    [9] Li, T.-P. and Ma, Y.Q.,
    Analysis methods for results in gamma-
    ray astronomy
    1983, ApJ 272,317

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁOD´ Z´ 2009
    1
    IceCube/AMANDA combined analyses
    for the search of neutrino sources at low energies
    Cecile´
    Roucelle
    , Andreas Gross
    , Sirin Odrowski
    , Elisa Resconi
    , Yolanda Sestayo
    for the IceCube Collaboration
    MPIK, Heidelberg, Germany
    See the special section of these proceedings
    Abstract. During the construction of IceCube, the
    AMANDA neutrino telescope has continued to ac-
    quire data and has been surrounded by IceCube
    strings. Since the year 2007, AMANDA has been
    fully integrated for data acquisition and joint Ice-
    Cube/AMANDA events have been recorded. Because
    of the finer spacing of AMANDA phototubes, the
    inclusion of AMANDA significantly extends the de-
    tection capability of IceCube alone for low energy
    neutrinos (100 GeV to 10 TeV). We present the results
    of two analyses performed on the 2007-2008 Icecube
    (22 string) and AMANDA data. No evidence of high
    energy neutrino emission was observed; upper limits
    are reported.In 2008-09, IceCube acquired data in a
    40 string configuration together with the last year of
    operation of AMANDA. Progress on the analysis of
    this new combined IceCube/AMANDA sample are
    presented as well. In addition, a novel method to
    study an extended region surrounding the most active
    parts of Cygnus with these datasets is described here.
    Keywords: Neutrino astronomy, galactic sources,
    IceCube, AMANDA, DeepCore
    I. INTRODUCTION
    Recent detections by Cherenkov telescopes provide
    evidence of particle acceleration up to TeV energies
    in astrophysical sources [1]. The TeV γ-ray emission
    from these sources could arise from the acceleration
    of electrons (production of γ-rays via inverse Compton
    scattering) or the acceleration of hadrons (production
    of γ-rays through the decay of neutral pions produced
    in pp/pγ interactions). In the later scenarios, the γ-
    ray production would be accompanied by the neutrino
    production since charged pions, like neutral pions, would
    be generated and decay within the source. The detection
    of high-energy neutrinos would thus be an unambiguous
    proof for the acceleration of hadrons in these sources.
    In particular, galactic TeV γ-ray sources present the
    bulk of their γ-ray emission at energies lower than a
    few TeV. The spectrum from these sources is soft with
    a typical spectral index (|Γ| > 2) and often exhibits
    an exponential cut-off at a few TeV. Both observations
    suggest a break in the neutrino spectrum below 100
    TeV. Accordingly, the flux from these sources would
    differ from the standard spectral index of -2 for neu-
    trino sources. Additionnaly, they represent “low energy”
    sources (TeV) for IceCube and would be challenging
    to detect. To enhance the sensitivity to this type of
    sources, an analysis comprising both the IceCube and
    the AMANDA detector has been performed. The higher
    density of optical modules in AMANDA than in Ice-
    Cube provides a sufficient increase in the number of
    hits that reconstruction of low energy, neutrino-induced
    events is possible. This increase in statistics particularly
    benefits searches for sources with steeply falling spectra
    (see Sec. III and Sec.IV). A first analysis has been
    made using the 22 string configuration of IceCube in
    combination with the AMANDA detector; the results
    are presented in this proceeding. A new sample of data
    has been collected with IceCube-40 and AMANDA and
    is under analysis. We present here the general scheme
    for this analysis, with particular emphasis on a specific
    development to enhance the detection sensitivity for
    extended active regions in the galactic plane.
    II. GALACTIC SOURCES : THE γ-ν CONNECTION
    Since neutrino and γ-rays are expected to be produced
    together in hadronic acceleration processes, the neutrino
    spectrum can be inferred from the observed γ-ray spec-
    trum of the source by a two-step procedure:
    1 - The γ-ray spectrum from a source is fitted as-
    suming a pp interaction model obtained using the
    parametrizations given in [4]. Possible γ-ray ab-
    sorption is estimated and corrected for before the
    fit.
    2 - With the obtained proton distribution and the
    target density, the expected neutrino spectrum is
    estimated.
    The Crab Nebula γ-ray energy spectrum has been mea-
    sured in details by the H.E.S.S. experiment [8]. It is
    described by a power law with spectral index (Γ) of
    -2.4 and has a γ-ray energy cutoff at ∼14 TeV. Although
    numerous arguments attribute the γ-ray production from
    this source to e
    +
    /e
    acceleration, its status as a standard
    candle argues for its use as a reference for neutrino
    astronomy. Moreover, the establishment of sufficiently
    low upper limits by IceCube on the neutrino emission
    could bring new constraints on the possible hadron ac-
    celeration at this source. Assuming that γ-rays from the
    Crab Nebula originate from hadronic processes (decay of
    π
    0
    mesons generated from pp interactions at the source)
    and that their absoption is negligible, the ν spectrum
    obtained is:
    Φ = 3 × 10
    −7
    e
    −E/7TeV
    (E/GeV)
    −2.4
    GeV
    −1
    cm
    −2
    s
    −1
    (1)

    2
    C. ROUCELLE et al. ICECUBE-AMANDA COMBINED ANALYSIS
    In the following, we use this computed spectrum as
    a reference (“Crab Nebula spectrum”) to estimate the
    sensitivity of analyses to low energy sources.
    III. ICECUBE-22/AMANDA: RESULTS
    During the two deployment seasons 2003-2004 and
    2004-2005 at the South Pole, the data acquisition system
    (DAQ) of AMANDA was significantly upgraded to pro-
    vide nearly deadtime-less operation and full digitization
    of the electronic readout [2]. This was achieved by using
    Transient Waveform Recorders (TWR). The new DAQ
    system allowed for the reduction of the multiplicity trig-
    ger threshold and, consequently, of the energy threshold
    to ∼50 GeV. By being optimally sensitive to neutrinos
    under 1 TeV, AMANDA thus complements IceCube well
    and was integrated into the full IceCube analysis starting
    in January 2006.
    A. Data sample and methods
    The IceCube 22-string run represents 276 days taken
    between May 2007 and April 2008. Within this pe-
    riod, the AMANDA detector was taking jointly with
    IceCube for 143 days. Nevertheless, since the 2006-07
    deployment season, every time the AMANDA detector is
    triggered, a readout request is sent to the IceCube detec-
    tor. Events are then merged for processing. The trigger
    rates are strongly dominated by downgoing, atmospheric
    muons produced in cosmic ray air showers above the
    detector. They outnumber atmospheric neutrinos by a
    factor ∼ 10
    6
    . This background is largely eliminated
    by limiting the analysis to upgoing muons using a fast
    reconstruction algorithm which is applied to all of the
    data. The selected events are then further pared down by
    applying a cpu-intensive, likelihood-based reconstruc-
    tion algorithm that accounts for the properties of the
    ice and then cutting on the fit direction and fit quality
    parameters. In this analysis, these cuts were optimized
    to obtain the best discovery potential for a source with a
    “Crab Nebula” spectrum (Eqn. 1). As low energy events
    are mainly due to the dominant atmospheric neutrino
    background, a significantly larger number of events is
    obtained with this selection than with other IceCube-22
    point source searches [12].
    In total, 8727 events are selected, of which 3430
    are combined IceCube/AMANDA events. Despite the
    smaller size of AMANDA (1/6 of the volume of
    IceCube-22) and its shorter livetime (less than 60%
    wrt. IceCube-22), the contribution of AMANDA to the
    combined detector sample, particularly at low energies,
    is clearly visible in the energy distribution simulated at-
    mospheric neutrinos retained at the final event selection
    in the analysis (Fig. 1). As a consequence, the sensitivity
    achieved with this approach for a source with a spectrum
    similar to the one expected for the Crab Nebula (Γ=-2.4;
    cut-off at 7 TeV) is better than the one achieved with
    the IceCube only analysis (Fig. 2). Even though, for a
    harder spectrum (Γ=-2;no cutoff), the standard IceCube
    only analysis remains better adapted.
    log (E/GeV)
    1
    1.5
    2
    2.5
    3
    3.5
    4
    4.5
    5
    5.5
    MC)
    ν
    #events (conv. atm
    0
    100
    200
    300
    400
    500
    combined (IC22+AMANDA)
    AMANDA
    IC22, no AMANDA
    Fig. 1. Event energy distribution for simulated atmospheric neutrinos
    at the final level of the galactic point source analysis normalized to the
    livetime of the IceCube 22 strings data taking (276 days) for IceCube
    only events and to the combined IceCube+AMANDA livetime (143
    days) for the AMANDA and combined events.
    log(E/GeV)
    2
    10
    3
    10
    4
    10
    5
    10
    6
    10
    ]
    -1
    s
    -2
    [GeV cm
    Φ
    2
    E
    -8
    10
    -7
    10
    -6
    10
    std. analysis, Crab-like spectrum
    Low E analysis, Crab-like spectrum
    -2
    spectrum
    std. analysis, E
    -2
    spectrum
    Low E analysis, E
    Fig. 2.
    Sensitivity for a source spectrum with Γ=-2 and a “Crab”
    spectrum (Γ=-2.4; cut-off at 7 TeV). This analysis (gray) is compared
    to the standard IceCube only analysis (black).
    B. Search on an a priori selected list of point sources
    With this dataset, a search for neutrino emission was
    performed for a list of four, preselected sources: the Crab
    Nebula, Cas A, SS 433 and LS I +61 303. For three of
    them, the γ-ray spectrum is known ([8]-[11]), so we
    optimized the analysis for the expected corresponding
    neutrino spectrum (for SS 433, which has no measured
    γ-ray spectrum, the optimisation was made with respect
    to a test spectrum with a spectral index Γ=-2.4 and a
    cut-off at 7 TeV). The test-statistic for the analysis is
    the log likelihood ratio of the signal hypothesis with
    best fit parameters to the pure background hypothesis.
    This method is widely used in IceCube [7]. This test-
    statistic provides an estimate for the significance of a
    deviation from background (pre-trial p-value) at a posi-
    tion in the sky. The post-trial p-value is then determined
    by applying the analysis to randomized samples. With
    this method, the lowest pre-trial p-value (p=0.14) was
    obtained for the Crab Nebula. This p-value or a lower
    one can be achieved in 37% of randomized samples.
    This excess is therefore not significant. The number of

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁOD´ Z´ 2009
    3
    signal events detected and their associated pre-trial p-
    values are summarized in the table below. Based on the
    γ-ray observations, the expected neutrino spectral index
    and possible cut-off energies have been calculated using
    the method described in Sec. II and are indicated in the
    same table.
    Source
    Γ
    ν
    ν cut-off
    Nb. of signal
    p-value
    events
    (pre-trial)
    Crab Nebula
    -2.39
    7 TeV
    3.3
    0.14
    Cas A
    -2.4
    -
    -1.9
    0.65
    SS433
    -
    -
    -0.9
    0.67
    LSI+61 303
    -2.8
    -
    -0.4
    0.47
    E [GeV]
    3
    10
    4
    10
    5
    10
    ]
    -1
    s
    -2
    dN/dE [GeVcm
    2
    E
    -9
    10
    -8
    10
    -7
    10
    -6
    10
    -5
    10
    Crab reference neutrino spectrum
    Crab (IC22+AMANDA upper limit)
    E [GeV]
    3
    10
    4
    10
    5
    10
    ]
    -1
    s
    -2
    dN/dE [GeVcm
    2
    E
    -13
    10
    -12
    10
    -11
    10
    CasA (IC22+AMANDA upper limit)
    CasA reference neutrino spectrum
    Fig. 3. Top: Crab upper limit obtained for this study compared to
    the reference neutrino spectrum computed for the Crab as in section II
    using [4] for modeling. Bottom: Cas A upper limit obtained for this
    study compared to its reference neutrino spectrum.
    Upper limits on the neutrino flux were derived from
    the number of events observed in the direction of the dif-
    ferent sources with this analysis. The limits obtained for
    the Crab Nebula and Cas A are presented in Fig. 3 and
    compared to their expected neutrino spectrum. The limit
    that can be set by this IceCube-22/AMANDA analysis
    is for example for the Crab Nebula a factor 18.9 above
    the expected reference spectrum. This calculation was
    also made for the case of Cas A (Fig. 3, bottom). This
    source was detected by HEGRA up to 10 TeV without
    evidence of high energy cut-off [9]. We extrapolate the
    power-law γ-ray spectrum given in [10] up to higher
    energies.
    Fig. 4. Galactic plane scan (longitude : 31.5
    < l < 214.5
    , latitude
    : -3
    < b < 3
    ) pretrial significance map for IceCube-22/AMANDA.
    The strongest excess at l=75.875
    , b=2.675
    (pre-trial p-value =
    0.0037). 95% of randomized datasets yielded a more significant excess.
    C. Galactic plane scan
    In addition to these sources, we performed an un-
    binned point source search of the galactic plane in
    the nominal field of view of IceCube (longitude :
    31.5
    <l<214.5
    , latitude -3
    <b<3
    ). The result of this
    search is shown Fig. 4. The most significant deviation
    from background observed in this galactic plane unbi-
    ased search is seen at l=75.875
    , b=2.675
    in galactic
    coordinates. The pre-trial p-value at this location is
    0.0037. For 95% of the randomized datasets (reproduc-
    ing a pure background hypothesis) an equal or lower
    probability is found and thus the observed excess is not
    significant.
    IV. ICECUBE-40/AMANDA: EXPECTATIONS
    A. Data sample
    For the dataset acquired between April 17, 2008 and
    February 2nd 2009 with the IceCube 40 strings configu-
    ration, the total livetime of the IceCube was 268.7 days,
    and the AMANDA sub-detector performed much better
    than for the 2007/8 season with a total livetime of 240
    days on the same period, corresponding to almost 90%
    of the IceCube livetime. As a consequence, even with
    the doubling of the size of IceCube, the relative number
    of combined IceCube-40/AMANDA events compared to
    the IceCube-40 only events remain comparable to the
    ratios obtained with the IceCube-22/AMANDA dataset.
    The data is still under processing for the selection of
    neutrino candidates and final exploitation will be made
    in the near future. Beyond replicating the galactic plane
    scan and the search for the same list of a priori selected
    sources with these new data, we will search for multiple
    unresolved sources in the Cygnus region applying a new
    analysis strategy.
    B. Extended sources: Multi-Point Source analysis
    A particular interest is given to active regions of
    the galactic plane, where several accelerators might
    contribute to a possible neutrino signal. The Cygnus
    region is a very active star-forming region located at

    4
    C. ROUCELLE et al. ICECUBE-AMANDA COMBINED ANALYSIS
    dfinal
    Entries 7898669
    Mean
    48.26
    RMS
    30.69
    angular distance (degrees)
    0
    20
    40
    60
    80
    100
    2
    2.5
    3
    3.5
    4
    4.5
    5
    dfinal
    Entries 7898669
    Mean
    48.26
    RMS
    30.69
    parejas final
    Fig. 5. Number of event pairs (distant of less than 2
    ) for the signal
    case divided by the average histogram of random cases with the MPS
    method for a simulated case presenting 3 sources (yielding each 8
    events in the detector) randomly distributed in a region of 11
    x7
    .
    galactic longitude 65
    < l < 85
    . Recently, the Mi-
    lagro collaboration measured both a diffuse TeV γ-ray
    emission and a bright, extended TeV source [5]. These
    observations suggest the presence of cosmic rays sources
    which accelerate hadrons that subsequently interact with
    the local, dense interstellar medium to produce γ-rays
    and possibly neutrinos through pion decay. Estimates of
    the neutrino emission from the zone of diffuse γ-ray
    emission are reported in [6].
    The current point source search method is optimized
    for resolveable sources. However, to study extended re-
    gions like the Cygnus region, this method is not optimal.
    A better analysis for these cases takes advantage of the
    possibility of clustering of neutrino events in the totality
    of the region to improve the detection probability. In
    this multi-point source (MPS) analysis, we construct a
    two-point correlation function in which each neutrino
    candidate that pointed inside the region of study is paired
    with all other neutrino candidates. A test statistic is then
    obtained from the number of “close” pairs for which
    the angular separation is at most 2 degrees, the bin
    size for achieving the best signal to noise ratio (for
    IceCube-22/AMANDA data). An excess in the number
    of these close pairs would indicate an emission from
    astrophysical sources in the chosen region. This method
    is sensitive not only to clustered signal that would come
    from a single source, but also would take advantage of
    the presence of a diffuse signal.
    To illustrate the potential of this method, we give
    an example of its performance for the IceCube-
    22/AMANDA configuration. Using the point-spread
    function obtained from the data (median value: 1.5
    ), we
    inserted simulated neutrino events from three possible
    sources in the IceCube-22/AMANDA dataset. Each sim-
    ulated source yielded eight events in the detector and was
    positioned randomly within a region of 11
    x7
    centered
    around Cygnus. Fig. 5 shows the histogram of event
    pairs for the signal case divided by the average histogram
    of random cases. The first bin thus corresponds to the ex-
    cess of “close pairs”. In order to evaluate the significance
    associated with this excess, the number of close pairs in
    10
    7
    scrambled sky maps is used. The excess obtained in
    this example has a p-value of 3×10
    −7
    , corresponding to
    a 5σ detection. For the same configuration, the standard
    point source analysis [12] is less sensitive as it would
    require 11 events from each of the sources to reach a
    detection at the 5σ level (instead of just 8). This analysis
    will be applied to the unblinded data for IceCube-
    22/AMANDA and IceCube-40/AMANDA in the near
    future. For IceCube-22/AMANDA, we will use a region
    surrounding the most active sources observed by Milagro
    on Cygnus to define our primaries (72
    < l < 83
    ; -3
    <
    b < 4
    ).
    V. CONCLUSION AND OUTLOOK
    Numerous galactic sources observed with γ-rays
    present a soft spectrum and possibly a cut off at an
    energy E <100 TeV. Under the hypothesis that accel-
    eration of hadrons explains the γ-ray emission, the as-
    sociated neutrino spectrum should exhibit a similar cut-
    off. The merging of the AMANDA and IceCube detector
    offers an enhancement in sensitivity for the search for
    these sources. The results of the IceCube-22/AMANDA
    configuration show no significant excess either for a
    systematic galactic plane scan on the parts visible for
    IceCube or for a list of a priori selected sources. The
    data acquired with the IceCube-40/AMANDA configura-
    tion are under study and an additional analysis allowing
    the investigation of the extended Cygnus region will be
    added. The AMANDA detector, which was shut down
    on May 15, 2009 as part of the startup of the physics run
    for the IceCube 59-string configuration detector, paved
    the road for the development of a nested, higher gran-
    ularity detector array within IceCube. A new detector
    array of this type, called “IceCube DeepCore”, is under
    construction [13]. It will consist of at least six strings
    instrumenting the deep ice (below 2100m) deployed in
    the center of IceCube and will be completed during the
    2009-2010 deployment season.
    REFERENCES
    [1] F. A. Aharonian et al., 2006a, ApJ, 636, 777
    [2] W. Wagner, [AMANDA Coll.], ICRC 2003
    [3] A. Gross, C. Ha, C. Rott, M. Tluczykont, E. Resconi,
    T. DeYoung, G. Wikstrom,¨ [IceCube Coll.], ICRC 2007,
    arXiv:0711.0353
    [4] S. R. Kelner et al. 2006, PhRvD, 74, 034018
    [5] A. A. Abdo et al, ApJ, 688, 1078, arXiv:0805.0417
    [6] S. Gabici, A. M. Taylor, R. J. White, S. Casanova, F. A.
    Aharonian, Astropart.Phys. 29 (2008) 180. arXiv:0806.2459
    [7] J.Braun et al., Astropart. Phys. 29 (2008) 299.
    [8] F. Aharonian et al. [H.E.S.S. coll], A&A 457 (2006)899.
    [9] F. Aharonian et al. [H.E.S.S. coll], A&A, 370 (2001)112
    [10] J. Albert et al., A&A, 474, (2007)937
    [11] J. Albert et al., 2006, Sci, 312, 1771
    [12] R. Abbasi et al. [IceCube coll], under pub., arXiv:0905.2253
    [13] C. Wiebusch [IceCube coll.] these proceedings

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁOD´ Z´ 2009
    1
    AMANDA 7-Year Multipole Analysis
    Anne Schukraft
    , Jan-Patrick Hulߨ
    for the IceCube Collaboration
    III. Physikalisches Institut, RWTH Aachen University, Germany
    see the special section of these proceedings
    Abstract. The multipole analysis investigates the
    arrival directions of registered neutrino events in
    AMANDA-II by a spherical harmonics expansion.
    The expansion of the expected atmospheric neutrino
    distribution returns a characteristic set of expansion
    coefficients. This characteristic spectrum of expan-
    sion coefficients can be compared with the expansion
    coefficients of the experimental data. As atmospheric
    neutrinos are the dominant background of the search
    for extraterrestrial neutrinos, the agreement of ex-
    perimental data and the atmospheric prediction can
    give evidence for physical neutrino sources or sys-
    tematic uncertainties of the detector. Astrophysical
    neutrino signals were simulated and it was shown
    that they influence the expansion coefficients in a
    characteristic way. Those simulations are used to
    analyze deviations between experimental data and
    Monte Carlo simulations with regard to potential
    physical reasons. The analysis method was applied on
    the AMANDA-II neutrino sample measured between
    2000 and 2006 and results are presented.
    Keywords: Neutrino astrophysics, Anisotropy,
    AMANDA-II
    I. INTRODUCTION
    The AMANDA-II neutrino detector located at South
    Pole was constructed to search for astrophysical neutri-
    nos. These neutrinos could originate from many different
    Galactic and extragalactic candidate source types such
    as Active Galactic Nuclei (AGN), supernova remnants
    and microquasars. The detection of neutrinos is based on
    the observation of Cherenkov light emitted by secondary
    muons produced in charged current neutrino interactions.
    This light is observed by photomultipliers deployed in
    the Antarctic ice. Their signals are used to reconstruct
    the direction and the energy of the primary neutrino.
    AMANDA-II took data between 2000 and 2006. The
    background of atmospheric muons is reduced by select-
    ing only upward-going tracks in the detector, as only
    neutrinos are able to enter the detector from below. This
    restricts the field of view to the northern hemisphere.
    The data is filtered and processed to reject
    misreconstructed downward-going muon tracks [1]. The
    final data sample contains 6144 neutrino induced events
    between a declination of 0
    and +90
    with a purity of
    > 95% away from the horizon.
    II. ANALYSIS PRINCIPLE
    The idea of this analysis is to search for deviations of
    the measured AMANDA-II neutrino sky map from the
    expected event distribution for atmospheric neutrinos,
    which constitute the main part of the data sample [2].
    A method to study such anisotropies is a multipole
    analysis, which was also used to quantify the Cosmic
    Microwave Background fluctuations. The analysis is
    based on the decomposition of an event distribution
    f(θ, φ) =
    P
    N
    events
    i=1
    δ(cos θ
    i
    − cos θ)δ(φ
    i
    − φ) into
    spherical harmonics Y
    m
    l
    (θ, φ), where θ and φ are the
    zenith and azimuth of the spherical analysis coordinate
    system. The expansion coefficients are
    a
    m
    l
    =
    Z
    0
    Z
    1
    −1
    d cos θ f(θ, φ)Y
    m∗
    l
    (θ, φ).
    (1)
    They provide information about the angular structure of
    the event distribution f(θ, φ). The index l corresponds
    to the scale of the angular structure δ ≈
    180
    l
    while
    m gives the orientation on the sphere. The expansion
    coefficients with m = 0 depend only on the structure in
    the zenith direction of the analysis coordinate system.
    Averaging over the orientation dependent a
    m
    l
    yields the
    multipole moments
    C
    l
    =
    1
    2l +1
    X
    +l
    m=−l
    |a
    m
    l
    |
    2
    .
    (2)
    They form an angular power spectrum characteristic for
    different input neutrino event distributions.
    The initial point of this analysis is the angular power
    spectrum of only atmospheric neutrino events. There-
    fore, neutrino sky maps containing 6144 atmospheric
    neutrino events according to the Bartol atmospheric
    neutrino flux model [3] are simulated and numerically
    decomposed with the software package GLESP [4].
    Statistical fluctuations are considered by averaging over
    1000 random sky maps, resulting in a mean ?C
    l
    ? and a
    statistical spread σ
    C
    l
    of each multipole moment.
    The same procedure is applied to simulated sky maps
    containing atmospheric and different amounts of signal
    neutrinos with a total event number of likewise 6144
    events. The influence of the signal neutrinos on the
    angular power spectrum is studied in terms of the pulls
    d
    l
    =
    ?C
    l
    ? − ?C
    l,atms
    ?
    σ
    C
    l,atms
    .
    (3)

    2
    SCHUKRAFT et al. AMANDA MULTIPOLE ANALYSIS
    l
    0
    10
    20
    30
    40
    50
    60
    70
    80
    l,atms
    C
    σ
    >)/
    l,atms
    > - <C
    l
    (<C
    -6
    -4
    -2
    0
    2
    4
    N
    Sources
    (#neutrinos)
    0 (0)
    200 (519)
    400 (1038)
    600 (1557)
    (a)
    l
    0
    10
    20
    30
    40
    50
    60
    70
    80
    0
    l,atms
    a
    σ
    >)/
    0
    l,atms
    > - <a
    0
    l
    (<a
    -6
    -4
    -2
    0
    2
    4
    6
    fraction (#neutrinos)
    0 (0)
    0.02 (122)
    0.04 (245)
    0.06 (368)
    0.08 (491)
    (b)
    Fig. 1. (a): Pull plot for the multipole moments C
    l
    of the isotropic point source model. Sources are simulated with a mean source strength
    µ = 5 and an E
    −2
    ν
    energy spectrum. The number of sources N
    sources
    on the full sphere is varied. The corresponding number of signal
    neutrinos on the northern hemisphere is given in brackets. The errors bars are hidden by the marker symbols. (b): Pull plot for the expansion
    coefficients a
    0
    l
    of the cosmic ray interaction model with the Galactic plane in Galactic coordinates. The fraction of neutrinos in the sky map
    originating from the Galactic plane is varied. The corresponding number of signal neutrinos is given in brackets. The errors bars are hidden by
    the marker symbols.
    III. SIGNAL SIMULATION
    The different models for candidate neutrino sources
    investigated in this analysis are:
    1) Isotropically distributed point sources
    2) A diffuse flux from FR-II galaxies and blazars [5]
    3) AGN registered in the Veron-Cetty´
    and Veron´
    (VCV) catalog [6]
    4) Galactic point sources such as supernova remnants
    or microquasars
    5) Cosmic rays interacting in the Galactic plane.
    All simulated pointlike neutrino sources are character-
    ized by a Poissonian distributed source strength with
    mean µ and an energy spectrum E
    −γ
    ν
    . The relative angu-
    lar detector acceptance depends on the neutrino energy
    and therefore on the spectral index of the simulated neu-
    trino source. Signal neutrinos are simulated according to
    this acceptance considering systematic fluctuations. The
    total number of signal neutrinos in a sky map of the
    northern hemisphere with N
    sources
    simulated sources on
    the full-sky is therefore given by ∼ 0.5 · µ · N
    sources
    .
    Additionally the angular resolution is taken into account.
    It dominates over the uncertainty between the neutrino
    and muon direction.
    The spectral index of pointlike sources is varied
    between 1.5 ≤ γ ≤ 2.3. As the spectral index of
    atmospheric neutrinos is close to 3.7, signal and back-
    round neutrinos underlie different angular detector ac-
    ceptances. Thus, additionally to the clustering of events
    around the source directions also the shape of the total
    angular event distribution is used to identify a signature
    of signal neutrinos in the angular power spectrum [7].
    Neutrinos from our Galaxy disturb the atmospheric
    event distribution by their bunching within the Galactic
    plane modeled by a Gaussian band along the Galactic
    equator. Neutrinos produced in cosmic ray interactions
    with the interstellar medium of our Galaxy are assumed
    to follow the E
    −2.7
    primary energy spectrum.
    A further topic (model 6) that can be studied with a
    multipole analysis are neutrino oscillations. The survival
    probability of atmospheric muon neutrinos depends
    on the neutrino energy and the traveling length of the
    neutrino as well as the mixing angle θ
    23
    and the squared
    mass difference ∆m
    2
    23
    . The traveling length can be
    expressed by the Earth’s radius and the zenith angle of
    the neutrino direction [7]. Thus, the neutrino oscillations
    disturb the angular event distribution of atmospheric
    neutrinos. With the assumption of sin
    2
    (2θ
    23
    ) ≈ 1 the
    squared mass difference remains for investigation. Due
    to the relatively high energy threshold of 50 GeV the
    effect is small.
    IV. EVALUATION OF THE POWER SPECTRA
    The deviations from a pure atmospheric angular power
    spectrum caused by signal neutrinos are studied by the
    pulls. These pulls are exemplarily shown in Fig. 1a for
    the model of isotropic point sources. The behaviour
    of the pulls is characteristic for each signal model.
    Different multipole moments carry different sensitivity
    to the neutrino signal. The absolute value of the pull
    increases linearly with the amount of signal neutrinos in
    the sky maps. Each pull has a predefined sign.
    The deviation of a particular sky map with multipole
    moments C
    l
    from the pure atmospheric expectation
    ?C
    l,atms
    ? is quantified by a significance indicator D
    2
    defined as
    D
    2
    =
    1
    l
    max
    l
    X
    max
    l=1
    sgn
    l
    · w
    l
    ·
    μ
    C
    l
    − ?C
    l,atms
    ?
    σ
    C
    l,atms
    2
    , (4)
    where l
    max
    determines the considered multipole mo-
    ments. The term in brackets is the pull between the

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁOD´ Z´ 2009
    3
    particular sky map and the mean of the atmospheric
    expectation as defined in Eq. 3. The factors
    w
    l
    =
    ?C
    l
    ? − ?C
    l,atms
    ?
    q
    σ
    2
    C
    l
    + σ
    2
    C
    l,atms
    (5)
    are defined to weight the pulls according to their ex-
    pected sensitivity to the signal. For each neutrino signal
    model one dedicated set of weights w
    l
    is determined.
    Due to the linear increase of the pulls with the signal
    strength the strength chosen to calculate w
    l
    is arbitrary.
    The weight factors w
    l
    carry the expected sign of the
    pulls. sgn
    l
    is the sign of the measured pull. Thus, the
    D
    2
    calculated for the particular sky map is increased if
    the observed deviation has the direction expected for the
    signal model and reduced otherwise.
    Due to the weighting of the pulls, the sensitivity
    becomes stable for high l
    max
    . A choice of l
    max
    = 100
    is sufficient to provide best sensitivity to all investigated
    signal models.
    The D
    2
    of a sky map is interpreted physically by the
    use of confidence belts. Therefore, 1000 sky maps for
    every signal strength within a certain range are simu-
    lated and the D
    2
    -value for each sky map is calculated
    separately to obtain the D
    2
    distributions. The calculation
    of the average upper limit at 90% confidence level
    assuming zero-signal is used to estimate the sensitivity
    of the analysis to different astrophysical models apriori.
    As the multipole analysis is applied to a wide range
    of astrophysical topics, the trial factor of the analysis
    becomes important. The trial factor raises with each new
    set of weights used to evaluate the experimental data.
    For this reason, models with almost similar weights are
    combined to a common set of weights and only six sets
    are remaining.
    If the signal signatures show up only in the zenith
    direction of the analysis coordinate system the expansion
    coefficients a
    0
    l
    are more sensitive than the multipole
    moments C
    l
    . The reason is, that the expansion coeffi-
    cients with m = 0 are independent from the azimuth
    φ and contain the pure information about the zenith
    direction θ. A signal only depending on θ causes only
    statistical fluctuations but no physical information in
    the other expansion coefficients. Therefore, the signal
    has only power in the a
    0
    l
    . The analysis method stays
    exactly the same in these cases, except that all C
    l
    are
    replaced by the a
    0
    l
    . This is related to the models of
    neutrinos from the Galactic plane and from sources of
    the VCV catalog, which show north-south-symmetries
    of the neutrino signals in Galactic and supergalactic
    coordinates, respectively. Unlike the multipole moments
    C
    l
    , the a
    0
    l
    do not average over different orientations.
    Therefore, the analysis of the a
    0
    l
    strongly depends on the
    used coordinate system. An example for pulls of a
    0
    l
    for
    the model of a diffuse neutrino flux from the Galactic
    plane is shown in fig. 1b. The characteristic periodic
    behavior of the pulls is explained by the symmetry
    properties of the spherical harmonics.
    l
    0
    2
    4
    6
    8
    10
    12
    14
    16
    18
    20
    l,atms
    C
    σ
    >)/
    l,atms
    - <C
    l
    (C
    -6
    -4
    -2
    0
    2
    4
    6
    Pulls experimental data
    Pulls for 600 mini sources (μ = 5)
    Pulls for 100000 nano sources (μ = 0.02)
    Fig. 2. Pull plot for the experimental multipole moments C
    l
    . Expected
    pulls for typical model parameters of isotropic point sources are shown
    for comparison. The error bars symbolize the statistical fluctuation
    expected for an atmospheric neutrino sky map.
    l
    0
    2
    4
    6
    8
    10
    12
    14
    16
    18
    20
    0
    l,atms
    a
    σ
    >)/
    0
    l,atms
    - <a
    0
    l
    (a
    -5
    -4
    -3
    -2
    -1
    0
    1
    2
    3
    4
    5
    Pulls experimental data
    Pulls
    ν
    from galactic plane (fraction 2%)
    Fig. 3. Pull plot for the experimental expansion coefficients a
    0
    l
    in
    Galactic coordinates. Expected pulls for typical parameters of cosmic
    ray interactions with the Galactic plane are shown for comparison.
    The error bars symbolize the statistical fluctuation expected for an
    atmospheric neutrino sky map.
    V. EXPERIMENTAL RESULTS
    The experimental data is analyzed in two steps. First,
    the experimental data is tested for its compatibility with
    the pure atmospheric neutrino hypothesis. Secondly, the
    experimental pulls are compared with the expectations
    for the different investigated neutrino models.
    The pulls of the experimental data are shown for the
    multipole moments C
    l
    in Fig. 2 and for the expansion
    coefficients a
    0
    l
    in Galactic coordinates in Fig. 3. To
    compare the measured data with the expected event dis-
    tribution, a D
    2
    is calculated for the multipole moments
    C
    l
    and the expansion coefficients a
    0
    l
    for transformations
    into equatorial, Galactic and supergalactic coordinates
    separately. As no signal model is tested sgn
    l
    = w
    l
    = 1
    is assumed. A comparison with the corresponding D
    2
    distributions results in the p-values giving the proba-
    bility to obtain a D
    2
    which is at least as extreme as
    the measured one assuming that the pure atmospheric
    neutrino hypothesis is true (Table I).

    4
    SCHUKRAFT et al. AMANDA MULTIPOLE ANALYSIS
    The statistical consistency of C
    l
    and a
    0
    l
    in equa-
    torial coordinates with the atmospheric expectation is
    marginal. Rotating to inclined coordinate systems, e.g.
    Galactic and supergalactic, the consistency improves.
    The deviation from the pure atmospheric expectation is
    not compatible with any of the signal models (see Fig.
    2, 3 for examples). The discrepancy may be attributed
    to uncertainties in the theoretical description of the
    atmospheric neutrino distribution, or to a contribution
    of unsimulated background of down-going muons mis-
    reconstructed as up-going, or to the modeling of prop-
    erties of the AMANDA detector.
    TABLE I
    P-VALUES FOR THE COMPATIBILITY OF EXPERIMENTAL DATA AND
    PURE ATMOSPHERIC NEUTRINO HYPOTHESIS.
    Observable
    p-value
    C
    l
    0.02
    a
    0
    l
    , Equatorial
    0.02
    a
    0
    l
    , Galactic
    0.15
    a
    0
    l
    , Supergalactic
    0.70
    The signal models are tested by calculating the D
    2
    -
    values of the experimental data using the corresponding
    sign and weight factors. As the observed deviations do
    not fit any of the investigated signal models the physical
    model parameters are constrained. Due to the observed
    systematic effects affecting mainly the multipole mo-
    ments C
    l
    and the equatorial expansion coefficients a
    0
    l
    no limits on the models analyzed in the corresponding
    coordinate systems (models 1, 2 and 6) are derived. The
    other models are less affected. The limits given below
    do not include these systematic effects.
    A limit on the source strength assuming the VCV
    source distribution (model 3) is calculated for those
    sources closer than 100 Mpc to the Earth. In this model
    all sources are expected to have the same strength and
    energy spectrum. For a typical spectral index of γ = 2
    the average source flux is limited by the experimental
    data to a differential source flux of dΦ/dE · E
    2
    ≤ 1.6 ·
    10
    −10
    GeV cm
    −2
    s
    −1
    sr
    −1
    in the energy range between
    1.6 TeV and 1.7 PeV.
    For the random Galactic sources (model 4), the
    number of sources is constrained assuming the same
    source strength and energy spectrum for all sources
    as well. For a spectral index of γ = 2, the limit
    on the number of sources is set by AMANDA to
    N
    sources
    ≤ 39 assuming a source strength of dΦ/dE ·
    E
    2
    ≤ 10
    −8
    GeV cm
    −2
    s
    −1
    sr
    −1
    or N
    sources
    . 4300 for
    sources with dΦ/dE · E
    2
    ≤ 10
    −10
    GeV cm
    −2
    s
    −1
    sr
    −1
    .
    For source fluxes in between the limit can be ap-
    proximated by assuming linearity between N
    sources
    and
    log
    ¡
    dΦ/dE · E
    2
    ¢
    .
    The differential flux limit obtained from the
    experimental data on the diffuse neutrino flux from
    cosmic ray interactions in the Galactic plane (model 5)
    is dΦ/dE · E
    2.7
    ≤ 3.2 · 10
    −4
    GeV
    1.7
    cm
    −2
    s
    −1
    sr
    −1
    .
    This flux limit is shown in Fig. 4 together with the
    results of two other AMANDA analyses and two
    Energy [GeV]
    2
    10
    3
    10
    4
    10
    5
    10
    6
    10
    7
    10
    8
    10
    9
    10
    ]
    -1
    sr
    -1
    s
    -2
    cm
    1.7
    [GeV
    2.7
    /dE * E
    Φ
    d
    -7
    10
    -6
    10
    -5
    10
    -4
    10
    -3
    10
    -2
    10
    -1
    10
    Limit Multipole 7yr
    Limit Multipole 4yr
    Limit J. Kelley 4yr AMANDA-II
    Prediction Gaisser, Halzen & Stanev
    Prediction Ingelman & Thunman
    Fig. 4. Limit of the 7yr multipole analysis on the diffuse neutrino flux
    from cosmic ray interactions in the Galactic plane in dependence of
    the valid energy range. The limit is compared with two other analyses
    [2], [8] and two theoretical predictions [9], [10].
    theoretical flux predictions. The seven year multipole
    analysis provides currently the best limit. However, it
    is still not in reach of the theoretical predictions.
    VI. CONCLUSION
    It is shown that the multipole analysis is sensitive to
    a wide range of physical topics. Its area of application
    is in particular the field of many weak sources in
    transition to diffuse fluxes. With the statistics of seven
    years of AMANDA data and improvements of the
    analysis technique the method is now restricted by
    systematic uncertainties in the atmospheric neutrino
    zenith distribution of the order of a few percent.
    Transforming to coordinate systems less affected by
    the equatorial zenith angle such as the Galactic and
    supergalactic system physical conclusions are still
    possible. A compatibility of the measurement with
    the background expectation of atmospheric neutrinos
    is observed. Current efforts to better understand the
    observed systematics would allow an application of the
    multipole analysis on future high statistic IceCube data.
    REFERENCES
    [1] R. Abbasi et al., Phys. Rev. D 79, 062001 (2009).
    [2] J.-P. Hulß,¨
    Ch. Wiebusch for the IceCube Collaboration, 30
    th
    International Cosmic Ray Conference (ICRC 2007), Merida,
    arXiv:0711.0353.
    [3] G. Barr et al., Phys. Rev. D 70, 023006 (2004).
    [4] Doroshkevich et al., International Journal of Modern Physics D,
    Vol 14, No. 2 (2005), http://www.glesp.nbi.dk/.
    [5] J. Becker, P. Biermann, W. Rhode, Astroparticle Physics, Vol
    23, No. 4 (2005).
    [6] M.-P. Veron-Cetty´
    , P. Veron,´
    A&A, 455, 733 (2006).
    [7] A. Schukraft, Multipole analysis of the AMANDA-II neutrino
    skymap, diploma thesis, RWTH Aachen (2009).
    [8] J. Kelley for the IceCube Collaboration, 29
    th
    International
    Cosmic Ray Conference (ICRC 2005), Pune, arXiv:0711.0353.
    [9] T. Gaisser, F. Halzen, T. Stanev, Phys. Rept., 258:173-236 (1995).
    [10] G. Ingelman, M. Thunman, arxiv:hep-ph/9604286 (1996).
    This contribution is supported by the German Academic Exchange
    Service (DAAD).

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    1
    Measurement of the atmospheric neutrino energy spectrum with
    IceCube
    Dmitry Chirkin
    for the IceCube collaboration
    University of Wisconsin, Madison, U.S.A.
    See the special section of these proceedings.
    Abstract
    . The IceCube detector, as configured dur-
    ing its operation in 2007, consisted of 22 deployed
    cables, each equipped with 60 optical sensors, has
    been the biggest neutrino detector operating during
    the year 2007, superseded only by its later config-
    urations. A high quality sample of more than 8500
    atmospheric neutrinos was extracted from this single
    year of operation and used for the measurement
    of the atmospheric muon neutrino energy spectrum
    from 100 GeV to 500 TeV discussed here. Several
    statistical techniques were used in an attempt to
    search for deviation of the neutrino flux from that
    of conventional atmospheric neutrino models.
    Keywords
    : atmospheric neutrinos, charm search,
    IceCube
    I. INTRODUCTION
    Most of the events recorded by the IceCube detector
    constitute the background of atmospheric muons that
    are produced in air showers. Once this background is
    removed the majority of events that remain are atmo-
    spheric neutrino events, i.e., (mostly) muons created
    by atmospheric neutrinos. Although much smaller, this
    also constitutes background for the majority of research
    topics in IceCube (e.g., extra-terrestrial neutrino flux
    searches), except one: the atmospheric neutrino study.
    As part of this study we verify that the atmospheric
    neutrinos observed by IceCube are consistent with pre-
    vious measurements at lower energies, and agree with
    the theoretical extrapolations at higher energies. Since
    much uncertainty remains in the description of the
    higher energy atmospheric neutrinos, this study could
    provide interesting constraints on (not yet observed)
    charm contribution to the atmospheric neutrino produc-
    tion. Since such charm contribution may affect the flux
    of atmospheric neutrinos in a way similar to extra-
    terrestrial diffuse contributions, we attempt to look for
    both simultaneously in a single likelihood approach.
    II. EVENT SELECTION
    For this analysis the new machine learning method
    (
    SBM
    ) described in [1] was employed. The quality
    parameters used with the event selection method of this
    paper include and build upon those discussed previously
    in [2]. Unfortunately the size limit of this proceeding
    precludes us from discussing all of the event selection
    quality parameters and techniques; instead we describe
    one new technique in detail below.
    78
    21
    46
    50
    72
    74
    29
    30
    40
    56
    59
    67
    73
    38
    47
    65
    39
    49
    66
    48
    57
    58
    x
    [
    m
    ]
    y
    [
    m
    ]
    -300
    -200
    -100
    0
    100
    200
    300
    400
    500
    600
    -100 0 100 200 300 400 500 600 700
    Fig. 1. View of the IceCube 22 string configuration, as used in the
    run of 2007. The size of the circle and color indicate the relative string
    weight, used to compute several quality parameters, such as the size
    of the veto region for contained events, or the total weight, which,
    much like the number of hit strings, gauges the size of an event and
    its importance for the analysis.
    Events in IceCube are normally formed by the DAQ
    by combining all hits satisfying the simple majority trig-
    ger. The simple majority trigger is defined to combine
    all hits, which belong to one or more hit sets of at
    least
    n
    different-channel hits within
    w
    ns of each other.
    Typically
    n = 8
    or more hits are required to be within
    w = 5
    us of each other to satisfy this trigger.
    The simple majority trigger combines hits into events
    only separating them in time. In IceCube a substantial
    fraction of events so formed turns out to consist of
    hits originating from two or more separate particles, or
    bundles of particles, typically unrelated to each other,
    traveling through well separated (in space) parts of the
    detector. In order to split up such events and to keep the
    rate of coincident (now in both time and space) events
    low, hits in the events were recombined via the use of
    the
    topological trigger
    . The definition of this trigger is
    very similar to that of the simple majority trigger given
    above: the
    topological trigger
    combines all
    topologically
    connected
    hits, which belong to one or more hit sets
    of at least
    n
    different-channel hits within
    w
    ns of each
    other. Two hits are called
    topologically connected
    if they
    satisfy all of the following (the numbers in italics show
    the values used in the present analysis):
    both hits originate on the detector strings

    2
    CHIRKIN
    et al.
    ATMOSPHERIC NEUTRINOS WITH ICECUBE
    1-cos(θ)
    entries
    10
    -1
    1
    10
    0
    0.2
    0.4
    0.6
    0.8
    1
    1.2
    preliminary
    Fig. 2. Zenith angle distribution of remaining data events in 275.5
    days of IceCube data (black) comparison with atmospheric neutrino
    prediction from simulation (red). Several double coincident air shower
    muon events remain at this level in simulation (shown in green).
    Vertically up-going tracks are at 0, horizontal tracks are at 1.
    if both hits are on the same string they should not
    be separated by more than
    30
    optical sensors
    the strings of both hits must be within
    500
    meters
    of each other
    the
    δt  δr/c
    must be less than
    1000
    ns.
    At least
    4
    topologically-connected hits within
    4
    us
    are required to form a topological triggered set, which
    is then passed through the simple majority trigger. Just
    like in the simple majority trigger, the hits not directly
    connected to each other can belong to the same event
    if they form topologically-connected sets satisfying the
    multiplicity condition with at least one and the same hit
    belonging to both sets.
    The required distance between the strings (
    500
    me-
    ters) was left intentionally high to allow easy scaling of
    the present analysis to higher-string IceCube detector
    configurations. Still, the rate of unrelated coincident
    events is much reduced via the use of the topological
    trigger. More importantly, the fraction of such events
    after the topological trigger stays at the same low level
    as the detector grows.
    An alternative approach to recognize coincident events
    by reconstructing them with double-muon hypothesis
    was tried in a separate effort. In the present work
    however it is believed that the topological trigger offers
    several crucial advantages:
    the separation of coincident events is performed at
    the hit selection level
    the method is faster as it does not require compli-
    cated dual-muon fits
    not only 2 but also 3 and more coincident events
    can be separated
    all of these are kept for the analysis (in the alter-
    native approach coincident events are thrown out)
    noise hits are cleaned very efficiently
    the rate of unresolved coincident events and coin-
    cident noise hits is kept at the same low level as
    the detector grows.
    The event selection resulted in 8548 events found in
    275.5 days of data of IceCube (see the 22-string config-
    neutrino signal
    corsika showers
    data, SBM step 1
    data, SBM step 2
    log
    10
    (E
    reco
    (COG)
    [
    GeV
    ])
    entries
    10
    -2
    10
    -1
    1
    10
    2 2.5 3 3.5 4 4.5 5 5.5 6 6.5 7
    preliminary
    Fig. 3. Reconstructed muon energy at the closest approach point to
    the center-of-gravity of hits in the event. Data distribution is shown at
    both steps 1 and 2 of the SBM event selection method [1]. After the
    90% purity level is reached in simulation (step 1) it is necessary to
    remove more events from data that do not look like well-reconstructed
    muons; this is achieved by comparing data events to simulated muon
    neutrino events (step 2).
    uration in Figure 1), or 31 events per day at
    >∼
    90%
    estimated (from simulation) purity level (contaminated
    by remaining atmospheric muon background). Compare
    this to expectation from simulation of 29.0 atmospheric
    neutrino events per day (Figure 2).
    III. ATMOSPHERIC NEUTRINO SPECTRUM
    UNFOLDING
    Figure 3 compares the measured muon energy distri-
    bution for conventional atmospheric neutrino simulation
    and data at
    >∼
    90% purity level. The difference between
    data at steps 1 and 2 of the SBM event selection is due to
    the presence of events that were unlike those simulated.
    Such events are removed at step 2 by comparing them
    to the events in the atmospheric neutrino simulation [1].
    At this time the difference between the two data curves
    should be treated as a measure of (at least some of) the
    systematic errors introduced by our simulation.
    The uncertainty in our measurement of muon energy
    is
    0.3 in
    log
    10
    (E
    µ
    )
    in a wide energy range (from 1
    TeV to 100 PeV). A larger smearing, estimated from
    neutrino simulation (based on [3]), is introduced when
    matching the muon energy at the location of the detector
    to the parent neutrino energy.
    We tried a variety of unfolding techniques to obtain
    the distribution of the parent muon neutrinos, including
    the SVD [4] with regularization term that was the
    second derivative of the unfolded statistical weight;
    and iterative Bayesian unfolding [5] with a 5-point
    spline fit smoothing function (with and without the
    smearing kernel smoothing). Since we are looking for
    deviations of the energy spectrum from the power law,
    the SVD with regularization term that is the second
    derivative of the log(flux) was selected as our method of
    choice. Additionally, we chose to include the statistical
    uncertainties of the unfolding matrix according to [6]
    (using the equivalent number of events concept as in
    [7]). The chosen method yielded the most consistent
    description of spectrum deviations that were studied;

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    3
    log
    10
    (Unfolded E
    pri
    [
    TeV
    ])
    Observed event counts
    unfolded event counts of simulated dataset
    0
    500
    1000
    1500
    2000
    2500
    3000
    -1 -0.5 0 0.5 1 1.5 2 2.5 3 3.5 4
    preliminary
    Fig. 4. Unfolded distribution of muon neutrino energies: the original
    distribution modeled according to [11] (red), median and 90% band
    of the unfolded result of 10000 simulated sets, drawn from the same
    simulation (blue dots/lines and black boxes, respectively). A small bias
    introduced by the regularization term shows up as a slight mismatch
    between the original and unfolded median bin values. Also shown is
    the distribution modeled according to [3] (green).
    unfolded spectrum of simulated dataset
    log
    10
    (Unfolded E
    pri
    [
    TeV
    ])
    (E/GeV)
    2
    ⋅Φ
    [
    GeV
    -1
    cm
    -2
    s
    -1
    sr
    -1
    ]
    10
    -9
    10
    -8
    10
    -7
    10
    -6
    10
    -5
    10
    -4
    10
    -1
    1
    10
    10
    2
    10
    3
    preliminary
    Fig. 5. Unfolded muon neutrino spectrum, averaged over zenith angle,
    same color designations as in Figure 4. The green points of [3] form a
    band as they are shown un-averaged, for each zenith angle separately.
    also errors estimated from half-width of the likelihood
    function were reasonable when compared to the spread
    of unfolded results in a large pool of simulated data sets
    (see Figures 4 and 5).
    It is possible to study the effect of small charm
    and
    E
     2
    isotropic diffuse contributions (as the two
    commonly studied deviations from the conventional neu-
    trino flux models). Injecting known amounts of such
    contributions into the simulated event sets one computes
    the 90% confidence belt as in [8], [9], [10] (shown in
    Figure 6 for statistical weight of events in one of the
    bins of the unfolded distribution). The following table
    summarizes the average upper limits for diffuse and
    RQPM (optimistic) charm models (using conventional
    neutrino flux description as in [11]):
    flux
    bin 8
    bin 9
    bin 10
    bin 11
    energies, TeV
    46.4
    100
    215
    1  10
    PeV
    E
     2
    ,
    10
     8
    ·
    5.48
    3.00
    3.00
    4.06
    RQPM (opt)
    ·
    0.74
    0.90
    1.34
    2.44
    number of events in bin 10
    signal strength
    10
    -8
    (E/GeV)
    2
    ⋅Φ
    [
    GeV
    -1
    cm
    -2
    s
    -1
    sr
    -1
    ]
    1
    10
    0
    2
    4
    6
    8 10 12 14 16 18 20
    preliminary
    Fig. 6.
    90% connfidence belt for
    E
     2
    isotropic diffuse flux
    contribution, calculated with 10000 independent simulated sets for bin
    10 (neutrino energies 215 TeV-1 PeV)
    10
    10
    2
    10
    3
    normalization
    spectral index correction
    -0.4
    -0.2
    0
    0.2
    0.4
    0.6
    0.8
    1
    1.2
    1.4
    preliminary
    Fig. 7. Likelihood model testing profile for a simulated spectrum with
    spectral index deviation of
    +0.2
    with respect to the reference model.
    The 90% confidence belt (shown as red contour) is very narrow and
    widens when systematical errors are taken into account.
    IV. LIKELIHOOD MODEL TESTING
    The likelihood model testing approach is well-suited
    to testing the data for specific deviations from the
    conventional flux model. This approach is based on the
    likelihood ordering principle of [8] and is easy employ
    when several deviations are tested for simultaneously
    [12]. This has recently been used in the analysis of the
    AMANDA data [13] and is also used in a similar study
    presented in [14].
    As an example, Figure 7 demonstrates the ability to
    measure the deviation of the conventional flux in overall
    normalization and spectral index (with 8548 neutrino
    events in the absence of systematical errors). Figure 8
    demonstrates the ability to discern simultaneous charm
    and diffuse
    E
     2
    contributions (assuming that the precise

    4
    CHIRKIN
    et al.
    ATMOSPHERIC NEUTRINOS WITH ICECUBE
    10
    20
    30
    40
    50
    60
    diffuse
    10
    -8
    (E/GeV)
    2
    ⋅Φ [
    GeV
    -1
    cm
    -2
    s
    -1
    sr
    -1
    ]
    ratio to RQPM(opt)
    0
    0.5
    1
    1.5
    2
    2.5
    3
    3.5
    4
    0
    2
    4
    6
    8
    10
    12
    14
    16
    preliminary
    Fig. 8. A 90% confidence belt for a simulated mixed contribution
    of 2
    ·
    RQPM (opt) charm expectation
    +6 · 10
     8
    E
     2
    isotropic
    (diffuse) component. This profile includes systematic errors on overall
    normalization and spectral index of the conventional neutrino flux
    (allowing them to vary freely).
    5
    10
    15
    20
    25
    30
    35
    diffuse
    10
    -8
    (E/GeV)
    2
    ⋅Φ [
    GeV
    -1
    cm
    -2
    s
    -1
    sr
    -1
    ]
    ratio to RQPM(opt)
    0
    0.25
    0.5
    0.75
    1
    1.25
    1.5
    1.75
    2
    0
    0.5
    1
    1.5
    2
    2.5
    3
    3.5
    4
    preliminary
    Fig. 9. 90% confidence level upper limit contours shown (in green) for
    11 independent simulated data sets (drawn from the same conventional
    flux parent simulation according to [11]), the “median” upper limit
    shown in red.
    normalization and spectral index of the conventional flux
    are also unknown). We estimate the median upper limits
    set by this method on both charm and diffuse
    E
     2
    components in Figure 9. We used the
    χ
    2
    with 2 degrees
    of freedom approximation to construct the confidence
    belts; the true 90% levels are even tigher than this (by
    factor
    ∼ 1.3  1.6
    ) due to high similarity of effects of
    both components on the eventual event distribution.
    V. MODEL REJECTION FACTOR
    This is a method that optimizes the placement of a
    cut on the energy observable to maximize sensitivity
    to an interesting flux contribution, discussed in [15].
    The model rejection factor (ratio of
    µ
    90
    to number of
    expected signal events for a given flux) computed from
    curves shown in Figure 10 achieves its optimal value
    with a cut of 224 TeV on the reconstructed muon energy.
    The corresponding best average upper limit (sensitivity,
    not including systematics) of
    2.14 · 10
     8
    is achieved.
    log
    10
    (E
    μ,reco
    [
    TeV
    ])
    cumulative counts above given E
    μ
    ,reco
    background (bartol)
    E
    -2
    10
    -8
    signal
    90
    >(b)
    0
    2
    4
    6
    8
    10
    12
    14
    16
    18
    20
    2
    2.2
    2.4
    2.6
    2.8
    3
    3.2
    3.4
    preliminary
    Fig. 10.
    Cumulative number of
    E
     2
    diffuse signal events shown
    in red, number of atmospheric neutrino events shown in blue, the
    corresponding average upper limit
    µ
    90
    is shown in green.
    VI. CONCLUSIONS
    We present a selection of 8548 muon neutrino events
    (with
    ∼<
    10% estimated contamintation from the mis-
    reconstructed air shower muon events) in 275.5 days of
    IceCube-22 data. An unfolding technique is selected and
    used to compute the average upper limit on diffuse and
    charm contributions. We found that the likelihood model
    testing and the model rejection factor methods both
    achieve (not surprisingly) somewhat better sensitivities.
    Since the study of systematic errors is (at the time
    of writing of this report) not yet completed, the average
    upper limits presented here do not contain systematic
    error effects, and the actual upper limits (or the unfolded
    spectrum) computed from the data are not yet shown.
    REFERENCES
    [1] D. Chirkin, et al.,
    A new method for identifying neutrino events
    in IceCube data
    , These proceedings
    [2] D. Chirkin, et al.,
    Effect of the improved data acquisition system
    of IceCube on its neutrino-detection capabilities
    , 30th ICRC,
    Merida, Mexico (arXiv:0711.0353)
    [3] M. Honda, et al., Physical Review D, V70, 043008 (2004)
    [4] V. Blobel,
    An unfolding method for high energy physics ex-
    periments
    , Advanced statistical techniques in particle physics
    conference, Durham, 2002
    [5] G.D’Agostini,
    A multidimensional unfolding method based on
    Bayes’ theorem
    , DESY 94-099 (1994)
    [6] R. Barlow and Ch. Beeston,
    Fitting using finite Monte Carlo
    samples
    , Computer Physics Communications 77 (1993) 219
    [7] G. Zech,
    Comparing statistical data to Monte Carlo simulation
    parameter fitting and unfolding
    , DESY 95-113 (1995)
    [8] G. Feldman and R. Cousins, Physical Review D, V57, 3873
    (1998)
    [9] K. Mu¨nich, J. Lu¨nemann
    Measurement of the atmospheric lepton
    energy spectra with AMANDA-II
    , 30th ICRC, Merida, Mexico
    (arXiv:0711.0353)
    [10] S. Gozzini,
    Search for Prompt Neutrinos with AMANDA-II
    , Ph.
    D. thesis, Johannes Gutenberg Universita¨t Mainz, 2008
    [11] G. Barr, et al., Physical Review D, V70, 023006 (2004)
    [12] G. Hill, et al.,
    Likelihood deconvolution of diffuse prompt and
    extra-terrestrial neutrino fluxes in the AMANDA-II detector
    , 30th
    ICRC, Merida, Mexico (arXiv:0711.0353)
    [13] R. Abbasi et al. (IceCube collaboration),
    Determination
    of the Atmospheric Neutrino Flux and Searches for New
    Physics with AMANDA-II
    , Accepted by Phys. Rev. D., 2009,
    (arXiv:0902.0675)
    [14] W. Huelsnintz and J. Kelley (IceCube collaboration),
    Search
    for quantum gravity with IceCube and high energy atmospheric
    neutrinos
    , these proceedings
    [15] G. Hill and K. Rawlins, Astroparticle physics 19 (2003), 393

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁOD´ Z´ 2009
    1
    Atmospheric Neutrino Oscillation Measurements with IceCube
    Carsten Rott
    (for the IceCube Collaboration
    )
    Center for Cosmology and AstroParticle Physics, Ohio State University, Columbus, OH 43210, USA
    See special section of these proceedings
    Abstract. IceCube’s lowest energy threshold for the
    detection of track like events (muon neutrinos) is
    realized in vertical events, due to IceCube’s geome-
    try. For this specific class of events, IceCube may
    be able to observe muon neutrinos with energies
    below 100 GeV at a statistically significant rate.
    For these vertically up-going atmospheric neutrinos,
    which travel a baseline length of the diameter of
    the Earth, oscillation effects are expected to become
    significant. We discuss the prospects of observing
    atmospheric neutrino oscillations and sensitivity to
    oscillation parameters based on a muon neutrino
    disappearance measurement performed on IceCube
    data with vertically up-going track-like events. We
    further discuss future prospects of this measurement
    and the impact of an IceCube string trigger con-
    figuration that has been active since 2008 and was
    specifically designed for the detection of these events.
    Keywords: Neutrino Oscillations IceCube
    I. INTRODUCTION
    The IceCube Neutrino Telescope is currently under
    construction at the South Pole and is about three quarters
    completed [1]. Upon completion in 2011, it will instru-
    ment a volume of approximately one cubic kilometer
    utilizing 86 strings, each of which will contain 60 Digital
    Optical Modules (DOMs). In total, 80 of these strings
    will be arranged in a hexagonal pattern with an inter-
    string spacing of about 125 m, and 17 m vertical sepa-
    ration between DOMs at a depth between 1450 m and
    2450 m. Complementing this 80 string baseline design
    will be a deep and dense sub-array named DeepCore [2].
    For this sub-array, six additional strings will be deployed
    in the center, in between the regular strings, resulting
    in an interstring-spacing of 72 m. DeepCore will be
    densely instrumented in the deep ice below 2100 m, with
    a vertical sensor spacing of 7 m. This array is specifically
    designed for the detection and reconstruction of sub-TeV
    neutrinos. Further, the deep ice provides better optical
    properties and the usage of high quantum efficiency
    photomultiplier tubes will enable us to study neutrinos
    in the energy range of a few tens of GeV. This makes
    DeepCore an ideal detector for the study of atmospheric
    neutrino oscillations [2].
    In this paper we present an atmospheric neutrino
    oscillation analysis in progress on data collected with the
    IceCube 22-string detector during 2007 and 2008. This
    is an update on a previous report [4], with a larger, more
    complete background simulation and hence re-optimized
    selection criteria. An alternative background estimation
    using the data itself is also discussed.
    The goal of this analysis is to measure muon neutrino
    µ
    ) disappearance as a function of energy for a constant
    baseline length of the diameter of the Earth by study-
    ing vertically up-going ν
    µ
    . Disappearance effects are
    expected to become sizable at neutrino energies below
    100 GeV in these vertical events. This energy range is
    normally hard to access with IceCube. However, due
    to IceCube’s vertical geometry, low noise rate, and low
    trigger threshold the observation of neutrino oscillations
    through ν
    µ
    disappearance seems feasible. Atmospheric
    neutrino oscillations have, as of today, not been observed
    with AMANDA or IceCube.
    Based on preliminary selection criteria, we show that
    IceCube has the potential to detect low-energy vertical
    up-going ν
    µ
    events and we estimate the sensitivity to
    oscillation parameters.
    II. ATMOSPHERIC NEUTRINO OSCILLATIONS
    Collisions of primary cosmic rays with nuclei in the
    upper atmosphere produce a steady stream of muon
    neutrinos from decays of secondaries (π
    ±
    , K
    ±
    ). These
    atmospheric neutrinos follow a steeply falling energy
    spectrum of index γ ? 3.7.
    In IceCube these muon neutrinos can be identified
    through the observation of Cherenkov light from muons
    produced in charged-current interactions of the neutrinos
    with the Antarctic ice or the bedrock below. The main
    difficulty in identifying these events stems from a large
    down-going high energy atmospheric muon flux, that
    could produce detector signatures consistent with those
    produced by up-going muons. These events are the
    background to this analysis.
    (GeV) (L = 12715 km )
    Muon Neutrino Energy E
    ν
    μ
    20
    40
    60
    80
    100 120 140 160 180 200
    )
    μ
    ν
    μ
    ν
    Survival Prob. P(
    0
    0.1
    0.2
    0.3
    0.4
    0.5
    0.6
    0.7
    0.8
    0.9
    1
    Vertical Muon Neutrino Survival Probability
    ν
    μ
    Survival Probablility
    2
    eV
    -3
    Δ
    m = 2.4
    ×
    10
    2
    2
    Θ
    = 1.0
    sin
    1200 600 400
    200
    L/E
    ν
    μ
    (km/GeV)
    Fig. 1. Muon neutrino survival probability under the assumption of
    effective 2-flavor neutrino oscillations ν
    µ
    ↔ ν
    τ
    as function of energy
    for vertically traversing neutrinos.

    2
    C. ROTT et al. ATM. OSCILLATION WITH ICECUBE
    Vertically up-going atmospheric neutrinos travel a
    distance of Earth diameter, which corresponds to a
    baseline length L of 12, 715 km. The survival probability
    for these muon neutrinos can be approximated using
    the two-flavor neutrino oscillation case and is shown in
    Figure 1 for maximal mixing and a ∆m
    2
    consistent with
    Super-Kamiokande [6] and MINOS [7] measurements. It
    illustrates the disappearance effect (large below energies
    of 100 GeV) we intend to observe.
    III. OSCILLATION ANALYSIS
    To probe oscillation effects, our selection criteria need
    to be optimized towards the selection of low-energy
    vertical muon events. The selection should also retain
    some events at higher energies (with no oscillation
    effects), that could be used to verify the overall normal-
    ization. Low energy vertical up-going muons in IceCube
    predominantly result in registered signals (”hits”) on a
    single string. The muon propagates very closely to one
    string, such that the Cherenkov light can be sampled
    well from even low-energy events. The probability of
    observing hits on a second string is very small due to
    the large interstring distance of 125 m, and is further
    suppressed through a local trigger condition known as
    HLC (Hard Local Coincidence). The HLC condition
    requires that a DOM only registers a hit if a (nearest
    or next-to-nearest) neighbor also registers a hit within
    1 µs. IceCube was operational in this mode for the 22
    and 40-string data.
    Given the nature of the signal events, the oscillation
    analysis can be performed very similarly on the different
    IceCube string configurations. To verify our understand-
    ing of the detector, we perform this analysis in steps.
    First, we use a subset of the 22-string configuration to
    develop and optimize the selection criteria, then cross
    check them on the full 22-string dataset and perform
    the analysis on the IceCube datasets acquired following
    the 22-string configuration.
    The IceCube 22-string configuration operated between
    May 31, 2007 and April 5, 2008. In this initial study,
    we analyze only a small subset of the data acquired over
    this period with a total livetime of 12.85 days, using ran-
    domly distributed data segments of up to 8 hour length
    collected during the period of 22-string operations. The
    dataset was triggered with the multiplicity eight DOM
    trigger and then preselected by a specific analysis filter
    running at the South Pole, selecting short track-like
    single string events. The filter requires after removal of
    potential noise hits, that all hits occur on a single string
    and that the time difference between the earliest and
    latest hit be less than 1000 ns. To partially veto down-
    going muon background it requires no hits in the top
    3 DOMs. Further, the hit time difference between at least
    two adjacent DOMs must be consistent with the speed
    of light within 25% tolerance, and the first DOM hit in
    time needs to be near the bottom or top within the series
    of DOMs hit on the single string. All filter selection
    criteria are designed to be directionally independent,
    so that vertical up-going events are collected as well
    as vertical down-going. The described analysis only
    uses the up-going sample collected by this filter. The
    down-going sample could be used in the future for
    flux normalization purposes, if we succeed in extracting
    a pure atmospheric neutrino sample against the large
    down-going atmospheric muon flux [3].
    To isolate our signal sample of vertical up-going
    ν
    µ
    events we apply a series of consecutive selection
    criteria. We require that the majority of time differences
    between adjacent DOMs are consistent with unscattered
    Cherenkov radiation (direct light) off a vertically up-
    going muon (L4). In addition, a maximum likelihood
    fit is applied requiring the muon to be reconstructed as
    up-going (L5). After these selection criteria, the dataset
    is still dominated by down-going muon background
    mimicking up-going events. This background is esti-
    mated using two CORSIKA [8] samples: one with an
    energy spectrum according to the Horandel¨
    polygonato
    model [5] and a second over-sampling at the high energy
    range. Simulations agree well with data in shape, but
    the normalization is found to be slightly high. Based on
    background and signal simulations (atmospheric ν
    µ
    were
    generated with ANIS [9]) we define a set of tight selec-
    tion criteria (that do not correlate strongly) and show
    good signal and background separation. These selection
    criteria are as follows: Event time length greater than
    400 ns (L6), mean charge per optical sensor larger than
    1.5 photo-electrons (pe), total charge collected during
    the first 500 ns larger than 12 pe (L7), and an inner string
    condition (the trigger string completely surrounded by
    neighboring strings) (L8). The tight selection criteria
    were independently optimized at level 5 in order to have
    high statistics and smoother distributions which would
    not be available at higher selection levels. Thereafter,
    we reject all events in the available background COR-
    SIKA sample corresponding to an equivalent detector
    livetime of at least two days, taking into account the
    oversampling. Using a conservative approach with two
    days of livetime equivalent we can set a 90%C.L. upper
    limit on the possible background contamination in the
    data sample of 14.8 events, in 12.85 days of livetime. In
    this sample we further expect 2.13 ± 0.07 (1.68 ± 0.06)
    signal events (with oscillation effects taken into account)
    from atmospheric neutrinos. See Table I for event counts
    as function of the selection criteria. Figure 2 shows the
    track length distribution after final selection criteria. The
    track length serves as an energy estimator working well
    at the energy range of interest since a muon travels
    roughly 5 m/GeV. As expected, short tracks show larger
    disappearance effects. Figure 3 shows the fraction of
    events selected by this analysis that are below a certain
    muon energy for different track lengths.
    The optimization and cross-check on the small sub-
    set of available data have been performed in a blind
    manner. One event was observed after final selection
    which is consistent with the prediction. This initial
    result indicates that we understand and model the low-

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁOD´ Z´ 2009
    3
    Track length (#DOMs)
    6
    8
    10
    12
    14
    16
    18
    20
    22
    Events
    0
    0.5
    1
    1.5
    2
    Atm. Nugen (no-osc)
    2.13
    Atm. Nugen (osc)
    1.68
    Data
    1
    Cut Level - 8 Track length (#DOMs) (L=12.85 days)
    Fig. 2.
    Expected track length of the signal, with and without
    oscillations taken into account, and compared to data after final
    selection criteria. .
    Track length (#DOMs)
    8
    9 10 11 12 13 14 15 16 17
    Fraction
    0
    0.05
    0.1
    0.15
    0.2
    0.25
    0.3
    0.35
    Tracklength vs. Muon Neutrino Energy at final selection cut level
    < 50 GeV
    Muon Neutrino Energy E
    ν
    μ
    [50,100] GeV
    Muon Neutrino Energy E
    ν
    μ
    > 100 GeV
    Muon Neutrino Energy E
    ν
    μ
    Fig. 3. Fraction of events in a given muon neutrino energy range as
    function of their track length defined by the number of DOMs hit at
    final selection.
    energy atmospheric neutrino region reasonably well. The
    analysis on the full dataset is in progress, including
    a larger background MC sample and a more detailed
    study of systematic uncertainties. Figure 4 shows the
    effective area for vertical up-going neutrinos in the 22-
    string detector at filter level and final selection.
    ν
    /GeV)
    log10(E
    1
    1.5
    2
    2.5
    3
    3.5
    )
    2
    Effective Area (m
    -9
    10
    -8
    10
    -7
    10
    -6
    10
    -5
    10
    -4
    10
    -3
    10
    Neutrino Effective Area - IceCube Preliminary
    Filter Level
    Final Cut Level
    Fig. 4. Average muon neutrino effective area for vertical up-going
    neutrinos (within 15 degree’s of vertical direction) as function of
    neutrino energy.
    IV. BACKGROUND ESTIMATION
    The background has been estimated using CORSIKA
    simulations. However, due to limited MC statistics there
    remains a large uncertainty at final selection.
    To cross-check the background estimation and to
    provide a second independent way to obtain a back-
    ground estimate, we use the data itself to determine the
    remaining background.
    Cut
    Corsika
    Sig. (with osc)
    Effect
    Data
    L3
    439 ± 2 · 10
    4
    20.3(17.3) ± 0.4
    15%
    331 · 10
    4
    L4
    54 ± 2 · 10
    3
    20.0(17.0) ± 0.3
    15%
    32 · 10
    3
    L5
    464 ± 175
    11.8(9.7) ± 0.2
    18%
    321
    L6
    351 ± 171
    10.7(8.8) ± 0.2
    18%
    207
    L7
    151 ± 41
    9.6(7.9) ± 0.2
    18%
    145
    L8
    0
    2.1(1.7) ± 0.08
    21%
    1
    TABLE I
    SUMMARY OF NUMBER OF EVENTS IN DATA AND AS PREDICTED BY
    SIMULATIONS AS FUNCTION OF THE SELECTION CRITERIA “CUT”
    LEVEL: L3 - INITIAL PROCESSING (TRIGGER, FILTER), L4/L5 -
    RECONSTRUCTED TRACK IS VERTICAL UP-GOING, L6/L7 - CHARGE
    BASED SELECTION CRITERIA, L8 - INNER STRINGS ONLY. SEE
    TEXT FOR DETAILED DESCRIPTION OF THE SELECTION CRITERIA.
    EFFECT REFERS TO THE SIZE OF THE DISAPPEARANCE EFFECT.
    The nature of the signal events (low energy vertical
    tracks on a single string) allows us to estimate the
    background based on the completeness of the veto region
    defined by the surrounding strings, using geometrical
    phase-space arguments.
    The total number of events observed is the sum of the
    passing signal events and background faking a signal.
    The two categories display very different behavior with
    respect to tightening the selection criteria. Signal events
    produce predominately real vertical tracks, so that the
    rate on strings regardless of their position is very similar
    (see Figure 5).
    Adjacent strings
    1
    2
    3
    4
    5
    6
    Events / 12.85 days
    1
    10
    2
    10
    3
    10
    4
    10
    5
    10
    6
    10
    7
    10
    8
    10
    9
    10
    Background (Cut Level 3)
    Background (Cut Level 4)
    Background (Cut Level 5)
    Vertical up Signal (Cut Level 4)
    Fig. 5.
    Number of events for 12.85 days of data at different cut
    levels as function of number of adjacent strings. The signal prediction
    is shown for comparison. Note that the number of adjacent strings does
    not affect the signal as those events are predominately single string
    events.
    Up-going ν
    µ
    of higher energies and non-vertical ν
    µ
    have a small impact on the overall rates. As selection
    criteria become more stringent, the rates on the strings
    become more homogeneous as they are dominated by
    “high quality” low-energy vertical muon neutrino events.
    Background behaves very differently under tightening
    selection criteria, as it becomes more difficult to produce
    a fake up-going track when the parameter space is taken
    away and the veto condition tends to have a larger
    impact.
    We determine the ratio between the average number

    4
    C. ROTT et al. ATM. OSCILLATION WITH ICECUBE
    of events observed on a string with n adjacent strings
    1
    and those with n + 1. At a low selection level, the rate
    on all strings is completely dominated by background.
    At high selection level, strings having less than four
    adjacent strings are also background dominated. We
    use these first three bins to scale the ratio distributions
    from an earlier selection level to the final selection
    level. Figure 6 shows the predicted number of events
    at next-to-final selection level (L7) obtained with this
    method. The background estimation method from data
    itself needs to be finalized, including a study of the
    systematic uncertainites. It provides a cross-check to the
    predictions from simulation and may ultimately be used
    as the preferred background estimation method in this
    analysis.
    Number adjacent strings
    1
    2
    3
    4
    5
    6
    Average number of events per string
    0
    10
    20
    30
    40
    50
    60
    Average Number of Events per String (L=12.85 days)
    Corsika Background
    Atm. Nugen + Corsika Background
    Background prediction from data
    Data
    Fig. 6. Average number of events per string at next-to-final selection
    level (L7) as function number of adjacent strings. Note that the right
    most bin corresponds to the final selection.
    V. DISCUSSION OF SENSITIVITY FOR 40-STRING
    AND FULL ICECUBE
    The IceCube 40-string dataset is in many ways su-
    perior to the 22-string dataset. The trigger system has
    been significantly improved over the 22-string detector
    through the addition of a string trigger [10], roughly
    doubling the vertical muon neutrino candidate events per
    string. In order to reject efficiently against down-going
    muon background, we require that a string be entirely
    surrounded by adjacent strings (inner strings criterion)
    as part of the final selection. The 40-string detector has
    about a factor of three more inner strings.
    Based on the selection criteria for the IceCube 22-
    string analysis, we have evaluated the sensitivity of the
    40-string detector with one year of data using a χ
    2
    -
    test on the track length distribution. Selection criteria
    are identical to those presented here, but the number of
    expected signal events is scaled according to expectation
    for the 40-string array. We expect about 400 signal
    events, based on the detector livetime, number of inner
    strings, and a factor two increase in number of events
    1
    We define adjacent strings as those that are within the nominal
    interstring-distance (roughly 125 m) of the hexagonal detector pattern.
    due to the string trigger. Figure 7 shows the expected
    sensitivity limits obtained in this way as function of
    the oscillation parameters. Systematic uncertainties are
    still being investigated and are not included; They are
    dominated by the atmospheric neutrino flux uncertainty,
    optical module sensitivity and ice effects.
    2
    23
    sin
    0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
    )
    4
    /c
    2
    (eV
    2
    m
    Δ
    0
    0.001
    0.002
    0.003
    0.004
    0.005
    0.006
    0.007
    0.008
    0.009
    0.01
    -3
    10
    -2
    10
    -1
    10
    1
    10
    2
    10
    Expected IceCube 40-string Sensitivity (no background)
    2
    Δχ
    95% C.L.
    90% C.L.
    1
    σ
    IceCube Preliminary
    Fig. 7.
    Expected constraints on oscillation parameters using the
    IceCube detector in the 40-string configuration under the assumption
    of zero background .
    VI. CONCLUSIONS
    Preliminary results obtained with a subset of the data
    collected with the IceCube 22-string configuration active
    during 2007 and 2008, suggest that IceCube may have
    sensitivity in the energy range where atmospheric oscil-
    lations become important. We estimate the sensitivity to
    oscillation parameters in the IceCube 40-string dataset
    and find that IceCube can potentially constrain them,
    pending the determination of the systematic uncertainties
    associated with the predicted distributions. Understand-
    ing of this energy region is also important for dark matter
    annihilation signals from the center of the Earth and
    further provides the groundwork for DeepCore, which
    will probe neutrinos at a similar and even lower energy
    range [2].
    REFERENCES
    [1] A. Achterberg et al. [IceCube Collaboration], Astropart. Phys.
    26, 155 (2006).
    [2] D. Grant et al. [IceCube Collaboration], Fundamental Neutrino
    Measurements with IceCube DeepCore, this proceedings.
    [3] I. F. M. Albuquerque and G. F. Smoot, Phys. Rev. D 64, 053008
    (2001).
    [4] C. Rott [IceCube Collaboration], “Neutrino Oscillation Measure-
    ments with IceCube,” arXiv:0810.3698.
    [5] J. R. Horandel,¨
    Astropart. Phys. 19 (2003) 193.
    [6] Y. Ashie et al. [Super-Kamiokande Collaboration], Phys. Rev. D
    71, 112005 (2005).
    [7] D. G. Michael et al. [MINOS Collaboration], Phys. Rev. Lett.
    97, 191801 (2006); P. Adamson et al. [MINOS Collaboration],
    Phys. Rev. Lett. 101, 131802 (2008).
    [8] D. Heck et al., Forschungszentrum Karlsruhe Report FZKA-
    6019, 1998.
    [9] A. Gazizov and M. P. Kowalski, Comput. Phys. Commun. 172,
    203 (2005).
    [10] A. Gross et al.[IceCube Collaboration], arXiv:0711.0353.

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    1
    Direct Measurement of the Atmospheric Muon Energy Spectrum
    with IceCube
    Patrick Berghaus
    for the IceCube Collaboration
    University of Wisconsin, Madison, USA
    see special section of these proceedings
    Abstract
    . Data from the IceCube detector in its
    22-string configuration (IC22) were used to directly
    measure the atmospheric energy spectrum near the
    horizon. After passage through more than 10 km
    of ice, muon bundles from air showers are reduced
    to single muons, whose energy can be estimated
    from the total number of photons registered in the
    detector. The energy distribution obtained in this way
    is sensitive to the cosmic ray composition around
    the knee and is complementary to measurements by
    air shower arrays. The method described extends
    the physics potential of neutrino telescopes and can
    easily be applied in similar detectors. Presented is
    the result from the analysis of one month of IC22
    data. The entire event sample will be unblinded once
    systematic detector effects are fully understood.
    Keywords
    : atmospheric muons, CR composition,
    neutrino detector
    I. INTRODUCTION
    While the primary goal of IceCube is the detection
    of astrophysical neutrinos, it also provides unique op-
    portunities for cosmic-ray physics [1]. One of the most
    important is the direct measurement of the atmospheric
    muon energy spectrum.
    As shown in figure 1 the energy spectrum of muons
    produced in cosmic-ray induced air showers has so far
    been measured only up to an energy of about 70 TeV
    [2]. The best agreement with theoretical models was
    found by the LVD detector, with the highest data point
    located at
    E
    µ
    = 40 TeV
    [3]. All these measurements
    have been performed using underground detectors. Their
    sensitivity was limited by the relatively small effective
    volume compared to neutrino telescopes.
    With a planned instrumented volume of one cubic
    kilometer, IceCube will be able to register a substantial
    amount of events even at very high energies, where the
    flux becomes very low. The limitation in measuring the
    muon spectrum is given by its high granularity, and
    consequent inability to resolve individual muons. Most
    air showers containing high energy muons will consist
    of bundles with hundreds or even thousands of tracks.
    Since the energy loss per unit length can be described by
    the equation
    dE/dx = a + bE
    , low-energy muons will
    contribute disproportionately to the total calorimetric de-
    tector response, which depends strongly on the energy of
    the primary, disfavoring the measurement of individual
    muon energies.
    10
    -2
    10
    -1
    10
    3
    10
    4
    10
    5
    3
    2
    1
    KM + ATIC2
    KM + GAMMA
    KM + ZS
    QGSJET II-03 + ZS
    SIBYLL 2.1 + ZS
    EPOS 1.61 + ZS
    L3+Cosmic, 2004
    CosmoALEPH, 2007
    LVD, 1998
    MSU, 1994
    Baksan, 1992
    Frejus, 1990
    Artyomovsk, 1985
    ||||
    MACRO best fit
    E
    μ
    , GeV
    E
    3
    D
    μ
    (E), cm
    -2
    s
    -1
    sr
    -1
    GeV
    2
    Fig. 1: Muon surface energy spectrum measurements
    compared to theoretical models [2].
    This problem can be resolved by taking advantage of
    the fact that low energy muons are attenuated by energy
    losses during passage through the ice. In this analysis,
    the emphasis was therefore set on horizontal events,
    where only the most energetic muons are still able to
    penetrate the surrounding material. The primary cosmic
    ray interaction in this region takes place at a higher
    altitude, and therefore in thinner air. The reinteraction
    probability for light mesons (pions and kaons) is smaller
    and the flux of muons originating in their decays is
    maximized.
    The main possibilities for physics investigations using
    the muon energy spectrum are:
    Forward production of light mesons at high ener-
    gies. While muon neutrinos at TeV energies mostly
    come from the process
    K → ν
    µ
    + X
    , for kine-
    matical reasons muons originate predominantly in
    pion decays
    π → ν
    µ
    + µ
    [4]. An estimate of
    the pion production cross section from accelerator
    experiments gives an uncertainty of
    δ(σ
    ) ≃ 15% + 12.2% · log
    10
    (E
    π
    /500 GeV)
    at
    x
    lab
    > 0.1
    above 500 GeV [5]. This value
    should also apply in good approximation to the
    conventional (non-prompt) muon flux.
    Prompt flux from charm meson decay in air show-
    ers [6]. Because of their short decay length, the

    2
    P. BERGHAUS
    et al.
    ICECUBE MUON ENERGY SPECTRUM
    reinteraction probability for heavy quark hadrons
    is negligible. The resulting muon energy spectrum
    follows the primary energy spectrum with a power
    law index of
    γ ≈  2.7
    and is almost constant over
    all zenith angles. Since the non-prompt muon flux
    from lighter mesons is higher near the horizon, this
    means that the relative contribution from charm is
    lowest, and very challenging to detect.
    Variations of the muon energy spectrum due to
    changes of the CR composition around the knee.
    Since the ratio of median parent cosmic ray and
    muon energy is
    ≤ 10
    at energies [7] above 1 TeV,
    a steepening of the energy-per-nucleon spectrum of
    cosmic rays at a few PeV will have a measurable ef-
    fect on the atmospheric muon spectrum at energies
    of hundreds of TeV. Comparison of the measured
    muon spectrum to various phenomenological com-
    position models was the main focus of this analysis.
    An additional benefit in the case of neutrino detectors
    is that a direct measurement of the muon flux will
    have important implications for neutrino analyses. By
    reducing the systematic uncertainties on atmospheric
    lepton production beyond 100 TeV, the detection po-
    tential for diffuse astrophysical fluxes will be enhanced.
    Also, atmospheric muons serve as a “test beam” that
    allows calibration of the detector response to high-
    energy tracks.
    II. COSMIC RAY COMPOSITION MODELS
    Starting from the hypothesis that most cosmic rays
    originate from Fermi acceleration in supernova shock
    fronts within our galaxy, the change in the energy
    spectrum can be explained by leaking of high energy
    particles. Since the gyromagnetic radius
    R =
    p
    eZB
    ≃ (10pc)
    E
    prim
    [P eV ]
    ZB[µG]
    depends on the charge
    Z
    of the particle, for a given
    energy nuclei of heavier elements are less likely to
    escape the galactic magnetic field than lighter ones.
    The general expression for the flux of primary nuclei
    of charge
    Z
    and energy
    E
    0
    is
    Z
    dE
    0
    = Φ
    0
    Z
    ?
    1+
    ?
    E
    0
    E
    trans
    ?
    ǫ
    c
    ?
     ∆γ
    ǫc
    where the transition energy
    E
    trans
    corresponds to
    E
    ˆ
    p
    · Z
    ,
    E
    ˆ
    p
    · A
    or simply
    E
    ˆ
    p
    for rigidity-dependent,
    mass-dependent and constant composition models. The
    parameter
    ǫ
    c
    determines the smoothness of the transition,
    and
    ∆γ
    the change in the power law index.
    Three alternative composition models have been pro-
    posed, which all can be fit reasonabkly well to the total
    cosmic ray flux in the region of the knee [8]. These are:
    Rigidity-Dependent
    ∆γ
    : This is the default com-
    position used in the IceCube downgoing muon
    simulation. It is also the one favored by current
    E
    μ
    [GeV]
    5
    2×10
    5
    3×10
    6
    10
    6
    2×10
    6
    3×10
    ]
    −1
    PeV)
    srad
    sec
    2
    [(km
    3
    (E/1PeV)
    μ
    N
    −7
    10
    −6
    10
    −5
    10
    Constant Composition
    Mass−Dependent
    Δγ
    Rigidity−Dependent
    Δγ
    Fig. 2: Atmospheric muon energy spectrum at surface
    level averaged over the whole sky as simulated with
    CORSIKA/SIBYLL.
    models of cosmic ray production and propagation
    in the galactic amgnetic field.
    Mass-Dependent
    ∆γ
    : An alternative model that
    also leads to a composition change around the knee.
    The change in the power law index does not depend
    on the charge, but on the mass of the nucleus. The
    best fit proposed in the original paper leads to a
    smaller value for the transition energy and a steeper
    spectrum after the cutoff.
    Constant Composition: Here, the composition of
    the primary cosmic ray flux does not change. The
    knee is explained by a common steepening in the
    energy spectrum for all primaries occurring at the
    same energy.
    The best measurement of the composition so far was
    done by KASCADE [9]. Its result was consistent with
    a steepening of the spectrum of light elements, but
    depended strongly on the hadronic interaction model
    used to simulate the air showers (SIBYLL or QGSJET).
    The influence of the three composition models on the
    muon energy spectrum is shown in figure 2. While the
    spectrum for the constant composition model gradually
    changes from
    E
     3.7
    to
    E
     4
    , the other two show a
    marked steepening corresponding to the cutoff in the
    energy per primary nucleon. By accurately measuring
    the muon energy spectrum, it is therefore possible to
    significantly constrain the range of allowed cosmic ray
    composition models in the knee region.
    III. ANALYSIS
    The data set used in this analysis is based on the
    IceCube online muon filter, designed to contain all
    track-like events originating from the region below
    70
    .
    It covers the period from June 2006 to March 2007
    with an integrated livetime of 275.6 days, during which
    IceCube was taking data with 22 strings (IC22). A
    number of quality cuts were applied in order to eliminate
    background from misreconstructed tracks and to reduce
    the median error in the zenith angle measurement to

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    3
    3
    ~e
    surf
    3.5
    4
    4.5
    5
    5.5
    /GeV (true)
    ,max
    μ
    E
    10
    log
    3
    3.5
    4
    4.5
    5
    5.5
    6
    [Hz]
    −8
    10
    −7
    10
    −6
    10
    −5
    10
    −4
    10
    −3
    10
    −2
    10
    Fig. 3: Relation between energy proxy
    surf
    and true
    surface energy of most energetic muon in shower. Here
    and in figure 4 the rigidity-dependent composition model
    was used.
    ≈ 0.7
    . The final sample corresponded to an event rate
    of 0.146 Hz.
    θ
    zen
    [deg]
    d
    vert
    [km]
    d
    slant
    [km]
    E
    thr
    µ
    [TeV ]
    0
    1.5
    1.5
    0.28
    70
    1.5
    4.39
    1.12
    70
    2.5
    7.31
    2.59
    85
    1.5
    17.21
    22.1
    85
    2.5
    28.68
    207
    TABLE I: Threshold energy for muons passing through
    ice. The energy values correspond to an attenuation of
    99.9%.
    To measure the single muon energy spectrum, it is
    necessary to reduce the background of high-multiplicity
    bundles, whose total energy depends primarily on the
    primary cosmic ray [10]. Since there is no possibility
    to accurately estimate the multiplicity of a downgoing
    muon bundle, the only way to obtain single muons is by
    selecting a region close to the horizon to which muons
    of lower energies cannot penetrate.
    The minimum energy required for muons passing
    through a distance
    d
    of ice can be approximated by the
    equation
    E
    cut
    (d) = (e
    bd
     1)a/b
    where
    a = 0.163GeVm
     1
    and
    b = 0.192 · 10
     3
    m
     1
    [11]. The resulting threshold energies corresponding to
    vertical tracks and for tracks at the top and bottom of
    the detector for angles near the horizon are shown in
    Table I.
    Two factors determine the upper energy bound of
    this analysis. One is the contribution from atmospheric
    neutrinos, which will eventually dominate the event
    sample at large depths. The other, and more important, is
    the finite zenith angle resolution. Using simulated data, it
    was determined that it effectively limits the measurement
    of the slant depth to a values below 15 km.
    N
    μ,detector
    1
    2
    3
    4
    5
    6
    7
    8
    9
    10
    fraction per bin
    −2
    10
    10
    −1
    1
    All
    ~e
    surf
    > 4.2
    ~e
    surf
    > 4.5
    ~e
    surf
    > 4.8
    Fig. 4: Simulated muon multiplicity for atmospheric
    showers at closest approach to the center of the InIce
    detector for different values of
    ǫ˜
    .
    μ
    /GeV)
    10
    (E
    log
    3.5
    4
    4.5
    5
    5.5
    6
    6.5
    fraction per bin
    0
    0.02
    0.04
    0.06
    0.08
    0.1
    0.12
    0.14
    0.16
    0.18
    0.2
    All Events
    4.2<
    ~e
    surf
    <4.3
    4.5<
    ~e
    surf
    <4.6
    4.8<
    ~e
    surf
    <4.9
    Fig. 5: True surface energy of most energetic muon in
    shower using rigidity-depending composition model for
    different values of
    surf
    , with fits to Gaussian function.
    All individual distributions are normalized to unity.
    Using the slant depth alone, the range of this analysis
    is therefore insufficient to probe the region beyond 100
    TeV. However, the reach can be extended by incorpo-
    rating information about the energy of the muon as it
    passes through the instrumented volume.
    For muon tracks in the detector, the energy resolution
    approaches
    ∆ log
    10
    (E) ≈ 0.3
    above 10 TeV [12]. This
    information can be combined with the slant depth to
    obtain a better estimate for the muon energy at the
    surface.
    A natural way to do this is by defining a surface
    energy proxy
    that behaves as
    exp(e˜
    surf
    ) ∝ log n
    γ
    · d
    slant
    where
    n
    γ
    represents the total number of photons
    measured by the detector. Figure 3 shows the re-
    sulting parameter, which has been linerly rescaled in
    such a way that its value corresponds to the mean
    log(E
    µ,surf
    /GeV )
    for any given bin, provided that
    the muon energy spectrum is reasonably close to the
    standard
    E
     3.7
    .
    An important criterion for the applicability of the

    4
    P. BERGHAUS
    et al.
    ICECUBE MUON ENERGY SPECTRUM
    ~e
    surf
    3
    3.5
    4
    4.5
    5
    5.5
    events per bin [Hz]
    −7
    10
    −6
    10
    −5
    10
    −4
    10
    −3
    10
    −2
    10
    Data
    Constant Composition
    Mass Dependent
    Δγ
    Rigidity Dependent
    Δγ
    Fit to Const. Comp.
    Fig. 6: Data from one month of IC22 at final cut level
    compared to simulated event rates and fit of empirical
    function
    exp(a + be˜ + ce˜
    2
    )
    to constant composition
    distribution.
    ~e
    surf
    3
    3.5
    4
    4.5
    5
    5.5
    Ratio Data/Simulation
    0
    0.2
    0.4
    0.6
    0.8
    1
    1.2
    1.4
    Constant Composition
    Mass Dependent
    Δγ
    Rigidity Dependent
    Δγ
    Fig. 7: Ratio of experimental to simulated
    surf
    distri-
    butions for one month of IC22 data. Uncertainties are
    statistical only and exclude systematic detector effects.
    energy proxy parameter is that over the entire range of
    measurement the muon multiplicity remains low, and the
    influence of high-multiplicity bundles small. Figure 4
    confirms that this is indeed the case. It should be noted
    here that the most energetic muon typically accounts
    for the dominant contribution to the total energy in the
    detector, such that other tracks in the bundle can be
    neglected.
    The spread in muon surface energies for a given value
    of
    surf
    is shown in Figure 5. Around the peak the
    distributions can be approximated by a Gaussian whose
    width lies in the range of
    ∆log
    10
    E ≈ 0.3  0.4
    .
    IV. RESULT
    Figure 6 shows simulated event rates in dependence
    of
    surf
    compared to data at final cut level. Almost over
    the entire range all three models can be approximated
    by the same empirical fit function. Only in the highest
    bin can a distinction be made.
    The experimental data agrees remarkably well with
    the simulation, as can be seen more clearly in Figure
    7. Despite the steeply falling distribution, the ratio of
    data to simulation remains very close to one over almost
    the entire range. For
    e˜ > 5
    , corresponding to
    E
    µ
    >
    100 TeV
    , the measurent is based on only three data
    events.
    Using the entire year of IC22 data, the predicted event
    yield for
    5.1 < e˜
    surf
    < 5.2
    based on the constant
    composition model corresponds to about 10 events. It
    is therefore unlikely, even neglecting systematic uncer-
    tainties, that any of the three models under considera-
    tion could definitely be excluded yet. This situation is
    expected to change as soon as 40-string data can be
    included in the analysis.
    V. CONCLUSION
    This result demonstrates the potential for an accurate
    measurement of the muon energy spectrum with large
    neutrino detectors. So far only one month of data has
    been considered in the analysis, corresponding to about
    10% of the entire event sample. Nevertheless, the mea-
    surement already covers an energy range almost a factor
    of three above that of the previous upper limit, with very
    good agreement between data and simulation.
    While it will be difficult to make a definitive statement
    about the cosmic ray composition around the knee based
    on IC22 data, it will be possible to confirm the validity
    of cosmic ray air shower models up to previously
    inaccessible energy ranges.
    At the time of writing, the instrumented volume
    of the detector has increased by a almost factor of
    three. Further enlargements are scheduled for the next
    few years. Future measurements of the muon energy
    spectrum will benefit from a larger effective area, and a
    substantial improvement in the angular resolution related
    to the longer lever arm for horizontal muon tracks within
    the detector.
    Once residual systematic detector uncertainties are
    resolved, a comprehensive analysis that accounts for
    both the energy spectrum of individual muons and the
    total shower energy in the detector will be feasible. The
    potential for such a combined measurement is unique to
    large volume detectors.
    REFERENCES
    [1] P. Berghaus [IceCube Collaboration],
    Proc. of ISVHECRI 2008
    ,
    arXiv:0902.0021 [astro-ph.HE].
    [2] A. A. Kochanov, T. S. Sinegovskaya and S. I. Sinegovsky,
    Astropart. Phys.
    30
    (2008) 219
    [3] M. Aglietta
    et al.
    [LVD Collaboration], Phys. Rev. D
    60
    (1999)
    112001
    [4] T. K. Gaisser, Nucl. Phys. Proc. Suppl.
    118
    (2003) 109.
    [5] G. D. Barr, T. K. Gaisser, S. Robbins and T. Stanev, Phys. Rev.
    D
    74
    (2006) 094009.
    [6] G. Gelmini, P. Gondolo and G. Varieschi, Phys. Rev. D
    67
    (2003)
    017301.
    [7] T.K. Gaisser, “Cosmic Ray and Particle Physics,” Cambridge
    University Press, 1990.
    [8] J. R. Ho¨randel, Astropart. Phys.
    19
    , 193 (2003).
    [9] T. Antoni
    et al.
    [The KASCADE Collaboration], Astropart. Phys.
    24
    , 1 (2005).
    [10] D. Chirkin [AMANDA Collaboration],
    Proc. of ICRC 2003
    .
    [11] D. Chirkin and W. Rhode, arXiv:hep-ph/0407075.
    [12] J. Zornoza, D. Chirkin [IceCube Coll.], ICRC 2007,
    arXiv:0711.0353 (pp. 63-66).

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁOD´ Z´ 2009
    1
    Search for Diffuse High Energy Neutrinos with IceCube
    Kotoyo Hoshina
    for the IceCube collaboration
    Department of Physics, University of Wisconsin, Madison, WI 53706, USA
    See the special section of these proceedings
    Abstract. We performed a search for diffuse high
    energy neutrinos using data obtained with the Ice-
    Cube 22 string detector during a period 2007-2008.
    In this analysis we used an E
    −2
    spectrum as a typical
    flux resulting from cosmic ray shock acceleration. Us-
    ing a likelihood track reconstruction, approximately
    5700 track-like neutrinos are extracted from 275.7
    days data at an estimated 95% purity level. The
    expected sensitivities obtained are in a range of
    2.2×10
    −8
    ∼ 2.6×10
    −8
    E
    −2
    GeV cm
    −2
    s
    −1
    sr
    −1
    with
    four different energy estimators. The analysis method
    and results are presented along with discussions of
    systematics.
    Keywords: IceCube neutrino diffuse
    I. INTRODUCTION AND DETECTION PRINCIPLE
    The IceCube neutrino observatory is the world’s
    largest neutrino telescope under construction at the ge-
    ographic South Pole. During 2007, it collected data
    with 1320 digital optical modules(DOM) attached to 22
    strings (with 60 optical modules per string). They are
    deployed in clear glacial ice at depths between 1450
    to 2450 meters beneath the surface, where the photon
    scattering and absorption are known by preceding in situ
    measurements [1]. When a neutrino interacts inside or
    close to the IceCube detector, DOMs capture Cherenkov
    photons from secondary charged particles with 10 inch
    photomultiplier tubes and generate digital waveforms. In
    most cases, we require at least 8 DOMs to be triggered
    within a 10 micro second time window. Once the trigger
    condition is satisfied, all digital waveforms are collected
    and then processed by online filtering programs to
    filter out background events. In this analysis we used
    275.7 days livetime of data and obtained 5718 candidate
    neutrino induced events after the final event selection.
    The event selection process is described in Section II.
    The event sample after the selection process mainly
    consists of atmospheric neutrinos. To separate extrater-
    restrial high-energy neutrinos from atmospheric neutri-
    nos, one can apply two types of analysis techniques.
    The first is a point source analysis that uses the di-
    rection of the neutrinos to survey high-density event
    spots (hotspots). The second, called a diffuse analysis,
    examines the energy spectrum itself and compares it to
    various physics models. Since the diffuse analysis does
    not require multiple events from an astrophysical source,
    it is possible to take into account faint sources that are
    not significant by themselves in a point source analysis.
    However, in general, a diffuse analysis requires a better
    detector simulation. While a point source analysis uses
    data to search for a hotspot, the diffuse analysis has
    to rely on simulated parameter distributions under an
    assumption of a physics model to test observed distri-
    butions in the data.
    In this analysis we assumed a Φ ∝ E
    −2
    energy spec-
    trum for neutrinos from astrophysical sources result-
    ing from shock acceleration processes [2]. Since the
    atmospheric neutrino flux has a much softer energy
    spectrum [3][4][5], the signal neutrinos may form a
    high-energy tail in an energy-related observable over
    atmospheric neutrinos. The search for an extraterrestrial
    neutrino component uses the number of events above
    an energy estimator cut after subtracting a calculated
    contribution from atmospheric neutrinos. The cut was
    optimized to produce the best limit setting sensitivity [6].
    Results and possible sources of systematics errors are
    discussed in Section IV.
    II. EVENT SELECTION
    Cosmic ray interactions in the atmosphere create
    pions, kaons and charmed hadrons which can later decay
    into muons and neutrinos. The primary background
    before the event selection is atmospheric muons travel-
    ing downward through the ice. Their intensity strongly
    depends on the zenith angle of the muon: it decreases
    as the zenith angle increases because a higher zenith
    angle results in a longer path length from the surface
    of the Earth to the IceCube detector. The largest zenith
    angle of atmospheric muons is around 85 degrees and
    their path length inside the Earth is over 20 km. The
    first filter is thus designed to select only upward going
    events. For estimation of the zenith angle, we used a log
    likelihood reconstruction. In this analysis, the minimum
    zenith threshold is 90 degrees.
    After the zenith angle filter is applied, the remaining
    data still contains many orders of magnitude more mis-
    reconstructed background than neutrino-induced events.
    They are downward going muons, but reconstructed
    as upward because of poor event quality (low num-
    ber of triggered DOM, grazing an edge of detector,
    etc) or two muons that passed through the detector
    within a trigger time window (coincidence muons) and
    mis-reconstructed as a single upward going muon
    1
    .
    These mis-reconstructed events are effectively rejected
    by checking fit quality parameters [7]:
    1
    This difficulty is mainly caused by scattering of photons in ice.
    The effective scattering length of Cherenkov photons in IceCube is
    around 30 m [1].

    2
    K.HOSHINA et al. SEARCH FOR DIFFUSE NEUTRINOS WITH ICECUBE
    NCh
    0 20 40 60 80 100 120 140160 180 200
    10
    -1
    1
    10
    10
    2
    10
    3
    Data (intg 5718)
    Coincidence muons (intg 37.0846)
    Corsika single muons (intg 19.7651)
    Atmospheric neutrinos (intg 5440.27)
    10
    -7
    E
    -2
    neutrinos (intg 195.185)
    (a) NCh
    Cos(Zenith)
    -1
    -0.8 -0.6
    -0.4 -0.2
    0
    0.2
    10
    10
    2
    Data (intg 5718)
    Coincidence muons (intg 37.0846)
    Corsika single muons (intg 19.7651)
    Atmospheric neutrinos (intg 5440.56)
    10
    -7
    E
    -2
    neutrinos (intg 200.232)
    (b) cosine zenith
    COGZ
    -500-400-300-200-100 0 100 200300 400 500
    1
    10
    10
    2
    Data (intg 5718)
    Coincidence muons (intg 37.0846)
    Corsika single muons (intg 19.7651)
    Atmospheric neutrinos (intg 5440.56)
    10
    -7
    E
    -2
    neutrinos (intg 200.231)
    (c) COGZ
    loglBayes32 - loglSPE32
    -20
    0
    20
    40
    60
    80
    100
    1
    10
    10
    2
    10
    3
    Data (intg 5661)
    Coincidence muons (intg 37.0846)
    Corsika single muons (intg 19.7651)
    Atmospheric neutrinos (intg 5410.45)
    10
    -7
    E
    -2
    neutrinos (intg 183.312)
    (d) log likelihood ratio
    log
    10
    (dEdX [GeV])
    -2
    -1
    0
    1
    2
    3
    4
    5
    10
    -1
    1
    10
    10
    2
    10
    3
    Data (intg 5718)
    Coincidence muons (intg 37.0846)
    Corsika single muons (intg 19.7651)
    Atmospheric neutrinos (intg 5440.09)
    10
    -7
    E
    -2
    neutrinos (intg 200.749)
    (e) log10(dEdX)
    E [GeV]
    10
    log
    1
    2
    3
    4
    5
    6
    7
    2
    m
    -3
    10
    -2
    10
    -1
    10
    1
    10
    2
    10
    3
    10
    4
    10
    Average(90-180deg)
    90 < zen < 110
    110 < zen < 145
    145 < zen < 180
    (f) Effective Area
    Fig. 1. (a - d) : Comparison of simulations and data for basic parameters after the purification process. COGZ(c) is the z-position of the
    center of charge-gravity of an event in the IceCube coordinate system. (d) : An alternative energy estimator. See Sec. III. (f) : Effective area
    of ν
    µ
    + ν
    µ
    after the event selection, in several zenith angle ranges.
    • Number of direct hits (NDir) : number of hits which
    are assumed to result mostly from unscattered
    Cherenkov photons
    • Projected length of direct hits (LDir) : Largest
    distance of a pair of projections from direct hit
    positions to a reconstructed track
    • Reduced log likelihood : log likelihood result of a
    reconstructed track divided by number of degrees
    of freedom
    • log likelihood ratio : difference of log likelihood
    parameters between a fit and a Bayesian fit which
    is forced to reconstruct as downward going
    • smoothness of hits : a parameter for how hits are
    generated smoothly along a reconstructed track
    • log likelihood ratio between single muon fit and
    Bayesian weighted double muon fit : similar param-
    eter as log likelihood ratio, but uses two Bayesian
    fits as a hypothesis of coincidence muons
    The “direct hits” are defined by the arrival times of
    photons at each DOM and a reconstruction. Once a
    reconstruction is determined, at each DOM, we obtain
    a minimum path and earliest possible arrival times of
    photons (geometrical hit times) from the Cherenkov light
    emission point. Some photons may take a longer path
    because of scattering, which result in a time delay from
    the geometrical hit time. In this analysis, we chose a
    time window of [-15ns, 75ns] from the geometrical hit
    time to accept a hit as a direct hit.
    The log likelihood ratio gives a comparison between
    two fits, a standard likelihood fit and a fit with a zenith-
    dependent weight which follows a zenith distribution of
    atmospheric muons. A reliable good quality fit should
    have a large ratio, while mis-reconstructed atmospheric
    muons have relatively smaller ratios.
    With these quality parameters, we defined a set of
    cut parameters to purify neutrino-induced events using
    Monte-Carlo simulation. For atmospheric muons, we
    generated 10 days of single unweighted CORSIKA
    muons, 5 × 10
    5
    events of energy weighted CORSIKA
    muons
    2
    , and 7.4 days of unweighted CORSIKA coin-
    cidence muons. For atmospheric neutrinos, 2.6 × 10
    7
    ν
    µ
    events were generated with an E
    −1
    spectrum and
    re-weighted with a conventional atmospheric neutrino
    flux [3] plus a prompt neutrino model [4][5]. The
    optimal cut is chosen to retain as many high energy
    neutrinos as possible while keeping purity of neutrinos
    above 95 %.
    The optimized cut parameter is then applied to data
    and compared with Monte-Carlo predictions. In order
    not to bias the analysis, the highest energy tails of
    both data and simulation were kept hidden from the
    analyzer during this final optimization process of cuts.
    The number of DOMs that has at least one hit (NCh)
    is used to determine the open window: we compared
    events which NCh less than 80. Small discrepancies
    2
    The power law index of the primary particle is changed to be harder
    by +1. The effective livetime varies in each primary energy bin, for
    example, 10 TeV weighted muons correspond to one year of effective
    livetime. The effective livetime also depends on zenith, e.g. a value of
    a year for muons around 70 degree.

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁOD´ Z´ 2009
    3
    between data and simulation around the threshold of
    some quality parameters were observed, mainly because
    of insufficient statistics of background coincident muon
    simulation. These events are removed by tightening the
    cut parameters moderately.
    Figs. 1 shows the comparison of data and simulation
    after the final event selection. The Monte Carlo simu-
    lation reproduces data well in most event variables, but
    discrepancies are still present in some depth dependent
    variables like COGZ, the z (vertical, or depth) coordinate
    of the center-of-gravity of the charge in the event (z
    = 0 in the center of the detector). This systematics is
    discussed in Sec. IV. The neutrino effective area of ν
    µ
    + ν
    µ
    after optimal quality cuts for 275.7 days of livetime
    of IceCube 22 strings is shown in Fig. 1f.
    III. ENERGY ESTIMATORS AND SENSITIVITY
    Unlike the previous detector AMANDA, the IceCube
    detector retains the original waveform by digitizing ana-
    log waveforms inside the DOM. This technology allows
    us to use charge information as an energy estimator.
    Recently, new techniques for energy reconstruction were
    developed using the charge information as well as the
    hit times. In this section, we compare the sensitivity of
    following energy estimators.
    • NCh : number of triggered DOMs. It is simple, but
    has a relatively strong connection with the track
    geometry and the ice layers where the muon passed
    through.
    • NPe : Total charge collected by all triggered DOMs
    of an event. Basically it is similar to NCh, but has
    a larger and smoother dynamic range than NCh.
    • dEdx : A table based energy reconstruction. Using
    a table generated by a photon propagation program
    (Photonics [8]), it estimates the energy deposit
    along a reconstructed track. The reconstruction
    takes into account the ice properties as a function
    of depth. [9]
    • MuE : a simple energy reconstruction. Similar to
    dEdx, but uses an homogeneous ice model instead
    of layered ice photonics tables. [10]
    To obtain sensitivities, we assumed no extra-terrestrial
    signal over a given energy threshold, then calculated
    the expected upper limit using the Feldman-Cousins
    method [11]. The Model Rejection Factor [6] is then
    optimized to have the best sensitivity for E
    −2
    test signal
    flux. Table I shows sensitivities at corresponding energy
    estimator thresholds. The average number of background
    neutrinos and Φ = 10
    −7
    E
    −2
    signal neutrinos above the
    threshold are also predicted.
    IV. RESULTS AND DISCUSSION
    Table I also lists the number of data events above
    the optimized energy thresholds for the four energy es-
    timators. We observed a statistically significant excess of
    data over the atmospheric neutrino prediction (including
    prompt atmospheric neutrinos) for all energy estimators
    except for dEdx. However, disagreements between data
    and simulation in depth dependences (for example, in
    COGZ in Fig. 1c) point to unresolved systematics in
    our simulation. In this section we discuss the effect of
    the COGZ problem to this analysis.
    The depth dependences in the optical properties of the
    glacial ice, reflecting changes in dust concentration due
    to climate variations when the ice was formed [1], are
    taken into account in the detector simulation. However,
    as Fig. 1c shows, these dependences are not fully repro-
    duced by the simulation. In this analysis, the discrepancy
    is most severe in the deep part of the detector, for COGZ
    < -250 m, which is also where most of the highest-
    energy events lie. The event excess we observed thus
    could be due to systematics rather than a signal flux.
    To test the hypothesis that the excess is due to
    inaccuracies in our simulation of depth dependences, we
    repeated the analysis on data from the shallow (COGZ
    > 0 m) part of the detector and from the deep (COGZ <
    0 m) part separately. Fig. 1 shows the COGZ distribution
    as a function of cosine zenith for the data, atmospheric
    neutrino simulation, and a subtraction of the simulation
    from data. To eliminate any bias from hard components
    like prompt neutrinos or extra-terrestrial neutrinos, we
    set an additional energy cut NCh < 50 to plot Fig.
    1. Fig. 2c indicates that the systematic problems are
    not specific to the highest energy events. Using events
    with COGZ > 0 m and cosine zenith less than -0.2, the
    data and simulation agrees relatively well. We performed
    the same procedures on the full dataset and no data
    excess is observed in any of the energy estimators. This
    result could be compared with the AMANDA diffuse
    analysis [12] because the majority of hits are recorded
    by DOMs at depths where AMANDA is deployed.
    Considering the sensitivities listed in Table II, this result
    is consistent with the current upper limit for diffuse
    muon neutrinos 7.4×10
    −8
    GeV cm
    −2
    s
    −1
    sr
    −1
    . On the
    other hand, at COGZ < 0 m with the same zenith cut, we
    observed an event excess with three energy estimators.
    Since the sensitivities of the lower COGZ sample are
    worse than the upper COGZ events, the event excess
    we observed with the full data set is highly likely due
    to systematics. Table II summarizes all numbers obtained
    from the two subsets.
    Some of the systematics issues will be resolved with
    ongoing calibration studies. Our description of the op-
    tical ice properties has larger uncertainties in the deep
    ice, where we so far have relied on extrapolations of
    the AMANDA measurements in the shallower ice [1],
    using measurements of dust concentration in Antarctic
    ice cores for the extrapolation. The ice core data indicate
    a strong improvement in ice clarity below AMANDA
    depths, with an estimated increase in average scatter-
    ing and absorption lengths of up to 40% at depths
    greater than 2100 m. With such different ice properties
    in the two parts of the detector, we are investigating
    our possibly increased sensitivity to systematic error
    sources that are present at AMANDA depths but become
    more significant in the deeper, clearer ice. We are also

    4
    K.HOSHINA et al. SEARCH FOR DIFFUSE NEUTRINOS WITH ICECUBE
    TABLE I
    SENSITIVITIES OF ICECUBE 22 STRINGS 275.7 DAYS WITH VARIOUS ENERGY ESTIMATORS. NO SYSTEMATICS ERROR INCLUDED.
    Estimator
    MRF (sensitivity)
    Energy Threshold
    Mean Background
    Mean Signal
    Data observed
    NCh
    0.22 (2.2 × 10
    −8
    E
    −2
    )
    NCh ≥ 99
    9.3
    29.4
    22
    NPe
    0.26 (2.6 × 10
    −8
    E
    −2
    )
    log10(NPe) ≥ 3.15
    6.6
    22.5
    10
    dEdx
    0.25 (2.5 × 10
    −8
    E
    −2
    )
    log10(dEdx) ≥ 1.4
    4.1
    19.8
    4
    MuE
    0.24 (2.4 × 10
    −8
    E
    −2
    ) log10(MuE) ≥ 5.05
    6.4
    28.4
    13
    TABLE II
    SENSITIVITIES OF ICECUBE 22 STRINGS 275.7 DAYS WITH ADDITIONAL COGZ CUT AND COSINE ZENITH CUT (COSθ < -0.2). NO
    SYSTEMATICS ERROR INCLUDED.
    Estimator
    COGZ cut
    MRF (sensitivity)
    Energy Threshold
    Mean Background
    Mean Signal
    Data observed
    NCh
    COGZ>0
    0.41 (4.1 × 10
    −8
    E
    −2
    )
    NCh ≥ 68
    7.9
    15.0
    3
    NPe
    COGZ>0
    0.54 (5.4 × 10
    −8
    E
    −2
    )
    log10(NPe) ≥ 2.85
    8.0
    11.3
    5
    dEdx
    COGZ>0
    0.50 (5.0 × 10
    −8
    E
    −2
    ) log10(dEdx) ≥ 0.97
    7.9
    12.2
    5
    MuE
    COGZ>0
    0.50 (5.0 × 10
    −8
    E
    −2
    )
    log10(MuE) ≥ 4.65
    9.9
    13.2
    7
    NCh
    COGZ<0
    0.47 (4.7 × 10
    −8
    E
    −2
    )
    NCh ≥ 80
    12.8
    15.7
    25
    NPe
    COGZ<0
    0.64 (6.4 × 10
    −8
    E
    −2
    )
    log10(NPe) ≥ 3.15
    2.4
    6.4
    4
    dEdx
    COGZ<0
    0.58 (5.8 × 10
    −8
    E
    −2
    ) log10(dEdx) ≥ 0.91
    15.5
    14.0
    14
    MuE
    COGZ<0
    0.62 (6.2 × 10
    −8
    E
    −2
    )
    log10(MuE) ≥ 5.00
    2.9
    7.1
    6
    cosine zenith
    -1 -0.9 -0.8 -0.7 -0.6 -0.5 -0.4 -0.3 -0.2 -0.1 0
    COGZ
    -500
    -400
    -300
    -200
    -100
    0
    100
    200
    300
    400
    500
    20
    40
    60
    80
    100
    120
    140
    160
    180
    200
    (a) Data
    cosine zenith
    -1 -0.9 -0.8 -0.7 -0.6 -0.5 -0.4 -0.3 -0.2 -0.1 0
    COGZ
    -500
    -400
    -300
    -200
    -100
    0
    100
    200
    300
    400
    500
    20
    40
    60
    80
    100
    120
    140
    160
    180
    200
    (b) Atmospheric ν
    cosine zenith
    -1 -0.9 -0.8 -0.7 -0.6 -0.5 -0.4 -0.3 -0.2 -0.1 0
    )
    ν
    data-simulation(Atms.
    -500
    -400
    -300
    -200
    -100
    0
    100
    200
    300
    400
    500
    0
    50
    100
    150
    200
    (c) Data - Atmospheric ν
    Fig. 2. (a,b) : Number of Low NCh events at final cut level in COGZ vs cosine zenith. Events contributing to the plot are limited to NCh <
    50. (c): Subtraction of plots (a) and (b). The boxes checked with x represent negative values.
    improving our photon propagation simulation to better
    reproduce the data in the clearest ice. This improved sim-
    ulation will be tested with data from in-situ light sources
    (LED flashers, nitrogen lasers) and well-reconstructed
    downward going muons.
    Among the four energy estimators, dEdx shows the
    most stable results. However, all the systematic prob-
    lems must be understood before we proceed to claim
    a physics result. The IceCube 22 string configuration is
    the first detector which allows a detailed study of Monte-
    Carlo simulation and the detector in the deep ice with
    reasonable statistics. These results will be essential not
    only for this analysis, but also for upcoming analysis
    with the IceCube 40 string configuration.
    V. CONCLUSION
    Using 275.7 days of upward going muon events
    collected by the IceCube 22 string configuration, we
    performed a search for a diffuse flux of high energy ex-
    traterrestrial muon neutrinos. The expected sensitivities
    are around 2.5×10
    −8
    GeV cm
    −2
    s
    −1
    sr
    −1
    for an E
    −2
    flux using four different energy estimators. We observed
    an excess of data over that expected from background
    above the best energy cut with some energy estimators.
    In order to test the geometric stability of this analysis, we
    performed the same analysis using two subsets of data
    divided by a threshold COGZ = 0 m. Having inconsistent
    results between these two subsets, the data excess we
    observed is highly likely dominated by systematics. With
    events at COGZ > 0 m, we observed no data excess with
    any of the energy estimators, which is consistent with the
    current upper limit on a diffuse flux of muon neutrinos
    obtained by the AMANDA diffuse analysis [12]. Many
    ongoing calibration studies will reveal the unknown
    systematics in the near future.
    REFERENCES
    [1] M Ackermann, et al. Journal of Geophysical Research, Vol. 111,
    D13203 (2006).
    [2] E. Waxman and J. Bahcall, Phys. Rev. D 59, 023002 (1998).
    [3] G.D. Barr, T.K. Gaisser, S. Robbins, and T. Stanev, Phys. Rev.
    D 74, 094009 (2006).
    [4] G. Firoentini, A. Naumov, and F.L Villante, Phys. Lett. B 510,
    173 (2001).
    [5] E.V. Bugaev et al. , Il Nuovo Cimento 12C, No. 1, 41 (1989).
    [6] G.C. Hill and K. Rawlins, Astropart. Phys. 19, 393 (2003).
    [7] J. Ahrens et al. , Nucl. Inst. Meth. A 524, 169 (2004)
    [8] Lundberg, J.; Miocinovic, P. Woschnagg, K. et al. Nucl. Inst.
    Meth. A 581, 619 (2007).
    [9] S. Grullon et al. ”Reconstruction of high energy muon events in
    icecube using waveforms,” in Proc. of 30th ICRC (2007).
    [10] J.D. Zornoza, D. Chirkin, ”Muon energy reconstruction and
    atmospheric neutrino spectrum unfolding with the IceCube de-
    tector”, in Proc. of 30th ICRC (2007).
    [11] G. Feldman and R. Cousins Phys. Rev. D 57 3873 (1998), G. C.
    Hill Phys. Rev. D 118101 (2003).
    [12] A. Achterberg et al., Phys. Rev. D, 76, 042008 (2007).

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    1
    A Search For Atmospheric Neutrino-Induced Cascades with
    IceCube
    Michelangelo D’Agostino
    for the IceCube Collaboration
    Department of Physics, University of California, Berkeley, CA 94720, USA
    See the special section of these proceedings.
    Abstract
    . The IceCube detector is an all-flavor
    neutrino telescope. For several years IceCube has
    been detecting muon tracks from charged-current
    muon neutrino interactions in ice. However, IceCube
    has yet to observe the electromagnetic or hadronic
    particle showers or “cascades” initiated by charged-
    or neutral-current neutrino interactions. The first
    detection of such an event signature will likely come
    from the known flux of atmospheric electron and
    muon neutrinos. A search for atmospheric neutrino-
    induced cascades was performed using a full year
    of IceCube data. Reconstruction and background
    rejection techniques were developed to reach, for the
    first time, an expected signal-to-background ratio
    1
    or better.
    Keywords
    : atmospheric, neutrino, IceCube
    IceCube is a cubic kilometer neutrino telescope cur-
    rently under construction at the geographical South Pole.
    With 59 of 86 strings of photomultiplier tubes currently
    embedded into Antarctica’s deep glacial ice, IceCube is
    already the world’s largest neutrino detector [1].
    IceCube detects high energy neturinos by observing
    Cherenkov light from the secondary particles produced
    in neutrino interactions in ice. In charged-current
    ν
    µ
    interactions, the outgoing energetic muon emits light
    along its track through the detector. A hadronic particle
    shower or cascade is also produced at the neutrino
    interaction vertex, but this is usually well outside of
    the instrumented detector volume. In charged-current
    ν
    e
    interactions, the outgoing electron initiates an electro-
    magnetic (EM) cascade which accompanies the hadronic
    cascade. Neutral-current interactions of any neutrino
    flavor produce hadronic cascades.
    At the energies relevant for atmospheric neutrinos,
    both hadronic and EM cascades develop over lengths of
    only a few meters. In a sparsely instrumented detector
    like IceCube, they look like point sources of Cherenkov
    light whose spherical wavefronts expand out into the
    detector. While muon tracks have been detected by
    neutrino telescopes, cascade detection has remained an
    elusive goal for high energy neutrino astrophysics.
    The well-studied atmospheric neutrino flux can serve
    as a calibration source for the cascade detection channel
    and should provide a valuable proof-of-principle for all-
    flavor detection. Once neutrino-induced cascades have
    been detected from the atmosphere, they should also
    open up a powerful channel for astrophysics analysis.
    Cut
    0.4
    0.5
    0.6
    0.7
    0.8
    0.9
    1
    1.1
    1.2
    N
    -2
    10
    -1
    10
    1
    10
    2
    10
    3
    10
    4
    10
    Atm.
    ν
    e
    +
    ν
    μ
    Cascades (MC)
    Data (10%)
    IceCube Preliminary
    Final Cut Variable Cumulative Distribution
    Fig. 1. Number of surviving events as a function of final cut strength
    for signal Monte Carlo and a 10% sample of the full one year dataset.
    Since cascades are topologically distinct from muons,
    they can be separated from the cosmic ray background
    over the entire
    of the sky [2].
    The challenge of separating a cascade signal from
    the overwhelming background of downgoing air-shower
    muons is significant. In its 22 string configuration,
    ∼10
    billion events triggered the IceCube detector in one year
    of operation. Of these, only
    ∼10
    ,000 are expected to
    be atmospheric neutrino-induced cascades. Because the
    atmospheric
    ν
    µ
    and
    ν
    e
    fluxes differ [3], these
    ∼10
    ,000
    events are unequally distributed among the different
    cascade signal classes. For each
    ν
    e
    , we expect
    ∼1
    .3
    ν
    µ
    neutral-current events and
    ∼2
    .9
    ν
    µ
    charged-current
    events where the hadronic cascade from the interaction
    vertex is inside the detector (so-called “starting events”).
    To begin the analysis, a fast filter was developed
    to run online at the South Pole to select promising
    candidate events for satellite transmission to the northern
    hemisphere. The filter selected events with a spheri-
    cal topology that were not good fits to relativistically
    moving tracks. After this online filter, each event was
    reconstructed according to track and cascade hypotheses
    using hit timing information, and well-reconstructed
    down-going tracks were thrown out.
    A new, analytic energy reconstruction method for
    cascades was developed that takes into account the
    significant depth variation of the optical properties of
    the glacial ice at the South Pole [4]. Several more
    topological variables with good separation power were
    also calculated for each event.
    The main background for neutrino-induced cascade
    searches comes from the stochastic energy losses suf-
    fered by cosmic ray muons as they pass through the ice
    surrounding the optical sensors. Two basic variables are

    2
    M. D’AGOSTINO
    et al.
    ATMOSPHERIC NEUTRINO-INDUCED CASCADES
    log(E) [GeV]
    3
    3.5
    4
    4.5
    5
    5.5
    6
    -4
    10
    -3
    10
    -2
    10
    -1
    10
    Atm.
    ν
    e
    +
    ν
    e
    Atm.
    ν
    μ
    +
    ν
    μ
    NC
    Atm.
    ν
    μ
    +
    ν
    μ
    CC
    True
    ν
    Surface Energy
    Fig. 2. Monte Carlo distributions of true neutrino energy at the earth’s
    surface for events surviving a final variable cut value of 0.73.
    employed to reduce this background. First, we measure
    how far inside the geometric volume of the detector
    the reconstructed cascade vertex lies. Muons with a
    large stochastic energy loss far inside the detector are
    more likely to leave early hits in outer sensors and
    can thus be rejected. Second, background separation
    becomes easier as the cascade energy increases. This is
    because the more energetic stochastic losses that mimic
    neutrino-induced cascades would have to come from
    more energetic muons, which are more likely to leave
    additional light that will allow for their identification.
    We therefore expect that more energetic cascades deep
    inside the detector will be the easiest signal to separate
    from background.
    Along these lines, several neural networks were
    trained on 12 topological and reconstruction-based vari-
    ables, including reconstructed energy and a measure of
    containment within the detector. The product of these
    variables is taken as the final discriminating cut variable.
    Figure 1 shows the number of remaining events as a
    function of the cut on this final variable for events that
    reconstruct above 5 TeV for signal Monte Carlo and a
    10% sub-sample of the available data.
    While nothing can yet be concluded from the 10%
    data sample alone, the full dataset, which will be pre-
    sented in this talk, may show signs of converging to
    the signal expectation. Figure 2 shows the true neutrino
    energy at the earth’s surface for the three classes of
    simulated signal.
    REFERENCES
    [1] A. Karle for the IceCube Collaboration, “IceCube: Construction
    Status and First Results,”
    arXiv:0812.3981
    , Dec. 2008.
    [2] A. Achterberg et al., “Search for neutrino-induced cascades from
    gamma-ray bursts with amanda,”
    The Astrophysical Journal
    , vol.
    664, no. 1, pp. 397–410, 2007.
    [3] G. D. Barr, T. K. Gaisser, P. Lipari, S. Robbins, and T. Stanev,
    “Three-dimensional calculation of atmospheric neutrinos,”
    Physi-
    cal Review D
    , vol. 70, no. 2, p. 023006, Jul. 2004.
    [4] E. Middell and M. D’Agostino, “Improved reconstruction of
    cascade-like events in icecube,”
    These Proceedings
    , 2009.

    PROCEEDINGS OF THE 31
    st
    ICRC, !OD´ Z´ 2009
    1
    First search for extraterrestrial neutrino-induced cascades with
    IceCube
    Joanna Kiryluk
    for the IceCube Collaboration
    Lawrence Berkeley National Laboratory and University of California Berkeley, Berkeley, CA 94720, USA
    see special section of these proceedings
    Abstract
    . We report on the first search for extra-
    terrestrial neutrino-induced cascades in IceCube.
    The analyzed data were collected in the year 2007
    when 22 detector strings were installed and oper-
    ated. We will discuss the analysis methods used to
    reconstruct cascades and to suppress backgrounds.
    Simulated neutrino signal events with a
    E
    −2
    energy
    spectrum, which pass the background rejection crite-
    ria, are reconstructed with a resolution
    ∆(log E) ∼
    0.27
    in the energy range from
    ∼ 20
    TeV to a few
    PeV. We present the range of the diffuse flux of
    extra-terrestrial neutrinos in the cascade channel in
    IceCube within which we expect to be able to put a
    limit.
    Keywords
    : extraterrestrial, neutrino, IceCube
    I. I
    NTRODUCTION
    IceCube is a 1 km
    3
    Cherenkov detector under con-
    struction at the South Pole. Its primary goals are to detect
    high energy extra-terrestrial neutrinos of all !avors in
    a wide energy range, from ∼100 GeV to ∼100 EeV,
    search for their sources, for example active galactic
    nuclei and gamma ray bursts, and to measure their
    diffuse !ux. When complete, the IceCube detector will
    be composed of 4800 Digital Optical Modules (DOMs)
    on 80 strings spaced 125 m apart. In addition there will
    be 6, more densely populated, Deep Core strings inside
    the IceCube detector volume. The array covers an area
    of one km
    2
    at depths from 1.45 to 2.45 km below the
    surface [1].
    High energy neutrinos are detected by observing the
    Cherenkov radiation from secondary particles produced
    in neutrino interactions inside or near the detector.
    Muon neutrinos in charged current (CC) interactions are
    identi"ed by the "nal state muon track [2]. Electron and
    tau neutrinos in CC interactions, as well as all !avor
    neutrinos initiating neutral current (NC) interactions
    are identi"ed by observing electromagnetic or hadronic
    showers (cascades). A 10 TeV cascade triggers IceCube
    optical modules out to a radius of about 130 m. Cascades
    can be reconstructed with good energy resolution, but
    limited pointing resolution. The good energy resolution
    and low background from atmospheric neutrinos make
    cascades attractive for diffuse extraterrestrial neutrino
    searches [3].
    We present expected sensitivities for the diffuse !ux
    of extra-terrestrial neutrinos in the cascade channel in
    IceCube. This work uses data collected in 2007 with
    the 22 strings that were deployed in IceCube at that
    time. The total livetime amounts to 270 days. Ten per
    cent of the data were used as a ”burn” sample to
    develop background rejection criteria. The results, after
    unblinding, will be based on the remaining 90% of the
    data, about 240 days.
    II. D
    ATA AND ANALYSIS
    Backgrounds from atmospheric muons, produced in
    interactions of cosmic rays with nuclei in the Earth’s
    atmosphere form a considerable complication in all neu-
    trino searches in IceCube. A "ltering chain developed
    using Monte Carlo simulations of muon background and
    neutrino signal was used to reject these backgrounds
    online and of!ine.
    The atmospheric muon background was simulated
    with CORSIKA [4]. In addition to the single muon
    events, which form the dominant background, an appro-
    priate number of overlaying events was passed through
    the IceCube trigger and detector simulator to obtain
    a sample of coincident muons. The coincident muon
    events make a few per cent contribution to the total
    trigger rate. The signal, electron neutrino events, was
    simulated using an adapted version of the Monte Carlo
    generator ANIS [5] for energies from 40 GeV to 1 EeV
    and with a E
    −2
    energy spectrum.
    All estimates for the number of signal events later in
    the text assume an E
    −2
    spectrum and !ux strength of:
    Φ
    model
    = 1.0 × 10
    −6
    (E/GeV)
    −2
    /(GeV s sr cm
    2
    ). (1)
    A. Online filtering
    The main physics trigger is a ”simple multiplicity
    trigger” (SMT), requiring photon signals in at least 8
    DOMs, with the additional requirement of accompany-
    ing hits in any of the ±2 neighboring DOMs, each above
    a threshold of 1/6 single photoelectron signal and within
    a 5 μs coincidence window. Averaging over seasonal
    changes of the trigger rate for IC22 was 550 Hz. The
    mean SMT rate is generally well reproduced by Monte
    Carlo simulation, which gives 565 Hz. Assuming the
    !ux given in Eq. 1, approximately 2.7 × 10
    3
    electron
    neutrino events and ∼ 1 × 10
    10
    background event are
    expected to trigger the detector in 240 days.
    The backgrounds are suppressed online with "rst-
    guess reconstruction algorithms [6]. A "rst guess track
    "t assumes that all hits can be projected onto a line,
    and that a particle producing those hits travels with
    velocity v
    line
    . In addition a simple cut on sphericity

    2
    JOANNA KIRYLUK
    et al.
    EXTRATRRESTRIAL CASCADES WITH ICECUBE
    Fig. 1.
    The reconstructed center-of-gravity (COG) x after online
    "ltering. Data is shown as continuous lines, background Monte Carlo
    is shown as dashed lines. Monte Carlo data is normalized to the
    experimental number of events.
    of the events (EvalRatio
    ToI
    ) is used to select events
    with hit topology consistent with cascades. Cut values
    used in online "lter are given in Table I. In the case of
    cascades, the online "lter reduced the SMT trigger rate
    to ∼ 20 Hz, or 3.5 % of the total trigger rate. Monte
    Carlo studies show that the "lter retains about 70% of
    the simulated signal and rejects 97.5% of the simulated
    background that trigger the detector. The Monte Carlo
    simulation thus underestimates the overall rate observed
    in the data. Otherwise main characteristics are well
    reproduced, Fig. 1 which shows the reconstructed center-
    of-gravity (COG) x position. The COG is calculated for
    each event as the signal amplitude weighted mean of all
    hit DOM positions.
    B. Offline filtering
    The data, after online "ltering and transfer from the
    South Pole, were passed through more sophisticated
    algorithms to reconstruct both muon tracks and cascades.
    This reconstruction uses the maximum-likelihood recon-
    struction algorithms described in [2], [6].
    Several cuts were applied sequentially, and the inter-
    mediate data sets are identi"ed as different levels. Level-
    1 is the trigger level and events passing the online
    "ltering correspond to Level-2. The rates at different
    levels are summarized in Table I.
    At Level-3 events were selected with (i) a recon-
    structed track zenith angle greater than 73
    and (ii)
    a difference Llh(track)-Llh(cascade) > −16.2 in the
    likelihood parameters of the track and cascade reon-
    structions to select cascade-like events. This selection
    was optimized for the combined ef"ciency (∼ 80%)
    in both atmospheric[7] and extra-terrestrial neutrino
    searches and keeps the data at this level common to
    both analyses. At Level-4 we require that all cascades
    originate inside the detector. In IceCube many muon
    tracks that radiate energetic bremsstrahlung or produce
    hits in DOMs close to the detector edges can mimic
    uncontained cascades. To remove this background of
    partial bright muon events we require that the four
    earliest hits in the event are inside the "ducial volume
    of the detector. The boundaries of the "ducial volume
    in x-y are shown in Fig.2 as continuous lines. In the
    z direction only an upper boundary was used. It was
    set at the position of the 8th DOM from the top.
    Approximately 1% of the background events (data and
    Monte Carlo) and ∼ 13% of the Monte Carlo signal
    events after online "ltering pass Level-3 selections and
    satisfy the "ducial volume requirement.
    At Level-5 we require that the number of hit DOMs
    (NCh) is greater than 20, that the reconstructed track
    zenith angle exceeds 69
    , and that the event duration,
    de"ned as a time difference between the last and "rst
    hit DOM, is less than 5 μs. The later cut removes long
    events, which are mostly coincident double or triple
    muon events typically with a high multiplicity of hit
    Fig. 2. a) The y versus x positions of the strings in the IC22 detector
    con"guration. b) The reconstructed center-of-gravity (COG) y versus
    x position from IC22 data after online "ltering. The continuous lines
    show the boundaries of the "ducial volume, which is used in the
    analyses to restrict the position of the "rst hits in the event.

    PROCEEDINGS OF THE 31
    st
    ICRC, !OD´ Z´ 2009
    3
    Fig. 3. The "ll-ratio versus the distance D (de"ned in the text) for signal Monte Carlo (left) muon background Monte Carlo (middle) and
    the data (right) for events with COG-z > −100 m (top) and COG-z < −100 m (bottom) The dashed lines show the background cut at level
    7 used in the analysis.
    DOMs.
    At Level-6 we require that the reconstructed cascade
    vertex positions x(y) and COG-x(y) agree to within 60
    meters, and that the reduced track and cascade recon-
    struction likelihood ratio is less than 0.95. For each event
    we apply the two track reconstruction algorithm and
    require that the reconstructed tracks coincide to within
    1μs. This selection mostly removes background events
    with coincident muon tracks which are well separated
    in time.
    At Level-7 stringent selections are made on the DOM
    multiplicity and the "ll-ratio. The "ll-ratio quanti"es
    the fraction of hit DOMs within a sphere around the
    reconstructed cascade vertex position with a radius
    2 × D, where D is the average displacement of the
    reconstructed cascade vertex with respect to the positions
    of the hit DOMs in the event. The "ll-ratio versus the
    distance D for signal Monte Carlo, muon background
    Monte Carlo, and the data for events with COG-z >
    −100m and COG-z < −100 m is shown in Fig.3. The
    presently used version of background Monte Carlo is in
    good agreement with the data for the top part of the
    detector, but not for the bottom part of the detector.
    In the bottom part of the detector, the clear ice (less
    absorption than at the top of the detector) makes some
    muons look like cascade (spherical shape and high DOM
    multiplicity). After applying the cuts on the "ll-ratio
    and the distance D, as shown by the dashed lines in
    Fig.3 , 135 events from the data burn sample and 11
    background Monte Carlo events remained. Almost all
    of them originate in the bottom part of the detector, as
    shown in Fig. 3. Remaining 11 Monte Carlo background
    events correspond to an expected ∼ 90 events for the 240
    days of the IC22 run.
    We placed a "nal Level-8 cut on the reconstructed
    energy, log E
    reco
    > 4.2, which rejects all remaining
    background events in the data burn sample and in the
    background Monte Carlo.
    III. R
    ESULTS
    The
    expected
    number
    of
    signal
    events
    (NSignal) from a diffuse !ux with a strength of
    10
    −6
    (E/GeV)
    −2
    /(GeV . s . sr . cm
    2
    ) is 52 ν
    e
    events
    for 240 days of livetime. Signal simulations show that
    events that pass all background rejection criteria are in
    the energy range from ∼ 20 TeV to a few PeV (with
    a mean energy of ∼ 160 TeV). The energy resolution
    is ∆(log E) ∼ 0.27, the x and y position resolution is
    ∼ 10 meters. The z position resolution is worse, 25 m,
    because of a small fraction of events that originated
    below the detector where no "ducial volume cut was
    applied.
    A burn sample of ∼ 10% of the total IC22 data set
    and the background Monte Carlo sample were used in
    developing background rejection criteria. The selections
    are such that all events in the burn sample and all
    background Monte Carlo events are rejected, whereas
    a signi"cant fraction of the signal Monte Carlo events
    are retained.
    The model rejection factor (MRF) de"ned as: MRF
    = ?μ
    90
    ? / NSignal, will be used to determine the !ux
    limit:
    Φ
    limit
    = MRF × f (E),
    (2)
    where f (E) is given by Eq.1.

    4
    JOANNA KIRYLUK
    et al.
    EXTRATRRESTRIAL CASCADES WITH ICECUBE
    TABLE I
    E
    VENT RATES AT DIFFERENT SELECTION LEVELS FOR EXPERIMENTAL DATA (BURN SAMPLE), ATMOSPHERIC MUONS BACKGROUND
    MONTE CARLO AND ν
    e
    SIGNAL MONTE CARLO ASSUMING THE FLUX Φ
    model
    = 1.0 × 10
    −6
    (E/GeV)
    −2
    /(GeV . s . sr . cm
    2
    )
    Level
    Selection Criteria
    Exp Data
    Tot Bg MC
    Signal MC ν
    e
    1
    Trigger
    580 Hz
    565 Hz
    2.7 × 10
    3
    × (240 days)
    −1
    2
    v
    line
    < 0.25 and EvalRatio
    ToI
    > 0.109
    20 Hz
    14 Hz
    1.8 × 10
    3
    × (240 days)
    −1
    3
    Zenith > 73
    and Llh(track) - Llh(cascade)> −16.2
    4 Hz
    2.8 Hz
    1.3 × 10
    3
    × (240 days)
    −1
    4
    Fiducial Volume (Fig.2)
    0.3 Hz
    0.15 Hz
    240× (240 days)
    −1
    5
    NCh > 20 and Zenith
    32iter
    > 69
    and EvtLength < 5μs
    0.02 Hz
    0.01 Hz
    165× (240 days)
    −1
    6
    |RecoX − COGX | < 60m and
    0.011 Hz
    0.004 Hz
    161× (240 days)
    −1
    |RecoY − COGY | < 60m and
    ReducedLlh(track) / ReducedLlh(cascade) > 0.95 and
    RecoTrack1(Time)-RecoTrack2(Time) < −1μs
    7
    Fill-Ratio (Fig.3)
    6.8 × 10
    −5
    Hz
    4.3 × 10
    −6
    Hz
    68× (240 days)
    −1
    8
    NCh> 60 and log E
    reco
    > 4.2
    0
    0
    52× (240 days)
    −1
    The analysis is limited by the currently available
    background Monte Carlo sample. It is not possible to
    subtract the simulated residual background contribution
    with suf"cient precision. Thus the sensitivities for the
    diffuse !ux of extraterrestrial neutrino signal, de"ned
    as the average upper limit at 90% CL and absence of
    signal [9], cannot be determined. To give an order of
    magnitude for the limit, a conservative estimate making
    no assumptions on background would be 4 × 10
    −8
    (5 ×
    10
    −7
    )(E/GeV)
    −2
    /(GeV . s . sr . cm
    2
    ) for a hypotheti-
    cal number of observed events after unblinding of 0 (20).
    Enclosing, we expect the !ux limit from
    this analysis to be of the same order as the
    limit on the diffuse !ux Φ
    limit
    =
    1.3 ×
    10
    −7
    (E/GeV)
    −2
    /(GeV . s . sr . cm
    2
    ) [10] in the
    cascade channel as obtained from 5 years of AMANDA
    data. Additional background Monte Carlo events
    are being generated and systematic uncertainties are
    currently being studied.
    R
    EFERENCES
    [1] IceCube, A. Achterberg
    et al.
    , Astropart. Phys.
    26
    , 155 (2006);
    R. Abbasi et al. Nucl. Instrum. Meth.
    A601
    (2009) 294.
    [2] AMANDA , J. Ahrens
    et al.
    , Nucl. Instrum. Meth.
    A524
    (2004)
    169.
    [3] M. Kowalski, JCAP 05 (2005) 010.
    [4] D. Heck
    et al.
    , Tech. Rep. FZKA, 6019 (1998).
    [5] A. Gazizov and M. Kowalski, Computer Physics Communica-
    tions Vol. 172
    3
    (2005) 203.
    [6] AMANDA, J. Ahrens
    et al.
    , Phys. Rev.
    D67
    (2003) 012003.
    [7] M. D’Agostino (for the IceCube Collaboration),
    A search for
    atmospheric neutrino-induced cascades with IceCube
    , these pro-
    ceedings.
    [8] IceCube, J. Ahrens
    et al.
    , IceCube Preliminary Design Document
    (2001).
    [9] G. Feldman and R. Cousins, Phys. Rev.
    D57
    (1998) 3873.
    [10] O. Tarasova, M. Kowalski and M. Walter (for the IceCube
    Collaboration), proceedings of the 30th International Cosmic Ray
    Conference (ICRC 2007), Merida, Yucatan, Mexico, 3-11 Jul
    2007; arXiv:0711.0353 [astro-ph] pages 83-86.

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    1
    Improved Reconstruction of Cascade-like Events in IceCube
    Eike Middell
    , Joseph McCartin
    and Michelangelo D’Agostino
    §
    for the IceCube Collaboration
    DESY, D-15735 Zeuthen, Germany
    Institut fu¨r Physik, Humboldt-Universita¨t zu Berlin, D-12489 Berlin, Germany
    Dept. of Physics and Astronomy, University of Canterbury, Private Bag 4800, Christchurch, New Zealand
    §
    Dept. of Physics, University of California, Berkeley, CA 94720, USA
    See the special section of these proceedings.
    Abstract
    . Cascade-like events are one of the main
    signatures in the IceCube neutrino detector. This sig-
    nature includes electromagnetic and hadronic parti-
    cle showers from charged or neutral current interac-
    tions and hence it provides sensitivity to all neutrino
    flavours. At energies below 10 PeV these cascades
    have characteristic lengths of only several meters.
    Compared to the dimensions of the detector they
    appear as point-like but anisotropic light sources.
    We present a new approach to the reconstruction of
    such events. A maximum likelihood algorithm that
    incorporates the results of detailed simulations of the
    light propagation in ice, allows for a significantly
    better analysis of the recorded photon intensities and
    arrival times. The performance of the algorithm is
    evaluated in a Monte Carlo study. It suggests that
    for cascades an angular resolution of
    30
    is possible.
    Keywords
    : IceCube, cascades, reconstruction
    I. INTRODUCTION
    The IceCube detector [1] is being built at the ge-
    ographical South Pole. It aims for the detection of
    neutrinos of cosmic origin, which could answer open
    questions in astroparticle physics such as the origin
    of cosmic rays and the nature of dark matter. In its
    originally planned setup the IceCube detector consists
    of 4800 digital optical modules (DOMs) on 80 strings.
    These are horizontally spaced by 125 m and located in
    depths ranging from 1
    .
    45 to 2
    .
    45 km, thereby spanning
    a volume of a cubic kilometer of glacial ice. In order
    to lower IceCube’s energy threshold down to 10 GeV,
    the DeepCore extension will arrange 6 additional strings
    in the center of the array. On these strings the DOMs
    are closer to each other and are located in depths with
    optimal optical properties.
    Each DOM contains a photomultiplier tube (PMT)
    and the necessary readout electronics. Two digitization
    devices allow for the measurement of time distributions
    and intensities of photon fluxes inside the detector: the
    Analog Transient Waveform Digitizer (ATWD) taking
    128 samples over the first 420 ns and the Flash Analog-
    to-Digital Converter (FADC) taking 256 samples in an
    interval of 6
    .
    4
    μ
    s [2]. Presently three quarters of the
    detector are successfully deployed and are taking data.
    Neutrinos can interact in the instrumented volume
    through neutrino-nucleon or neutrino-electron scattering.
    The former process dominates. One exception is the
    resonant scattering of anti-electron neutrinos on atomic
    electrons at energies of 6
    .
    3 PeV, known as the Glashow
    resonance. The neutrino interaction is not detected di-
    rectly but it can produce charged particles which emit
    Cherenkov light in the transparent detector medium. The
    possible final states of a neutrino interaction depend on
    the flavour and interaction type. For neutrino astronomy
    the most prominent neutrino signature is formed by final
    states with an emerging muon. They allow to deduce
    the neutrino direction and provide large effective areas
    because of the large range of the muon.
    The signatures of interest here are neutrino induced
    electromagnetic and hadronic particle showers. Such
    cascades can originate from all neutrino flavours and oc-
    cur in many of the interaction scenarios. Assuming that
    the neutrinos were generated in pion decays one expects
    a flavour ratio at the source of
    ν
    e
    :
    ν
    μ
    :
    ν
    τ
    =
    1 : 2 : 0.
    Due to neutrino oscillations this ratio is transformed to
    1 : 1 : 1 before detection, which makes the sensitivity to
    all flavours important.
    Furthermore, electromagnetic cascades allow for a
    good energy reconstruction, since the number of emit-
    ted photons scales linearly with the deposited energy.
    Hadronic cascades appear similar to electromagnetic
    ones, with the small correction that for the same de-
    posited energy there are about 20% fewer photons pro-
    duced [3].
    Below 10 PeV cascades have characteristic lengths of
    several meters. Compared to the distances between the
    DOMs they appear as point-like light sources. Nev-
    ertheless, the angular emission profile of a cascade
    is anisotropic: the photons originate from one point
    but they are preferably emitted in the direction of the
    Cherenkov angle
    Θ
    c
    =
    41
    [4]. Therefore, close to the in-
    teraction vertex the neutrino direction can in principle be
    derived from the angular distribution of the Cherenkov
    photons. For the large spacing of the DOMs this ability is
    impaired due to the strong light scattering in the ice [5].
    Because of this inherent difficulty of reconstructing the
    direction of particle showers in ice, studies of these
    events have been restricted to the search for a diffuse
    flux of neutrinos. In this situation even a rough estimate

    2
    E.MIDDELL
    et al.
    IMPROVED RECONSTRUCTION OF CASCADE-LIKE EVENTS IN ICECUBE
    on the neutrino detection would enhance the possibilities
    of this detection channel.
    II. NEW APPROACH TO CASCADE RECONSTRUCTION
    The existing maximum likelihood reconstruction for
    cascades [3] does not account for the inhomogeneity of
    the ice and does not try to reconstruct the neutrino
    direction. It also does not exploit all the capabilities of
    the IceCube DAQ.
    The aim of the current work is to use all relevant
    information in the waveforms captured by the DOMs
    to reconstruct the incident neutrino in a cascade-like
    event. The point-like but directed cascade can be fully
    described by 7 parameters: the time and vertex
    (
    t
    ,
    x
    ,
    y
    ,
    z
    )
    of the neutrino interaction, the deposited energy
    E
    and
    the direction of the neutrino. The latter is described
    by the two angles zenith
    Θ
    and azimuth
    φ
    . The re-
    construction searches for the set of these parameters
    c
    = (
    t
    ,
    x
    ,
    y
    ,
    z
    ,
    E
    ,
    Θ
    ,
    Φ)
    that fits the observation best.
    A good understanding of the optical properties of
    the glacial ice is crucial to the IceCube experiment.
    The instrumented volume is pervaded with dust layers
    that track historic climatological changes. Since the
    propagation of light in such an inhomogeneous medium
    cannot be treated analytically, the Photonics Monte Carlo
    package [6], [7] has been used. Its simulation results
    are available in tabulated form. For a given setup of
    a light source and a DOM these tables allow to make
    predictions for the mean expected amplitude
    ?
    μ
    (
    c
    )
    ?
    and
    the photon arrival time distribution
    p
    (
    t
    d
    ,
    c
    )
    , where
    t
    d
    denotes the delay time. For a photon with speed
    c
    ice
    that is emitted at
    (
    t
    e
    ,?
    x
    e
    )
    and recorded at
    (
    t
    r
    ,?
    x
    r
    )
    the time
    t
    d
    =
    t
    r
    t
    e
    |
    ?
    x
    r
     ?
    x
    e
    |
    /
    c
    ice
    denotes the additional time the
    photon takes to reach the receiver over a scattered path
    rather than a straight line. Scattering in ice can cause
    delay times up to a few microseconds. Depending on
    orientation and distance of the cascade with respect to
    the DOM the arrival time distributions differ in shape
    (compare Figure 1).
    With the tabulated quantities the expected amplitude
    in a time interval
    [
    t
    1
    ,
    t
    2
    ]
    calculates to:
    μ
    (
    c
    ) =
    f
    ?
    μ
    (
    c
    )
    ?
    Z
    t
    2
    t
    1
    p
    (
    t
    d
    ,
    c
    )
    dt
    d
    +
    R
    noise
    (
    t
    2
    t
    1
    )
    (1)
    Two small corrections are applied to the prediction of
    Photonics. A constant rate
    R
    noise
    accounts for noise
    hits and a factor
    f
    corrects for deviations from the
    mean amplitude due to the PMT response and charge
    reconstruction, which is not modelled by Photonics.
    With this prediction a likelihood description of the
    measurement is possible. Assuming a Poisson process
    for every distinct
    1
    sample
    i
    taken by the ATWD and
    the FADC in DOM
    o
    , one can compare the measured
    amplitude
    n
    oi
    to the mean expectation
    μ
    oi
    and construct
    the likelihood:
    L
    =
    o
    ,
    i
    μ
    oi
    (
    c
    )
    n
    oi
    n
    oi
    !
    exp
    {
    μ
    oi
    (
    c
    )
    }
    .
    (2)
    0
    500
    1000
    1500
    2000
    2500
    3000
    3500
    4000
    delay time [ns]
    0.0000
    0.0005
    0.0010
    0.0015
    0.0020
    dP/dt [1/ns]
    distance= 100mdistance= 100m
    toward
    backward
    0
    500
    1000
    1500
    2000
    2500
    3000
    3500
    4000
    delay time [ns]
    0.0000
    0.0001
    0.0002
    0.0003
    0.0004
    0.0005
    0.0006
    0.0007
    0.0008
    dP/dt [1/ns]
    distance= 300mdistance= 300m
    toward
    backward
    Fig. 1. Tabulated delay time distributions for a DOM at 100 m and
    300 m distance to the cascade. The distributions are shown for two
    orientations of the cascade, pointing either toward or away from the
    DOM. Photons are increasingly delayed if they either travel larger
    distances or have to be backscattered to reach the DOM.
    By taking the negative logarithm and rearranging the
    terms one obtains:
    log(
    L
    ) =
    o
    ?
    μ
    o
    ?
    n
    o
    log
    ?
    μ
    o
    ?
    i
    n
    oi
    log
    ?
    μ
    oi
    ?
    μ
    o
    ?
    ?
    (3)
    where
    ?
    μ
    o
    ?
    =
    i
    μ
    oi
    and
    n
    o
    =
    i
    n
    oi
    . The combinatorial
    term from the Poisson probability has been omitted since
    it does not depend on the reconstruction hypothesis.
    A considerable speedup in the computation results
    from the fact that in the sum over the samples
    i
    only time
    intervals with
    n
    oi
    >
    0 contribute. Hence, periods in the
    DOM readout with no measured charge can be ignored.
    Practically this is implemented in two steps: first the
    waveform is scanned for pulses, then these pulses are
    used to calculate the likelihood.
    The cascade reconstruction is performed by searching
    numerically for the minimum of
    log(
    L
    )
    , which is a
    function of the seven cascade parameters. This mini-
    mization is seeded with the time, vertex and direction
    estimates that one obtains from calculating the center
    of gravity and tensor of inertia of the hit pattern. These
    calculations are implemented in IceCube’s first-guess re-
    construction algorithms. The number of triggered DOMs
    provides a rough estimate of the deposited energy. The
    minimization is done by MINUIT with a simplex algo-
    rithm that is executed iteratively to improve the result
    stepwise.
    The problem can be significantly simplified if the
    vertex and the time of the interaction are already known
    (e.g. when they are determined by another method) and
    the orientation of the cascade is neglected. Then the
    likelihood, which now only depends on the cascade
    energy, provides an energy reconstruction that benefits
    1
    In the first 420 ns the readout windows of the ATWD and FADC
    overlap. One has to choose between both measurements. Because of
    its precision, the samples from the ATWD are preferred.

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    3
    60
    40
    20 0 20 40 60
    x
    reco
    x
    true
    [m]
    0
    500
    1000
    1500
    2000
    =0.18m
    =6.16m
    60
    40
    20 0 20 40 60
    z
    reco
    z
    true
    [m]
    0
    500
    1000
    1500
    2000
    2500
    3000
    3500
    4000
    =0.17m
    =3.98m
    Fig. 2.
    Offsets between the reconstructed and the true
    x
    and
    z
    coordinates obtained from an iterative minimization of the 7 dimensional
    likelihood. Only cascades are selected, whose reconstructed vertex is contained in IC40. The width
    σ
    of a fitted Gaussian defines the resolution,
    which is better for
    z
    because of the denser DOM spacing along the string.
    4
    3
    2
    10 1
    2
    3
    4
    log
    10
    (E
    reco
    /E
    true
    )
    0
    1000
    2000
    3000
    4000
    5000
    6000
    =0.05
    =0.13
    2
    4
    6
    8 10
    log
    10
    (E
    true
    /GeV)
    2
    4
    6
    8
    10
    log
    10
    (
    E
    reco
    /GeV
    )
    Fig. 3. Left: Offset between the reconstructed and the deposited logarithmic energy for the same event sample. Right: Comparison between
    the reconstructed and the deposited logarithmic energy. The deviation from the identity line above 10 PeV illustrates the increasing impact of
    saturation effects on the energy reconstruction.
    from the improved light-propagation model. In this case,
    the search for the minimum is reduced to a numerical
    root finding problem:
    ∂(
    log(
    L
    ))
    E
    =
    o
    μ
    o
    n
    o
    1
    +
    R
    noise
    Δ
    t
    μ
    o
    ?
    =0
    (4)
    where
    Δ
    t
    denotes the readout window length.
    III. RESULTS
    The reconstruction algorithm has been tested with a
    simulated electron neutrino dataset for IceCube in its
    year 2008 configuration with 40 strings. The primary
    neutrinos have energies in the range from 10
    1
    .
    7
    GeV to
    10
    10
    GeV and are weighted to an
    E
    2
    spectrum. For
    the simulation of showers the parametrization derived
    in [4] and implemented in Photonics is used. Lower
    energetic showers (
    <
    PeV) are represented as point-like
    light source with an anisotropic emission profile. At PeV
    energies the cascade is split up into several cascades to
    simulate the elongation due to the LPM effect.
    To be part of the further on used event selection, an
    event has to trigger the detector, the reconstruction must
    converge (fulfilled by 79% ) and the reconstructed vertex
    has to be located inside the geometric boundaries of the
    detector (fulfilled by 38%).
    To evaluate the resolution of the reconstruction the
    distribution of offsets between the reconstructed and
    the true vertex coordinates and energies are shown in
    Figures 2 and 3. The obtained vertex resolutions are
    about 7 m in
    x
    and
    y
    and 4 m in
    z
    . This is an improvement
    with respect to the existing likelihood reconstruction [8].
    For the same dataset and selection criteria it yields
    resolutions of 15m in
    x
    and
    y
    and 8m in
    z
    . The better
    resolution in
    z
    results from the smaller distances of only

    4
    E.MIDDELL
    et al.
    IMPROVED RECONSTRUCTION OF CASCADE-LIKE EVENTS IN ICECUBE
    1.0
    0.5 0.0 0.5 1.0
    cos(
    )
    500
    1000
    1500
    2000
    2500
    med = 0.71
    2345678
    log
    10
    (E
    true
    )
    0
    10
    20
    30
    40
    50
    60
    70
    80
    90
    median
    (
    )[
    deg
    ]
    Fig. 4. Left: Distribution of the cosine of the angle between the reconstructed and the true direction. The angular resolution is given by the
    median. Right: Angular resolution as a function of the energy.
    17 m between the DOMs on one string.
    The result of the energy reconstruction is shown in
    Figure 3. A resolution of
    σ(log
    10
    (
    E
    reco
    /
    E
    true
    )) =
    0
    .
    13
    has been obtained. For large photon fluxes, which can
    originate from highly energetic or nearby cascades, the
    saturation of the PMT limits the recorded charge. This
    affects the energy reconstruction as can be seen in the
    right plot of Figure 3. Above 10 PeV the reconstructed
    energy is systematically too low due to the saturation.
    A useful measure for the angular resolution is the
    median of the cos(Ψ) distribution, where
    Ψ
    is the
    angle between the true and the reconstructed direction.
    For all events that fulfill the selection criterion this
    distribution is plotted in the left plot of Figure 4. A
    study of the energy dependence suggests that for the
    interesting energy range of 10 TeV to 10 PeV an angular
    resolution of 30
    35
    is possible (right plot in Figure
    4). At energies above 10 PeV, the LPM effect leads
    to an elongation of the cascade and the reconstruction
    hypothesis of a point-like light source becomes no longer
    applicable.
    IV. SUMMARY AND OUTLOOK
    A maximum likelihood reconstruction for cascade-like
    events has been developed. It takes into account the full
    recorded waveform information as well as the ice proper-
    ties. A simulation study for the 40 string detector geom-
    etry of the year 2008 demonstrates the feasibility of an
    angular resolution of down to 30
    . Compared to muons
    this is still a very limited precision, but it can provide
    new opportunities for neutrino searches with cascade-
    like events. With the angular resolution achieved, the
    discrimination between upward and downward going
    neutrinos becomes possible as well as the identification
    of neutrinos originating from the galactic plane. With the
    DeepCore extension a further improvement is expected.
    The achieved results have to withstand further ver-
    ification. The next step is to test the performance of
    the algorithm on measurements with LED and laser
    light sources in the detector and muon events with
    bright bremsstrahlung cascades. Several possibilities to
    enhance the algorithm exist. A different description of
    saturated DOMs in the likelihood could improve the
    performance at higher energies. It will be investigated
    if the shape of the likelihood could be used to estimate
    the error of the reconstruction. Finally, the presented ap-
    proach can be extended to reconstruct combined events
    with more than one light source in the detector.
    REFERENCES
    [1] A. Achterberg et al., Astropart.Phys.26:155-173 (2006)
    [2] R. Abbasi et al., Nucl.Instrum.Meth.A601:294-316 (2009)
    [3] M. Kowalski, PhD-Thesis, Humboldt University to Berlin (2004)
    [4] C. Wiebusch, PhD-Thesis, RWTH Aachen, PTHA 95/37 (1995)
    [5] M. Ackermann et al., J. Geophys. Res., 111, D13203 (2006)
    [6] http://photonics.sourceforge.net
    [7] J. Lundberg et al., Nucl.Instrum.Meth.A581:619-631 (2007)
    [8] J. Kiryluk et al., in Proc. of 30th ICRC (2007)

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    1
    Searches for neutrinos from GRBs with the IceCube 22-string
    detector and sensitivity estimates for the full detector
    A. Kappes
    ∗†
    , P. Roth
    , E. Strahler
    , for the IceCube Collaboration
    §
    Physics Dept. University of Wisconsin, Madison WI 53703, USA
    affiliated with Universita¨t Erlangen-Nu¨rnberg, D-91058 Erlangen, Germany
    Physics Dept. University of Maryland, College Park MD 20742, USA
    §
    see special section of these proceedings
    Abstract
    . This contribution presents results of
    searches with IceCube in its 22-string configuration
    for neutrinos from 41 stacked gamma-ray bursts
    (GRBs) detected in the northern sky by satellites like
    Swift. In addition, the capabilities of the full 80-string
    detector based on a detailed simulation are discussed.
    GRBs are among the few potential source classes for
    the highest energy cosmic rays and one of the most
    puzzling phenomena in the universe. In their ultra-
    relativistic jets, GRBs are thought to produce neutri-
    nos with energies well in excess of 100 TeV. However,
    up to now, no such neutrino has been observed.
    IceCube, currently under construction at the South
    Pole, is the first km
    3
    scale neutrino telescope. As
    such it will have a significantly improved sensitivity
    compared to the precursor class of 0.01 km
    3
    neutrino
    telescopes.
    Keywords
    : Gamma-Ray Bursts, Neutrinos, IceCube
    I. INTRODUCTION
    Gamma-ray Bursts (GRBs) have been proposed as a
    plausible source of the highest energy cosmic rays [1]
    and high energy neutrinos [2]. The prevalent belief is
    that the progenitors of so called
    long-soft
    GRBs are very
    massive stars that undergo core collapse leading to the
    formation of a black hole.
    Short-hard
    GRBs are believed
    to be the product of the merger of binary compact
    objects such as neutrons stars and black holes leading to
    the creation of a single black hole. Material is ejected
    from the progenitor in ultra-relativistic jets. In these jets,
    electrons and baryons are accelerated to high energies,
    where the synchrotron radiation from the electrons is
    observed as the prompt
    γ
    -ray signal. Neutrinos are
    predicted to be produced in the interaction of accelerated
    baryons with matter or photons in various phases of
    the GRB:
    TeV precursor
    —while the jet burrows through
    the envelope of the progenitor of a long-soft burst [3];
    PeV prompt
    —in coincidence with the observed
    γ
    -ray
    signal [2];
    EeV early afterglow
    —as the jet collides with
    interstellar material or the progenitor wind in the early
    afterglow phase [4].
    IceCube is a high energy (
    E ? 1
    TeV) neutrino tele-
    scope currently under construction at the South Pole [5].
    When completed, the deep ice component of IceCube
    will consist of 5160 digital optical modules (DOMs)
    arranged in 86 strings frozen into the ice, at depths
    ranging from 1450 m to 2450 m. Each DOM contains
    a photo-multiplier tube and supporting hardware inside
    a glass pressure sphere. The total instrumented volume
    of IceCube will be
    ∼ 1 km
    3
    . The DOMs indirectly detect
    neutrinos by measuring the Cherenkov light from sec-
    ondary charged particles produced in neutrino-nucleon
    interactions. Presently, 59 strings are installed and col-
    lect data continuously. Construction is scheduled for
    completion by 2011. AMANDA-II, IceCube’s prede-
    cessor array, operated between January 2000 and May
    2009. It consisted of 677 optical modules arranged on
    19 strings with an instrumented volume approximately
    60 times smaller than that of IceCube. Searches with
    AMANDA-II for neutrinos in coincidence with GRBs
    have been reported with negative results [6], [7].
    The two main channels for detecting neutrinos with
    IceCube are the muon and the cascade channels. Charged
    current interactions of
    ν
    µ
    produce muons that, at TeV
    energies, travel for several kilometers in ice and leave a
    track-like light pattern in the detector. The detectors are
    mainly sensitive to up-going muon neutrinos as the Earth
    can be used to shield against the much larger flux of (up-
    going) atmospheric muons. Searches for neutrinos from
    GRBs in the muon channel benefit from good angular
    resolution (
    ∼ 1
    for
    E
    ν
    > 1
    TeV) and from the long
    range of high energy muons. Therefore, we use this
    channel in our analyses.
    II. ICECUBE 22-STRING RESULTS
    In our analyses, we search the IceCube 22-string con-
    figuration data, collected between May 2007 and April
    2008, for muon neutrinos from GRBs in the northern
    hemisphere. In [8] further analyses using IceCube 22-
    string data are presented which extend the muon neutrino
    search to GRBs in the southern sky and use the cascade
    channel to search for neutrinos of all flavors from GRBs
    in both hemispheres, respectively.
    We perform our searches both in the prompt (defined
    by the observed
    γ
    -ray emission) and the precursor (100 s
    before the prompt time window) time windows. In
    order to account for alternative emission scenarios, an
    additional search is conducted in an extended window
    from
     1
    h to
    +3
    h around the burst. The data outside

    2
    A. KAPPES
    et al.
    GRB SEARCHES WITH ICECUBE
    E
    ν
    (GeV)
    3
    10
    4
    10
    5
    10
    6
    10
    7
    10
    8
    10
    9
    10
    )
    -2
    dN/dE (GeV cm
    ×
    2
    E
    -8
    10
    -7
    10
    -6
    10
    -5
    10
    -4
    10
    -3
    10
    -2
    10
    10
    -1
    1
    10
    2
    10
    3
    10
    4
    10
    41 individual bursts
    Sum of 41 individual bursts
    Average WB burst
    Sum of 41 WB bursts
    preliminary
    Fig. 1.
    Neutrino spectra for all 41 GRBs investigated in the
    analyses. The fluences were calculated for each burst individually using
    measured bursts parameters following [11]. For comparison, average
    Waxman-Bahcall GRB fluences (WB, [2]) are shown.
    these windows (off-time data) is used to estimate the
    background in the search windows.
    To prevent bias in our analyses, the data within the
     1
    h to
    +3
    h window (on-time data) are kept blind
    during optimization. Only low level quantities of the on-
    time data are examined in order to determine stability.
    The remaining, usable off-time data amounts to 269
    days of livetime. Of the 48 northern hemisphere bursts
    detected by satellites (mainly Swift [9]), 7 do not have
    quality IceCube data associated with them during the
    prompt/precursor emission windows. For all remaining
    41 GRBs, tests show no indications of abnormal behav-
    ior of the detector.
    As customary, we use the Waxman-Bahcall model as
    a benchmark for neutrino production in GRBs. The orig-
    inal calculation with this model [2] used average GRB
    parameters as measured by BATSE [10]. It was refined
    by including specific details for individual GRBs [11].
    Our neutrino calculations follow the latter prescription.
    For many GRBs the available information is incomplete.
    In that case we use average parameters in the modeling
    of the neutrino flux. The individual burst neutrino spectra
    are displayed in Fig. 1.
    Tracks are reconstructed using a log-likelihood recon-
    struction method [12]. A fit of a paraboloid to the region
    around the maximum in the log-likelihood function
    yields an estimate of the uncertainty on the reconstructed
    direction. Initially, candidate neutrino events are out-
    numbered (by several orders of magnitude) by down-
    going atmospheric muons that are mis-reconstructed as
    up-going events. Application of data selection criteria
    allows us to extract a high-purity sample of up-going
    (atmospheric, and potentially astrophysical) neutrinos.
    In order to determine our detector response to the
    expected GRB neutrinos, we simulate these signal events
    using ANIS [13]. Background from atmospheric muons
    is simulated with CORSIKA [14]. Propagation of neu-
    trinos and muons through the Earth and ice are per-
    formed with ANIS and MMC [15]. The photon signal
    at the DOMs is determined from a detailed simulation
    [16] of the propagation of Cherenkov light from muons
    and showers through the ice. The simulation of the DOM
    Fig. 2. The SVM classifier distribution of off-time data, simulated
    backgrounds (dashed), and simulated GRB muon neutrinos (solid).
    The lower frame displays the MDF resulting from a cut on the SVM
    classifier. The vertical dashed line shows the final tightened cut at 0.25.
    response takes into account the DOM’s angular accep-
    tance and includes a simulation of the DOM electronics.
    The DOM output is then processed with a simulation of
    the trigger. Afterwards, the simulated events are treated
    in the same way as the real data.
    A. Binned analysis
    We perform a binned analysis searching for emission
    during the prompt phase. After a loose preselection of
    events, various quality parameters are combined using a
    machine learning algorithm. The algorithm used was a
    Support Vector Machine (SVM) [17] with a radial basis
    function kernel. The SVM was trained using the off-
    time filtered data as background and all-sky neutrino
    simulation weighted to the sum of the individual burst
    spectra as signal. The optimum SVM parameters (kernel
    parameter, cost factor, margin) were determined using a
    coarse, and then fine, grid search with a 5-fold cross
    validation technique at each node [18].
    The resulting SVM classification of events is shown
    in Fig. 2. The final cut on this parameter is optimized to
    detect a signal fluence with at least 5
    σ
    (significance)
    in 50% of cases (power) by minimizing the Model
    Discovery Factor (MDF). The MDF is the ratio be-
    tween the signal fluence required for a detection with
    the specified significance and power and the predicted
    fluence [19]. The angular cut around each GRB is then
    calculated to keep 3/4 of the remaining signal after the
    cut on the SVM classifier. In this way, there is one
    cut on the SVM classifier for all GRBs, but different
    angular cuts around each GRB according to the angular
    resolution of the detector in that direction. The SVM cut
    that returns the best sensitivity is at 0.22. This cut lies
    directly on a discontinuity in the MDF curve, and so it
    is tightened away from that discontinuity so that a 1
    σ
    underestimation of the background level will not lead to
    a discovery claim more significant than is appropriate.
    B. Unbinned likelihood analysis
    We compare the performance of the binned analysis
    for the prompt emission to that of an unbinned likelihood

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    3
    analysis. Furthermore, we use the unbinned method to
    look for neutrino emission in the precursor and extended
    time window. The unbinned method used here is similar
    to that described in [20]. The signal,
    S (?x
    i
    )
    , and back-
    ground,
    B(?x
    i
    )
    , PDFs are formed from a product of a
    directional, time and an energy PDF.
    Signal PDF:
    The directional signal PDF is a two-
    dimensional Gaussian distribution with the two widths
    being the major and minor axes of the
    error ellipse
    of the paraboloid fit. The time PDF is flat over the
    respective time window and falls off on both sides with a
    Gaussian distribution of variable width depending on the
    duration of the emission. The energy PDF is determined
    from the distribution of an energy estimator [21] for each
    GRB individually. The signal PDFs of the GRBs are
    combined using a weighted sum [22]
    S
    tot
    (?x
    i
    )=
    ?
    N
    GRBs
    j=1
    w
    j
    S
    j
    (?x
    i
    )
    ?
    N
    GRBs
    j=1
    w
    j
    ,
    (1)
    where
    S
    j
    (?x
    i
    )
    is the signal PDF of the
    j
    th GRB and
    w
    j
    is a weight that for of the prompt and precursor window
    is proportional to the expected number of events in the
    detector according to the calculated fluences. In the case
    of the extended window we use
    w
    j
    = 1
    for all GRBs.
    Background PDF:
    For the directional background
    PDF obtained from the off-time data, the detector asym-
    metries in zenith and azimuth are taken into account by
    evaluating the data in the detector coordinate system.
    The time distribution of the background during a GRB
    can be assumed to be constant, yielding a flat time PDF.
    The energy PDF is determined in the same way as for
    the signal PDF with the spectrum corresponding to the
    Bartol atmospheric neutrino flux [23].
    All PDFs are combined in a log-likelihood ratio
    ln(R) = ? n
    s
    ? +
    ?
    N
    i=1
    ln
    ?
    ?n
    s
    ? S
    tot
    (?x
    i
    )
    ?n
    b
    ? B(?x
    i
    )
    +1
    ?
    (2)
    where the sum runs over all reconstructed tracks in the
    final sample. The variable
    ?n
    b
    ?
    is the expected mean
    number of background events, which is determined from
    the off-time data set. The mean number of signal events,
    ?n
    s
    ?
    , is a free parameter which is varied to maximize
    equation 2 in order to obtain the best estimate for the
    mean number of signal events,
    ?nˆ
    s
    ?
    .
    To determine whether a given data set is compatible
    with the background-only hypothesis,
    10
    8
    background
    data sets for the on-time windows are generated from
    off-time data by randomizing the track times while
    taking into account the downtime of the detector. For
    each of these data sets the
    ln(R)
    value is calculated.
    The probability for a data set to be compatible with
    background is given by the fraction of background data
    sets with an equal or larger
    ln(R)
    value.
    The analysis is performed on a high-purity up-going
    neutrino sample after tight selection criteria have been
    applied. The unbinned likelihood method requires an
    ∼ 1.8
    times lower fluence for a
    detection than the
    E
    ν
    (GeV)
    4
    10
    5
    10
    6
    10
    7
    10
    8
    10
    9
    10
    )
    -2
    fluence (GeV cm
    ×
    2
    E
    -5
    10
    -4
    10
    -3
    10
    -2
    10
    -1
    10
    1
    10
    2
    10
    E
    ν
    (GeV)
    4
    10
    5
    10
    6
    10
    7
    10
    8
    10
    9
    10
    )
    -2
    fluence (GeV cm
    ×
    2
    E
    -5
    10
    -4
    10
    -3
    10
    -2
    10
    -1
    10
    1
    10
    2
    10
    preliminary
    Fig. 3. Light lines—Predicted fluences from the 41 northern hemi-
    sphere GRBs for different emission models: prompt (solid, sum of
    individual spectra as plotted in Fig. 1) and precursor (dashed, [3]).
    Dark lines—90% C.L. upper limits on the neutrino fluences obtained
    with the unbinned likelihood analysis.
    binned method. The former is therefore used for the
    results presented in this paper.
    The unblinding procedure involves applying the like-
    lihood method to the on-time data set after neutrino can-
    didate event selection. For all three emission scenarios
    the best estimate for the number of signal events (
    ?nˆ
    s
    ?
    )
    is zero and hence consistent with the null hypothesis.
    Figure 3 displays preliminary 90% C.L. upper limits
    for the 41 GRBs on the fluence in the prompt phase
    (sum of individual spectra as plotted in Fig. 1) of
    3.7 ×
    10
     3
    ergcm
     2
    (72 TeV – 6.5 PeV) and on the fluence
    from the precursor phase [3] of
    1.16 × 10
     3
    erg cm
     2
    (2.2 TeV – 55 TeV), where the quoted energy ranges
    contain 90% of the expected signal events in the detector.
    The limits obtained are not strong enough to constrain
    the models. The preliminary 90% C.L. upper limit for
    the wide time window is
    2.7 × 10
     3
    erg cm
     2
    (3 TeV –
    2.8 PeV) assuming an
    E
     2
    flux.
    III. ICECUBE SENSITIVITY STUDY FOR THE FULL
    DETECTOR
    Previous studies have estimated the sensitivity of the
    completed IceCube to neutrino fluxes from GRBs [24],
    [25]. We present new results using updated information
    about the detector, improved simulation, and more ac-
    curate calculation of the backgrounds.
    We utilize the same methods as in the 22-string search
    to study the sensitivity of the full 86 string detector.
    We generate a set of fake GRBs by sampling from the
    populations observed by the Swift [9] and Fermi [26]
    satellites and taking their observation rates into account.
    We distribute these bursts isotropically over the sky and
    randomly in time to produce a set of 142 fake GRBs
    in the northern sky for a detector livetime of one year,
    with 6840 s of total emission during the prompt phase.
    Currently, we use the average Waxman-Bahcall GRB
    neutrino flux for all bursts [2]. In the future, it will
    be replaced by individual spectra. To test the precursor
    phase, we assume each burst has such emission [3]

    4
    A. KAPPES
    et al.
    GRB SEARCHES WITH ICECUBE
    10
    E
    ν
    (GeV)
    log
    2
    3
    4
    5
    6
    7
    8
    9
    )
    2
    effective area (m
    μ
    ν
    -5
    10
    -4
    10
    -3
    10
    -2
    10
    -1
    10
    1
    10
    2
    10
    3
    10
    4
    10
    5
    10
    0
    °
    <
    δ
    < 90
    °
    0
    °
    <
    δ
    < 19
    °
    19
    °
    <
    δ
    < 42
    °
    42
    °
    <
    δ
    < 90
    °
    preliminary
    Fig. 4. Muon neutrino effective area for the full IceCube detector as a
    function of energy. Solid line is averaged over the half sky, while dot-
    dashed, dotted, and dashed lines represent the most horizontal, middle,
    and most vertical thirds of the northern sky in
    cos δ
    , respectively.
    lasting for 100 s immediately preceeding the observed
    photons.
    As no off-time data is available to determine the
    background, it is simulated. Atmospheric muons and
    neutrinos are generated over the full sky and propagated
    to the detector in the same manner as outlined in
    section II. The geometry of the full detector is simu-
    lated in determining the response to Cherenkov photons.
    Signal and background are filtered with cuts on quality
    parameters to create a sample of well-reconstructed,
    seemingly upgoing events. Further event selection with
    a machine learning algorithm [27] is then performed
    to remove the remaining misreconstructed downgoing
    muons. Afterwards, no atmospheric muons remain in the
    sample due to the limited amount of Monte Carlo. As
    no real data is available for comparison, the exact purity
    of the remaining background sample is unknown, but
    is estimated to consist of
    > 95%
    atmospheric neutrinos
    below the horizon, while retaining a large fraction of
    GRB signal neutrinos. The effective area for different
    declination bands is shown in Fig. 4. Given the detector
    angular resolution of
    ∼ 1
    , we select a search bin radius
    of
    2
    around each fake GRB location, retaining 70–90%
    of signal neutrinos (depending on declination) while
    dramatically reducing the isotropic background rate. The
    background is then rescaled to match the emission time
    window for each burst.
    First results of this study indicate that we will be able
    to detect neutrinos from GRBs in either phase at the
    level in greater than 90% of potential experiments
    within the first few years of operating the full detector.
    In the event of non-detection, we will be able to set strict
    upper limits well below the fluences predicted by these
    models.
    IV. CONCLUSIONS
    We have presented results of searches for muon neu-
    trinos from GRBs with the 22-string configuration of
    the IceCube detector. These searches covered several
    time windows corresponding to the various phases of the
    predicted emission. In all cases, the data were consistent
    with the background only hypothesis. Hence, we place
    upper limits on the muon neutrino fluences from the
    different phases, which, however, are not tight enough
    to constrain any model yet.
    We are also performing a detailed sensitivity study
    for the full 80-string IceCube detector. The preliminary
    results of this study show that IceCube will be able to
    detect the neutrino flux predicted by the leading models
    with a high level of significance within the first few
    years of operation or, in the event of no observation,
    place strong constraints on emission of neutrinos from
    GRBs.
    V. ACKNOWLEDGMENTS
    A. Kappes acknowledges the support by the EU Marie
    Curie OIF Program.
    REFERENCES
    [1] E. Waxman,
    Phys. Rev. Lett.
    , vol. 75, p. 386, 1995.
    [2] E. Waxman and J. N. Bahcall,
    Phys. Rev. Lett.
    , vol. 78, p. 2292,
    1997.
    [3] S. Razzaque, P. Meszaros, and E. Waxman,
    Phys. Rev.
    , vol. D68,
    p. 083001, 2003.
    [4] E. Waxman and J. N. Bahcall,
    ApJ
    , vol. 541, p. 707, 2000.
    [5] A. Karle, (IceCube Coll.)
    et al.
    , Preprint arXiv:0812.3981, 2008.
    [6] A. Achterberg, (IceCube Coll.)
    et al.
    ,
    ApJ
    , vol. 664, p. 397, 2007.
    [7] ——,
    ApJ
    , vol. 674, p. 357, 2008.
    [8] K. Hoffman, K. Meagher, P. Roth, I. Taboada, (IceCube Coll.)
    et al.
    , in
    this proceedings, ID515
    , 2009.
    [9] D. N. Burrows
    et al.
    ,
    Space Sci. Rev.
    , vol. 120, p. 165, 2005.
    [10] G. J. Fishman
    et al.
    , in
    Proc. GRO Science Workshop
    , GSFC,
    1989, p. 2.
    [11] D. Guetta
    et al.
    ,
    Astropart. Phys.
    , vol. 20, p. 429, 2004.
    [12] J. Ahrens, (AMANDA Coll.)
    et al.
    ,
    Nucl. Inst. Meth.
    , vol. A524,
    p. 169, 2004.
    [13] A. Gazizov and M. O. Kowalski,
    Comp. Phys. Comm.
    , vol. 172,
    p. 203, 2005.
    [14] D. Heck
    et al.
    , “Corsika: A Monte Carlo code to simulate
    extensive air showers,” FZKA, Tech. Rep., 1998.
    [15] D. Chirkin and W. Rhode, Preprint hep-ph/0407075, 2004.
    [16] J. Lundberg
    et al.
    ,
    Nucl. Inst. Meth.
    , vol. A581, p. 619, 2007.
    [17] C. Cortes and V. Vapnik,
    Machine Learning
    , vol. 20, no. 3, p.
    273, 1995.
    [18] C. Hsu, C. Chang, and C. Lin, “A practical guide to support
    vector classification,” Department of Computer Science, National
    Taiwan University, Tech. Rep., 2003.
    [19] G. C. Hill, J. Hodges, B. Hughey, A. Karle, and M. Stamatikos,
    in
    Proc. PHYSTAT O5: Statistical Problems in Particle Physics
    ,
    Oxford, United Kingdom, Sep 2005.
    [20] J. Braun, J. Dumm, F. de Palma, C. Finley, A. Karle, and
    T. Montaruli,
    Astropart. Phys.
    , vol. 29, p. 299, 2008.
    [21] J. Zornoza, D. Chirkin, (IceCube Coll.)
    et al.
    , in
    Proc. Interna-
    tional Cosmic Ray Conference (ICRC’07)
    , Me´rida, Mexico, Aug.
    2007.
    [22] R. Abbasi, (HiRes Coll.)
    et al.
    ,
    ApJ
    , vol. 636, p. 680, 2006.
    [23] G. D. Barr
    et al.
    ,
    Phys. Rev.
    , vol. D70, p. 023006, 2004.
    [24] J. Ahrens, (IceCube Coll.)
    et al.
    ,
    Astropart. Phys.
    , vol. 20, p.
    507, 2004.
    [25] A. Kappes, M. Kowalski, E. Strahler, I. Taboada, (IceCube Coll.)
    et al.
    , in
    Proc. International Cosmic Ray Conference (ICRC’07)
    ,
    Me´rida, Mexico, Aug. 2007.
    [26] A. A. Moiseev
    et al.
    ,
    Nucl. Inst. Meth.
    , vol. A588, p. 41, 2008.
    [27] D. Chirkin, (IceCube Coll.)
    et al.
    , in
    this proceedings, ID0466
    ,
    2009.

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    1
    Search for neutrinos from GRBs with IceCube
    K. Meagher
    , P. Roth
    , I. Taboada
    , K. Hoffman
    , for the IceCube Collaboration
    Physics Dept. University of Maryland, College Park MD 20742, USA
    School of Physics and Center for Relativistic Astrophysics, Georgia Institute of Technology, Atlanta, GA 30332, USA
    see special section of these proceedings
    Abstract
    . Gamma-ray bursts (GRBs) are one of the
    few potential sources for the highest energy cosmic
    rays and one of the most puzzling phenomena in
    the universe. In their ultra relativistic jets, GRBs
    are thought to produce neutrinos with energies well
    in excess of 100 TeV. IceCube, a neutrino telescope
    currently under construction at the South Pole, will
    have improved sensitivity to these yet unobserved
    neutrinos. This contribution describes the methods
    used for all IceCube neutrino searches from GRBs
    triggered by satellites. We also present the status
    of three searches for neutrinos in coincidence with
    GRBs. The first search seeks to extend existing Ice-
    Cube 22-string
    ν
    µ
    searches to the high background
    southern hemisphere bursts. A second search looks
    for neutrino-induced cascades with the 22-string con-
    figuration of IceCube. Another
    ν
    µ
    search is planned
    for the 40-string configuration of IceCube, and its
    status is presented here. This paper is a companion of
    another ICRC IceCube contribution that summarizes
    the IceCube 22-string northern hemisphere
    ν
    µ
    GRB
    search results and the expected capabilities of the
    completed 86-string detector.
    Keywords
    : Gamma-Ray Bursts, Neutrinos, IceCube
    I. INTRODUCTION
    Gamma-ray Bursts (GRBs) have been proposed as
    one of the most plausible sources of the highest energy
    cosmic rays [1] and high energy neutrinos [2]. The
    prevalent belief is that the progenitors of so called
    long-
    soft
    GRBs are very massive stars that undergo core
    collapse leading to the formation of a black hole.
    Short-
    hard
    GRBs are believed to be the product of the merger
    of binary compact objects such as neutrons stars and
    black holes leading to the creation of a single black
    hole. Material is ejected from the progenitor in ultra-
    relativistic jets which then produce the observed burst
    of
    γ
    -rays and accelerate particles, including baryons, to
    high energy. Neutrinos are predicted to be produced in
    multiple scenarios: while the jet burrows through the
    envelope of the progenitor of a long-soft burst [3] (TeV
    precursor), in coincidence with the observed
    γ
    -ray signal
    [2] (prompt) and as the jet collides with interstellar
    material or the progenitor wind in the early afterglow
    phase [4] (EeV early afterglow.)
    We use the Waxman-Bahcall model as a benchmark
    for neutrino production in GRBs. The original calcula-
    tion with this model used average GRB parameters as
    measured by BATSE [2]. It was refined by including
    specific details for individual GRBs [5]. Our neutrino
    calculations follow this latter prescription. For many
    GRBs the available information is incomplete. In that
    case we use average parameters in the modeling of the
    neutrino flux.
    IceCube is a high energy (
    E ? 1
    TeV) neutrino
    telescope currently under construction at the South Pole
    [6]. The total instrumented volume of IceCube will
    be
    ∼ 1km
    3
    . IceCube indirectly detects neutrinos by
    measuring the Cherenkov light from secondary charged
    particles produced in neutrino-nucleon interactions. A
    total of 5160 Digital Optical Modules (DOMs) arranged
    in 86 strings frozen in the ice are planned. The results
    presented here correspond to the 22- and 40-string
    configurations. AMANDA-II [7], IceCube’s predecessor
    array, had an instrumented volume
    ≈ 60
    times smaller
    than that of the full IceCube. Searches of neutrinos
    in coincidence with GRBs by AMANDA have been
    reported with negative results [8], [9].
    The two main channels for detecting neutrinos with
    IceCube are the muon and the cascade channels. Charged
    current interactions of
    ν
    µ
    produce muons that, at TeV en-
    ergies, travel for several kilometers in ice. For the muon
    channel the detectors are mainly sensitive to up-going
    muons as the Earth can be used to shield against the
    much larger flux of down-going atmospheric muons. Be-
    cause the neutrino-induced muon spectrum from GRBs
    is expected to be much harder than cosmic-ray induced
    muons GRB neutrino searches can be extended to the
    southern hemisphere are shown in section III. Searches
    for neutrinos from GRBs in the muon channel benefit
    from good angular resolution (
    ∼ 1
    for
    E
    ν
    > 1
    TeV)
    and from the long range of high energy muons. In
    the cascade channel the detectors are sensitive to all
    neutrino flavors through various interaction channels. In
    this case almost all of the neutrino energy is deposited
    in a narrow cylinder of O(10 m) in length; point-like
    compared to IceCube dimensions. The cascade channel
    analyses benefit from good energy resolution (
    ∼ 0.1
    in
    log
    10
    E
    ) and from 4
    π
    sr sensitivity. Complex event
    topologies can also arise from
    ν
    τ
    -induced events for
    energies above
    ?
    1 PeV [10].

    2
    K. MEAGHER
    et al.
    GRB SEARCHES WITH ICECUBE
    II. SATELLITE TRIGGERED SEARCHES FOR
    NEUTRINOS IN COINCIDENCE WITH GRBS
    There are several methods for searching for neutrinos
    from GRBs. The present contribution and its companion
    [11] are
    satellite triggered searches
    . A list of GRB times
    and sky localizations is obtained from satellites, such
    as Swift, Fermi and others. From the perspective of
    IceCube, the ideal GRB that is a source of neutrinos
    has a high photon fluence, a well measured spectrum,
    redshift and other electromagnetic properties and is lo-
    calized with higher accuracy than the pointing resolution
    of IceCube (
    ∼ 1
    for the completed detector). Therefore
    wide field of view searches are preferable even at the
    expense of reduced sensitivity. In that respect, Fermi is
    the main source of GRBs expected to produce neutrinos.
    Fermi started operations in summer 2008, before this
    time, the main source of GRBs for study was Swift.
    To avoid potential biases, all satellite triggered
    searches are conducted using
    blind
    analysis methods.
    The
    on-time
    window around each GRB is left unex-
    amined, except for low level quantities that allow to
    establish the stability of the detector. The length of the
    on-time window depends on the analysis. The remainder
    of the data collected by IceCube, or
    off-time
    window,
    are used to measure the background experimentally. The
    on-time window is studied (
    unblinded
    ) only once the
    analysis procedure has been fully established.
    Searches for GRB neutrinos are performed if the
    detector is determined to have been in a period of
    stable operation according to general data requirements
    developed and shared by the IceCube collaboration.
    In addition the time difference between consecutive
    events is calculated. At trigger level and for initial
    event selection criteria the event rate in IceCube is
    dominated by atmospheric muons produced in cosmic
    ray showers. Given uncorrelated cosmic rays the time
    difference between consecutive events is expected to fall
    exponentially with time and the time constant should
    correspond to the inverse of the detector event rate.
    Finally a histogram of the frequency of number of
    events in 10 s bins is fitted with a Gaussian distribution.
    Deviations from a normal distribution, measured by a
    reduced
    χ
    2
    , indicate periods of high or low detector
    event rate. Only GRBs corresponding to stable detector
    periods are considered.
    Neutrinos are simulated using an implementation of
    the ANIS code [12] and atmospheric muons using
    the CORSIKA air shower simulation package [13].
    Propagation of neutrinos and muons through the Earth
    and ice are performed with ANIS and MMC [14].
    The photon signal in the DOMs is determined from a
    detailed simulation [15] of the propagation of Cherenkov
    light from muons and showers through the ice. This is
    followed by a simulation of the DOM electronics and
    the trigger. The DOM signals are then processed in the
    same way as the data. The theoretical models tested have
    been corrected to take into account neutrino oscillations.
    III. ICECUBE 22-STRING SOUTHERN HEMISPHERE
    MUON SEARCH
    In this analysis we search for muon neutrinos emit-
    ted in the prompt phase from GRBs in the southern
    hemisphere (negative declination). We use filtered data
    collected with the IceCube detector in its 22-string
    configuration between May 2007 and April 2008 for
    bursts with declination
    >  40
    . Very low level data
    taken within two hours of a burst trigger is used for those
    with declination
    <  40
    . In both cases, the data taken
    ±
    20 minutes from the burst trigger is considered to be
    the on-time window. Following the stability procedure
    described in section II we find that two of the 42
    southern hemisphere bursts do not pass the data quality
    criteria or have missing data during the prompt emission
    windows. For the remaining 40 GRBs, these tests show
    no indications of abnormal behavior of the detector.
    Tracks are reconstructed using a log-likelihood re-
    construction method [16]. A fit of a paraboloid to
    the region around the minimum in the log-likelihood
    function yields an estimate of the uncertainty on the
    reconstructed direction. Various quality parameters and
    energy related parameters are derived from the results of
    some other reconstructions discussed in [16] and [17].
    The track quality and energy related variables are
    combined using a machine learning algorithm. The algo-
    rithm used was a Support Vector Machine (SVM) [18]
    with a radial basis function kernel. One SVM was trained
    for the filtered dataset after a loose preselection of events
    and another was trained on the low level dataset. In
    both cases, off-time background data is taken as the
    background and all-sky neutrino simulation is used as
    the signal. The result is an SVM classification between
    -1 (background-like) and 1 (signal-like) for all events.
    An unbinned likelihood method like the one described
    in [19] was used to search each on-time window. This
    method avoids using restrictive selection criterion to
    throw away events but instead uses probability density
    functions (PDFs) to evaluate whether events are more
    likely to be signal or background. The signal,
    S (?x
    i
    )
    ,
    and background,
    B(?x
    i
    )
    , PDFs are each the product of a
    time, a space, and an SVM PDF.
    The space signal PDF is a two-dimensional Gaussian
    determined from the paraboloid fit. The time PDF is flat
    over the respective time window and falls off on both
    sides with a Gaussian distribution with width equal to
    the time window length. The SVM PDF is determined
    from the SVM classifier distribution for simulated signal
    events.
    For the space background PDF the detector asymme-
    tries in zenith and azimuth are taken into account by
    evaluating the off-time data in the detector coordinate
    system. The time distribution of the background during
    a GRB is flat over the entire on-time window. The
    SVM PDF is again determined from the SVM classifier
    distribution of off-time background data.
    All PDFs are combined in an extended log-likelihood
    function [20] where the sum runs over all reconstructed

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    3
    Fig. 1.
    The calculated neutrino fluxes for all burst triggers taken
    during IceCube’s 22-string operations in different declination bands.
    Fig. 2. The 5
    σ
    50% MDF for each burst in the southern hemisphere
    muon neutrino search.
    tracks in the final sample. The variable
    ?n
    b
    ?
    is the
    expected mean number of background events, which is
    determined from the off-time data set. The mean number
    of signal events,
    ?n
    s
    ?
    , is a free parameter which is varied
    to maximize the expression
    ln (R(?n
    s
    ?)) = ? n
    s
    ? +
    ?
    N
    i=1
    ln
    ?
    ?n
    s
    ? S
    tot
    (?x
    i
    )
    ?n
    b
    ? B
    tot
    (?x
    i
    )
    +1
    ?
    (1)
    in order to obtain the best estimate for the mean number
    of signal events,
    ?n
    ˆ
    s
    ?
    .
    To determine whether a given data set is compatible
    with the background-only hypothesis,
    10
    8
    background
    data sets for the on-time windows are generated from
    off-time data by randomizing the track times while
    taking into account the downtime of the detector. For
    each of these data sets the
    ln(R)
    value is calculated.
    The probability for a data set to be compatible with
    background is given by the fraction of background data
    sets with an equal or larger
    ln(R)
    value. The sensitivity
    of each search is determined by injecting simulated
    signal events into these randomizations and observing
    the resultant
    ln(R)
    distribution. This allows for the
    calculation of the Model Detection Factor (MDF) for
    each analysis (figure 2). The MDF is the ratio of the
    lowest signal fluence required for a detection with the
    required significance and power to the predicted fluence
    [21].
    IV. ICECUBE 22-STRING CASCADE SEARCH
    We are currently conducting a search for neutrino-
    induced cascades in the prompt phase for 81 GRBs at
    all declinations in coincidence with data from stable
    detector periods (see section II for details) collected
    with IceCube in its 22-string configuration. The on-time
    period for this analysis is
    ±
    1 hour. The off-time is
    the remainder of the data collected by IceCube with
    22 strings between May 2007 and April 2008 with a
    livetime of
    269 days.
    The analysis proceeds in three steps. First, a prelimi-
    nary selection of cascade-like events is performed online
    at the South Pole. Second, the South Pole filtered data
    are reconstructed by minimizing log-likelihood functions
    that take into account the propagation of photons through
    ice from the source to the digital optical modules [16].
    The reconstructions are performed for both a muon
    hypothesis and a cascade hypothesis. The muon hy-
    pothesis reconstruction provides a position, a direction,
    time and several quality parameters that describe how
    appropriately the muon hypothesis fits the data. The
    cascade hypothesis reconstruction provides a candidate
    neutrino interaction vertex, time, cascade energy and
    quality parameters. After the reconstruction further se-
    lection criteria are applied:
    • L
    µ
     L
    cascade
    >  16.2
    . The difference in the
    log-likelihood quality parameters for the muon and
    cascade reconstruction identifies events that are
    better described by the cascade hypothesis.
    • θ
    µ
    > 73
    . Events that match a down-going muon
    are rejected. Here the
    θ
    µ
    = 0
    represents a vertical
    down-going muon.
    • L
    cascade
    /(Nhit  5) < 8.0
    . Cascade events that
    are low energy or too far from the detector are re-
    constructed poorly. We use the cascade hypothesis
    reduced log-likelihood quality parameter to select
    well reconstructed cascade events.
    • N
    1hit
    /N
    hit
    < 0.1
    . This quantity is a simple cas-
    cade energy proxy because it is equivalent to the
    surface to volume ratio of a spherical pattern of
    light.
    N
    1hit
    measures the number of DOMs in an
    event that have only one hit (typically one photo-
    electron),
    N
    hit
    measures the total number of hits.
    For the optimization of the selection criteria we are
    currently using the Waxman-Bahcall spectrum [2] for
    the expected
    ν
    e
    + ν¯
    e
    signal. After applying the selection
    criteria described above, we expect 0.36 (
    ν
    e
    + ν¯
    e
    ) from
    81 GRBs. Because Swift is the main source of GRBs for
    this analysis, we expect the typical GRB to be about one
    order of magnitude dimmer than what was assumed for
    the Waxman-Bahcall spectrum
    1
    . If a detailed per-burst
    simulation is performed we expect a significantly lower
    signal rate. After applying the selection criteria described
    above
    ≈ 1.5 × 10
    5
    events remain in the off-time data.
    1
    The Waxman-Bahcall model assumed BATSE average GRB pa-
    rameters, especially
    z
    GRB
    = 1
    , while Swift’s mean observed redshift
    is significantly higher.

    4
    K. MEAGHER
    et al.
    GRB SEARCHES WITH ICECUBE
    For the third and final part of the analysis we dis-
    criminate signal from background with a neural net-
    work that uses the parameters described above plus the
    reconstructed energy of the cascade hypothesis and a
    topological parameter that discriminates long (muon)
    from spherical (cascade) events. A cut on the neural
    network parameter provides the final discrimination be-
    tween signal and background.
    V. ICECUBE 40-STRING MUON SEARCH
    IceCube began operating with 40 strings on April 5
    2008 and continues to collect data in this configuration at
    the time of writing. During this time IceCube remained
    extremely stable and maintained a livetime of approx-
    imately 95%. These additional strings give IceCube a
    fiducial volume of approximately
    0.5
    km
    3
    making it
    the largest neutrino detector to date. This section will
    cover the analysis of the northern hemisphere bursts. An
    analysis of the southern hemisphere bursts will follow.
    To date there have been 116 northern hemisphere
    GRBs reported via GCN circulars during 40-string op-
    erations. The launch of the Fermi Gamma-Ray Space
    Telescope with the Gamma-Ray Burst Monitor (GBM)
    has greatly increased the number of bursts available for
    analysis. However, the GBM bursts are usually poorly
    localized and have 1 sigma uncertainties spanning from
    1 to 15 degrees. In addition there are several bursts
    detected by other satellites of the InterPlanetary Net-
    work (IPN), including the brightest burst in the sample,
    GRB080408B, which result in a total of 48 bursts with
    localization uncertainties of larger than one degree. In
    order to search regions of the sky larger than IceCube’s
    angular resolution of approximately 1.5 degrees, new
    methods must be utilized. Expanding on the unbinned
    likelihood analysis presented in section III, an extended
    source hypothesis must be created that takes into account
    both the GRB’s localization error and IceCube’s angular
    uncertainty:
    S
    space
    (?x
    i
    )=
    1
    N
    ?
    d
    Ω · e
    ( ?
    r ?
    rγ )2
    2 σγ
    +
    ( ?
    r  ?
    rν )2
    2 σν
    (2)
    where
    ?r
    γ
    and
    σ
    γ
    are the location and uncertainty of the
    GRB as reported in the GCN circular,
    ?r
    ν
    and
    σ
    ν
    are the
    uncertainty of the IceCube neutrino candidate, and
    N
    is
    a normalization.
    Sensitivity studies are currently being performed and
    will be available soon.
    VI. CONCLUSIONS
    Satellite triggered searches for neutrinos in coinci-
    dence with GRBs use many common techniques. The
    southern hemisphere
    ν
    µ
    search is a first attempt to ex-
    tend IceCube’s sensitivity to GRBs into the higher back-
    ground region above the horizon. The cascade search
    provides sensitivity to all neutrino flavors over 4
    π
    sr.
    The 40-string search provides greater sensitivity due to
    IceCube’s growing effective area and greater number of
    burst triggers from Fermi.
    10
    2
    10
    3
    10
    4
    10
    5
    10
    6
    10
    7
    10
    8
    10
    9
    E
    [GeV]
    10
    -16
    10
    -14
    10
    -12
    10
    -10
    10
    -8
    10
    -6
    10
    -4
    10
    -2
    10
    0
    E
    2
    F
    [
    GeVcm
    2
    ]
    Total Waxman & Bahcall
    Avg Waxman & Bahcall
    Total Individual Spectra
    Fig. 3. The Calculated Neutrino Spectrum for 102 of the 116 northern
    hemisphere bursts for which spectral information was available. The
    Sum of the Neutrino spectrum is plotted along with the Average
    Waxman and Bahcall spectrum for a single burst and for 102 bursts.
    10
    2
    10
    3
    10
    4
    10
    5
    10
    6
    10
    7
    10
    8
    10
    9
    E
    [GeV]
    10
    -4
    10
    -3
    10
    -2
    10
    -1
    10
    0
    10
    1
    10
    2
    10
    3
    10
    4
    10
    5
    10
    6
    A
    eff
    [
    m
    2
    ]
    ☎☎☎☎☎☎
    100000.
    .
    .
    .
    .
    .
    086420 <<<<<<CCCCCCoooooossssss((((((
    ✆✆✆✆✆✆
    )))))) <<<<<<
    ☎☎☎☎☎
    +000000
    .
    .
    .
    .
    .
    .
    864202
    Fig. 4. Effective Area of 40 string IceCube.
    REFERENCES
    [1] E. Waxman,
    Phys. Rev. Lett.
    , vol. 75, p. 386, 1995.
    [2] E. Waxman and J. N. Bahcall,
    Phys. Rev. Lett.
    , vol. 78, p. 2292,
    1997.
    [3] S. Razzaque, P. Meszaros, and E. Waxman,
    Phys. Rev.
    , vol. D68,
    p. 083001, 2003.
    [4] E. Waxman and J. N. Bahcall,
    ApJ
    , vol. 541, p. 707, 2000.
    [5] D. Guetta
    et al.
    ,
    Astropart. Phys.
    , vol. 20, p. 429, 2004.
    [6] A. Karle, (IceCube Coll.)
    et al.
    , in
    Proc. International Cosmic
    Ray Conference (ICRC’09)
    , Lodz´, Poland, 2009.
    [7] E. Andre´s, (AMANDA Coll.)
    et al.
    ,
    Nature
    , vol. 410, p. 441,
    2001.
    [8] A. Achterberg, (IceCube Coll.)
    et al.
    ,
    ApJ
    , vol. 664, p. 397, 2007.
    [9] ——,
    ApJ
    , vol. 674, p. 357, 2008.
    [10] T. Deyoung, S. Razzaque, and D. F. Cowen,
    Astroparticle
    Physics
    , vol. 27, pp. 238–243, Apr. 2007.
    [11] A. Kappes, P. Roth, E. Strahler, (IceCube Coll.)
    et al.
    , in
    Proc.
    International Cosmic Ray Conference (ICRC’09)
    , Lodz´, Poland,
    2009.
    [12] A. Gazizov and M. O. Kowalski,
    Comp. Phys. Comm.
    , vol. 172,
    p. 203, 2005.
    [13] D. Heck
    et al.
    ,
    Technical Report FZKA
    , vol. 6019, 1998.
    [14] D. Chirkin and W. Rhode, “hep-ph/0407075.”
    [15] J. Lundberg
    et al.
    ,
    Nucl. Inst. Meth.
    , vol. A581, p. 619, 2007.
    [16] J. Ahrens, (AMANDA Coll.)
    et al.
    ,
    Nucl. Inst. Meth.
    , vol. A524,
    p. 169, 2004.
    [17] J. Zornoza, D. Chirkin, (IceCube Coll.)
    et al.
    , in
    Proc. Interna-
    tional Cosmic Ray Conference (ICRC’07)
    , Merida, Mexico, Aug.
    2007.
    [18] C. Cortes and V. Vapnik,
    Machine Learning
    , vol. 20, no. 3, p.
    273, 1995.
    [19] J. Braun, J. Dumm, F. de Palma, C. Finley, A. Karle, and
    T. Montaruli,
    Astropart. Phys.
    , vol. 29, p. 299, 2008.
    [20] R. J. Barlow,
    Statistics
    . Wiley, 1989.
    [21] G. C. Hill, J. Hodges, B. Hughey, A. Karle, and M. Stamatikos,
    in
    Proc. PHYSTAT O5: Statistical Problems in Particle Physics
    ,
    Oxford, United Kingdom, Sep 2005.

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁOD´ Z´ 2009
    1
    Search for GRB neutrinos via a (stacked) time profile analysis
    Martijn Duvoort
    ?
    and Nick van Eijndhoven
    ?
    for the IceCube collaboration
    y
    ?
    University of Utrecht, The Netherlands
    y
    see special section of these proceedings
    Abstract. An innovative method to detect high-
    energy neutrinos from Gamma Ray Bursts (GRBs) is
    presented. The procedure provides a good sensitivity
    for both prompt, precursor and afterglow neutrinos
    within a 2 hour time window around the GRB
    trigger time. The basic idea of the method consists
    of stacking of the observed neutrino arrival times
    with respect to the corresponding GRB triggers. A
    possible GRB neutrino signal would manifest itself
    as a clustering of signal candidate events in the
    observed time profile. The stacking procedure allows
    to identify a signal even in the case of very low rates.
    We outline the expected performance of analysing
    four years of AMANDA data (2005–2008) for a
    sample of 130 GRBs. Because of the extreme optical
    brightness of GRB080319B, it might be that this
    particular burst yielded multiple detectable neutrinos
    in our detector. As such, the method has also been
    applied to the data of this single burst time profile.
    The results of this analysis are presented in a separate
    section.
    Keywords: GRB Neutrinos AMANDA/IceCube
    I. INTRODUCTION
    Gamma Ray Bursts are among the most promising
    sources for high-energy neutrino detection: the accurate
    localization and timing information presently available,
    enable very effective background reduction for high-
    energy neutrino detectors. Yet, no previous search for
    GRB neutrinos has led to a discovery [1], [2]. As most
    models of a GRB jet predict neutrino formation simul-
    taneous with the prompt ? emission, previous analyses
    aim to discover neutrinos that arrive simultaneously with
    the prompt photons. However, it might be that the main
    GRB neutrino signal is not simultaneous with the prompt
    gammas, either in production or arrival at the Earth.
    A variety of the models predict the formation of high-
    energy neutrinos at different stages in the evolution of
    a GRB. Afterglow models predict a significant neutrino
    flux a few seconds after the prompt emission [3], [4].
    The existence of multiple colliding shells in a GRB
    jet [5] may also lead to a time difference between
    high-energy gamma emission and neutrinos. Even if the
    neutrinos and photons are produced at the same stage
    of the evolution of the GRB jet, a time difference at
    the observer may be present: as the jet evolves, it will
    become transparent for photons at a later stage than for
    neutrinos. Therefore, neutrinos might be able to escape
    the source region well before the high-energy photons.
    This will depend heavily on the actual stage in the
    evolution of the jet.
    For our analysis we use the data of the AMANDA-
    II detector at the South Pole [6] to look for a neutrino
    signal. Our analysis method is aimed to be less model
    dependent than previous GRB analyses. It is insensitive
    to a possible time difference between the arrival of the
    prompt photons and the high-energy neutrino signal. We
    limit the dependence on the expected neutrino spectrum
    by not using any energy dependent selection criteria.
    We only use directional selection parameters based on
    the reconstructed muon track, resulting from an incident
    muon neutrino [6]. As the detectable number of signal
    neutrinos in our detector per GRB is very low [7]
    (˝ 1), our method is designed to allow for gaining
    sensitivity to a GRB neutrino signal by stacking neutrino
    data of multiple GRBs around their trigger time. Those
    stacked time profiles can be analysed using the same
    techniques as the time profile of a single GRB. We
    first outline the analysis method itself, then we give the
    results of applying this method to GRB080319B, the
    most luminous GRB observed to date.
    II. THE ANALYSIS METHOD
    We look for signal events correlated with the GRB
    direction and time. As the background of our detector,
    which consists of cosmic ray events, is not correlated,
    we start by filtering the data for a GRB coincidence, both
    spatially and temporally. The exact selection parameters
    we use are optimized as outlined in section III.
    The GRB data that passes the cuts has a certain time-
    distribution with respect to the GRB trigger time. The
    background events that pass the cuts will be uniformly
    distributed in time with respect to the GRB trigger. A
    possible GRB signal will be clustered in time. Note that
    this argument also holds for the case of stacking multiple
    GRB time windows, which is the main purpose of this
    analysis method. Here we assume that the intrinsic time
    difference between photons and neutrinos is a charac-
    teristic feature for all GRBs in our sample. Obviously
    we aim to have all GRB signal neutrinos ending up in
    the same time-bin. Therefore, the usage of a too small
    time bin will reduce the sensitivity as signal entries will
    end up in different bins. Using a too large time bin
    also reduces the sensitivity as background entries will
    start to dominate the bins. We estimate the timespread
    of the neutrino signal to be of the same order of the
    observed photonic GRB duration: the T
    90
    time, defined
    as the time in which 5%  95% of the GRB fluence
    was detected. This is a safe estimate as the intrinsic

    2
    DUVOORT et al. GRB NEUTRINO SEARCH USING TIME PROFILE ANALYSIS
    Fig. 1. The ? distribution for randomizing 13 entries in 120 bins.
    (108 randomizations)
    timespread of the neutrino signal will not be larger than
    that of the photons: as the source region will always be
    more opaque for photons than for neutrinos, the photon
    signal will spread more in time than the neutrino signal.
    We have chosen a conservative bin size of 60 s, resulting
    in 120 time bins in our 2 hour window.
    The probability of observing a certain time distribu-
    tion given a uniform background distribution, of in total
    n entries divided over m time bins, is given by the
    multinomial distribution [8]:
    p(n
    1
    ; n
    2
    ; :::n
    m
    jnm) =
    n!
    n
    1
    ! ? ? ? n
    m
    !
    p
    n
    1
    1
    ???p
    n
    m
    m
    ? p: (1)
    Here p
    i
    is the probability of an entry ending up in bin
    i. In case of a uniform background this is simply m
     1
    .
    The n
    i
    represents the number of entries in bin i. We
    derive the bayesian ? ?  10 log p [9]:
    ?=  10
    "
    log n! +
    X
    m
    k=1
    (n
    k
    log p
    k
     log n
    k
    !)
    #
    : (2)
    If the observation is due to the expected background,
    a low ? value will be obtained. Deviations from the
    expected background will result in increased ? values.
    We intend to compare the ? value of the observed
    data, including a possible signal, with the distribution
    of uniform background ? values. We obtain such back-
    ground sets by (uniformly) randomizing the entries in
    the two hour time window, keeping the total number
    of entries constant to what we find in the data. In
    case of a large signal contribution, this may result in
    underestimating the significance of the signal. However,
    for such a high signal contribution we will be able to
    claim discovery anyway. To claim a discovery we require
    at least a 5˙ level, which means that only a fraction of
    5:73?10
     7
    (the corresponding P-value) of all the ?s of
    the various background sets is allowed to exceed some
    threshold ?
    0
    . In case the ? value of our observed data
    is larger than ?
    0
    , we have a discovery.
    In order to reach the necessary accuracy, we perform
    10
    8
    randomizations of all the data events that pass the
    criteria and calculate the ? value of each randomization
    to obtain a background ? distribution. In figure 1 we
    give one example of our parameter space. Here n = 13
    entries exist in our simulated observation time window
    of 120 bins.
    Fig. 2. The time distribution of 3 signal events (at t = 0) and 10
    randomly distributed events in a 2 hour window using 60 s bin size.
    As an example, one might observe a time distribu-
    tion, consisting of 3 signal events in a single bin and
    10 randomly distributed background entries, which is
    shown in figure 2. The ? value associated with this
    distribution equals 186:15. When comparing with the
    background ? distribution of figure 1 it becomes clear
    that this corresponds to a P-value of 1:13 ? 10
     3
    above
    the observed ?
    0
    = 186:15. For a 5˙ discovery we need
    this fraction to be less than 5:73 ? 10
     7
    . Therefore,
    observing a time profile like figure 2 will not result in
    a significant discovery.
    III. OPTIMIZATION OF THE SELECTION PARAMETERS
    The significance of our observation is determined by
    the method outlined in section II. Before we do this
    we need to optimize the directional parameter values
    which we use for selecting the final event sample by
    means of a blind analysis. In order to stay comparable
    to previous analyses, we will use the standard Model
    Discovery Factor (MDF) [10], [11] to determine the
    optimum of our parameter space. At those optimal
    settings the standard Model Rejection Factor (MRF) [12]
    is calculated.
    The average expected number of background counts
    per time bin ?
    b
    is calculated by simply dividing the
    total number of observed entries in the time window
    by the number of time bins. This is justified by the
    assumption that the expected signal is much smaller
    than the background ?
    b
    ˛?
    s
    . We optimize the selection
    parameters for a 5˙ discovery. The significance we use
    in the calculation of the MDF is corrected for a trial
    factor due to the number of bins.
    By systematically going through the grid of our
    parameter space, we reach the parameter values cor-
    responding to a minimum MDF, i.e. we optimize our
    analysis for discovery. In case of no discovery, the MRF
    at these settings will provide a flux upper limit. Since we
    optimize our parameters on the randomized data itself,
    our background set consists of randomized background
    plus signal entries. For parameter settings where less
    than four entries pass, the ? statistics cannot result in
    a discovery: all possible P-value exceed 5:73 ? 10
     7
    .
    Therefore, we require that at our optimal thresholds,
    at least four events pass our filter. This is achieved by
    slightly relaxing the selection criteria.

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁOD´ Z´ 2009
    3
    IV. THE GRB080319B AMANDA ANALYSIS
    Even though the expected number of signal neutrinos
    for an average GRB is extremely low, the atypical
    GRB080319B might yield an unusually strong neutrino
    signal justifying an individual neutrino analysis. The
    analysis of IceCube data [13] was confined to 10 minutes
    around the GRB trigger time. No neutrino signal was
    found. We analyse a larger data block from one hour
    before till one hour after the trigger time, and use the
    same spectrum as in [13] for our optimization and limit:
    dN
    ?
    dE
    ?
    =
    8
    <
    :
    6:620 ? 10
     16
    ? E
    0:59
    ?
    if E
    ?
    ? E
    1
    ;
    0:768 ? E
     2:145
    ?
    if E
    1
    ? E
    ?
    ? E
    2
    ;
    6:690 ? E
     4:145
    ?
    if E
    ?
    ? E
    2
    ;
    (3)
    with the fluence, dN
    ?
    =dE
    ?
    , in (GeV cm
    2
    )
     1
    and the
    break energies: E
    1
    = 322:064 TeV; E
    2
    = 2952:35 TeV.
    For our time profile we use a bin size of 60 s,
    roughly the T
    90
    of this burst. While we expect this to be
    wide enough for the GRB neutrino signal to fit in one
    bin, a possible neutrino signal can be spread over two
    adjacent bins. This obviously lowers the significance of
    the observation. Therefore, in case we do not find a 5˙
    result with our initial analysis, we compensate for this
    binning effect by performing our analysis a second time,
    where we shift our bins by half a binwidth.
    Using a simulated neutrino fluence following the
    spectrum (3), we obtain the optimum of our parameter
    space following the method of section III. We find at
    the optimal parameter settings a 5˙ MDF of 123:65
    and have six events passing the filter. Based on the
    GRB spectrum we expect 0:064 signal entries to pass
    the filter. We find (at 90% confidence level) an MRF of
    38:8 for the GRB spectrum. Likewise, using a generic
    E
     2
    spectrum, we obtain at these settings a limit of
    E
    2
    dN
    ?
    =dE
    ?
    = 1:11 ? 10
     2
    (GeV cm
    2
    s)
     1
    at 90%
    confidence level. Note that these limits are conservative
    as the ? statistics we use to claim discovery is more
    sensitive than the Poisson statistics on which the MRF
    is based.
    The previous analysis of IceCube data [13] quotes
    a sensitivity at a fluence of 22:7 times the expected
    spectrum at 90% C.L. for prompt emission. We find a
    90% C.L. limit at 38:8 times the expected spectrum for a
    neutrino signal arriving in the central bin. This difference
    Fig. 3. The 90% C.L. upper limits on the fluence of GRB080319B
    with respect to the calculated neutrino fluence (3) for both this analysis
    and the IceCube analysis in its 9 string configuration.
    can be seen in figure 3, where the limits of both analyses
    are given.
    The neutrino effective areas for the AMANDA detec-
    tor for this analysis are given in figure 4. It is given at
    Fig. 4. The neutrino effective area for the position of GRB080319B,
    both at trigger level and at final cut level.
    both trigger level and at the level where all our selections
    have been applied. The ratio between the 2 histograms is
    the signal passing rate, which, for the GRB spectrum (3),
    equals 49:4%.
    The time profile we find after unblinding is consistent
    with the background-only hypothesis. Repeating the
    analysis with shifted time bins does not change this.
    V. THE STACKING ANALYSIS
    In this section we present the expected results of
    analysing the stacked AMANDA data of 130 GRBs
    between 2005 and 2008. These well-localized bursts are
    all in the Northern hemisphere to reduce the background
    due to atmospheric events. The time profile of each GRB
    is sampled to form a stacked time profile.
    Due to the different redshifts of the GRBs in the
    sample, the effect of cosmological time dilation on the
    intrinsic time difference between photons and neutrinos
    will result in a timespread on the arrival of the neutrino
    signal. This spread will increase for larger time differ-
    ences. We compensate for this by enlarging the bin sizes
    for bins further away from the trigger. Each bin will
    be enlarged by a factor of hzi + 1, where hzi is the
    average redshift of the GRBs in our sample. We choose
    to have our central bin range from h T
    90
    i ˇ  30 s
    to hT
    90
    i ˇ 30 s, allowing for a scatter in the neutrino
    arrival time of the average length of the photon signal.
    The second bin is a factor of hzi+ 1 ˇ 3 larger than the
    maximum scatter we allow in the center bin and ranges
    [30; 120] s (and [ 120;  30] s). The next bin is again a
    factor of 3 larger.
    The fact that the bins in our time window have
    unequal sizes does not influence our method. It is simply
    taken into account by using, for each bin, the correct
    p
    i
    , the probability for an entry to fall in that bin, see
    equation (2). Let us consider the same time profile as
    above (figure 2) with these new bin settings. This leads
    to the time profile as given in figure 5. Because our
    time window now has variable binning, the configuration
    itself changed significantly with respect to the regular

    4
    DUVOORT et al. GRB NEUTRINO SEARCH USING TIME PROFILE ANALYSIS
    Fig. 5.
    The time distribution of 3 signal events and 10 randomly
    distributed events in a 2 hour window. Here we use nine variable time
    bins as explained in the text.
    case of figure 2. Hence the new ? value of our obser-
    vation (63:44) differs from the previously found value.
    The background ? distribution of figure 1 will change
    accordingly. Following our example, one can study the
    n
    signal
    P-value
    P-value
    regular 60 s bins
    variable bins
    1
    6:32 ? 10
     2
    2:36 ? 10
     1
    2
    9:67 ? 10
     3
    3:77 ? 10
     2
    3
    1:13 ? 10
     3
    3:14 ? 10
     3
    4
    2:71 ? 10
     5
    1:80 ? 10
     4
    5
    1:1 ? 10
     6
    7:8 ? 10
     6
    6
    1 ? 10
     8
    2:2 ? 10
     7
    TABLE I
    COMPARISON BETWEEN THE SIGNIFICANCE FOR THE CASE OF THE
    TIME PROFILE OF FIGURE 2 AND THE SAME SITUATION USING THE
    VARIABLE BINS AS IN FIGURE 5. HERE WE VARY THE AMOUNT OF
    SIGNAL ENTRIES IN THE CENTER BIN n
    signal
    ; THE 10
    BACKGROUND ENTRIES ARE LEFT UNTOUCHED.
    effect of the variable binning on the significance of our
    time profile for different signal strengths. From table I
    one can see that introducing variable bin sizes slightly
    lowers the significance of our observations (their P-
    value) for a signal falling in the center bin.
    For the optimization of the selection parameters we
    use both a Waxman-Bahcall and a generic E
     2
    spec-
    trum. Again, we optimize for discovery using the stan-
    dard MDF. As a result from the various binsizes, the
    limit of this analysis depends on the bin size, and
    therefore depends on the time difference with the GRB
    trigger. The sensitivity of this analysis for each bin in our
    Time range
    WB spectrum at 1 PeV
    E
    2
    dN
    ?
    =dE
    ?
    w.r.t. GRB trigger
    (GeV cm2 s sr)
     1
    (GeV cm2 s sr)
     1
    [ 30; 30] s
    2:9 ? 10
     8
    1:55 ? 10
     8
    ? [30; 120] s
    3:0 ? 10
     8
    1:58 ? 10
     8
    ? [120; 390] s
    3:3 ? 10
     8
    1:76 ? 10
     8
    ? [390; 1200] s
    4:2 ? 10
     8
    2:24 ? 10
     8
    ? [1200; 3600] s
    5:8 ? 10
     8
    3:09 ? 10
     8
    TABLE II
    THE 90% C.L. SENSITIVITY OF THE STACKING ANALYSIS FOR
    EACH TIME BIN, FOR BOTH THE WAXMAN-BAHCALL (WB) AND A
    GENERIC E
     2
    SPECTRUM.
    time window are given in table II. For the central bin we
    Fig. 6. The sensitivity of this analysis for both a generic E
     2
    and
    a Waxman-Bahcall (WB) source spectrum (90% C.L.).
    have shown the limits in figure 6. Note that these limits
    only apply to a neutrino signal arriving simultaneously
    with the prompt photon emission.
    VI. DISCUSSION
    Currently, the most restrictive muon neutrino up-
    per limit has been determined by AMANDA at
    E
    2
    dN
    ?
    =dE
    ?
    ? 1:7 ? 10
     8
    GeV cm
     2
    s
     1
    sr
     1
    based
    on a sample of over 400 GRBs and for the Waxman-
    Bahcall spectrum at 1 PeV [2]. For our analysis no
    energy dependent selection parameters are used and the
    optimum of the selection parameters is independent of
    the source spectrum we use. As such, our analysis is
    less model dependent and it allows for a possible time
    difference between photons and neutrinos. Furthermore,
    the stacking procedure provides sensitivity even in the
    case of very low individual GRB rates. As such, the
    present analysis has the potential of detecting precursor
    and afterglow neutrinos in addition to prompt ones.
    The method may also be used to analyse the data of
    individual GRBs. By construction our method is slightly
    less sensitive compared to a model dependent analysis
    of a single time bin. The effective area of the complete
    IceCube detector will be at least ˘150 times larger than
    AMANDA’s [14]. Applying our analysis on one year
    data of the full IceCube, would result in a sensitivity
    well below the predicted Waxman-Bahcall spectrum.
    REFERENCES
    [1] Achterberg, A., et al. 2008, Astrophys. J. 674, 357
    [2] Achterberg, A., et al. 2007, Astrophys. J. 664, 397
    [3] Waxman, E., & Bahcall, J. N. 2000, Astrophys. J., 541, 707
    [4] Dai, Z. G., & Lu, T. 2001, Astrophys. J. 551, 249
    [5] Dar, A., & De Rujula, A. 2001, arXiv:astro-ph/0105094
    [6] Ahrens, J., et al. 2004, Nucl. Instr. Meth. A 524, 169
    [7] Halzen, F., & Hooper, D. W. 1999, Astrophys. J. 527, L93
    [8] Gregory, P. C. 2005, Bayesian Logical Data Analysis for the
    Physical Sciences, Cambridge University Press, UK, 2005
    [9] van Eijndhoven, N. 2008, Astropart. Phys. 28, 540
    [10] Punzi, G. 2003, Statistical Problems in Particle Physics, Astro-
    physics and Cosmology, 79
    [11] Hill, G. C., Hodges, J., Hughey, B., Karle, A., & Stamatikos, M.
    2006, Statistical Problems in Particle Physics, Astrophysics and
    Cosmology, 108
    [12] Hill, G. C., & Rawlins, K. 2003, Astropart. Phys. 19, 393
    [13] IceCube Collaboration: R. Abbasi 2009, arXiv:0902.0131
    [14] Ahrens, J., et al. 2004, Astropart. Phys. 20, 507

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    1
    Optical follow-up of high-energy neutrinos detected by IceCube
    Anna Franckowiak
    , Carl Akerlof
    §
    , D. F. Cowen
    ∗†
    , Marek Kowalski
    , Ringo Lehmann
    ,
    Torsten Schmidt
    and Fang Yuan
    §
    for the IceCube Collaboration
    and for the ROTSE Collaboration
    Humboldt Universita¨t zu Berlin
    Pennsylvania State University
    University of Maryland
    §
    University of Michigan
    see the special section in this proceedings
    Abstract
    . Three-quarters of the 1 km
    3
    neutrino
    telescope IceCube is currently taking data. Current
    models predict high-energy neutrino emission from
    transient objects like supernovae (SNe) and gamma-
    ray bursts (GRBs). To increase the sensitivity to such
    transient objects we have set up an optical follow-
    up program that triggers optical observations on
    multiplets of high-energy muon-neutrinos. We define
    multiplets as a minimum of two muon-neutrinos from
    the same direction (within 4
    ) that arrive within a
    100 s time window. When this happens, an alert
    is issued to the four ROTSE-III telescopes, which
    immediately observe the corresponding region in
    the sky. Image subtraction is applied to the optical
    data to find transient objects. In addition, neutrino
    multiplets are investigated online for temporal and
    directional coincidence with gamma-ray satellite ob-
    servations issued over the Gamma-Ray Burst Coor-
    dinate Network. An overview of the full program is
    given, from the online selection of neutrino events to
    the automated follow-up, and the resulting sensitivity
    to transient neutrino sources is presented for the first
    time.
    Keywords
    : Neutrinos, Supernovae, Gamma-Ray
    Bursts
    I. INTRODUCTION
    When completed, the in-ice component of IceCube
    will consist of 4800 digital optical modules (DOMs)
    arranged on 80 strings frozen into the ice, at depths
    ranging from 1450m to 2450m [1]. Furthermore there
    will be six additional strings densely spaced at the
    bottom half of the detector. The total instrumented
    volume of IceCube will be 1 km
    3
    . Each DOM contains
    a photomultiplier tube and supporting hardware inside a
    glass pressure sphere. The DOMs indirectly detect neu-
    trinos by measuring the Cherenkov light from secondary
    charged particles produced in neutrino-nucleon interac-
    tions. IceCube is most sensitive to neutrinos within an
    energy range of TeV to PeV and is able to reconstruct
    the direction of muon-neutrinos with a precision of
    1
    .
    The search for neutrinos of astrophysical origin is among
    the primary goals of the IceCube neutrino telescope.
    Source candidates include galactic objects like super-
    nova remnants as well as extragalactic objects like
    Fig. 1: Neutrino event spectrum in the IceCube detector,
    from kaon and pion decay in the supernovae-jet model
    of Ando and Beacom [5].
    Active Galactic Nuclei and Gamma-Ray Bursts [9] [10].
    Offline searches for neutrinos in coincidence with GRBs
    have been performed on AMANDA and IceCube data.
    They did not lead to a detection yet, but set upper
    limits to the predicted neutrino flux [13]. While the
    rate of GRBs with ultra-relativistic jets is small, a much
    larger fraction of SNe not associated with GRBs could
    contain mildly relativistic jets. Such mildly relativistic
    jets would become stalled in the outer layers of the
    progenitor star, leading to essentially full absorption of
    the electromagnetic radiation emitted by the jet. Hence,
    with the postulated presence of mildly relativistic jets
    one is confronted with a plausible but difficult-to-test hy-
    pothesis. Neutrinos may reveal the connection between
    GRBs, SNe and relativistic jets. As was recently shown,
    mildly relativistic jets plowing through a star would be
    highly efficient in producing high-energy neutrinos [5]–
    [7]. The predicted neutrino spectrum follows a broken
    power law and Fig. 1 shows the expected signal spectrum
    for neutrinos produced in kaon and pion decay in the
    source, simulated using the full IceCube simulation
    chain. The expected number of signal events is small
    and requires efficient search algorithms to reduce the
    background of atmospheric neutrinos (see section II).
    An optical follow-up program has been started which
    enhances the sensitivity for detecting high-energy neutri-
    nos from transient sources such as SNe. In this program,
    the direction of neutrinos are reconstructed online, and
    if their multiplicity pass a certain threshold, a Target-

    2
    AUTHOR
    et al.
    PAPER SHORT TITLE
    of-Opportunity (ToO) notice is sent to the ROTSE-III
    network of robotic telescopes. These telescopes monitor
    the corresponding part of the sky in the subsequent
    hours and days and identify possible transient objects,
    e.g. through detection of rising supernova light-curves
    lasting several days. If in this process a supernova is
    detected optically, one can extrapolate the lightcurve or
    afterglow to obtain the explosion time [2]. For SNe,
    a gain in sensitivity of about a factor of 2-3 can be
    achieved through optical follow-up observations of neu-
    trino multiplets [4]. In addition to the gain in sensitivity,
    the follow-up program offers a chance to identify its
    transient source, be it a SN, GRB or any other transient
    phenomenon.
    II. NEUTRINO ALERT SYSTEM
    IceCube’s optical follow-up program has been operat-
    ing since fall of 2008. In order to match the requirements
    given by limited observing time at the optical telescopes,
    the neutrino candidate selection has been optimized to
    obtain less than about 25 background multiplets per year.
    The trigger rate of the 40 string IceCube detector is about
    1000 Hz. The muon filter stream reduces the rate of
    down-going muons created in cosmic ray showers dra-
    matically by limiting the search region to the Northern
    hemisphere and a narrow belt around the horizon. The
    resulting event stream of 25 Hz is still dominated by
    misreconstructed down-going muons. Selection criteria
    based on on track quality parameters, such as number
    of direct hits
    1
    , track length and likelihood of the recon-
    struction, yield a reduced event rate of 1 event/(10 min).
    The optimized selection criteria are relaxed to improve
    the signal efficiency, 50% of the surviving events are
    still misreconstructed down-going muons, while 50%
    are atmospheric neutrinos. During the antarctic summer
    2008/2009, 19 additional strings were deployed, which
    have been included in the data taking since end of
    April 2009. To take into account an enhancement in the
    rate due to the increased detector volume, the selection
    criteria have been adjusted and will yield a cleaner event
    sample containing only 30% misreconstructed muons.
    From this improved event sample, neutrino multiplet
    candidates with a time difference of less than 100 s
    and with an angular difference (or ’space angle’) of less
    than 4
    are selected. The choice of the time window
    size is motivated by jet penetration times. Gamma-ray
    emission observed from GRBs has a typical length of
    40 s, which roughly corresponds to the duration of
    a highly relativistic jet to penetrate the stellar enve-
    lope. The angular difference is determined by IceCube’s
    angular resolution. Assuming single events from the
    same true direction, 75% of all doublets are confined
    to a space angle of 4
    after reconstruction. Once a
    multiplet is found, a combined direction is calculated
    as a weighted average of the individual reconstructed
    1
    Hits that are measured within
    [-15ns
    ,75ns] from the predicted
    arrival time of Cherenkov photons, without scattering, given by the
    track geometry.
    event directions, with weights derived from the estimated
    direction resolution of each track. The resolution of the
    combined direction is up to a factor of
    1/√
    2
    better
    than that of individual tracks. The multiplet direction
    is sent via the network of Iridium satellites from the
    South Pole to the North, where it gets forwarded to the
    optical telescopes. At this point in time, due to limited
    parallelization of the data processing at the South Pole a
    delay of 8 hours is accumulated. In the near future, the
    online processing pipeline will be upgraded, reducing
    the latency drastically to the order of minutes.
    A total of 14 alerts have passed the selection criteria and
    were sent to the telescopes within 7 months of operation.
    III. OPTICAL FOLLOW-UP OBSERVATIONS
    At the moment IceCube alerts get forwarded to
    the Robotic Optical Transient Search Experiment
    (ROTSE) [3]. Additions to the list of participating tele-
    scopes are planned. ROTSE-III is dedicated to observa-
    tion and detection of optical transients on time scales of
    seconds to days. The original emphasis was on GRBs
    while it more recently has also started a very successful
    SN program. The four ROTSE-III telescopes are in-
    stalled around the world (in Australia, Namibia, the USA
    and Turkey). The ROTSE-III equipment is modest by
    the standards of modern optical astronomy, but the wide
    field of view and the fast response permit measurements
    inaccessible to more conventional instruments. The four
    0.45 m robotic reflecting telescopes are managed by a
    fully-automated system. They have a wide field of view
    (FOV) of
    1.85
    × 1.85
    imaged onto a 2048
    ×
    2048
    CCD, and operate without filters. The cameras have a
    fast readout cycle of 6 s. The limiting magnitude for a
    typical 60 s exposure is around 18.5 mag, which is well-
    suited for a study of GRB afterglows during the first
    hour or longer. The typical full width at half maximum
    (FWHM) of the stellar images is smaller than 2.5 pixels
    (8.1 arcseconds). Note that ROTSE-IIIs FOV matches
    the size of the point spread function of IceCube well.
    Once an IceCube alert is received by one of the tele-
    scopes, the corresponding region of the night sky will
    be observed within seconds. A predefined observation
    program is started: The prompt observation includes
    thirty exposures of 60 seconds length
    2
    . Follow-up
    observations are performed for 14 nights. Eight images
    with 60 seconds exposure time are taken per night. The
    prompt observation is adjusted to the typical rapidly de-
    caying lightcurve of a GRB afterglow, while the follow-
    up observation of 14 days permits the identification of
    an increasing SN lightcurve. Once the images are taken,
    they are automatically processed at the telescope site.
    Once the data is copied from the telescopes, a second
    analysis is performed off-line, combining the images
    from all sites. Image subtraction is performed according
    2
    Once the delay caused by data processing at the South Pole (see
    section II) is reduced to the order of minutes, the prompt observation
    will include ten short observations of 5 seconds, ten observations of
    20 seconds and twenty long exposures of 60 seconds.

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    3
    the methods presented in [8]. Here the images of the
    first night serve as reference, while the images from the
    following nights are used to search for the brightening
    of a SN lightcurve.
    IV. SENSITIVITY
    The sensitivity of the optical-follow up program is
    determined by both IceCube’s sensitivity to high energy
    neutrino multiplets and ROTSE-III’s sensitivity to SNe.
    We will distinguish two cases: The first being that no
    optical counterpart is observed over the course of the
    program (assuming 25 alerts per year) and the second
    that a SN is identified in coincidence.
    A. No Optical Counterpart Discovered
    With no coincident SN observed, one obtains an upper
    limit on the average number of SNe that could produce
    a coincidence:
    N
    IC
    ROTSE
    < 2.44
    (for 90% confidence
    level). Constraints on a given model are obtained by de-
    manding that the model does not predict a number in ex-
    cess of the SN event upper limit. We construct a simple
    model based on Ando & Beacom type SNe [5]. We in-
    troduce two parameters: The first is the rate of SNe pro-
    ducing neutrinos
    ρ = (4/3π)
     1
    10
     4
    ρ
    SN
    2 e 4
    Mpc
     3
    yr
     1
    .
    Note that
    ρ
    SN
    2 e 4
    = 1
    corresponds to one SN per year
    in a 10 Mpc sphere, about the rate of all core-collapse
    SNe in the local Universe [12]. Since we expect only
    a subset of SNe to produce high energy emission, one
    can assume
    ρ
    SN
    2 e 4
    < 1
    . The second parameter is the
    hadronic jet energy
    E
    jet
    = 3 · 10
    51
    ǫ
    jet
    3 e51
    ergs
    and we
    choose to scale the flux normalization of the model
    of Ando & Beacom,
    F
    0
    , by
    ǫ
    jet
    3 e51
    . Fig. 2 shows the
    constraints that one can place on the density and jet
    kinetic energy in the
    E
    jet
     ρ
    plane. The basic shape of
    the constraints that can be obtained in the
    E
    jet
     ρ
    plane
    can be understood from the following considerations.
    The number of neutrinos depends on the jet energy and
    the distance:
    N
    ν
    ∝ ǫ
    jet
    3 e51
    · r
     2
    .
    The program requires
    at least
    N
    ν,min
    = 2
    detected neutrinos in IceCube. A
    SN with jet energy
    ǫ
    jet
    3 e51
    produces
    N
    ν,min
    neutrinos if
    it is closer than
    r
    max
    :
    N
    ν,min
    ∝ ǫ
    jet
    3 e51
    · r
     2
    max
    ,
    which
    yields
    r
    max
    ∝ (ǫ
    jet
    3 e51
    )
    1/2
    . The volume
    V
    limited by
    r
    max
    contains
    N
    SN
    ∝ ρ
    SN
    2 e 4
    · r
    3
    max
    SN that can pro-
    duce two neutrinos. Therefore the number of detection
    N
    SN
    IC/ROTSE
    is given by
    N
    SN
    IC/ROTSE
    ∝ ρ
    SN
    2 e 4
    ·(ǫ
    jet
    3 e51
    )
    3/2
    .
    For normalization we use Ando & Beacom-like SNe,
    which occur at a rate of
    ρ
    SN
    2 e 4
    = 1
    with GRB-like
    energies (
    ǫ
    jet
    3 e51
    = 1
    ) and yield
    N
    SN,AB
    IC/ROTSE
    = 200
    expected IceCube/ROTSE coincidences per year.
    N
    SN
    IC/ROTSE
    = N
    SN,AB
    IC/ROTSE
    ρ
    SN
    2 e 4
    · (ǫ
    jet
    3 e51
    )
    3/2
    (1)
    According to [11] a non-detection limits the number of
    IceCube/ROTSE coincidences at a 90% confidence level
    to
    N
    SN
    IC/ROTSE
    < 2.44
    . Using Eq. 1 one obtains the
    two-dimensional constraints on density and hadronic jet
    energy for this model:
    ρ
    SN
    2 e 4
    jet
    3 e51
    )
    3/2
    < 2.44/N
    SN,AB
    IC/ROTSE
    < 0.012,
    (2)
    Fig. 2: Sensitivity in the
    E
    jet
     ρ
    plane after one year
    of operation of the 40 string IceCube detector (dashed
    line—90% CL; solid line—one coincident detection per
    year).
    which is a reasonably good representation of the two
    dimensional constraints for not too small densities
    ρ
    SN
    2 e 4
    >∼
    10
     3
    . For GRB-like energies (
    ǫ
    jet
    3 e51
    = 1
    ), it
    follows that at most one out of 80 SNe produces Ando &
    Beacom-like jets in its core. Phrased in absolute terms, if
    no SN will be detected, the rate of SNe with a mildly rel-
    ativistic jet should not exceed
    ρ = 3.1·10
     6
    Mpc
     3
    yr
     1
    (at 90% confidence level) in our program. The cut-off at
    small densities visible in Fig. 2 is due to ROTSE-III’s
    limiting magnitude. The sphere (i.e. effective volume)
    within which ROTSE-III can detect Supernovae has a
    radius of about 200-300 Mpc. ROTSE-III effectively
    cannot probe SN subclasses that occur less then once
    per year within this sphere.
    B. Significance in case of a detection
    Next we address the case that a SN was detected
    in the follow-up observations. The task mainly consists
    of computing the significance of the coincidence. We
    compute this for one year of data and 25 alerts. Each
    alert leads to the observation of a
    ∆Ω = 1.85
    ×1.85
    =
    3.4
    square degree field, hence over the course of the
    year ROTSE-III covers a fraction of the sky given
    by
    ∆Ω/4π × N
    alerts
    = 2.1 · 10
     3
    . Next assume that
    the time window for a coincident of an optical SN
    detection and candidate neutrino multipet is given by
    ∆t
    d
    , the accuracy with which we can determine the
    initial time of the supernova explosion. Studying the
    lightcurve of supernova SN2008D, which has a known
    explosion start-time given by an initial x-ray flash, we
    have developed an accurate way to estimate
    ∆t
    d
    from
    a SN lightcurve [2]. We fit the light curve data to a
    model that postulates a phase of blackbody emission
    followed by a phase dominated by pure expansion of the
    luminous shell. Explosion times can be determined from
    the lightcurve with an accuracy of less than 4 hours. A
    detailed description of this method can be found in [2].

    4
    AUTHOR
    et al.
    PAPER SHORT TITLE
    The number of accidental SNe found will be propor-
    tional to
    ∆t
    d
    and the total number of SNe per year that
    ROTSE-III would have sensitivity to detect, if surveying
    the sky at all times,
    N
    ROTSE
    ≈ 10
    4
    . Putting all this
    together the number of random coincidences is:
    N
    bg
    = N
    alerts
    N
    ROTSE
    ∆Ω
    ×
    ∆t
    d
    yr
    = 0.056
    ∆t
    d
    d
    .
    (3)
    For
    N
    bg
    ≪ 1
    this corresponds to the chance probability
    p = 1  exp( N
    bg
    ) ≈ N
    bg
    of observing at least one
    random background event. For
    ∆t
    d
    = 1d
    and no other
    information, the observation of a SN in coincidence with
    a neutrino signal would have a significance of about 2
    σ
    .
    The significance can be improved by adding neutrino
    timing information as well as the distance information
    of the object found. We first discuss the extra timing
    information. So far we have only required that two
    neutrinos arrive within 100 s to produce an alert. Thus,
    in the analysis presented above, the significance for two
    events 1 s apart would be the same as for 99 s difference.
    Since the probability
    p
    t
    to find a time difference less
    than
    ∆t
    ν
    due to a background fluctuation is given by
    p
    t
    = ∆t
    ν
    /100s
    assuming a uniform background, we in-
    clude the time difference in the chance probability. Next
    we discuss the use of the SN distance. One can safely
    assume that there will be a strong preference for nearer
    SNe, since these are most likely to lead to a neutrino
    flux large enough to produce a multiplet in IceCube.
    Using the distance
    d
    SN
    as an additional parameter one
    can compute the probability to observe a background SN
    at a distance
    d ≤ d
    SN
    . The probability is given by the
    ratio of SNe observed by ROTSE-III within the sphere
    d
    SN
    to all SNe:
    p
    d
    = N
    ROTSE
    (d)/N
    ROTSE
    . In case of
    a detection both
    d
    SN
    and
    ∆t
    ν
    will be available. We use
    a simple Monte Carlo to obtain the significance of this
    detection. For example the detection of two neutrinos
    with a temporal difference of
    ∆t
    ν
    =
    10 s in coincidence
    with a SN in
    d
    SN
    =
    20 Mpc distance has a p-value of
    5 · 10
     4
    , which corresponds to
    3.5σ
    , assuming a total of
    N
    alerts
    = 25
    alerts found in the period of one year.
    V. COINCIDENCES WITH GCN-GRBS
    According to current models, about every 15-20th
    GRB that can be detected by IceCube will produce
    a neutrino doublet. Hence there is a small possibility
    that we will find a doublet in coincidence with a GCN
    alert, a case that we consider separately here. The
    significance of such a coincidence can be estimated with
    calculation analogous to Eq. 3. The number of accidental
    coincidences with a time difference less than
    ∆t
    is given
    by:
    N
    bg
    = N
    alerts
    N
    GCN
    ∆Ω
    ×
    ∆t
    yr
    = 3.2 · 10
     8
    ∆t
    1s,
    .
    (4)
    where we have assumed 200 GCN notices and 30
    multiplets a year. A coincidence occurs whenever the
    neutrinos and the GRB overlap within predefined win-
    dows in direction and time. For illustrative purposes, if
    we choose a 1.5-degree directional window and a 4-
    hour time window (corresponding roughly to IceCube’s
    point spread function and to GRB observations and
    modeling), Eq. 4 yields an expected background count
    of
    N
    BG
    = 4.7 · 10
     4
    . This corresponds to a
    3.5σ
    effect, or equivalently the expectation of a false positive
    from background once every 2100 years. We can further
    reduce the expected background by assuming that the
    neutrino signal is most likely to be emitted at the same
    time as the gamma rays. Since the background multiplets
    will be distributed uniformly across the 4-hour window,
    we can multiply the chance probability above by the
    factor
    p
    t
    =
    ?
    ?
    ?
    ?
    t
    GRB
     t
    ν
    4
    hours
    ?
    ?
    ?
    ?
    (5)
    where the absolute value is taken since we assume
    the neutrinos are equally likely to be emitted before
    the gamma-rays as they are after. Note that our flat
    probability assumption for the relative emission times of
    gamma rays and neutrinos from GRBs can, of course, be
    modified to follow any particular theoretical model. With
    all these assumptions, if we observe a coincidence that
    is 300 seconds from the GRB onset time, the chance
    probability is then given by
    N
    BG
    · p
    t
    = 4.7 · 10
     4
    ·
    300/14400 = 9.8 · 10
     6
    , which corresponds to a
    4.4σ
    result.
    VI. CONCLUSION
    We have presented the setup and performance of
    IceCube’s optical follow-up program, which was started
    in October 2008. The program increases IceCube’s sen-
    sitivity to transient sources such as SNe and GRBs and
    furthermore allows the immediate identification of the
    source. Non-detection of an optical counterpart allows
    the calculation of a limit on model parameters such as
    jet energy and density of SN accompanied by jets.
    In addition multiplets of neutrinos are tested for coinci-
    dences with GCN messages. Even a single coincidence
    detection would be significant.
    VII. ACKNOWLEDGMENTS
    A. Franckowiak and M. Kowalski acknowledge the
    support of the DFG. D. F. Cowen thanks the Deutscher
    Akademischer Austausch Dienst (DAAD) Visiting Re-
    searcher Program and the Fulbright Scholar Program.
    REFERENCES
    [1] A. Achterberg et al. [IceCube Collaboration], Astropart.Phys.
    26:155-173, 2006
    [2] D. F. Cowen, A. Franckowiak and M. Kowalski, arXiv:0901.4877
    [3] C. W. Akerlof
    et al.
    , PASP 115:132-140, 2003
    [4] M. Kowalski, A. Mohr, Astropart.Phys. 27:533-538,2007
    [5] S. Ando, J. Beacom, Phys.Rev.Lett. 95:061103,2005
    [6] S. Razzaque, P. Meszaros, E. Waxman, Phys.Rev.Lett.
    93:181101, 2004
    [7] S. Horiuchi, S. Ando, Phys.Rev. D77:063007,2008.
    [8] F. Yuan, C. W. Akerlof, Astropart.Phys. 677:808-812,2008
    [9] J. Becker, Phys.Rept. 458:173-246,2008
    [10] F. Halzen, D. Hooper, Rept.Prog.Phys. 65:1025-1078, 2002
    [11] G. J. Feldman, R. D. Cousins, Phys.Rev. D57:3873-3889,1998
    [12] S. Ando, F. Beacom, H. Yuksel, Phys.Rev.Lett. 95:171101, 2005
    [13] A. Achterberg et al. [IceCube Collaboration], APJ 674:357-370,
    2007

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁOD´ Z´ 2009
    1
    Results and Prospects of Indirect Searches for Dark Matter with
    IceCube
    Carsten Rott
    and Gustav Wikstrom¨
    for the IceCube collaboration
    .
    Dept. of Physics and Center for Cosmology and Astro-Particle Physics, Ohio State University, Columbus, OH 43210, USA
    Oskar Klein Centre and Dept. of Physics, Stockholm University, SE-10691 Stockholm, Sweden
    See special section of these proceedings
    Abstract. Dark matter could be indirectly detected
    through the observation of neutrinos produced as
    part of its self-annihilation process. Possible signa-
    tures are an excess neutrino flux from the Sun, the
    center of the Earth or from the galactic halo, where
    dark matter could be gravitationally trapped. We
    present a search for muon neutrinos from neutralino
    annihilations in the Sun performed on IceCube data
    collected with the 22-string configuration. No excess
    over the expected atmospheric background has been
    observed and upper limits at 90% confidence level
    have been obtained on the annihilation rate and
    converted to limits on WIMP-proton cross-sections,
    for neutralino masses in the range of 250 GeV to
    5 TeV. Further prospects for the detection of dark
    matter from the Sun, the Earth, and the galactic halo
    will be discussed.
    Keywords: Dark Matter, Neutrinos
    I. INTRODUCTION
    The existence of dark matter can be inferred from
    a number of observations, among them rotational pro-
    files of galaxies, large scale structures, and WMAP’s
    anisotropy measurement on the cosmic microwave back-
    ground. Weakly Interacting Massive Particles (WIMPs),
    ‘cold’ thermal relics of the Big Bang, are leading dark
    matter candidates. Besides overwhelming observational
    evidence for its existence, the properties of dark matter
    can only be understood through detection of direct or
    indirect signals from its interactions or through the
    production at collider experiments. In the Minimal Su-
    persymmetric Standard Model (MSSM) the neutralino
    is a promising WIMP particle. It is stable and can
    annihilate pair-wise into Standard Model particles [1].
    Galactic WIMPs could be gravitationally captured in
    the Sun or the Earth and accumulated in their cores
    [2]. Among the secondary products from the WIMP
    annihilations we expect neutrinos, which could escape
    from the center of the Sun or Earth and be detected in
    neutrino telescopes. Neutrinos are also expected from
    annihilations in the galactic halo. In IceCube [3] we
    observe Cherenkov light from relativistic muons in ice.
    The data analysis is focused on selecting upward-going
    events in order to separate muons from neutrino in-
    teractions from background muons created in cosmic-
    ray air showers. In this paper we present a search
    for a neutralino annihilation signal from the Sun with
    the IceCube 22-string detector. Future sensitivities of
    IceCube to this signal are discussed, as well as the
    prospects of observing annihilation signals from the
    Earth or the galactic halo.
    II. THE ICECUBE NEUTRINO TELESCOPE
    The IceCube Neutrino Telescope is a multipurpose
    detector under construction at the South Pole, which
    is currently about three quarter completed [3]. Upon
    completion in 2011, IceCube will instrument a volume
    of approximately one cubic kilometer of ice utilizing
    86 strings, each instrumented with 60 Digital Opti-
    cal Modules (DOMs). Eighty of these strings will be
    arranged in a hexagonal pattern with an inter-string
    spacing of about 125 m and with 17 m vertical separation
    between DOMs, at a depth between 1450 m and 2450 m.
    Complementing this 80 string baseline design will be
    a deep and dense sub-array named DeepCore [4] that
    will be formed out of seven regular IceCube strings
    in the center of the array together with six additional
    strings deployed in between them. In this way, the sub-
    array will achieve an interstring-spacing of 72 m. The
    six additional DeepCore strings will have a different
    distribution of their 60 DOMs, optimizing their design
    towards a lower energy threshold. The optical sensors
    will have a vertical spacing of 7 m, will be deployed in
    deep transparent ice
    1
    and will consist of high quantum
    efficiency photomultiplier tubes (HQE PMTs). This will
    enable us to study neutrinos at energies down to a few 10
    GeV. DeepCore will be an extremely interesting detector
    for the study of WIMPs.
    III. ANALYSIS OF 22-STRING DATA
    The 2007 dataset, consisting of 104.3 days livetime
    with the Sun below the horizon recorded with the
    IceCube 22-string detector, was searched for a neutrino
    signal from the Sun [5]. The event sample was reduced
    in steps from 4.8 · 10
    9
    to 6946 events at final level,
    which constitutes the expected sample of atmospheric
    1
    The deep ice is clearer, with a scattering length roughly twice
    that of the upper part of the IceCube detector. In addition, the deeper
    location (below 2000 m) provides an improved shielding of cosmic
    ray backgrounds.

    2
    C. ROTT et al. DARK MATTER IN ICECUBE
    cos(
    Ψ
    )
    0.99
    0.992
    0.994
    0.996
    0.998
    1
    Events
    0
    5
    10
    15
    20
    25
    30
    35
    cos(
    Ψ
    )
    0.99
    0.992
    0.994
    0.996
    0.998
    1
    Events
    0
    5
    10
    15
    20
    25
    30
    35
    Data
    Background
    WIMP 1000 GeV, hard
    Fig. 1. Cosine of the angle to the Sun, Ψ, for data (squares) with
    one standard deviation error bars, and the atmospheric background
    expectation (dashed line). Also shown is a simulated signal (m
    χ˜
    0
    1
    =
    1000 GeV, hard spectrum) scaled to the found upper limit of µ
    s
    = 6.8
    events.
    neutrinos with a contamination of atmospheric muons.
    Since the analysis is based on comparing the shape of
    the angular distribution of signal and background (see
    section III-B), there is no need to achieving a high purity
    atmospheric neutrino sample at final cut level. Filtering
    was based on log-likelihood muon track reconstructions,
    geometry, and time evolution of the hit pattern. Events
    were required to have a good quality track reconstruction
    with a zenith angle in the interval 90
    to 120
    . Multi-
    variate training and selection was done with the help of
    Support Vector Machines [6]. At the final stages in the
    analysis, randomized real data were used to model the
    atmospheric background.
    A. Simulations
    Five WIMP masses: 250, 500, 1000, 3000, and 5000
    GeV were simulated using WimpSim [7] in two an-
    nihilation channels, bb (soft channel), and W
    +
    W
    (hard channel), representing the extremes of the neu-
    trino energy distributions. Single and coincident shower
    atmospheric muon backgrounds were simulated using
    CORSIKA [8]. The atmospheric neutrino background
    was simulated [9] following the Bartol flux [10].
    Charged particle propagation [11] and photon propa-
    gation [12], using ice measurements [13], were also
    simulated.
    B. Results
    The final data sample was used to test the hypothesis
    that it contains a certain signal level, against the null
    hypothesis of no signal. The shape of the angular distri-
    bution of events with respect to the Sun was used as a
    test statistic (see Figure 1). The background-only p.d.f.
    was constructed from data with randomized azimuth
    angles, while the p.d.f.s for the different signal models
    tested were obtained from Monte Carlo. A limit was
    set on the relative strength of the signal p.d.f. using
    Neutralino mass (GeV)
    10
    2
    10
    3
    10
    4
    10
    )
    -1
    y
    -2
    Muon flux from the Sun (km
    10
    2
    3
    10
    10
    4
    5
    10
    lim
    CDMS(2008)+XENON10(2007)
    SI
    SI
    <
    σ
    σ
    lim
    CDMS(2008)+XENON10(2007)
    SI
    SI
    < 0.01xσ
    σ
    10
    2
    10
    3
    10
    4
    10
    10
    2
    3
    10
    10
    4
    5
    10
    2
    < 0.20
    0.05 <
    Ω
    χ
    h
    BAKSAN 1978-1995
    MACRO 1989-1998
    SUPER-K 1996-2001
    AMANDA-II 2001 (hard)
    IceCube-22 2007 (soft)
    IceCube-22 2007 (hard)
    IceCube-80+DeepCore 1825d sens. (hard)
    thr
    = 1 GeV
    Indirect searches - E
    μ
    Fig. 2. Upper limits at the 90% confidence level on the muon flux
    from neutralino annihilations in the Sun for the soft (bb) and hard
    (W
    +
    W
    ) annihilation channels, adjusted for systematic effects, as
    a function of neutralino mass [5]. For neutralino masses below m
    W
    τ
    +
    τ
    is used as the hard annihilation channel. The lighter [green] and
    darker [blue] shaded areas represent MSSM models not disfavored by
    direct searches [20], [21] based on σ
    SI
    and 100· σ
    SI
    , respectively. A
    muon energy threshold of 1 GeV was used when calculating the flux.
    Also shown are the limits from BAKSAN [15], MACRO [16], Super-
    K [17], and AMANDA [18], and the expected sensitivity of IceCube
    with DeepCore.
    a Feldman-Cousins [14] confidence interval construc-
    tion. These limits were transformed to a limit on the
    muon flux above 1 GeV, which is shown in Figure 2
    together with previous limits [15], [16], [17], [18],
    MSSM models [19], and a conservative estimate of
    the full IceCube sensitivity including DeepCore. The
    models shown are those not excluded by CDMS [20] and
    XENON10 [21] based on the spin-independent WIMP-
    proton cross-section. Models in the darker region require
    a factor of 100 increase in sensitivity of direct detection
    experiments in order to be probed by them. Assuming
    that WIMPs are in equilibrium in the Sun, the limit
    on the muon flux can be converted to a limit on the
    spin-dependent WIMP-proton cross-section [22]. These
    limits are shown in Figure 3 together with previous
    limits [17], [20], [23], [24], MSSM models, and the
    IceCube sensitivity.
    IV. E
    ARTH
    WIMP
    S
    Dark matter could also be gravitationally trapped at
    the center of the Earth. Such scenarios are generally
    not favored due to its less efficient capture of dark
    matter. However, from an experimental point of view,
    the searches from dark matter from the center of the
    Earth are still of interest due to the many unknowns that
    plague the relic, capture and annihilation processes that
    enter into the calculation of the expected fluxes. In order
    to search for an indirect signal from dark matter anni-
    hilation from the Earth, IceCube uses muon neutrinos
    ν
    µ
    that interact in or below the IceCube detector. They
    produce vertically up-going track-like events, that point
    back to the center of Earth. We have designed a string
    trigger [25] for IceCube that is specifically optimized for

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁOD´ Z´ 2009
    3
    Neutralino mass (GeV)
    10
    2
    10
    3
    10
    4
    10
    )
    2
    Neutralino-proton SD cross-section (cm
    10
    -41
    -40
    10
    -39
    10
    -38
    10
    -37
    10
    -36
    10
    -35
    10
    -34
    10
    -33
    10
    lim
    CDMS(2008)+XENON10(2007)
    SI
    SI
    <
    σ
    σ
    lim
    CDMS(2008)+XENON10(2007)
    SI
    SI
    < 0.01xσ
    σ
    2
    < 0.20
    0.05 <
    Ω
    χ
    h
    CDMS (2008)
    COUPP (2008)
    KIMS (2007)
    SUPER-K 1996-2001
    IceCube-22 2007 (soft)
    IceCube-22 2007 (hard)
    IceCube-80+DeepCore 1825d sens. (hard)
    10
    2
    10
    3
    10
    4
    10
    10
    -41
    -40
    10
    -39
    10
    -38
    10
    -37
    10
    -36
    10
    -35
    10
    -34
    10
    -33
    10
    10
    2
    10
    3
    10
    4
    10
    10
    -41
    -40
    10
    -39
    10
    -38
    10
    -37
    10
    -36
    10
    -35
    10
    -34
    10
    -33
    10
    Fig. 3.
    Upper limits at the 90% confidence level on the spin-
    dependent neutralino-proton cross-section σ
    SD
    for the soft (bb) and
    hard (W
    +
    W
    ) annihilation channels, adjusted for systematic effects,
    as a function of neutralino mass [5]. The lighter [green] and darker
    [blue] shaded areas represents MSSM models not disfavored by direct
    searches [20], [21] based on σ
    SI
    and 100 · σ
    SI
    , respectively. Also
    shown are the limits from CDMS [20], COUPP [24], KIMS [23] and
    Super-K [17], and the expected sensitivity of IceCube with DeepCore.
    /(GeV))
    primary
    log10(E
    1.5
    2
    2.5
    3
    3.5
    4
    Efficiency increase
    1
    10
    Fig. 4.
    Impact of the string trigger for the detection of vertically
    up-going muon neutrinos as function of their energy. The efficiency
    increase is shown compared to IceCube’s multiplicity eight DOM
    trigger.
    this class of events. A similar trigger has also been active
    in AMANDA. The IceCube string trigger, which selects
    events with a cluster of hits on a single string, has been
    active since spring 2008. It requires 5 DOMs to be above
    threshold in a series of 7 consecutive DOMs, within a
    time window of 1.5 µs. Due to the low noise environ-
    ment and this special trigger, IceCube has an energy
    threshold for these vertical events that can reach below
    100 GeV. The increase in efficiency to these events,
    over the default DOM multiplicity trigger condition, is
    shown in Figure 4. Based on selection criteria optimized
    for these vertically up-going events [30], we will derive
    a sensitivity for the detection of a possible additional
    muon flux. These results will be shown at the time of
    the conference.
    Interpreting a possible muon flux (induced from muon
    neutrino interactions in or below the IceCube detector)
    from WIMP annihilation in the Earth is somewhat more
    complicated compared to the solar WIMP searches. The
    escape velocity is relatively small (v ? 15 km/s at
    the center) and capture is only possible for low speed
    WIMPs unless its mass is nearly identical to that of one
    of the nuclear species in the Earth. WIMPs are typically
    only expected to be captured after they are bound to
    the solar system due to previous scattering in the Sun;
    such capture mechanisms are described in [26], [27],
    [28]. Contrarily to the Sun, capture and annihilation of
    WIMPs are generally not in equilibrium in the Earth.
    Hence, the expected flux of neutrinos from dark matter
    annihilations strongly depends on how much dark matter
    was previously accumulated. Models that enhance the
    collection of dark matter by the Earth therefore also
    significantly boost expected signals. One such example
    is an expected boost due to lower velocity WIMPs in
    the galactic halo from previous dwarf mergers. Such
    scenarios could boost fluxes at neutrino telescopes by
    a few orders of magnitude [29].
    Such examples show that big uncertainties remain
    in the overall flux predictions for neutrinos from the
    center of the Earth. IceCube, with the combination of
    DeepCore, is an ideal instrument to look for such signals.
    V. HALO WIMPS
    Besides searches for indirect signals from dark matter
    annihilation in the center of the Sun and Earth, another
    promising way is to look directly at the galactic halo.
    Such a signal could be seen in neutrinos as a large scale
    flux anisotropy that peaks towards the Galactic Center.
    IceCube has in the past not performed a dedicated search
    for such signals. However, theoretical predictions indi-
    cate that such a search can provide stringent limits [31],
    [32] on the dark matter self-annihilation cross-section.
    They are complementary to Solar WIMP searches, as
    they probe the dark matter self-annihilation cross-section
    directly.
    The analysis for a neutrino flux anisotropy is still on-
    going on the IceCube dataset. We perform this analysis
    on data collected with the IceCube 40-string configura-
    tion. Neutrino-induced muon events are being used to
    search for a neutrino flux anisotropy towards the direc-
    tion of the Galactic Center. In its current configuration,
    IceCube can only access up-going muon neutrinos for
    the energy range of interest (around and below a TeV)
    with sufficient background rejection. The region closest
    towards the Galactic Center, accessible in IceCube with
    up-going events, is therefore near the horizon. It covers,
    in part, a distance of about 30
    towards the Galactic
    Center. Using a declination band simplifies the back-
    ground estimation, as an on and off-source comparison
    can be performed. Second order effects need to be taken

    4
    C. ROTT et al. DARK MATTER IN ICECUBE
    into account; these include uneven detector exposure-
    times, as the reconstruction efficiency is a function of
    the azimuth angle for the tracks in the same declination
    band. We plan to present the sensitivity using this
    method at the time of the conference for different signal
    distributions within the declination bands [33].
    For the future, DeepCore is especially promising for
    the halo WIMP searches, as it lowers the neutrino energy
    threshold, holds promises for cascade reconstruction and
    will allow observation of the entire sky. The lower
    energy threshold will increase expected signal rates,
    especially to WIMPs with masses of a few hundred GeV
    and scale with the increase in neutrino effective area.
    Leading dark matter candidates have masses in the sub-
    TeV range, so the expected neutrino energy spectrum
    is at the low energy end of IceCube’s sensitivity. The
    detection of low energy cascades caused by ν
    τ
    , ν
    e
    charged current interactions or neutral current of all
    neutrino flavors is especially interesting, as the atmo-
    spheric neutrino background to this signal is much lower
    than the muon neutrino background. Even a very limited
    angular resolution for these cascades, which IceCube
    might be able to achieve, would benefit the analysis, as
    it is looking for a large scale anisotropy. The usage of
    surrounding IceCube strings as veto against down-going
    muons in DeepCore is expected to give large reductions
    in this background and enable us to study the entire
    sky. Simple veto methods have achieved background
    reductions of four orders of magnitude with excellent
    signal retention and have potential for greater than 6
    orders of magnitude rejection utilizing reconstruction
    veto methods [34]. Since the Galactic Center, for which
    the largest flux from dark matter annihilation is expected,
    is located in the southern hemisphere, this will benefit
    the analysis in particular.
    Expected neutrino fluxes from dark matter self-
    annihilations in the galactic halo are generally small.
    Results from PAMELA [35] and Fermi [36] might indi-
    cate larger than usual self-annihilation cross-sections of
    the halo dark matter, this could either be due to unusually
    large boost factors (clumpiness) well above expectations
    from dark matter halo simulations, or due to an enhance-
    ment in the self-annihilation cross-section (for example
    Sommerfeld enhancement). Lepton results could also
    be entirely explainable by astronomical sources (for
    example pulsars [37]). Regardless of what the source
    of the recent excess is, it only shows there remains a
    large uncertainty in any flux predictions for neutrinos
    from dark matter annihilations or decays in the galactic
    halo. This, it will be important to check for any such
    signals with neutrinos.
    VI. SUMMARY
    IceCube has set the best limits to date on WIMP
    annihilation in the Sun using 22-string data from 2007.
    Using data from the completed 86-string detector, which
    will include the DeepCore low-energy extension, im-
    provements of an order of magnitude are expected.
    Searches for signals from the Earth and the galactic halo
    are also expected to give interesting results.
    REFERENCES
    [1] G. Jungman, M. Kamionkowski and K. Griest, Phys. Rep. 267,
    195 (1996).
    [2] W. H. Press and D. N. Spergel, Astrophys. J 296, 679 (1985).
    [3] A. Achterberg et al., Astropart. Phys. 26, 155 (2006).
    [4] D. Cowen for the IceCube coll., Proc. of NEUTEL09, (2009).
    [5] R. Abbasi et al., Phys. Rev. Lett. 102, 201302 (2009).
    [6] S. S. Kerthi et al., Neural Comp. 13, 637 (2001).
    [7] M. Blennow, J. Edsjo,¨ T. Ohlsson, JCAP 01, 021 (2008).
    [8] D. Heck et al., FZKA Report 6019 (1998).
    [9] A. Gazizov and M. Kowalski, Comput. Phys. Commun. 172, 203
    (2005).
    [10] G. D. Barr et al., Phys. Rev. D 70, 023006 (2004).
    [11] D. Chirkin and W. Rhode, hep-ph/0407075v2.
    [12] J. Lundberg et al., Nucl. Instr. Meth. A 581, 619 (2007).
    [13] M. Ackermann et al., J. Geophys. Res. 111, 02201 (2006).
    [14] G. J. Feldman and R. D. Cousins, Phys. Rev. D 57, 3873 (1998).
    [15] M. M. Boliev et al., Nucl. Phys. Proc. Suppl. 48, 83 (1996).
    [16] M. Ambrosio et al., Phys. Rev. D 60, 082002 (1999).
    [17] S. Desai et al., Phys. Rev. D 70, 083523 (2004).
    [18] M. Ackermann et al., Astropart. Phys. 24, 459 (2006).
    [19] P. Gondolo et al., JCAP 0407, 008 (2004).
    [20] Z. Ahmed et al., astro-ph/0802.3530.
    [21] J. Angle et al., Phys. Rev. Lett. 100, 021303 (2008).
    [22] G. Wikstrom¨ and J. Edsjo,¨ JCAP 04, 009 (2009).
    [23] H. S. Lee et al., Phys. Rev. Lett. 99, 091301 (2007).
    [24] E. Behnke et al., Science 319, 933 (2008).
    [25] A. Gross et al. for the IceCube coll., astro-ph/0711.0353.
    [26] J. Lundberg and J. Edsjo,¨ Phys. Rev. D 69, 123505 (2004).
    [27] A. Gould, Astrophys. J. 328, 919 (1988).
    [28] A. H. G. Peter, astro-ph/0902.1348.
    [29] T. Bruch, A. H. G. Peter, J. Read, L. Baudis and G. Lake, astro-
    ph/0902.4001.
    [30] C. Rott, for the IceCube coll., these proceedings.
    [31] H. Yuksel, S. Horiuchi, J. F. Beacom and S. Ando, astro-
    ph/0707.0196.
    [32] J. F. Beacom, N. F. Bell and G. D. Mack, Phys. Rev. Lett. 99,
    231301 (2007).
    [33] G. Mack, and C. Rott, forthcoming.
    [34] D. Grant et al., for the IceCube coll., these proceedings.
    [35] O. Adriani et al., Nature 458, 607 (2009).
    [36] Fermi/LAT Collaboration, arXiv:0905.0025.
    [37] H. Yuksel, M. D. Kistler and T. Stanev, astro-ph/0810.2784.

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    1
    Search for the Kaluza-Klein Dark Matter with the
    AMANDA/IceCube Detectors
    Matthias Danninger
    and Kahae Han
    for the IceCube Collaboration
    Department of Physics, Stockholm University, AlbaNova, S-10691 Stockholm, Sweden
    Department of Physics and Astronomy, University of Canterbury, Pr. Bag 4800 Christchurch, New Zealand
    See special section of these proceedings
    Abstract
    . A viable WIMP candidate, the lightest
    Kaluza-Klein particle (LKP), is motivated by theories
    of universal extra dimensions. LKPs can scatter off
    nuclei in large celestial bodies, like the Sun, and
    become trapped within their deep gravitational wells,
    leading to high WIMP densities in the object’s core.
    Pair-wise LKP annihilation could lead to a detectable
    high energy neutrino flux from the center of the Sun
    in the IceCube neutrino telescope.
    We describe an ongoing search for Kaluza-Klein
    solar WIMPs with the AMANDA-II data for the
    years 2001-2003, and also present a UED dark matter
    sensitivity projected to 180 days from a study of
    data taken with the combined AMANDA II and
    IceCube detector in the year 2007. A competitive
    sensitivity, compared to existing direct and indirect
    search experiments, on the spin-dependent cross
    section of the LKP on protons is also presented.
    Keywords
    : Kaluza-Klein, Dark Matter, IceCube
    I. INTRODUCTION
    Kaluza-Klein weakly interacting massive particles
    (WIMP) arising from theories with extra dimensions
    have come under increased scrutiny [1] alongside WIMP
    candidates from supersymmetric particle theories, e.g.
    the neutralino.
    Several analyses [2], [3] performed on the data from the
    AMANDA-II and the IceCube detectors have already put
    limits on the neutralino induced muon flux from the Sun
    comparable to that of direct detection experiments. The
    first excitation of the Kaluza Klein (KK) photon,
    B
    (1)
    ,
    in the case of Universal Extra Dimensions (UED) with
    one extra dimension, annihilates to all standard model
    particles. This results in the production of a detectable
    flux of muon neutrinos in the IceCube detector.
    B
    (1)
    is often referred to as the LKP - lightest Kaluza-
    Klein Particle. KK-momentum conservation leads to the
    stability of the LKP, which makes it a viable dark matter
    candidate. Compared to neutralino WIMPs, LKPs come
    from a relatively simple extension of the Standard Model
    and, consequently, branching ratios (see Table I) and
    cross sections are calculated with fewer assumptions and
    parameter-dependences. This feature allows us to per-
    form a combined channel analysis for an LKP particle.
    Another consequence of the simple UED model is that
    with the assumption of a compactified extra dimension
    TABLE I
    POSSIBLE CHANNELS FOR THE PAIR ANNIHILATION OF
    B
    (1)
    B
    (1)
    AND BRANCHING RATIOS OF THE FINAL STATES. FIGURES TAKEN
    FROM [20].
    Annihilation Process
    Branching ratio
    B
    (1)
    B
    (1)
    → ν
    e
    ν
    e
    ,
    ν
    µ
    ν
    µ
    ,
    ν
    τ
    ν
    τ
    0.012
    → e
    +
    e
    ,
    µ
    +
    µ
    ,
    τ
    +
    τ
    0.20
    → uu
    ,
    cc
    ,
    tt
    0.11
    → dd
    ,
    ss
    ,
    bb
    0.07
    scale of around 1TeV, the particle takes a much narrower
    range of masses [1] from the relic density calculation -
    ranging from
    600
    GeV to
    800
    GeV and
    500
    GeV to
    1500
    GeV if coannihilations are accounted for [4]. Moreover,
    collider search limits rule out LKP masses below
    300
    GeV [5], [6].
    In this paper we describe an ongoing solar WIMP
    analysis with the (2001) AMANDA data. Furthermore,
    we derive for the combined geometry of
    22
    IceCube
    strings (IC22) and AMANDA (to be referred to as the
    combined analysis in the rest of the paper) the projected
    sensitivity on the muon flux and spin-dependent (SD)
    cross section obtained for LKP WIMPs with data from
    the year 2007.
    The AMANDA-II detector, a smaller predecessor of
    IceCube with
    677
    OMs on
    19
    strings, ordered in a
    500
    m by
    200
    m diameter cylindrical lattice, has been
    fully operational since 2001 [7]. The IceCube Detector,
    with its
    59
    th
    string deployed this season, is much larger
    with increased spacing between the strings and will
    have a total instrumented volume of
    1
    km
    3
    [8]. The set-
    up in 2007 for the combined analysis consisted of
    22
    IceCube strings, and the
    19
    AMANDA strings, with
    a separate trigger and data acquisition system. The
    detector geometry for both AMANDA-II and IC22 is
    shown in Fig. 2a.
    II. SIMULATIONS
    A solar WIMP analysis can be thought of as using
    the Earth as its primary physical filter for data, as one
    only looks at data collected when the Sun is below
    the horizon at the South Pole,
    Θ
    ∈ [90
    , 113
    ]
    . Sin-
    gle
    1
    ,
    µ
    single
    , and coincident
    2
    ,
    µ
    coin
    , atmospheric muons
    that come from cosmic ray showers in a zenith angle
    1
    atmospheric muons from single CR showers
    2
    atmospheric muons from coincident CR showers

    2
    M. DANNINGER
    et al.
    KALUZA-KLEIN DARK MATTER DETECTION IN AMANDA/ICECUBE
    -200
    -100
    0
    100
    200
    300
    400
    500
    0
    100
    200
    300
    400
    500
    600
    String Position North [m]
    String Position East [m]
    IC 22
    AMANDA
    Fig. 1. Top view of the 2007 IceCube+AMANDA detector configura-
    tion. The IceCube-22 strings (squares) enclose the AMANDA strings
    (circles).
    range
    Θ
    µ
    of
    [0
    , 90
    ]
    , constitute the majority of the
    background, whereas the near-isotropic distribution of
    atmospheric neutrinos,
    ν
    atm
    , will form an irreducible
    background. The atmospheric muon backgrounds are
    generated using CORSIKA [9] with the Ho¨randel CR
    composition model [10]. For the atmospheric neutrino
    background, produced according to the Bartol model
    [11], ANIS [12] is used. For the combined analysis,
    the simulated
    µ
    single
    background has a detector-livetime
    of
    1.2
    days,
    µ
    coin
    of
    7.1
    days and
    ν
    atm
    of
    9.8
    years.
    WIMPSIM [13], [14] was used to generate the signal
    samples for LKP WIMPs, consisting of
    2
    million events
    per channel for WIMP masses varying from
    250
    GeV
    to
    3000
    GeV. Individual annihilation channels (three
    ν
    ’s,
    τ
    , and t,b,c quarks), contributing to
    ν
    µ
    ’s at the detector,
    were generated for the combined analysis, as well as
    for the AMANDA only analysis (in the latter case for
    the energy range from
    500
    GeV to
    1000
    GeV). Muon
    and Cˇ erenkov light propagation in Antarctic ice were
    simulated using IceCube/AMANDA software such as
    MMC [15], PTD and photonics [16]. Finally, AMASIM
    for AMANDA and ICESIM for the combined detector
    were used to simulate the detector response. The signal
    detection efficiency of the two detector configurations is
    given by the effective volume,
    V
    eff
    , which is defined for
    a constant generation volume,
    V
    gen
    , by
    V
    eff
    = V
    gen
    ·
    N
    obs
    N
    gen
    ,
    (1)
    where
    N
    obs
    is the number of observed LKP events and
    N
    gen
    the number of generated LKP events, undergoing
    charged-current interaction within
    V
    gen
    .
    V
    eff
    is a good
    quantity to compare LKP detectability at trigger level
    for the two analyses, shown in Fig. 2a.
    After deadtime correction, 142.5 days of data when the
    Sun was below the horizon were available in 2001 with
    a total number of
    1.46 · 10
    9
    recorded events for the
    AMANDA analysis. The combined analysis is utilizing
    a projected total livetime of
    180
    days of data for the
    calculated sensitivities in this paper.
    The main purpose of the Monte Carlo (MC) simulations
    of the various background sources is to show that a good
    agreement with experiment is achieved, demonstrating
    a sufficient understanding of the detector. Thus, it
    is viable to assume that the LKP signal samples are
    simulated correctly within the AMANDA/IceCube
    simulation-chain and can be used to select the different
    cut parameters for the higher cut levels
    L2, L3
    and
    L4
    ,
    because their difference from background in different
    parameter distributions can be clearly identified. The
    actual cut value of each cut level is obtained by
    maximizing the efficiency function, or a figure-of-
    merit, for simulated LKP signals and the experimental
    background sample, which consists of data taken when
    the Sun was above the horizon and therefore contains
    no solar WIMP signal. Setting cut values based on
    experimental background datasets has the advantage
    that possible simulation flaws are minimized.
    III. FILTERING
    LKP signals are point sources with very
    distinct directional limitations (zenith angle theta,
    Θ
    zen
    = 90
    ± 23
    ). Hence, the general strategy of
    filtering for both analyses is to apply strict directional
    cuts in early filter levels.
    L0
    and
    L1
    consist of
    calibration, reconstruction and making a simple angular
    cut of
    Θ
    zen
    > 70
    on the first-guess reconstructed
    track. This leads to a passing efficiency of around
    0.7
    for all LKP signal samples, and reduction of around
    0.002
    for both, data and muon background. All events
    passing the
    L0 + L1
    level are reconstructed using
    log-likelihood methods (llh).
    L2
    is a two dimensional
    cut on the reconstructed llh-fit zenith angle (
    Θ
    zen,llh
    )
    within
    Θ
    and the estimated angular uncertainty of
    the llh track.
    L3
    picks reconstructed tracks, which
    are nearly horizontal and pass the detector, to further
    minimize vertical tracks associated with background
    events. The multivariate filter level,
    L4
    , consists of
    two different multivariate analysis routines from the
    TMVA [17] toolkit, namely a support vector machine
    (SVM) together with a Gaussian fit-function and a
    neural network (NN). The input variables for the
    two algorithms are obtained by choosing parameters
    with low correlation but high discrimination power
    between background and signal. The individual output
    parameters are combined in one multivariate cut
    parameter
    Q
    NN
    · Q
    SVM
    .
    IV. SENSITIVITY
    After the
    L4
    cut
    3
    , the muon background reduction is
    better than a factor
    1.16 · 10
     7
    , which implies that the
    final sample is dominated by
    ν
    atm
    background. The solar
    3
    starting with
    L4
    , only the combined analysis is discussed

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    3
    LKP mass in [GeV ]
    500
    1000
    1500
    2000
    2500
    3000
    ]
    3
    effective Volume in [ km
    -3
    10
    -2
    10
    -1
    10
    AMANDA trigger level
    IceCube-22+AMANDA trigger level
    IceCube-22+AMANDA final level
    (a) Effective volume
    LKP mass in [ GeV ]
    500
    1000
    1500
    2000
    2500
    3000
    ]
    -1
    year
    -2
    Muon Flux in [ km
    3
    10
    4
    10
    Muon Flux IceCube-22+AMANDA(2007)180d sens.
    (b) Muon flux sensitivity
    Fig. 2. Fig.2a shows the effective volume as a function of LKP mass at trigger level and final cut level for the IceCube-22+AMANDA analysis
    and at trigger level only for the AMANDA analysis. Fig.2b demonstrates the projected sensitivity to
    180
    days of livetime on the muon flux
    from LKP annihilations in the Sun as a function of LKP mass for the IceCube-22+AMANDA detector configuration.
    search looks for an excess in neutrino events over the
    expected background in a specifically determined search
    cone towards the direction of the Sun with an opening
    angle
    Ψ
    . Events with a reconstructed track direction
    pointing back towards the Sun within an angle
    Ψ
    are
    kept, where
    Ψ
    is optimized to discriminate between the
    ν
    atm
    background and a sum of all seven LKP channels,
    weighted with the expected branching ratios as listed in
    Table I.
    The expected upper limit, or sensitivity, for an expected
    number of background events
    n
    Bg
    is
    µ
    90%
    s
    (n
    Bg
    ) =
    ?
    n
    obs=0
    µ
    90%
    s
    (n
    obs
    )
    (n
    Bg
    )
    n
    obs
    (n
    obs
    )!
    e
     n
    Bg
    ,
    (2)
    where
    µ
    90%
    s
    (n
    obs
    )
    is the Feldman-Cousins upper limit
    for the number of observed events,
    n
    obs
    [18]. The model
    rejection factor [19]
    MRF =
    µ
    90%
    s
    n
    s
    ,
    (3)
    is used to determine the optimum opening angle
    Ψ
    of the solar search cone. Here,
    n
    s
    is the number of
    surviving LKP events within
    Ψ
    .
    Under the assumption of no signal detection, it is
    possible to derive the Feldman-Cousins sensitivity
    discussed above for the combined detector with a total
    projected livetime of
    T
    live
    = 180
    days. The expected
    number of events after cut level
    L4
    are estimated from
    a processed subset of observational data with a detector
    livetime of
    5.61
    days. The results are then extrapolated
    to the total livetime
    T
    live
    , yielding an expectation of
    7140
    events. The corresponding expectation from the
    simulated background samples,
    n
    Bg,MC
    , normalized
    to the data at filter level
    L1
    and extrapolated to
    T
    live
    , is
    633(µ
    coin
    ) + 1038(µ
    single
    ) + 5340(ν
    atm
    ) =
    7011(n
    Bg,MC
    )
    .
    The expected sensitivity on the neutrino-to-muon
    conversion rate
    90%
    ν →µ
    is given by
    90%
    ν →µ
    =
    µ
    90%
    s
    V
    eff
    · T
    live
    ,
    (4)
    where the effective volume
    V
    eff
    is given by eq. 1. For
    each annihilation channel, one can separately calculate
    the
    V
    eff
    within the solar search cone, determined by the
    combined signal p.d.f.,
    f
    all
    S
    (x|Ψ)
    , and thereby determine
    a
    90%
    ν →µ
    for each channel. Additionally, the combined
    effective volume,
    V
    eff ,LKP
    , for the expected
    ν
    LKP
    spec-
    trum is given by the sum of the individual
    V
    eff
    per
    channel, weighted with the respective branching ratio of
    each channel. For the neutrino-to-muon conversion rate
    per single channel, the expected limit on the annihilation
    rate in the core of the Sun per second is given by,
    90%
    A
    = (c
    1
    (ch, m
    B
    (1)
    ))
     1
    ·
    90%
    ν →µ
    ,
    (5)
    where
    c
    1
    (ch , m
    B
    (1)
    )
    is an LKP annihilation channel (
    ch
    )
    and energy dependent constant. The sensitivity to the
    muon flux at a plane at the combined detector is derived
    via the calculation chain
    90%
    ν →µ
    → 
    90%
    A
    → Φ
    90%
    µ
    and
    is performed using the code described in [13], [14]. The
    results for the final
    V
    eff
    and the predicted sensitivity to a
    muon flux resulting from LKP induced annihilations in
    the Sun for the combined IC22 and AMANDA detector
    2007 with a total livetime of
    180
    days are presented in
    figures 2a and 2b. From the derived
    ν
    -to-
    µ
    conversion
    rate,
    90%
    ν →µ,LKP
    , we can calculate the sensitivity for the
    annihilation rate in the Sun per second,
    90%
    A,LKP
    .

    4
    M. DANNINGER
    et al.
    KALUZA-KLEIN DARK MATTER DETECTION IN AMANDA/ICECUBE
    LKP mass (GeV)
    2
    10
    3
    10
    4
    10
    )
    2
    LKP - proton SD cross-section (cm
    -44
    10
    -43
    10
    -42
    10
    -41
    10
    -40
    10
    -39
    10
    -38
    10
    -37
    10
    -36
    10
    -35
    10
    -34
    10
    2
    < 0.1161)
    lim
    precision data + WMAP (0.1037 <
    Ω
    h
    SD
    SD
    <
    σ
    σ
    IceCube-22+AMANDA (2007) 180d sens.
    CDMS (2008)
    COUPP (2008)
    KIMS (2007)
    Fig. 3. Theoretically predicted spin-dependent
    B
    (1)
    -on-proton elastic scattering cross sections are indicated by the shaded area [22]. The cross-
    section prediction vary with the assumed mass of the first KK excitation of the quark, constrained by
    0.01 ≤ r = (m
    q
    (1)
     m
    B
    (1)
    )/m
    B
    (1)
    0.5
    . The current ‘best’ limits, set by direct search experiments are plotted together with the sensitivity of the combined detector IceCube-
    22+AMANDA. The region below
    m
    B
    (1)
    = 300
    GeV is excluded by collider experiments [5], [6] and
    m
    B
    (1)
    > 1500
    GeV is strongly
    disfavored by WMAP observations [23].
    In [20], it is shown that the equilibrium condition be-
    tween
    A,LKP
    and the capture rate
    C
    is met by LKPs
    within the probed mass range. Furthermore, the capture
    rate of LKPs in the Sun is entirely dominated by the
    spin-dependent component of the
    B
    (1)
    -on-proton elastic
    scattering [21]. Consequently, presuming an equilibrium
    of
    A,LKP
    = C
    , the sensitivity for the spin-dependent
    elastic scattering cross section
    4
    of
    B
    (1)
    can be calculated
    as,
    σ
    H ,SD
    ≃ 0.597·10
     24
    pb
    ?
    m
    B
    (1)
    1TeV
    ?
    2
    ·
    90%
    A ,LKP
    s
     1
    ?
    .
    (6)
    The estimated sensitivity for the spin-dependent cross
    section for LKPs is displayed in figure 3, along with the
    most recently published limits from direct search ex-
    periments. The theoretical spin-dependent cross section
    predictions (shaded area) for LKPs are taken from [22]
    and are plotted for different predictions for the mass of
    the first KK-excitation of the quark.
    V. CONCLUSION AND OUTLOOK
    We showed that a competitive result on the spin-
    dependent cross-section of LKP-on-proton scattering
    can be obtained with the combined geometry of
    AMANDA-II and IceCube-22, which explores parts
    of the unrejected regions in the theoretically predicted
    LKP-region.
    We also described the ongoing solar WIMP analysis
    4
    The local density of DM in our galaxy is taken to match the mean
    density
    ρ
    DM
    = 0.3
    GeV/c
    2
    cm
    3
    , and the rms velocity is set to
    v =
    270
    km/s.
    of the AMANDA-II data taken during 2001. This
    will be extended to include 2002 and 2003 data.
    Furthermore, as the energy signature of
    ν
    µ
    ’s induced
    by LKP annihilations in the Sun is very hard, the
    fullsized IceCube-80 detector will markedly improve
    the sensitivity and set strong limits on LKP WIMP
    theories.
    REFERENCES
    [1] D.Hooper
    et al.
    ,
    The PAMELA and ATIC Signals from Kaluza-
    Klein Dark Matter
    http://arxiv.org/abs/0902.0593.
    [2] J.Braun
    et al.
    , (Searches for WIMP DM from the Sun with
    AMANDA) ICRC (2009).
    [3] R.Abbasi
    et al.
    , (the IceCube collaboration) arXiv:astro-
    ph/0902.2460 (14 Feb 2009).
    [4] T.Appelquist
    et al.
    ,
    Physical Review D
    64
    (2001) 035002.
    [5] I.Gogoladze
    et al.
    ,
    Physical Review D
    74
    (2006) 093012.
    [6] T.Appelquist
    et al.
    ,
    Physical Review D
    67
    (2003) 055002.
    [7] E.Andres
    et al.
    , (the AMANDA collaboration)
    Astrop. Phys.
    13
    (2000) 1.
    [8] A.Achterberg
    et al.
    , (the IceCube collaboration)
    Astrop. Phys.
    26
    (2006) 155.
    [9] D.Heck
    et al.
    ,
    FZKA Report
    6019
    , Forschungszentrum Karlsruhe
    (1998).
    [10] J.H
    randel,
    Astrop. Phys.
    19
    (2003) 193.
    [11] G.D.Barr
    et al.
    ,
    Phys. Review
    D70
    (2004) 023006.
    [12] A.Gazizov
    et al.
    ,
    Comput.Phys.Commun.
    172
    (2005) 203.
    [13] J.Edsj
    ,
    WimpSim Neutrino Monte Carlo
    http://www.physto.se/
    edsjo/wimpsim/.
    [14] M.Blennow
    et al.
    ,
    JCAP
    21
    (2008).
    [15] D.Chirkin,
    Cosmic ray energy spectrum measurement with
    AMANDA
    PhD Thesis, UC Berkley (2003).
    [16] J.Lundberg
    et al.
    ,
    Nucl.Instr.Meth.
    A581
    (2007).
    [17] A.H
    cker
    et al.
    , http://tmva.sourceforge.net.
    [18] G.J.Feldmann
    et al.
    ,
    Phys. Rev.
    D57
    (1998) 7.
    [19] G.C.Hill
    et al.
    ,
    Astrop. Phys.
    19
    (2003) 393.
    [20] D.Hooper
    et al.
    , arXiv:hep-ph/0208261v3 (16 Feb 2003).
    [21] D.Hooper
    et al.
    , arXiv:hep-ph/0409272v1 (23 Sep 2004).
    [22] S.Arrenberg
    et al.
    , arXiv:hep-ph/0805.4210v1 (28 May 2008).
    [23] M.Tegmark
    et al.
    ,
    Physical Review D
    74
    (2006) 123507.

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    1
    Searches for WIMP Dark Matter from the Sun with AMANDA
    James Braun
    and Daan Hubert
    for the IceCube Collaboration
    Dept. of Physics, University of Wisconsin, Madison, WI 53706, USA
    Vrije Universiteit Brussel, Dienst ELEM, B-1050 Brussels, Belgium
    See the special section of these proceedings
    Abstract
    . A well-known potential dark matter sig-
    nature is emission of GeV - TeV neutrinos from
    annihilation of neutralinos gravitationally bound to
    massive objects. We present results from recent
    searches for high energy neutrino emission from
    the Sun with AMANDA, in all cases revealing no
    significant excess. We show limits on both neutralino-
    induced muon flux from the Sun and neutralino-
    nucleon cross section, comparing them with recent
    IceCube results. Particularly, our limits on spin-
    dependent cross section are much better than those
    obtained in direct detection experiments, allowing
    AMANDA and other neutrino telescopes to search a
    complementary portion of MSSM parameter space.
    Keywords
    : AMANDA WIMP Neutralino
    I. INTRODUCTION
    Weakly interacting massive particles (WIMPs) with
    electroweak scale masses are currently a favored expla-
    nation of the missing mass in the universe. Such particles
    must either be stable or have a lifetime comparable to
    the age of the universe, and they would interact with
    baryonic matter gravitationally and through weak inter-
    actions. The minimal supersymmetric standard model
    (MSSM) provides a natural candidate, the lightest neu-
    tralino [1]. A large range of potential neutralino masses
    exists, with a lower bound on the mass of the lightest
    neutralino of 47 GeV imposed by accelerator-based
    analyses [2], while predictions based on the inferred dark
    matter density suggest masses up to several TeV [3].
    Searches for neutralino dark matter include
    direct
    searches for nuclear recoils from weak interaction of
    neutralinos with matter [4], [5] and
    indirect
    searches
    for standard model particles produced by neutralino
    annihilation. Particularly, a fraction of neutralinos inter-
    acting with massive objects would become gravitation-
    ally bound and accumulate in the center. If neutralinos
    comprise dark matter, enough should accumulate and an-
    nihilate to produce an observable neutrino flux. Searches
    for a high energy neutrino beam from the center of
    the Earth [6] and the Sun [7], [8], [9], [10], [11] have
    yielded negative results. Observations of a cosmic ray
    electron-positron excess by ATIC [12], PPB-BETS [13],
    Fermi [14], and HESS [15], along with the anomalous
    cosmic ray positron fraction reported by PAMELA [16],
    could be interpreted as an indirect signal of dark matter
    annihilation in our galaxy [17].
    Here we present searches for a flux of GeV–TeV
    neutrinos from the Sun using AMANDA. We improve on
    the sensitivity of the previous AMANDA analysis [11]
    significantly and extend the latest results from IceCube
    [7] to lower neutralino masses. We observe no neutralino
    annihilation signal and report limits on the neutrino-
    induced muon flux from the Sun and the resulting limits
    on neutralino-proton spin-dependent cross section.
    II. NEUTRINO DETECTION WITH AMANDA
    The detection of neutrino fluxes above
    ∼ 50
    GeV is a
    major goal of the Antarctic Muon And Neutrino Detector
    Array (AMANDA). AMANDA consists of 677 optical
    modules embedded 1500 m to 2000 m deep in the ice
    sheet at the South Pole, arranged in 19 vertical strings
    and occupying a volume of
    ∼ 0.02
    km
    3
    . Each module
    contains a 20 cm diameter photomultiplier tube (PMT)
    optically coupled to an outer glass pressure sphere.
    PMT pulses (“hits”) from incident Cherenkov light are
    propagated to surface electronics and are recorded as
    an event when 6–7 hits on any one string or 24 total
    hits occur within 2.5
    µ
    s. The vast majority of the
    O(
    10
    9
    ) events recorded each year are downgoing muons
    produced by cosmic ray air showers in the atmosphere
    above the South Pole. Relativistic charged leptons pro-
    duced near the detector via charged-current neutrino
    interactions similarly trigger the detector, with several
    thousand atmospheric neutrino induced muon events
    recorded per year. The hit leading edge times, along
    with the known AMANDA geometry and ice properties
    [18], allow reconstruction of muon tracks with median
    accuracy 1.5
    – 2.5
    , dependent on zenith angle.
    AMANDA operated in standalone from 2000–2006
    and is currently a subdetector of the much larger (
    km
    3
    ) IceCube Neutrino Observatory [19], scheduled
    for completion in 2011. The optical module density of
    AMANDA is much higher than that of IceCube, making
    AMANDA more efficient for low-energy muons (
    ? 300
    GeV) which emit less Cherenkov light.
    III. DATA SELECTION AND METHODS
    We describe two separate searches for Solar neu-
    tralinos in this proceeding. First, we present a search
    using a large data sample from 2000–2006 prepared
    for a high energy extraterrestrial point source search
    [20], [21]. We also present a search using data from
    2001–2003, optimized to retain low energy events [22].
    Both analyses are done in two stages; first, neutrino

    2
    BRAUN
    et al.
    AMANDA DARK MATTER SEARCHES
    induced muon events are isolated from the much larger
    background of downgoing muons, then a search method
    is used to test for an excess at the location of the Sun.
    A. Data Selection
    While the Sun is above the horizon, neutrino-induced
    muons from the Sun are masked by the much larger
    background of downgoing cosmic ray muons; thus, we
    select data during the period when the Sun is below the
    horizon (Mar. 21 – Sept. 21), resulting in 953 days live-
    time from 2000–2006 and 384 days from 2001–2003. In
    both analyses, neutrino events are isolated by selecting
    well reconstructed upgoing muon tracks. Events are first
    reconstructed with fast pattern matching algorithms, and
    events with zenith angles
    θ < 80
    (
    θ < 70
    for the
    2001-2003 analysis) are discarded, eliminating the vast
    majority of downgoing muons. The remaining events
    are reconstructed with a more computationally intensive
    maximum-likelihood reconstruction [23] accurate to 1.5
    – 2.5
    , and again events with
    θ < 80
    are discarded.
    O(
    10
    6
    ) misreconstructed downgoing muon events re-
    main per year, and these are reduced by cuts on
    track quality parameters such as track angular uncer-
    tainty [24], the smoothness (evenness) of hits along
    the track [23], and the likelihood difference between
    the maximum-likelihood track and a forced downgoing
    likelihood fit using the zenith distribution of downgoing
    muons as a prior [23]. For the 2000–2006 analysis, 6595
    events remain after quality cuts, dominantly atmospheric
    neutrinos [20], reduced to 4665 events by requiring dates
    when the Sun is below the horizon. Zenith distributions
    from 2000–2006 are shown in figure 1.
    We consider neutralino masses from 100 GeV to 5
    TeV and two extreme annihilation channels:
    W
    +
    W
    (
    τ
    +
    τ
    for 50 GeV) and
    b
    ¯
    b
    , which produce high and
    low energy neutrino spectra, respectively, relative to the
    neutralino mass. The fraction of signal events retained
    depends on the neutrino energy spectrum and varies from
    17% for a 5 TeV neutralino mass and
    W
    +
    W
    channel
    to 1% for 100 GeV and
    b
    ¯
    b
    channel, relative to trigger
    level, in the 2000–2006 analysis.
    The 2001–2003 analysis is a dedicated neutralino
    search, unlike the 2000-2006 analysis, and more con-
    sideration is given to low energy events. Twelve event
    observables are considered, and selection criteria based
    on these observables are optimized separately for three
    signal classes, dependent on neutralino mass and anni-
    hilation channel, to maximize retention of signal events.
    The signal classes are shown below along with the
    number of events passing selection criteria.
    Class
    Channel
    m
    χ
    Final Events
    A
    W
    +
    W
    250 GeV – 5 TeV
    670
    b
    ¯
    b
    500 GeV – 5 TeV
    B
    W
    +
    W
    100 GeV
    504
    b
    ¯
    b
    250GeV
    C
    τ
    +
    τ
    50 GeV
    398
    b
    ¯
    b
    50, 100 GeV
    cos
    θ
    -1
    -0.5
    0
    0.5
    1
    Events
    1
    10
    10
    2
    3
    10
    10
    4
    10
    5
    6
    10
    10
    7
    8
    10
    9
    10
    Downgoing Muons
    Misreconstructed Muons
    Atmospheric Neutrinos
    Filtered Data
    Final Sample
    Neutralino Signal
    Fig. 1. Reconstructed zenith angles of data (circles) at trigger level,
    filter level (
    θ < 80
    ), and final cut level, true (fine dotted) and
    reconstructed (solid) zenith angles of CORSIKA [31] downgoing muon
    simulation at trigger level, reconstructed zenith angles of ANIS [28]
    atmospheric neutrino simulation at trigger level, filter level, and final
    cut level (dashed), and reconstructed zenith angles of a neutrino signal
    from the Sun at final cut level (dash-dotted).
    The selection is more efficient than the 2000-2006
    analysis, with 21% of signal retained for 5 TeV,
    W
    +
    W
    channel, to 4% for 100 GeV,
    b
    ¯
    b
    channel. The 2001–
    2003 analysis additionally considers 50 GeV neutralino
    masses, with a signal efficiency of 1%–3%.
    B. Search Method
    Both analyses use maximum-likelihood methods [25]
    to search for an excess of events near the location of
    the Sun. The data is modeled as a mixture of
    n
    s
    signal
    events from the Sun and background events from both
    atmospheric neutrinos and misreconstructed downgoing
    muons. The signal likelihood for the i
    th
    event is
    S
    i
    =
    1
    2πσ
    2
    i
    e
    ψ
    2
    i
    2 σ
    2
    i
    ,
    (1)
    where
    ψ
    i
    is the space angle difference between the event
    and the Sun, and
    σ
    i
    is the event angular uncertainty
    [24]. The background likelihood
    B
    i
    is obtained from
    the zenith distribution of off-source data. The full-data
    likelihood over all
    N
    data events is
    L = P (Data|n
    s
    ) =
    ?
    N
    i=1
    ?
    n
    s
    N
    S
    i
    + (1 
    n
    s
    N
    )B
    i
    ?
    (2)

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    3
    and is numerically maximized to find the best fit event
    excess
    n
    ˆ
    s
    . The likelihood ratio
     2 log
    L(0)
    L(nˆ
    s
    )
    is approxi-
    mately
    χ
    2
    distributed and provides a measure of signif-
    icance. Event upper limits are set from this likelihood
    using the Feldman-Cousins unified construction [26].
    C. Signal Simulation and Systematic Uncertainties
    Neutrino energy distributions at Earth from neutralino
    annihilation in the Sun are generated by DarkSUSY
    [27]. For the 2000–2006 analysis, neutrino events are
    generated with ANIS [28], with muons propagated using
    MMC [29], then reweighted to the energy distributions
    described above. For 2001–2003, the DarkSUSY energy
    distributions are sampled by WimpSimp [30], and muons
    are propagated with MMC.
    Uncertainties in our signal simulation are dominated
    by uncertainties in optical module sensitivity and photon
    propagation in ice. These uncertainties are constrained
    by comparing the trigger rate of CORSIKA [31] down-
    going muon simulation using various hadronic models
    with the observed AMANDA trigger rate. The effect on
    signal prediction is measured by shifting the simulated
    optical module efficiency by these constraints and is
    10% for
    m
    χ
    = 5
    TeV,
    W
    +
    W
    channel, to 21% for
    m
    χ
    = 100
    GeV,
    b
    ¯
    b
    channel. Other sources of uncertainty
    include event selection (4%–8%) and uncertainty in neu-
    trino mixing angles (5%). For the 2000–2006 analysis,
    uncertainties total 13%–24% and are included in the
    limit calculation using the method of Conrad
    et al.
    [32]
    as modified by Hill [33]. Uncertainties for 2001–2003
    total 23%–38% and are included in the limits assuming
    the worst case.
    IV. RESULTS
    The search methods are applied to the final data, and
    both analyses reveal no significant excess of neutrino-
    induced muons from the direction of the Sun. A Sun-
    centered significance skymap from the 2000–2006 anal-
    ysis (figure 2) shows a
    0.8σ
    deficit from the direction of
    the Sun. For the 2001–2003 analysis, a deficit of events
    is observed in classes A and C, and a small excess is
    seen in class B. Each excess or deficit is within the 1
    σ
    range of background fluctuations.
    Upper limits on the neutralino annihilation rate in the
    Sun are calculated from the event upper limit
    µ
    90
    by
    A
    =
    4πR
    2
    µ
    90
    N
    A
    ρT
    L
    V
    ef f
    ?
    ?
    m
    χ
    0
    σ
    νN
    dN
    ν
    dE
    dE
    ?
     1
    ,
    (3)
    where
    R
    is the Earth-Sun radius,
    N
    A
    is the Avogadro
    constant,
    ρ
    is the density of the detector medium,
    T
    L
    is the livetime, and
    σ
    νN
    is the neutrino-nucleon cross
    section. The muon neutrino energy spectrum
    dN
    ν
    dE
    for a
    given annihilation channel is obtained from DarkSUSY
    and includes absorption and oscillation effects from tran-
    sit through the Sun and to Earth. The energy-averaged
    effective volume
    V
    ef f
    is obtained from simulation. Lim-
    its on muon flux are given by
    Φ
    µ
    =
    A
    4πR
    2
    ?
    m
    χ
    1 GeV
    dN
    µ
    dE
    dE,
    (4)
    and limits on neutralino-proton cross section are calcu-
    lated according to [34]. These quantities are tabulated in
    table I for the more restrictive of the two analyses. Muon
    flux limits, assuming a 1 GeV threshold on muon energy,
    and spin-dependent cross section limits are shown in
    figure 3 for both analyses.
    V. DISCUSSION
    These limits extend the latest IceCube limits to lower
    neutralino masses and are now beginning to exclude
    neutralino spin-dependent cross sections allowed by
    direct detection experiments (figure 3). A 1000-fold
    improvement over current direct-detection limits [4], [5]
    does not significantly constrain allowed spin-dependent
    cross sections; thus, neutrino telescopes will continue to
    observe a complementary portion of MSSM parameter
    space over the next several years. IceCube is currently
    operating with 59 strings and will contain 86 strings
    when complete in 2011. The DeepCore extension to
    IceCube [35], six strings with tighter string spacing (72
    m), tighter optical module spacing (7 m), and higher
    PMT quantum efficiency, will be complete in 2010.
    DeepCore will significantly enhance the sensitivity of
    IceCube to low energy muons, extending the reach of
    IceCube to lower neutralino masses.
    REFERENCES
    [1] M. Drees and M. M. Nojiri, Phys. Rev. D
    47
    , 376 (1993).
    [2] C. Amsler
    et al., Phys. Lett. B
    667
    , 1 (2008).
    [3] R. C. Gilmore, Phys. Rev. D
    76
    , 043520 (2007).
    [4] Z. Ahmed
    et al.
    , Phys. Rev. Lett.
    102
    , 011301 (2009).
    [5] J. Angle
    et al.
    , Phys. Rev. Lett.
    100
    , 021303 (2008).
    [6] A. Achterberg
    et al., Astropart. Phys.
    26
    , 129 (2006).
    [7] R. Abbasi
    et al.
    , arXiv:0902.2460.
    [8] M. M. Boliev
    et al.
    , Nucl. Phys. Proc. Suppl.
    48
    , 83 (1996).
    [9] M. Ambrosio
    et al.
    , Phys. Rev. D
    60
    , 082002 (1999).
    [10] S. Desai
    et al.
    , Phys. Rev. D
    70
    , 083523 (2004).
    [11] M. Ackermann
    et al.
    , Astropart. Phys.
    26
    , 155 (2006).
    [12] J. Chang
    et al., Nature
    456
    , 362 (2008).
    [13] S. Torii
    et al., arXiv:0809.0760 (2008).
    [14] A. A. Abdo et al., Phys. Rev. Lett.
    102
    , 181101 (2009).
    [15] F. Aharonian et al., arXiv:0905.0105.
    [16] O. Adriani et al., Nature
    458
    , 607 (2009).
    [17] L. Bergstro¨m, J. Edsjo¨, and G. Zaharijas, arXiv:0905.0333
    (2009).
    [18] M. Ackermann
    et al., J. Geophys. Res.
    111
    , D13203 (2006).
    [19] J. Ahrens
    et al., Astropart. Phys.
    20
    , 507 (2004).
    [20] R. Abbasi
    et al.
    , Phys. Rev. D
    79
    , 062001 (2009).
    [21] J. Braun, PhD Thesis, University of Wisconsin (2009).
    [22] D. Hubert, PhD Thesis, Vrije Universiteit Brussel (2009).
    [23] J. Ahrens
    et al., Nucl. Inst. Meth. A
    524
    , 169 (2004).
    [24] T. Neunho¨ffer, Astropart. Phys.
    25
    , 220 (2006).
    [25] J. Braun et al., Astropart. Phys.
    29
    , 299 (2008).
    [26] G.J. Feldman and R.D. Cousins, Phys. Rev. D
    57
    , 3873 (1998).
    [27] P. Gondolo
    et al., JCAP
    0407
    , 008 (2004).
    [28] A. Gazizov and M. Kowalski, Comput. Phys. Commun.
    172
    , 203
    (2005).
    [29] D. Chirkin and W. Rhode, arXiv:hep-ph/0407075 (2004).
    [30] L. Bergstro¨m, J. Edsjo¨, and P. Gondolo, Phys. Rev. D
    58
    , 103519
    (1998).
    [31] D. Heck
    et al., FZK, Tech. Rep. FZKA 6019 (1998).
    [32] J. Conrad
    et al., Phys. Rev. D
    67
    , 012002 (2003).
    [33] G.C. Hill, Phys. Rev. D
    67
    , 118101 (2003).
    [34] G. Wikstro¨m and J. Edsjo¨, JCAP
    04
    , 009 (2009).
    [35] D. Cowen
    et al., in
    Proc. of NEUTEL09
    , Venice, (2009).
    [36] H. S. Lee
    et al.
    , Phys. Rev. Lett.
    99
    , 091301 (2007).
    [37] E. Behnke
    et al.
    , Science
    319
    , 933 (2008).

    4
    BRAUN
    et al.
    AMANDA DARK MATTER SEARCHES
    Fig. 2. Sun-centered skymap of event excesses from the 2000–2006 analysis.
    m
    χ
    (GeV )
    Channel
    V
    ef f
    (m
    3
    )
    µ
    90
    A
    (s
     1
    )
    Φ
    µ
    (km
     2
    y
     1
    )
    σ
    SI
    (cm
    2
    )
    σ
    S D
    (cm
    2
    )
    50
    τ
    +
    τ
    4.31 × 10
    3
    6.2
    2.11 × 10
    25
    1.21 × 10
    5
    1.84 × 10
     40
    4.80 × 10
     38
    b
    ¯
    b
    8.62 × 10
    2
    8.4
    1.32 × 10
    27
    1.32 × 10
    6
    1.15 × 10
     38
    3.01 × 10
     36
    100
    W
    +
    W
    2.87 × 10
    4
    4.5
    1.88 × 10
    23
    6.75 × 10
    3
    3.40 × 10
     42
    1.52 × 10
     39
    b
    ¯
    b
    8.65 × 10
    3
    4.5
    1.42 × 10
    25
    4.94 × 10
    4
    2.56 × 10
     40
    1.14 × 10
     37
    200
    W
    +
    W
    3.42 × 10
    5
    4.0
    9.81 × 10
    21
    1.09 × 10
    3
    4.23 × 10
     43
    2.98 × 10
     40
    b
    ¯
    b
    9.80 × 10
    3
    4.5
    1.29 × 10
    24
    1.13 × 10
    4
    5.56 × 10
     41
    3.92 × 10
     38
    500
    W
    +
    W
    1.31 × 10
    6
    3.7
    2.07 × 10
    21
    5.39 × 10
    2
    3.51 × 10
     43
    3.81 × 10
     40
    b
    ¯
    b
    8.87 × 10
    4
    4.0
    8.52 × 10
    22
    2.12 × 10
    3
    1.45 × 10
     41
    1.57 × 10
     38
    1000
    W
    +
    W
    2.18 × 10
    6
    3.6
    1.39 × 10
    21
    4.18 × 10
    2
    7.82 × 10
     43
    1.01 × 10
     39
    b
    ¯
    b
    2.14 × 10
    5
    4.0
    2.89 × 10
    22
    1.26 × 10
    3
    1.63 × 10
     41
    2.10 × 10
     38
    2000
    W
    +
    W
    2.38 × 10
    6
    3.6
    1.56 × 10
    21
    3.90 × 10
    2
    3.19 × 10
     42
    4.52 × 10
     39
    b
    ¯
    b
    3.53 × 10
    5
    3.9
    1.46 × 10
    22
    9.10 × 10
    2
    2.98 × 10
     41
    4.23 × 10
     38
    5000
    W
    +
    W
    2.07 × 10
    6
    3.6
    2.20 × 10
    21
    3.94 × 10
    2
    2.66 × 10
     41
    3.97 × 10
     38
    b
    ¯
    b
    4.59 × 10
    5
    3.7
    8.91 × 10
    21
    7.17 × 10
    2
    1.08 × 10
     40
    1.61 × 10
     37
    TABLE I
    EFFECTIVE VOLUME, EVENT UPPER LIMIT, AND PRELIMINARY LIMITS ON NEUTRALINO ANNIHILATION RATE IN THE SUN,
    NEUTRINO-INDUCED MUON FLUX FROM THE SUN, AND SPIN-INDEPENDENT AND SPIN-DEPENDENT NEUTRALINO-PROTON CROSS SECTION
    FOR A RANGE OF NEUTRALINO MASSES, INCLUDING SYSTEMATICS.
    Neutralino mass m
    χ
    (GeV)
    Preliminary
    10
    2
    10
    3
    10
    4
    10
    )
    -1
    y
    -2
    Muon flux from the Sun (km
    10
    2
    3
    10
    10
    4
    5
    10
    6
    10
    10
    2
    10
    3
    10
    4
    10
    10
    2
    3
    10
    10
    4
    5
    10
    6
    10
    lim
    CDMS(2008)+XENON10(2007)
    SI
    SI
    <
    σ
    σ
    lim
    CDMS(2008)+XENON10(2007)
    SI
    SI
    < 0.001
    σ
    σ
    2
    < 0.20
    0.05 <
    Ω
    χ
    h
    BAKSAN 1978-1995
    MACRO 1989-1998
    SUPER-K 1996-2001
    IceCube-22 2007 (soft)
    IceCube-22 2007 (hard)
    IceCube-80+DeepCore 1800d sens. (hard)
    AMANDA 2001-2003 (soft)
    AMANDA 2001-2003 (hard)
    AMANDA 2000-2006 (soft)
    AMANDA 2000-2006 (hard)
    thr
    = 1 GeV
    Indirect searches - E
    μ
    Neutralino mass m
    χ
    (GeV)
    Preliminary
    10
    2
    10
    3
    10
    4
    10
    )
    2
    (cm
    SD
    σ
    Neutralino-proton SD cross-section
    10
    -41
    -40
    10
    -39
    10
    -38
    10
    -37
    10
    -36
    10
    -35
    10
    -34
    10
    -33
    10
    -32
    10
    -31
    10
    lim
    CDMS(2008)+XENON10(2007)
    SI
    SI
    <
    σ
    σ
    lim
    CDMS(2008)+XENON10(2007)
    SI
    SI
    < 0.001
    σ
    σ
    2
    < 0.20
    0.05 <
    Ω
    χ
    h
    CDMS (2008)
    COUPP (2008)
    KIMS (2007)
    SUPER-K 1996-2001
    IceCube-22 2007 (soft)
    IceCube-22 2007 (hard)
    IceCube-80+DeepCore 1800d sens. (hard)
    AMANDA 2001-2003 (soft)
    AMANDA 2001-2003 (hard)
    AMANDA 2000-2006 (soft)
    AMANDA 2000-2006 (hard)
    10
    2
    10
    3
    10
    4
    10
    10
    -41
    -40
    10
    -39
    10
    -38
    10
    -37
    10
    -36
    10
    -35
    10
    -34
    10
    -33
    10
    -32
    10
    -31
    10
    Fig. 3. Preliminary limits on neutrino-induced muon flux from the Sun (left) along with limits from IceCube [7], BAKSAN [8], MACRO [9],
    and Super-K [10], and limits on spin-dependent neutralino-proton cross section (right) along with limits from CDMS [4], IceCube [7], Super-K
    [10], KIMS [36], and COUPP [37]. The green shaded area represents models from a scan of MSSM parameter space not excluded by the
    spin-independent cross section limits of CDMS [4] and XENON [5], and the blue shaded area represents allowed models if spin-independent
    limits are tightened by a factor of 1000. The projected sensitivity of 10 years operation of IceCube with DeepCore is shown in both figures.

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    1
    The extremely high energy neutrino search with IceCube
    Keiichi Mase
    , Aya Ishihara
    and Shigeru Yoshida
    for the IceCube Collaboration
    The department of Physics, Chiba University, Yayoi-tyo 1-33, Inage-ku, Chiba city, Chiba 263-8522, Japan
    See the special section of these proceedings
    Abstract
    . A search for extremely high energy (EHE)
    cosmogenic neutrinos has been performed with Ice-
    Cube. An understanding of high-energy atmospheric
    muon backgrounds that have a large uncertainty is
    the key for this search. We constructed an empirical
    high-energy background model. Extensive compar-
    isons of the empirical model with the observational
    data in the background dominated region were
    performed, and the empirical model describes the
    observed atmospheric muon backgrounds properly.
    We report the results based on the data collected
    in 2007 with the 22 string configuration of IceCube.
    Since no event was found after the search for the
    EHE neutrinos, a preliminary upper limit on an
    E
     2
    flux of
    E
    2
    φ
    ν
    e
    µ
    τ
    ≤ 5.6 × 10
     7
    GeV cm
     2
    s
     1
    sr
     1
    (90% C.L.) is placed in the energy range
    10
    7.5
    < E
    ν
    < 10
    10.6
    GeV.
    Keywords
    : neutrinos, IceCube, extremely high en-
    ergy
    I. INTRODUCTION
    Extremely high energy cosmic-rays (EHECRs) with
    energies above
    10
    11
    GeV are observed by several ex-
    periments. Although there is an indication that EHECRs
    are associated with the matter profile of the universe
    [1], their origin is still unknown. The detection of cos-
    mogenic EHE neutrino signals with energies greater than
    10
    7
    GeV can shed light on their origin. The cosmogenic
    neutrinos [2] produced by the GZK mechanism [3]
    carry information on the EHECR source evolution and
    the maximum energy of EHECRs at their production
    site [4]. Thus, EHE neutrinos can provide fundamental
    information about how and where the EHECRs are
    produced.
    The detection of EHE neutrinos has been an exper-
    imental challenge because the very small intensities of
    expected EHE neutrino fluxes require a huge effective
    detection volume. The IceCube neutrino observatory,
    currently under construction at the geographic South
    Pole, provides a rare opportunity to overcome this diffi-
    culty with a large instrumental volume of 1 km
    3
    .
    The backgrounds for the EHE neutrino signals are
    atmospheric muons. The large amount of atmospheric
    muons come vertically, while the signal comes primar-
    ily from zenith angles close to the horizon, reflecting
    competitive processes of generation of energetic sec-
    ondary leptons reachable to a detector and absorption of
    neutrinos due to an increase of the cross-sections. The
    atmospheric muon backgrounds drop off rapidly with
    increasing energy. Therefore, a possible EHE neutrino
    flux will exceed the background in the EHE region (∼
    >
    10
    8
    GeV). The signal is separated from the backgrounds
    by using angle and energy information.
    II. THE EHE EVENTS AND THE ICECUBE DETECTOR
    At extremely high energies, neutrinos are mainly
    detected via secondary muons and taus induced during
    the propagation of EHE neutrinos in the earth [5]. These
    particles are seen in the detector as a series of energetic
    cascades from radiative energy loss processes such as
    pair creation, bremsstrahlung and photonuclear interac-
    tions rather than as minimum ionizing particles. These
    radiative energy losses are approximately proportional
    to the energies of the muon and tau, making it possible
    to estimate its energy by observing the energy deposit
    in the detector.
    The Cherenkov light from the particles generated
    through the radiative processes are observed by an
    array of Digital Optical Modules (DOMs) which digitize
    the charges amplified by the enclosed 10” Hamamatsu
    photomultiplier tubes (PMTs) with a gain of
    ∼ 10
    7
    . The
    total number of photo-electrons (NPE) detected by all
    DOMs is used to estimate the energy of particles in this
    analysis. It is found that NPE is a robust parameter for
    estimating the particle energy.
    The data used in this analysis were taken with the
    22 string configuration of IceCube (IC22). Each string
    consists of 60 DOMs and 1320 DOMs in total with 22
    strings. The data taking began May, 2007, and continued
    to April, 2008. This analysis used a specific filtered data
    to select high energy events, which requires a minimum
    number of 80 triggered DOMs. The total livetime is
    242.1 days after removing data taken with unstable
    operation. The event rate at this stage is
    ∼1
    .5 Hz with
    a 16% yearly variation. Then, 6516 events with NPE
    greater than
    10
    4
    (corresponding to CR primary energy
    of about
    10
    7
    GeV and neutrino energy of about
    10
    6
    GeV
    (with
    E
     2
    flux)) are selected and used for the further
    analysis.
    III. BACKGROUND MODELING
    A. Construction of the empirical model
    Bundles of muons produced in CR air showers are the
    major background for the EHE signal search. Multiple
    muon tracks with a small geometrical separation resem-
    ble a single high energy muon for the IceCube detector.
    An understanding of the high energy atmospheric muon
    backgrounds is essential for the EHE signal search.

    2
    K. MASE
    et al.
    EHE NEUTRINO SEARCH WITH ICECUBE
    However, the backgrounds at the relevant energy range
    (
    > 10
    7
    GeV) is highly uncertain because of the poorly
    characterized hadronic interactions and composition of
    the primary CR where no direct measurement is avail-
    able.
    Therefore, we constructed an empirical model based
    on the Elbert model [6], optimizing the model to match
    the observational data reasonably in the background
    dominant energy region (
    10
    4
    <
    NPE
    < 10
    5
    ). The model
    is then extrapolated to higher energies to estimate the
    background in the EHE signal region. (See Fig. 1)
    The original Elbert model gives a number of muons
    for a CR primary energy
    E
    0
    such as
    N
    µ
    =
    E
    T
    E
    0
    A
    2
    cos θ
    ?
    AE
    µ
    E
    0
    ?
     α
    ?
    1 
    AE
    µ
    E
    0
    ?
    β
    ,
    (1)
    where
    A
    is the mass number of primary CRs with energy
    of
    E
    0
    , and
    θ
    is the zenith angle of a muon bundle.
    α
    ,
    β
    and
    E
    T
    are empirical parameters. The energy weighted
    integration of the formula relates the total energy carried
    by a muon bundle
    E
    µ
    B ,surf
    to the primary CR energy
    E
    0
    as,
    E
    µ
    B ,surf
    ?
    E
    0
    /A
    E
    surf
    th
    dN
    µ
    dE
    µ
    E
    µ
    dE
    µ
    ≃ E
    T
    A
    cos θ
    α
    α  1
    AE
    surf
    th
    E
    0
    ?
     α+1
    (
    ,
    2)
    where
    E
    surf
    th
    is a threshold energy of muons contributing
    to a bundle at surface and depends on the zenith angle.
    A surface threshold is related to a threshold energy at
    the IceCube depth
    E
    in ice
    th
    , by assuming a proportional
    energy loss to the bundle energy during propagation.
    This threshold at the IceCube depth is independent of
    zenith angle.
    With help of a Monte-Carlo (MC) simulation for the
    detector response as well as the measured CR flux, it is
    possible to predict the NPE distribution for certain
    α
    and
    E
    in ice
    th
    parameters. The CR flux used in this analysis
    is taken from the compilation of several experimental
    observations in Ref. [7]. The detector response includ-
    ing the Cherenkov photon emission, the propagation in
    the detector volume and the PMT/DOM response is
    simulated with the IceCube simulation program. The
    α
    and
    E
    in ice
    th
    parameters are, then, optimized to express
    the observed NPE distributions. The best optimized
    parameters are derived as
    α = 1.97
    and
    E
    in ice
    th
    = 1500
    GeV.
    With this empirical model, a simple simulation is
    feasible rather than simulating all muon tracks in a
    bundle, where the multiplicity can reach ten thousand for
    CR primary energies of
    10
    11
    GeV. Therefore, a bundle
    is replaced by a single track with the same energy as
    the entire bundle. It is shown in the next section that
    this substitution works well to express the observational
    data.
    Data generated with CORSIKA [8] (with the SIBYLL
    high energy hadronic interaction model) are also used.
    However, the extensive resources required for MC gen-
    eration precludes production of MC data with energy
    above
    10
    10
    GeV. Therefore, the CORSIKA data are
    mainly used to confirm the empirical model in the back-
    ground dominant energy region and provide redundant
    tools to study systematic uncertainty on the background
    estimation.
    The relation between CR primary energy and the NPE
    (which is the empirical model itself) is independently
    verified by using information from coincident events
    with the in-ice and surface detectors. The surface de-
    tectors can estimate the CR primary energy and the
    in-ice detectors give NPE. The relation is found to be
    consistent with the empirical model we derived.
    B. Comparison between observational data and MC
    An extensive comparison between the empirical
    model and the observational data was performed. The
    empirical model is found to describe the observational
    data reasonably in most cases. However, a significant
    difference was found in the
    z
    position (depth) of the
    center of gravity of the event (CoGZ) distribution. Many
    events are found in the deep part of the detector for the
    empirical model, while the events concentrate more at
    the top for the observational data. The difference is only
    seen for the vertical muons. This is probably due to the
    simple single muon substitution for the muon bundles in
    the empirical model. The more energetic single muons
    penetrate into the deep part, while many low energy
    muons in the bundles lose energies at the top of the
    detector for the vertical case. However, for the inclined
    cases, the bundles are already attenuated before coming
    to the detector, giving reasonable agreement between the
    observational data and the empirical model. Therefore,
    vertical events whose reconstructed zenith angles are less
    than 37
    are not used in this analysis. A simple algorithm
    is used for the angle reconstruction, based on the time
    sequence of the first pulses recorded by DOMs.
    Several distributions for the observational data and
    MC data after removing the vertical events are shown
    in Fig. 1 as well as the expected GZK cosmogenic
    neutrino signal [4]. As seen in the figure, the empirical
    model describes the observational data reasonably. The
    observed CoGZ distribution is also well represented by
    the empirical model after removing vertical events. The
    observed data are bracketed by the pure CORSIKA
    (SIBYLL) proton and iron simulation as expected.
    Some up-going events are seen in the observational
    data, though this is consistent with the empirical back-
    ground model. It is found that they are horizontally mis-
    reconstructed. On the other hand, fewer horizontal events
    are found for the CORSIKA data sets. This is because
    the CORSIKA data exhibit a better angular resolution
    of 1.4
    (one sigma) compared to the empirical model of
    2.5
    . The angular resolution for the observational data
    is estimated with help of the IceTop geometrical recon-
    struction. The estimated resolution is 2.5
    and consistent
    with the one of the empirical model. Another difference

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    3
    (total Npe)
    10
    log
    4
    4.5
    5
    5.5
    6
    6.5
    Event number in 242 days
    -3
    10
    -2
    10
    -1
    10
    1
    10
    2
    10
    3
    10
    Total
    Total Npe
    Npe distributiondistribution
    cos(zenith angle)
    -1 -0.8 -0.6 -0.4 -0.2 -0 0.2 0.4 0.6 0.8
    Event number in 242 days
    -3
    10
    -2
    10
    -1
    10
    1
    10
    2
    10
    3
    10
    Zenith angle distribution
    CoGZ [m]
    -400
    -200
    0
    200
    400
    600
    Event number in 242 days
    -2
    10
    -1
    10
    1
    10
    2
    10
    3
    10
    CoGZ distribution
    Fig. 1. The total NPE, zenith angle and CoGZ distributions between observational and MC data. The black dots represents observational data,
    green lines for empirical model (The shade expresses the uncertainty of the model), red for proton (CORSIKA, SIBYLL) and magenta for iron
    (CORSIKA, SIBYLL). The expected signal from GZK neutrinos[4] is also plotted with blue lines.
    between the observational data and the CORSIKA data
    is found in the CoGZ distribution. The CORSIKA data
    concentrate more at the top of the detector especially
    for vertical events. The CORSIKA data also show a
    narrower distribution in the relation of CR primary
    energy and the NPE. All these facts seem to indicate that
    the bundles in CORSIKA consist of more lower energy
    muon tracks compared to the observational data, leading
    to bundles with less stochastic energy losses. In order to
    confirm this hypothesis, more specific investigation is
    needed.
    The GZK signal events populate the EHE region
    and tend to be horizontal, as described in a previous
    section. This allows one to discriminate them from the
    background. The signal is also concentrated in the deep
    part of the detector because of the more transparent ice
    there.
    IV. SEARCH FOR EHE NEUTRINO SIGNAL
    Using the empirical background model, the EHE sig-
    nal search was performed based on the NPE and zenith
    angle information. The selection criteria are determined
    by using only MC data sets that are optimized with the
    observational data in the background dominated energy
    region (
    10
    4
    NPE
    ≤ 10
    5
    ), following a blind analysis
    procedure.
    It is found that the large spread of mis-reconstructed
    events extended to the signal region. We found that
    the angular resolution is related to the CoGZ position.
    Events whose CoGZs are at the bottom of the detector
    (CoGZ
    < 
    250 m) and which pass through the edge
    or outside of the bottom detector are significantly mis-
    reconstructed horizontal. When an inclined track reaches
    at the edge of the bottom part of the detector, there is
    no more detector below, so that the hit timing pattern
    resembles a horizontal track. The very clean ice at the
    bottom part of the detector and the biggest dust layer at
    middle enhance this effect. Therefore, the data sample
    is divided into two by the CoGZ position as follows.
    region A:
    250
    <
    CoGZ
    < 
    50 m, and CoGZ
    >
    50 m
    region B:
    CoGZ
    < 
    250 m, and
    50
    <
    CoGZ
    <
    50 m
    A clear difference between the backgrounds and the
    signal is seen in the zenith angle and total NPE relations
    as shown in Fig. 2. The atmospheric background muon
    distribution shows a steep fall in NPE and peaks in the
    vertical direction, while the GZK signal is mainly hor-
    izontal and at higher NPE, allowing the discrimination
    of the backgrounds by rejecting low NPE events and
    vertically reconstructed events. It is also obvious that
    the large spread in zenith angle direction for region B
    due to mis-reconstructed events.
    The selection criteria to separate signal from back-
    ground are determined for region A and B separately.
    The criteria are determined at first for each zenith angle
    bins, requiring the background level to be negligible
    compared to the signal (
    10
     4
    events per 0.1 cos(zenith
    angle) bin per 242.1 days). After the optimization for
    each zenith angle bin, the determined cut-offs in NPE
    are connected with contiguous lines as shown in Fig. 2.
    The expected numbers of signal and background
    events with the selection criteria are summarized in
    Table I.
    TABLE I
    EXPECTED EVENT NUMBER
    Models
    Expected events in 242.1 days
    GZK1 [4]
    0.16
    ±
    0.00 (stat.)
    +0.03
     0.05
    (sys.)
    Atm. muon
    (6.3
    ±
    1.4 (stat.)
    +6.4
     3.9
    (sys.))
    ×10
     4
    The effective area for each neutrino flavor averaged
    over all solid angles with the selection criteria is shown
    in Fig. 3.
    V. RESULTS
    The EHE neutrinos are searched for by applying the
    selection criteria determined in the previous section to
    the 242.1 days of observed data taken in 2007.
    Since no event is found after the search, a 90 %
    C.L. upper limit for all neutrino flavors (assuming full
    mixing neutrino oscillations) is placed with the quasi-
    differential method based on the flux per energy decade
    (
    ∆ log
    10
    E
    = 1.0) described in Ref. [9]. A 90 % C.L.
    preliminary upper limit for an
    E
     2
    spectrum is also
    derived as
    E
    2
    φ
    ν
    e
    µ
    τ
    ≤ 5.6 × 10
     7
    GeV cm
     2
    s
     1

    4
    K. MASE
    et al.
    EHE NEUTRINO SEARCH WITH ICECUBE
    log (total Npe)
    4
    4.5
    5
    5.5
    6
    6.5
    cos(zenith angle)
    -1
    -0.8
    -0.6
    -0.4
    -0.2
    -0
    0.2
    0.4
    0.6
    0.8
    1
    -6
    10
    -5
    10
    -4
    10
    -3
    10
    -2
    10
    -1
    10
    1
    10
    2
    10
    3
    10
    4
    10
    Obs. data
    region A
    log (total Npe)
    4
    4.5
    5
    5.5
    6
    6.5
    cos(zenith angle)
    -1
    -0.8
    -0.6
    -0.4
    -0.2
    -0
    0.2
    0.4
    0.6
    0.8
    1
    -6
    10
    -5
    10
    -4
    10
    -3
    10
    -2
    10
    -1
    10
    1
    10
    2
    10
    3
    10
    4
    10
    atmospheric muons
    region A
    (total Npe)
    10
    log
    4
    4.5
    5
    5.5
    6
    6.5
    cos(zenith angle)
    -1
    -0.8
    -0.6
    -0.4
    -0.2
    -0
    0.2
    0.4
    0.6
    0.8
    1
    -6
    10
    -5
    10
    -4
    10
    -3
    10
    -2
    10
    -1
    10
    1
    10
    2
    10
    3
    10
    4
    10
    GZK signals
    region A
    log (total Npe)
    4
    4.5
    5
    5.5
    6
    6.5
    cos(zenith angle)
    -1
    -0.8
    -0.6
    -0.4
    -0.2
    -0
    0.2
    0.4
    0.6
    0.8
    1
    -6
    10
    -5
    10
    -4
    10
    -3
    10
    -2
    10
    -1
    10
    1
    10
    2
    10
    3
    10
    4
    10
    Obs. data
    region B
    log (total Npe)
    4
    4.5
    5
    5.5
    6
    6.5
    cos(zenith angle)
    -1
    -0.8
    -0.6
    -0.4
    -0.2
    -0
    0.2
    0.4
    0.6
    0.8
    1
    -6
    10
    -5
    10
    -4
    10
    -3
    10
    -2
    10
    -1
    10
    1
    10
    2
    10
    3
    10
    4
    10
    atmospheric muons
    region B
    log (total Npe)
    4
    4.5
    5
    5.5
    6
    6.5
    cos(zenith angle)
    -1
    -0.8
    -0.6
    -0.4
    -0.2
    -0
    0.2
    0.4
    0.6
    0.8
    1
    -6
    10
    -5
    10
    -4
    10
    -3
    10
    -2
    10
    -1
    10
    1
    10
    2
    10
    3
    10
    4
    10
    GZK signals
    region B
    Fig. 2. The zenith angle Vs total NPE. The top plots are for region A and the bottom ones for region B. The plots are for the observational
    data, the background from the empirical model and the GZK signal[4] from left.
    10
    (Energy/GeV)
    log
    6 6.5 7 7.5 8 8.5 9 9.5 10 10.5
    ]
    2
    Effective area [m
    1
    10
    10
    2
    3
    10
    10
    4
    ν
    e
    ν
    μ
    ν
    τ
    Fig. 3. The effective area for each flavor neutrino after applying the
    signal selection criteria averaged over all solid angles. Blue dotted line
    represents
    ν
    e
    , black solid line for
    ν
    µ
    and red dashed line for
    ν
    τ
    .
    sr
     1
    , where 90 % of the events are in the energy range
    of
    10
    7.5
    < E
    ν
    < 10
    10.6
    GeV, taking the systematics into
    account. These preliminary limits as well as results of
    several model tests are shown in Fig. 4. The derived limit
    is comparable to the Auger [13] and HiRes [16] limit.
    The AMANDA limit [12] for an
    E
     2
    flux is better than
    the limit by this analysis. This is because AMANDA has
    a better sensitivity for lower energy and the livetime is
    about twice as much as this analysis.
    The systematics such as detector sensitivity, neu-
    trino cross-section, hadronic interaction model, yearly
    variation are currently being investigated. The biggest
    uncertainty comes from the NPE difference observed by
    the absolutely calibrated light source in situ, and it is
    estimated to be on the order of 30 %. These systematics
    are included in the upper limit calculation. The details
    of the systematics estimation as well as more detail
    of this analysis will be presented in another paper in
    preparation.
    ACKNOWLEDGEMENTS
    We acknowledge the U. S. National Science Founda-
    tion, and all the agencies to support the IceCube project.
    This analysis work is particularly supported by the Japan
    Society for the Promotion of Science.
    10
    (Energy) [GeV]
    log
    6.5
    7
    7.5
    8
    8.5
    9
    9.5
    10 10.5
    11
    ]
    -1
    sr
    -1
    s
    -2
    dF/dE) [GeV cm
    2
    (E
    10
    log
    -9
    -8.5
    -8
    -7.5
    -7
    -6.5
    -6
    -5.5
    -5
    -4.5
    -4
    IC22
    Z-burst
    GZK 1
    GZK 2
    GZK 3
    Auger
    ANITA
    -2
    )
    Auger (E
    HiRes
    -2
    )
    RICE (E
    -2
    )
    AMANDA (E
    -2
    )
    IC22 (E
    IceCube preliminary
    Fig. 4. The preliminary upper limit by the IC-22 EHE analysis (red
    solid) (all flavor) with the systematics taken into account. The thick
    long dashed green line represents GZK model 1 [4], light blue is for
    GZK model 2 [10], blue line is for GZK model 3 [17] and yellow for
    Z-burst model [11]. The dotted green line is 90 % C.L. upper limit for
    GZK model 1 by this analysis. The upper limit by other experiments
    are also shown with dashed lines [12], [13], [14], [15], [16]. Limits
    from other experiments are converted to all flavors where necessary
    assuming full mixing neutrino oscillations.
    REFERENCES
    [1] The Pierre Auger Collaboration, Science
    318
    , 896 (2007).
    [2] V. S. Beresinsky and G. T. Zatsepin, Phys. Lett.
    28B
    , 423 (1969).
    [3] K. Greisen, Phys. Rev. Lett.
    16
    , 748 (1966); G. T. Zatsepin and
    V. A. Kuzmin, Pisma Zh. Eksp. Teor. Fiz.
    4
    , 114 (1966) [JETP.
    Lett.
    4
    , 78 (1966)].
    [4] S. Yoshida and M. Teshima, Prog. Theor. Phys.
    89
    , 833 (1993).
    [5] S. Yoshida
    et al.
    , Phys. Rev. D
    69
    , 103004 (2004).
    [6] T. K. Gaisser,
    Cosmic Rays and Particle Physics
    (Cambridge
    University Press, Cambridge, England, 1990) p206.
    [7] M. Nagano and A. A. Watson, Rev. Mod. Phys.
    72
    , 689 (2000).
    [8] D. Heck
    et al.
    , Report FZKA 6019, (Forschungszentrum Karl-
    sruhe 1998).
    [9] X. Bertou
    et al.
    , Astropart. Phys.,
    17
    , 183 (2002).
    [10] O. E. Kalashev, V. A. Kuzmin, D. V. Semkoz, and G. Sigl, Phys.
    Rev. D
    66
    063004 (2002).
    [11] S. Yoshida, G. Sigl and S. Lee, Phys. Rev. Lett.
    81
    , 5505 (1998).
    [12] Andrea Silvestri (IceCube), Proc. 31st ICRC (2009).
    [13] J. Abraham
    et al.
    (Auger), Phys. Rev. Lett.
    100
    , 21101 (2008).
    [14] P. W. Gohram
    et al.
    (ANITA), arXiv:0812.2715v1 (2008).
    [15] I. Kravchenko
    et al.
    (RICE), Phys. Rev. D
    73
    , 082002 (2006).
    [16] R. U. Abbasi
    et al.
    (HiRes), ApJ
    684
    , 790 (2008).
    [17] R. Engel
    et al.
    , Phys. Rev. D
    64
    , 093010 (2001)

    PROCEEDINGS OF THE 31
    st
    ICRC, ?OD´ Z´ 2009
    1
    Study of very bright cosmic-ray induced muon bundle signatures
    measured by the IceCube detector
    Aya Ishihara
    ?
    for the IceCube Collaboration
    y
    ?
    Department of Physics, Chiba University, Chiba 263-8522, Japan
    y
    See the special section of these proceedings.
    Abstract
    . We present the study of cosmic-ray in-
    duced atmospheric muon signatures measured by
    the underground IceCube array, some of which
    coincide with signals in the IceTop surface detector
    array. In this study, cosmic-ray primary energies are
    associated with the total number of photoelectrons
    (NPEs) measured by the underground IceCube opti-
    cal sensors with two methods. We found that multiple
    muons that produce
    10
    4
    ˘ 10
    5
    NPEs in the IceCube
    detector in 2008 is corresponding to the cosmic-ray
    primary energies of
    10
    7
    ˘ 10
    9
    GeV.
    This association allows us to study cosmic-ray
    physics using photon distributions observed by the
    underground detector that are characterized by the
    properties of muon bundles. It is observed that
    the detailed NPE space distributions in longitudinal
    and lateral directions from muon tracks display the
    ranging-out effect of low energy muons in each
    muon bundle. The distributions from 2008 high
    energy muon data samples taken with the IceCube
    detector are compared with two different Monte
    Carlo simulations. The ?rst is an extreme case that
    assumes a single high energy muon in which nearly
    all of the energy loss is due to stochastic processes
    in the ice. The other uses the CORSIKA program
    with SYBILL and QGSJET-II high energy hadron
    interaction models, in which approximately half of
    the energy loss is due to ionization of low energy
    muons.
    Keywords
    : IceCube, muon-bundle, high-energy
    I. INTRODUCTION
    Bundles of muons produced in the forward region
    of cosmic-ray air showers appear as bright signals in
    Cherenkov detectors. The multiple-muon tracks with
    a small geometrical separation (called `muon-bundles`)
    resemble a muon with a higher energy. Understanding
    of the background muon bundles using a full air shower
    MC simulation in the high energy range above 10
    7
    GeV is limited because the calculation involves poorly
    characterized hadronic interactions and a knowledge on
    the primary cosmic ray composition at energies where
    there is no direct measurement available. The experi-
    mental measurement of atmospheric muons provides an
    independent probe of the hadronic interactions and the
    primary cosmic-ray compositions.
    The IceCube neutrino observatory [1] provides a rare
    opportunity to access the primary cosmic-ray energies
    beyond accelerator physics. The IceCube detector lo-
    cated at the geographic South Pole consists of an array of
    photon detectors which contains a km
    3
    ?ducial volume
    of clean glacier ice as a Cherenkov radiator. Half of the
    ?nal IceCube detector (IC40) was deployed by the end
    of austral summer of 2008. The IC40 detector consists
    of 40 strings of cable assemblies with an intra-string
    spacing of 125 m. Each string has 60 optical sensors
    (DOMs) spacing at intervals of ˘17 m and stretching
    between depths of ˘1450 m and ˘2450 m in the glacial
    ice. DOMs are also frozen into tanks located on the
    surface near the top of each string. The ice-?lled tanks
    constitute an air shower array called IceTop [2]. IceTop
    can act as an independent air-shower array to measure
    cosmic-ray spectra as well as trigger simultaneously
    with the underground detector. This provides a reliable
    method to study the atmospheric muon bundles.
    The data taking with the IC40 detector con?guration
    was performed from April, 2008 through March, 2009.
    The high energy muon-bundle (HEMu) sample consists
    of events which measure between 6:3?10
    2
    and 6:3?10
    4
    photo-electrons (PEs) in at least 50 underground DOMs.
    An IceTop coincidence (HECoinc) sample is a subset of
    the HEMu sample with the additional requirement that
    IceTop can successfully reconstruct the air shower event.
    Similarly, samples (called VHEMu and VHECoinc) with
    higher NPE threshold of 7:0 ? 10
    3
    are studied. De?ni-
    tions of samples are summarized in Table I.
    Data studied in this paper is taken in the period of July
    to December 2008 with a livetime of 148.8 days. Event
    distributions of the samples are presented in Fig. 1.
    TABLE I
    DEFINITIONS OF SAMPLE CONDITIONS.
    threshold NPE value
    IceTop coincidence required
    HEMu
    6:3 ? 10
    2
    no
    HECoinc
    6:3 ? 10
    2
    yes
    VHEMu
    7:0 ? 10
    3
    no
    VHECoinc
    7:0 ? 10
    3
    yes
    II. COSMIC-RAY ENERGY AND UNDERGROUND
    BRIGHTNESS RELATION
    Because the energy losses of muon-bundles are indica-
    tors of their energies and multiplicities, measurements of
    the total energy deposit of muons (E
    loss
    ) in the detection
    volume is important for understanding of the nature of
    muon-bundles. Here, we use the total number of photo-
    electrons recorded by the all underground DOMs (NPE)
    as an indicator of E
    loss
    . The effective light deposit from

    2
    VERY HIGH ENERGY MUONS IN ICECUBE
    log
    10
    (NPE)
    2.8 3 3.2 3.4 3.6 3.8 4 4.2 4.4 4.6 4.8
    Number of events
    10
    -1
    1
    10
    10
    2
    10
    3
    10
    4
    10
    5
    10
    6
    10
    7
    Declination at a depth of 1500m [deg]
    0
    5
    10
    15 20
    25 30
    35 40
    45
    10
    -1
    1
    10
    10
    2
    10
    3
    10
    4
    10
    5
    10
    6
    10
    7
    Fig. 1. Atmospheric muon event distributions from 2008 sample as a
    function of NPE (left) and reconstructed zenith angle ? (right). Filled
    square denotes HEMu and triangles are HECoinc. Inverse triangles and
    open circles are that of VHE samples as de?ned in table I. Coincidence
    samples show a high detection ef?ciency for vertical events and the
    ef?ciency drops with zenith angles. Event rates decreased by ˇ 2.5
    orders of magnitude when NPE is increased by an order of magnitude.
    bundles can be parameterized with an effective track
    length l
    0
    as [3], [4],
    NPE ˘ l
    0
    (?N
    ?
    + ˘?E
    ?
    ) / E
    loss
    :
    (1)
    Here, N
    ?
    and ?E
    ?
    indicate multiplicities and energy
    sum of underground muons respectively. ? and ˘ are
    ionization and radiative energy loss coef?cients assumed
    to be constant with energy. Primary cosmic-ray energies
    are related to the NPE with two methods. The ?rst
    method is to directly relate the underground NPEs with
    IceTop cosmic-ray energy reconstruction results. The
    other is to construct an empirical model to characterize
    the event frequencies of underground NPEs from the
    experimentally measured cosmic-ray surface ?uxes [5].
    The former method has the advantage that both cosmic-
    ray energy and underground brightness are consistently
    measured quantities, while the directional acceptance is
    limited to near vertical. The latter method requires a
    model assumption in the underground bundle spectra
    shape but full angular acceptance is available.
    A. IceTop coincidence signals
    Figure 2 shows the measured underground NPE distri-
    bution as a function of cosmic-ray energies reconstructed
    by the IceTop air-shower array. The energy determina-
    tion method by the IceTop array is described in [6]. A
    clear correlation exhibits that bright underground events
    are associated with the high energy cosmic-ray induced
    air showers and each NPE region roughly corresponds
    to different cosmic-ray energy regimes. For example, it
    shows that the cosmic-ray primary energy of ˘ 3:0?10
    7
    GeV are associated with 10
    4
    NPE underground events.
    As shown in Fig. 1, because of the IceTop coincidence
    condition, most of events in this sample is near vertical.
    B. The empirical model
    A high energy muon empirical model is constructed as
    in [7]. In the model construction, the amount of energy
    log
    10
    IceTop CR Energy/GeV
    56789
    InIce NPE
    10
    log
    3
    3.5
    4
    4.5
    1
    10
    10
    2
    10
    3
    Fig. 2.
    Event distributions of HECoinc sample as a function of
    NPE and IceTop reconstructed primary cosmic-ray energies. A clear
    correlation is observed.
    that goes to muon-bundle from cosmic-ray primaries
    is expressed in terms of energy weighed integral of
    the Elbert formula [8]. Because a major part of NPEs
    from muon tracks is expected to be due to the radiative
    processes in the very bright events, it is assumed in
    the model that the NPEs from the ionization is neg-
    ligible compared to the stochastic energy losses,
    i.e.
    N
    ?
    = 1 in Eq. 1. We then ?t experimental data with
    this model by varying ?E
    ?
    in Eq. 1 until it reproduces
    the experimentally observed NPE event rates. The total
    energy in the bundle ?E
    ?
    is carried by a single muon
    and the muon is simulated with [9]. The model is
    constructed based on the data sample taken in 2007.
    The present sample from 2008 under study separately
    con?rms the agreement as shown in Fig. 3 above the
    NPE threshold of 7:0 ? 10
    3
    . Below the threshold value,
    the model assumption that nearly all energy losses are
    due to radiative processes is expected to fail. The relation
    between the true cosmic-ray energy and NPE is shown
    in Fig. 4. The relation shows reasonable agreement with
    the experimentally measured relation shown in Fig. 2 in
    the overlapped acceptance region. An extrapolation of
    the relation indicates that corresponding primary cosmic-
    ray energy is increased to 10
    9
    GeV for the muon bundle
    signals with 10
    5
    underground NPE.
    III. ENERGY LOSSES OF MUONS IN BUNDLES
    A. Muon spectra in bundles
    Average muon spectra in a bundle for different to-
    tal NPE range from CORSIKA MC simulation using
    SYBILL and QGSJET-II as high energy interaction
    models with iron primaries and corresponding single
    muon energy distribution from the empirical model are
    shown in Fig 5. The plot shows that the number of
    muons reaching the IceCube depth from CORSIKA
    simulations increase with their total NPE. While there is
    a large difference between the muon bundle spectra from
    the CORSIKA full air-shower simulations and the high
    energy single muon empirical model, both describe the
    NPE event rates with a reasonable agreement (Fig 3).
    There is no signi?cant difference in muon spectra from

    PROCEEDINGS OF THE 31
    st
    ICRC, ?OD´ Z´ 2009
    3
    log
    10
    (NPE)
    3
    3.2
    3.4
    3.6
    3.8
    4
    4.2
    4.4
    4.6
    4.8
    Number of events
    10
    10
    2
    10
    3
    10
    4
    10
    5
    10
    6
    Declination at a depth of 1500m [deg]
    0
    5
    10
    15
    20
    25
    30
    35
    40
    45
    10
    10
    2
    10
    3
    10
    4
    10
    5
    10
    6
    10
    7
    10
    8
    Fig. 3.
    Event distributions as a function of NPE (left) and recon-
    structed zenith angle ? (right). Squares and inverse triangles denote
    2008 high energy event sample as in Fig. 1. Filled histograms are
    from the Monte Carlo simulation of the high energy muon empirical
    model as described in the text. Dark and light colored histograms
    are from CORSIKA MC simulation using SYBILL and QGSJET-II
    as high energy interaction models with iron primaries respectively.
    Event distributions from proton primaries highly underestimate the
    event rates. It can be seen that all of three MC simulation gives a
    reasonable agreement with experimental observation.
    log
    10
    cosmic-ray energy /GeV
    6
    6.5
    7
    7.5
    8
    8.5
    9
    9.5
    NPE
    10
    log
    4
    4.2
    4.4
    4.6
    4.8
    5
    1
    10
    10
    2
    Fig. 4.
    The correlation between primary cosmic-ray energy to
    underground NPE from MC simulation with the high energy muon em-
    pirical model. A consistent relation obtained with IceTop/underground
    coincidence measurement is obtained.
    SYBILL and QGSJET-II high energy interaction models
    with iron primary below 4:0 ? 10
    4
    NPE, but they
    exhibits some difference for the brighter events which
    approximately corresponds to the primary cosmic ray
    energies above ˘ 10
    8
    GeV.
    The fact that the event rates as a function of the
    total NPE appear consistent among the three estimations
    with different muon bundle models indicates that the
    NPEs of an event insensitive to the energy spectra of
    muon bundles. It implies that to distinguish whether
    the observed photon emission is dominated by either
    the ?rst or the second term in Eq. 1 is dif?cult with
    the total NPE. This indicates that the NPE measure is
    a systematically robust variable when used in analysis
    as in [7]. On the other hand, to evaluate muon bundle
    structure in each event, this variable is not suf?cient.
    The nature of muon bundles, such as the muon spectra
    log
    10
    muon energy at depth /GeV
    01234567
    Number of muons at IceCube depth
    0
    50
    100
    150
    200
    250
    SYBILL / Iron
    QGS-II / Iron
    4.0 < log
    10
    NPE < 4.4
    4.4 < log
    10
    NPE < 4.6
    4.6 < log
    10
    NPE < 5.0
    emperical
    single muon model
    ´
    100
    Fig. 5. Average muon MC-truth energy spectra in a bundle in different
    NPE range are shown for SYBILL, QGSJET-II with iron primaries
    and the empirical single muon model which is multiplied by 100 for
    a better visibility. Each of solid and dashed lines represents different
    NPE regions which approximately correspond to different cosmic-ray
    primary energies as shown in Fig. 4. In the brightest events, both
    CORSIKA high-energy models predicts more than 5,000 muons in
    a bundle reaching the underground detector. The muon in the single
    muon empirical model has energies between 100 TeV and 10 PeV.
    as in Fig. 5, is expected to appear in more detailed NPE
    distributions along the muon bundle tracks.
    B. The lateral and longitudinal NPE distributions
    The NPE distributions as functions of distances along
    and perpendicular to the track are shown in Fig. 6.
    In the plots, only vertically reconstructed events (? ?
    15 degrees) are used. Vertical tracks are suitable for
    measurement of detailed longitudinal development of
    the energy losses because the DOM separation in the
    z direction is only 17 m compared to 125 m in x-y
    direction. The detected Cherenkov photon pro?le shows
    a good correlation with the depth dependence of the
    measured optical properties of glacier ice. Fig. 6a shows
    a typical 3-dimensional NPE distributions of an observed
    high energy muon-bundle track. The lower panels shows
    averaged NPE distributions in the 2D plane from vertical
    VHEMu events for 2008 data, SYBILL-iron and the
    empirical model. There are visible differences in the
    2D light deposit distributions between data and models
    which give similar NPE. The detailed NPE distributions
    can be further examined as a function of longitudinal
    distances along tracks at various lateral distances as
    shown in the Fig. 7. Each solid line denotes different
    lateral distance with a 50 m interval and the distributions
    correspond to the slices along the longitudinal distances
    in the left panel of the Fig. 6b. It can be seen that
    the NPE observed by each DOM decreases rapidly
    with lateral distances. The closest longitudinal NPE
    distribution (? 50 m) shows that at the upper IceCube
    detector ˘800 NPEs are observed in each DOM and
    gradually decreased to ˘300 NPEs at the bottom of
    detector. This is expected to be due to ranging-out of
    low energy muons in bundles as they travel through the
    detector. This clearly shows that the longitudinal NPE
    pro?les close to the track is sensitive to the muon energy
    loss pro?le. The effect is less visible when photons

    4
    VERY HIGH ENERGY MUONS IN ICECUBE
    (a)
    lateral distance [m]
    0
    200
    400
    600
    800
    1000
    longitudinal distance [m]
    -500
    0
    500
    1
    10
    10
    2
    10
    3
    2008 data
    lateral distance [m]
    0
    200
    400
    600
    800
    1000
    longitudinal distance [m]
    -500
    0
    500
    1
    10
    10
    2
    10
    3
    CORSIKA SYBILL
    lateral distance [m]
    0
    200
    400
    600
    800
    1000
    longitudinal distance [m]
    -500
    0
    500
    1
    10
    10
    2
    10
    3
    emperical single muon model
    (b)
    Fig. 6.
    The lateral and longitudinal NPE distributions from high
    energy muon-bundle events which produces very bright event signa-
    tures. (a) Left: A typical NPE space distributions of a bright event in
    2008. The size of squares indicates log
    10
    NPE. Solid line indicates
    the reconstructed direction. There is a loss of photons due to a dusty
    layer of ice positioning around z = -100 m. Right: The NPEs from
    each DOM are plotted as functions of distances perpendicular to and
    along the reconstructed track. Filled bins are the position where the
    DOMs exist in this lateral and longitudinal two dimensional space and
    z-axis indicates measured NPEs. When there is more than one DOMs
    in a bin, NPE averages are calculated. (b) An averaged lateral and
    longitudinal NPE distribution of vertical bright events. Left: Vertically
    reconstructed VHEMu sample. Middle: CORSIKA-SYBILL with iron
    primary. Right: the high energy single muon empirical model.
    propagated more than 50 m from the track where the
    effects of ice properties begin to dominate. The effect
    of the ice layers with different scattering/absorption
    properties highly modi?es the lateral NPE distributions
    in this case. The distributions of NPEs close to tracks
    are suitable to study muon-bundle properties and NPEs
    at distance re?ects the nature of photon propagation
    through the ice.
    IV. OUTLOOK
    The various parts of lateral and longitudinal pro?les
    of the NPE distributions in 2-dimensional space are
    governed by the nature of muon bundles and optical
    properties of the ice in different way. Speci?cally, de-
    tailed study of the longitudinal NPE pro?les at different
    lateral distances is important for a better understanding
    of both the muon-bundle and ice property modeling.
    The contributions from ionization and radiative en-
    ergy losses in the obtained lateral and longitudinal
    NPE distributions are not distinguishable so far. This
    is because longitudinal NPE pro?les shown in Fig. 7
    are obtained from multiple events and stochastic nature
    of energy losses are averaged out. However, a large
    Longitudinal distance[m]
    -600
    -400
    -200
    0
    200
    400
    600
    average
    10
    NPE per DOM
    10
    2
    10
    3
    0m < lateral distance < 50m
    50m < lateral distance < 100m
    100m < lateral distance < 150m
    Fig. 7.
    Averaged longitudinal NPE distributions of the vertical
    VHEMu event sample. Each solid line denotes longitudinal NPE
    distributions at various lateral distances with an interval of 50 m.
    From the top line to the bottom, the intervals corresponding to each
    line are 0 m˘50 m, 50 m˘100 m, 100 m˘150 m, 150 m˘200 m,
    200 m˘250 m and 250 m˘300 m respectively. A clear NPE devel-
    opments in both longitudinal and lateral directions are visible.
    difference between ionization and radiative energy losses
    is expected to appear in the event-by-event ?uctuations
    of longitudinal/lateral NPE distributions. The sizes of
    ?uctuations from stochastic energy losses are evaluated
    in [4] using the MMC program [10] and the ?uctuations
    from ionization are expected to be /
    p
    N
    ?
    .
    The deviations of NPE along track from an average
    NPE per DOM are contributed from variations of ice
    properties. Because the ice properties does not ?uctuate
    an event-by-event basis, it is possible to distinguish the
    variation due to ice properties and the ?uctuation due
    to stochastic energy losses. The variations in NPEs near
    the tracks where less affected from ice properties and
    also the event-by-event NPE ?uctuation at given depth
    are expected to be sensitive parameters to the stochastic
    part of the muon bundle energy losses.
    V. ACKNOWLEDGMENTS
    We acknowledge U.S. National Science Foundation,
    and all the agencies to support the IceCube project.
    This analysis work is particularly supported by the Japan
    Society for the Promotion of Science.
    REFERENCES
    [1] J. Ahrens
    et al.
    , Astropart. Phys.
    20
    507 (2004);
    http://icecube.wisc.edu/.
    [2] T. Gaisser, in Proceedings of the 30th ICRC, Merida (2007).
    [3] Ph.D. thesis, C.H. Wiebusch, Physikalische Institute, RWTH
    Aachen (1995)
    [4] Ph.D. thesis, Predrag Miocino‹
    vic,´ UC Berkeley (2001)
    [5] M. Nagano and A. A. Watson, Rev. Mod. Phys.
    72
    , 689 (2000).
    [6] F. Kislat, S. Klepser, H. Kolanoski and T. Waldenmaier for the
    IceCube collaboration, these proceedings
    [7] K. Mase, A. Ishihara and S. Yoshida for the IceCube collaboration,
    these proceedings
    [8] T. K. Gaisser,
    Cosmic Rays and Particle Physics
    , (Cambridge
    University Press, 1990).
    [9] http://www.ppl.phys.chiba-u.jp/JULIeT/
    [10] D. Chirkin and W. Rhode, arXiv:hep-ph/0407075

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁOD´ Z´ 2009
    1
    Search for High Energetic Neutrinos from Supernova Explosions
    with AMANDA
    Dirk Lennarz
    ¤
    and Christopher Wiebusch
    ¤
    for the IceCube Collaboration
    y
    ¤
    III. Physikalisches Institut, RWTH Aachen University, 52056 Aachen, Germany
    y
    See the special section of these proceedings
    Abstract. Supernova explosions are among the most
    energetic phenomena in the known universe. There
    are suggestions that cosmic rays up to EeV energies
    might be accelerated in the young supernova shell
    on time scales of a few weeks to years, which would
    lead to TeV neutrino radiation. The data taken
    with the AMANDA neutrino telescope in the years
    2000 to 2006 is analysed with a likelihood approach
    in order to search for directional and temporal
    coincidences between neutrino events and optically
    observed extra-galactic supernovae. The supernovae
    were stacked in order to enhance the sensitivity. A
    catalogue of relevant core-collapse supernovae has
    been created. This poster presents the results from
    the analysis.
    Keywords: AMANDA, high energy neutrino astron-
    omy, supernova
    I. INTRODUCTION
    Almost a hundred years after their discovery, the
    acceleration mechanisms and sources of the cosmic
    rays remain an unsolved problem of modern astronomy.
    Neutrino astronomy can be an important contribution to
    the solution of this problem. Young supernovae in con-
    nection with a pulsar have been proposed as a possible
    source of cosmic rays with energies up to the ankle. This
    pulsar model can be directly tested by measuring high
    energetic (TeV) neutrino radiation on time scales of a
    few weeks to years after the supernova [1][2].
    The AMANDA-II neutrino telescope is located in the
    clear ice at the geographic South Pole and was fully
    operational since 2000. It reconstructs the direction of
    high energetic neutrinos by measuring Cherenkov light
    from secondary muons. The main background are muons
    and neutrinos produced in air showers in the atmosphere.
    This analysis uses 7 years of AMANDA data taken
    during the years 2000-2006 with a total live-time of 1386
    days. The data reconstruction and filtering is described in
    [3] and the final event sample contains 6595 events. The
    contamination of mis-reconstructed atmospheric muon
    events is less than 5% for a declination greater than 5
    ±
    .
    II. PULSAR MODEL
    The liberation of rotational energy from a pulsar can
    accelerate particles to relativistic energies. Secondary
    particles, for example pions, are created in the interac-
    tion with the expanding supernova envelope and decay
    into neutrinos and other particles. In this analysis the
    log10(seconds)
    1
    2
    3
    4
    5
    6
    7
    8
    9
    10
    Arbitrary units
    -3
    10
    -2
    10
    -1
    10
    1
    Fig. 1. Typical supernova neutrino model light curve
    pulsar model as described in [2] is used. Thermonuclear
    supernovae have no pulsar inside the envelope and are
    therefore not considered by this model.
    The phase of powerful, high energetic neutrino emis-
    sion is limited by two characteristic times: the time
    at which the pion decay time becomes less than the
    time between two nuclear collisions (t
    ¼
    ) and the time at
    which the density of the envelope is sufficiently small
    for accelerated particles to escape into the interstellar
    space without interaction (t
    c
    ). The supernova neutrino
    luminosity as a function of time (model light curve) is
    given by:
    L(t) =
    Ã
    1 ¡ exp
    Ã
    ¡
    μ
    t
    c
    t
    2
    !!
    ¢
    1
    1+(
    t
    ¼
    t
    )
    3
    ¢ ¸L
    0
    μ
    1+
    t
    ¿
    ¡2
    ;
    (1)
    where ¸ is the fraction of the total magnetic dipole
    luminosity L
    0
    (in erg/s) that is transferred to accelerated
    particles and ¿ the characteristic pulsar braking time.
    The shape and length of the model light curve depend
    on the supernova envelope mass (M
    e
    ), uniformity (de-
    scribed by a parameter called ») and expansion velocity
    (V ), the pulsar braking time and the maximum pion
    energy. An E
    ¡2
    neutrino energy spectrum is assumed
    with an energy cutoff at 10
    14
    eV. Fig. 1 shows a typical
    model light curve for t
    ¼
    ¼ 8 £ 10
    3
    s and t
    c
    ¼ 2 £ 10
    6
    s.
    These values are obtained by choosing M
    e
    = 3M
    ˉ
    ,
    » = 1, V = 0:1c and ¿ = 1year.

    2
    LENNARZ et al. AMANDA SUPERNOVA SEARCH
    III. SUPERNOVA CATALOGUES
    For this analysis a catalogue of supernovae was cre-
    ated. It combines three different electronically avail-
    able and regularly updated SN catalogues [4][5][6].
    A comparison of the three catalogues revealed some
    inconsistencies in the listed information. A consistent
    selection was made with special attention to the objects
    mistaken for a supernova observation, the total number
    of supernovae and the supernova positions.
    Fig. 2 shows the distribution of the 4805 supernovae
    observed between 1885 and 2008. The clearly visible
    structure around the celestial equator are supernovae
    found by the Sloan Digital Sky Survey-II supernova
    survey. The nearest and best visible supernova for
    AMANDA was SN2004dj in NGC 2403 at a distance
    of approximately 3.33 Mpc.
    -150
    -100
    -50
    0
    50
    100
    150
    -90
    90
    +90°
    -90°
    +180°
    -180°
    Fig. 2. Distribution of observed supernovae in equatorial coordinates
    with the galactic plane indicated as dashed line. Due to the background
    from atmospheric muons only supernovae in the northern hemisphere
    are relevant.
    This analysis searches for directional and temporal co-
    incidences between neutrinos and supernovae. Therefore
    additional input has to be quantified for each supernova.
    Firstly, the expected neutrino flux has to be determined
    from an accurate distance. The supernova distance can
    be identified with the distance to the host galaxy and can
    be estimated from the redshift. The redshift estimate is
    replaced by a measured distance (e.g. Cepheid variables
    or Tully-Fisher relation) if available. This improves the
    distance accuracy for nearby supernovae, which are most
    relevant.
    Secondly, the explosion date is needed for the temporal
    correlation, but only the date of the optical maximum
    or the discovery date is available. From some well
    observed SNe (e.g. 1999ex and 2008D) it is known that
    the optical maximum occurs around 15-20 days after
    the explosion, which is used as a benchmark. Fig. 3
    shows the difference between the date of discovery and
    the date of maximum for those cases where the light
    curve was fitted to a template and the date of maximum
    extrapolated backwards in time or found on old photo
    plates. The majority of the supernovae are discovered
    within 20 days after the optical maximum. Hence, the
    discovery is assumed to be typically 20 days after the
    optical maximum. The uncertainty of the explosion date
    is accounted for in the likelihood approach.
    Number of days
    1
    10
    2
    10
    Number of SNe
    0
    2
    4
    6
    8
    10
    Fig. 3.
    Number of days between the optical maximum and the
    discovery if the supernova was discovered after the maximum. A linear
    one day binning is shown on an logarithmic x-axis.
    Thirdly, for the individual supernova the needed input
    for the pulsar model is not available. Therefore, all
    supernovae are treated equally. The influence of the
    model light curve on the analysis is tested by defining
    two additional sets of parameters which result in light
    curves with very short and long neutrino emission.
    Hence, altogether three different model light curves
    (typical, short, long) are used. The width of the plateau
    that can be seen in Fig. 1 is 12 days for the typical, 1
    day for the short and 76 days for the long light curve.
    The most realistic assumption for the supernovae in the
    catalogue is that they have individual realisations of the
    parameters of the pulsar model and therefore individual
    light curves between the extreme cases.
    IV. LIKELIHOOD APPROACH
    A new likelihood approach was developed for this
    analysis [7]. Its principal idea is to compare all neutrino
    events from the experimental data sample to every rele-
    vant supernova and evaluate the likelihood ratio (LHR)
    between the hypothesis that this event is signal and
    the hypothesis of being background. This yields a large
    value for a good and small value for a bad match. The
    LHR for all events is summed in order to obtain a
    cumulative estimator, called Q:
    Q =
    X
    events
    P
    SN
    p(~ajSN)p(SN)
    p(~ajBG)p(BG)
    ;
    (2)
    where ~a are characteristic observables of the event.
    The advantage of this likelihood definition is that it
    can be extended to a stacking analysis. Q automatically
    assigns a small weight to irrelevant combinations of

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁOD´ Z´ 2009
    3
    neutrinos and supernovae, while relevant ones receive a
    larger weight. Thus, all supernovae from the catalogue
    can be used in the analysis and no optimisation on the
    number of sources is needed. Q is a sum of likelihood
    ratios and therefore its absolute value contains no phys-
    ical information.
    The probabilities in the likelihood sum are constructed
    from properties of AMANDA, the experimental data
    sample and the considered model light curve. p(BG) is
    the probability to have background and is an unknown
    but constant factor. This probability is eliminated by
    redefining Q to Q ¢ p(BG).
    p(~ajBG) is the probability that, assuming an event is
    background, it is observed at its specific time and from
    its specific direction. It is factorised into a temporal
    and an angular part. The temporal part corresponds to
    the AMANDA live-time. However, it cancels out with
    the corresponding temporal part of p(~ajSN). AMANDA
    does not distinguish between signal and background
    neutrinos and was obviously taking data when the event
    was measured. The angular probability is constructed
    with the normalised zenith angle distribution of the
    experimental data sample (see Fig. 2 in [3]). The azimuth
    probability is constant, because AMANDA is completely
    rotated each day and the azimuth is randomised for the
    relevant time scales of this analysis.
    The supernova signal probability consists of p(~ajSN)
    and p(SN). The first part depends on the specific event
    and is the probability that an event from a supernova is
    observed at a given time from a given direction. p(SN)
    is the probability to observe a signal from that supernova
    and is estimated for each supernova.
    For p(~ajSN) two terms are considered. p(ªjSN) is
    the probability that a neutrino from a supernova is
    reconstructed with an angular difference ª relative to
    the supernova direction. This probability is calculated
    from the point-spread function, which is obtained from
    Monte Carlo simulations. The second term p(t; t
    SN
    jSN)
    yields the probability that a neutrino arrives with a time
    offset t ¡ t
    SN
    from the explosion date. This probability
    is taken from a likelihood light curve.
    In order to be less model dependent three generic
    likelihood light curves (typical, short, long) are con-
    structed. They are inspired by the model light curves and
    constructed conservatively in order to not miss signal by
    accidentally looking too early or too late. Hence, if the
    date of the optical maximum is known, the starting time
    for the likelihood light curves (t = 0) is defined to be
    30 days earlier. In case only the date of the discovery
    is known a 50 days earlier starting time is used. This
    makes sure that the explosion is not missed, because the
    time shift to the explosion date is overestimated by about
    15 days for the optical maximum and up to 35 days for
    the date of discovery. The likelihood light curves consist
    of a half Gaussian for t < 0, a plateau for t > 0 and
    another half Gaussian after the plateau. The length of the
    plateau is the full width at 90% of the model light curves
    and enlarged by the uncertainty of the explosion day.
    This uncertainty is bigger if only the date of discovery
    is known. The width of the Gaussian after the plateau
    is the full width at half maximum (FWHM) after the
    plateau of the model light curves. Fig. 4 shows the
    typical likelihood light curve for the date of discovery.
    Days
    -60 -40 -20 0
    20 40 60 80 100 120 140
    Probability
    0
    2
    4
    6
    8
    10
    12
    -3
    ×10
    Fig. 4. Typical likelihood light curve
    p(SN) depends on the supernova neutrino luminosity,
    distance and direction. The absolute value of p(SN) is
    determined by the supernova neutrino luminosity and is
    a free parameter of this analysis. However, the absolute
    normalisation is not required, because a constant factor
    results in a rescaling of Q and hence only relative values
    are important. All supernovae are assumed to have the
    same neutrino luminosity at source. p(SN) decreases like
    the flux with the square supernova distance. AMANDA
    is not equally sensitive to neutrinos from all directions.
    Therefore the angular acceptance for different supernova
    directions is taken into account.
    V. SIGNAL AND BACKGROUND SIMULATION
    Q distributions for signal and background simulations
    are used to construct confidence belts with the Feldman-
    Cousins approach to the analysis of small signals [8].
    Each simulated data sample contains 6595 signal or
    background events like the experimental data set. Back-
    ground events are simulated with the zenith angle dis-
    tribution of the experimental data and the AMANDA
    live-time. For the signal simulation a model light curve
    and the AMANDA angular and temporal acceptance is
    simulated. The angular acceptance includes a random
    simulation of assumed systematic uncertainties of the
    measured rate of high energetic muon neutrinos [3].
    The simulation of the temporal and angular acceptance
    reduces signal events from days with low live-time or
    unfavourable supernova directions.
    The confidence belts are used to estimate the sen-
    sitivity of the analysis. The sensitivity for the long
    model light curve is not compatible compared to [3].
    Furthermore, if the supernovae have short model light
    curves, the sensitivity is comparable for the short and

    4
    LENNARZ et al. AMANDA SUPERNOVA SEARCH
    typical likelihood light curves. The typical pulsar model
    is best detected with the typical likelihood light curve.
    Therefore the experimental data is analysed with the
    typical likelihood light curve, because it can cover a
    larger range of possible parameters.
    VI. EXPERIMENTAL RESULT
    Analysing the experimental AMANDA data with the
    typical likelihood light curve yields:
    Q
    Exp
    typical
    = 0:0059 :
    (3)
    Fig. 5 shows this value in a Q distribution for back-
    ground only. The p value of obtaining a Q value equal
    or bigger than 0.0059 is 73.0%. Hence, the Q value is
    consistent with background and no deviation from the
    background only hypothesis is found. Therefore upper
    limits for the three model light curves are derived.
    Value of Q
    0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016
    Occurence
    1
    10
    2
    10
    p value = 73.0%
    Fig. 5. Q value from analysing the experimental data sample with
    the typical likelihood light curve (horizontal line) in a distribution for
    background only.
    90% upper limits on the signal strength are derived
    from the Feldman-Cousins confidence belts. With the
    help of the signal simulation explained above this value
    converts to the sum of neutrinos from all supernovae.
    Assuming the above model ranking of sources and the
    stacking this limit can also be interpreted as a limit on
    the number of neutrinos from SN2004dj. Tab. I shows
    the obtained upper limits for the typical, short and long
    pulsar model light curve.
    Pulsar model
    All SNe
    SN2004dj
    Typical
    < 5:4
    < 1:0
    Short
    < 4:1
    < 0:9
    Long
    < 67:3
    < 5:9
    TABLE I
    90% UPPER LIMITS ON THE NUMBER OF NEUTRINOS FROM ALL
    SUPERNOVA AND FROM SN2004DJ.
    The event numbers can be converted to a flux by
    integrating the AMANDA neutrino effective area with
    the expected signal energy spectrum (E
    ¡2
    spectrum with
    cutoff at 10
    14
    eV). Assuming the typical pulsar model
    for all supernovae and taking the average of the effective
    area over all directions, the 90% upper limit on the flux
    from all supernovae for the plateau of powerful neutrino
    radiation (12 days) is:
    dE
    ¢ E
    2
    < 5:2 £ 10
    ¡6
    GeV
    cm
    2
    s
    :
    (4)
    Using the effective area for the direction of SN2004dj,
    the corresponding 90% upper limit for SN2004dj is:
    dE
    ¢ E
    2
    < 8:4 £ 10
    ¡7
    GeV
    cm
    2
    s
    :
    (5)
    These limits are valid in the energy range from 1.1
    TeV to 84.0 TeV.
    Assuming that the energy range of the pulsar model
    as described in [2] can be extended to higher energies,
    the limits would improve by about 30% and are then
    valid in the energy range from 1.7 TeV to 2 PeV.
    VII. CONCLUSION
    For the first time the neutrino emission from young
    supernova shells was experimentally investigated. In
    the context of the pulsar model no deviation from the
    background only hypothesis was found.
    For a galactic supernova the expected flux from the
    pulsar model should be sufficient to be detectable by
    IceCube, the AMANDA successor. The sensitivity of
    this analysis might be enhanced by using an energy
    estimator in the likelihood and the individual event re-
    construction error instead of the energy averaged point-
    spread function.
    REFERENCES
    [1] V. S. Berezinsky and O. F. Prilutsky, Pulsars and Cosmic Rays in
    the Dense Supernova Shells, A&A, 66 (1978), no.3, pp. 325-334
    [2] H. Sato, Pulsars Covered by the Dense Envelopes as High-
    Energy Neutrino Sources, Progr. Theor. Phys., 58 (1977), no.2,
    pp. 549-559
    [3] J. Braun et al., Search for point sources of high energy neutrinos
    with final data from AMANDA-II, Phys. Rev. D, 79 (2009), no.6,
    pp. 062001-1 - 062001-15
    [4] List
    of
    supernovae
    from
    the
    CBAT,
    http://cfa-
    www.harvard.edu/iau/lists/Supernovae.html
    [5] Asiago Supernova Catalogue, http://cdsarc.u-strasbg.fr/viz-
    bin/Cat?B/sn
    [6] Sternberg Astronomical Institute (SAI) Supernova Catalogue
    http://www.sai.msu.su/sn/sncat/
    [7] D. Lennarz, Search for High Energetic Neutrinos from Supernova
    Explosions with the AMANDA Neutrino Telescope, Diploma
    thesis, RWTH Aachen University (2009)
    [8] G. J. Feldman and R. D. Cousins, Unified Approach to the
    Classical Statistical Analysis of Small Signals, Phys. Rev. D,
    57 (1998), no.7, pp. 3873-3889

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁOD´ Z´ 2009
    1
    Search for Ultra High Energy Neutrinos with AMANDA
    Andrea Silvestri
    for the IceCube Collaboration
    Department of Physics and Astronomy, University of California, Irvine, CA 92697, USA.
    See the special section of these proceedings.
    Abstract. We present results from the search for
    diffusely distributed Ultra High Energy (UHE) neu-
    trinos performed on data collected in 2003-2005 with
    the AMANDA experiment. At energies above a few
    PeV the Earth is opaque to neutrinos, therefore
    neutrinos must be differentiated from downward
    going cosmic ray induced (bundles of) muons. A
    search for a diffuse flux of UHE neutrinos shows
    no events, leading to a flux limit, summed over
    all flavors E
    2
    Φ
    ν
    ≤ 8.4 × 10
    −8
    GeV cm
    −2
    s
    −1
    sr
    −1
    (90% confidence level) for 10
    15.2
    eV < E
    ν
    < 10
    18.8
    eV. This limit is the most stringent placed to date.
    A number of model predictions different from the
    E
    −2
    spectrum have been tested and some have been
    rejected at a 90% C.L. We show that these results
    can also place a limit on the flux from point sources
    in the Southern Sky as a function of declination and
    valid in the same energy range.
    Keywords: Diffuse sources, high energy neutrinos,
    AMANDA
    I. INTRODUCTION
    Neutrino production from Active Galactic Nuclei
    (AGN), and other astrophysical sources have been exten-
    sively modeled during the past two decades, as described
    in [1], [2], [3]. Super-massive black holes hosted in
    the AGNs would accelerate, via a first-order Fermi
    mechanism, charged particles to ultra high energies. The
    collision of ultra-relativistic protons with the photon
    field in the AGN, via pγ and pp-interactions, would then
    produce high-energy neutrinos. The predicted intensity
    of neutrinos from these astronomical sources can reach
    the Earth and be detected by underground neutrino
    telescopes. Other theoretical calculations as presented
    in [4], [5], [6] and [2], derive an upper bound to
    the expected neutrino fluxes from high energy cosmic
    ray observations. These predictions, based on a model-
    independent approach, provide also a target for neutrino
    detector sensitivities. The predicted upper bound (ν
    µ
    and ν¯
    µ
    combined) for an E
    −2
    spectrum is E
    2
    Φ
    W B
    ν
    2 × 10
    −8
    ξ
    z
    GeV cm
    −2
    s
    −1
    sr
    −1
    , where ξ
    z
    accounts for
    cosmological model and source evolution. Using cos-
    mological dependence and source evolution that follows
    star formation rate over cosmological time gives ξ
    z
    ∼ 3.
    Assuming ν
    µ
    e
    = 0.5 at the source, produced from pγ
    and pp-collisions, the upper bound on the total flux for
    all ν− flavors becomes E
    2
    Φ
    W B
    ν
    ∼ 9 × 10
    −8
    GeV cm
    −2
    s
    −1
    sr
    −1
    .
    II. ANALYSIS
    The analysis results presented in this paper incorpo-
    rate 3-year of AMANDA data collected in 2003-2005
    (detector live-time of 507 days) and are based on work
    for one year analysis [7] of 2003 data. The Antarctic
    Muon And Neutrino Detector Array (AMANDA) [8],
    is the first neutrino telescope constructed in transparent
    ice, and deployed between 1500 m and 2000 m beneath
    the surface of the ice at the geographic South Pole
    in Antarctica. The AMANDA detector uses the Earth
    to filter out muons generated in the atmosphere on
    the Northern hemisphere and to search for point and
    diffuse sources of neutrinos with upward going direction
    at TeV to 100 TeV energies. However, at energies
    above PeV the Earth is opaque to neutrinos, therefore
    ν’s must be differentiated from the large background
    (billions of events per year) of downward going cosmic
    ray induced (bundles of) muons, which constitutes the
    primary challenge of this analysis. AMANDA has been
    taking data with the same detector configuration since
    2000, and the data acquisition electronics was upgraded
    in 2003 by recording full waveforms of the photo-
    electron (p.e.) pulses from the photomultiplier tubes
    (PMT) using Transient Waveform Recorders (TWR) [9].
    The entire 2003 data set of the AMANDA TWR tech-
    nology was processed, calibrated and analyzed to per-
    form an atmospheric neutrino analysis and a search for
    point sources in the Northern hemisphere sensitive at
    TeV energies [10], [11], which demonstrated the basic
    capabilities of the novel system to reproduce comparable
    physics results of the standard system of the AMANDA
    detector. After demonstrating the physics performance of
    the TWR technology, the analysis is performed to search
    for diffusely distributed neutrinos above PeV energies.
    The full waveforms from the PMT’s provide far more
    information on the light distribution from complex high
    energy events. However, the new technology produced
    ∼ 85 TB in 3 years, more than an order of magnitude
    increase w.r.t. the standard AMANDA system. To meet
    the challenge of large data structure and to simulate
    comparable data volume new analysis strategies were
    developed using high performance computing resources.
    The resources required for this analysis exceeded 2M
    CPU hours.
    The UHE analysis is performed by using the infor-
    mation of multiple p.e.’s from the PMT waveforms. The
    initial level of the analysis is defined by eliminating
    over 90% of the background by retaining only events

    2
    A. SILVESTRI et al. SEARCH FOR UHE NEUTRINOS
    NN
    2
    0
    0.2
    0.4
    0.6
    0.8
    1
    1.2
    arbitrary unit
    10
    −2
    10
    −1
    1
    10
    10
    2
    3
    10
    data 507 days
    bg mc
    signal mc
    ν
    /GeV)
    10
    (E
    log
    5
    6
    7
    8
    9
    10
    11
    12
    )
    2
    (m
    eff
    A
    10
    −2
    10
    −1
    1
    10
    10
    2
    3
    10
    10
    4
    ν
    μ
    ν
    e
    ν
    τ
    ν
    μ
    ν
    e
    ν
    τ
    ν
    μ
    ν
    e
    ν
    τ
    Fig. 1. Left panel: the neural network (NN2) distribution plotted for data, background and signal simulation. Right panel: detector effective
    area for muon, electron and tau neutrinos as a function of neutrino energy.
    with large number of p.e. pulses recorded in the array.
    After this level the analysis is refined by developing
    two independent neural networks. The first neural net-
    work mostly incorporates variables from reconstructed
    events, i.e. the reconstructed zenith angle of the events,
    which can separate downward going muons from signal
    mostly concentrated at the horizon. The second neural
    network uses primarily time dependent variables, like
    spread in leading edge of the arrival time and time-over-
    threshold values of the p.e. pulses, which at the higher
    level of the analysis better discern signal from high
    energy bundles of atmospheric muons. Variables were
    developed that exploit the full PMT waveforms, which
    in turn strongly correlate to signal features and better
    separate signal-like background events. Selection criteria
    based on single variable discriminators, like the number
    of photon-electrons were tested, but demonstrated not
    to be efficient for retaining signal events. Therefore a
    new set of variables were developed, which depend on
    timing and energy of typical signal events [7]. The new
    developed variables which use multiple photon-electrons
    in the PMT waveforms are the fluctuation of the time-
    over-threshold incorporated in the standard deviation of
    the tot’s (σ
    tot
    ), the mean of the leading edge times of
    the photon-electron pulses (µ
    le
    ), and the fluctuation from
    the standard deviation of the leading edge times (σ
    le
    ).
    Simulation of signal shows that distant UHE ν events
    may not deposit much light in the detector, but the spread
    in leading edge arrival time σ
    le
    and time-over-threshold
    σ
    tot
    values is large compared to typical background
    events. Background events tend to have large number of
    muons with relatively small lateral dimensions, which
    traverse through or close to the detector. Consequently,
    the arrival time of background photons shows little
    spread in time. On the other hand, signal events with
    comparable values of number of photon-electrons do
    not pass close to the detector. Therefore, these events
    differ from background because the photons show large
    variability in the arrival times. To further improve back-
    ground rejection, neural networks were developed and
    trained. The most powerful neural network included
    variables that measure the mean spread in leading edge
    times and fluctuations of time-over-threshold values. The
    search is performed with a blind analysis, i.e. 20% of
    the data sample is used to compare data with simulation
    while cut optimization is based on simulation solely.
    Once the analysis criteria are established, the cuts are
    frozen and applied to the remaining 80% of the data.
    Fig. 1 left panel shows the neural network (NN
    2
    ) before
    the final cut level for the combined data set, for the
    simulated atmospheric background and for the neutrino
    signal following an E
    −2
    spectrum.
    Background simulation was performed by generating
    primary cosmic ray using the CORSIKA package [12],
    propagating particles through the ice with the pro-
    gram MMC [13], recording detector response using
    the program AMASIM [14] with description of depth-
    dependent properties of Antarctic ice [15], and including
    proper treatment of waveform data. Similarly, neutrino
    signal simulation was performed for all flavors using the
    program ANIS [16]. Background simulation was biased
    in energy and spectrum towards high energy events to
    accommodate available computing resources. The final
    cut on NN
    2
    was determined by evaluating the model re-
    jection factor (MRF), as described in [17], and comput-
    ing the minimum of the ratio MRF = ?µ
    90
    (n|b)?/n
    sig
    .
    The ?µ
    90
    (n|b)? is the average 90% C.L. upper limit,
    determined by using the Feldman-Cousins method [18],
    computed over the Poisson probabilities for the exper-
    iment repeated many times, and n
    sig
    is the number of
    signal events for a given model. The minimum of the
    MRF determines the cut which is set to NN
    2
    > 0.85.
    After the final cut one experimental event is observed
    over a detector live-time of 507 days, consistent with
    0.9 (−0.9, +1.3) events from background expectation.
    III. RESULTS
    The search for a diffusely distributed flux of UHE
    neutrinos shows no signal events, leading to a prelimi-

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁOD´ Z´ 2009
    3
    ν
    (GeV))
    10
    (E
    log
    3
    4
    5
    6
    7
    8
    9
    10
    11
    12
    13
    )
    -1
    sr
    -1
    s
    -2
    dN/dE (GeV cm
    2
    E
    10
    -8
    10
    -7
    10
    -6
    preliminary
    AMANDA (this work)
    AMANDA [24]
    AMANDA [26]
    IceCube [29]
    Rice [23]
    HiRes [25]
    Auger [28]
    Anita [27]
    Waxman and Bahcall x 3/2
    AMANDA (this work)
    Mannheim 95 RL
    Mannheim 95 PG
    Waxman and Bahcall 99
    MPR 00
    Stecker 05
    ESS 01
    AMANDA (this work)
    Mannheim 95 RL
    Mannheim 95 PG
    Waxman and Bahcall 99
    MPR 00
    Stecker 05
    ESS 01
    Fig. 2. Experimental limits of this analysis on a diffuse E
    −2
    ν-flux for all flavors as a function of neutrino energy, thick solid line. Solid
    lines represent experimental limits from other experiments [23], [24], [25], [26], [27], [28], [29]. Dotted curves represent model predictions for
    a diffuse ν-flux, and predictions have been adjusted for all flavor neutrino contribution, where necessary.
    nary flux limit, summed over all flavors
    E
    2
    Φ
    ν
    ≤ 8.4 × 10
    −8
    GeV cm
    −2
    s
    −1
    sr
    −1
    (1)
    at 90% C.L. for the energy interval 10
    15.2
    eV < E
    ν
    <
    10
    18.8
    eV, defined by the 90% containment of the final
    neutrino energy distribution, which has a median energy
    of ?E
    ν
    ? = 4 × 10
    16
    eV. Fig. 1 right panel shows the
    detector effective area for all flavor neutrinos for an E
    −2
    spectrum as a function of neutrino energy, which for
    muon neutrinos reaches 100 m
    2
    for 100 PeV and rapidly
    increases with energy.
    The limits are computed including the contribution of
    systematic uncertainties by using the method described
    in [19]. Tab. I summarizes the different sources of
    systematic uncertainties which impact background and
    signal simulation in this analysis: The same numbers as
    given in the table is then repeated in the text. Differences
    in the simulation for cosmic ray composition by generat-
    TABLE I
    SUMMARY OF SYSTEMATIC UNCERTAINTIES ESTIMATED FROM
    DIFFERENT SOURCES IMPACTING BACKGROUND AND SIGNAL
    SIMULATION.
    Source
    bg
    signal (E
    −2
    )
    CR comp. and inter. models
    ±80%
    -
    detector sensitivity
    ±15%
    ±15%
    year-to-year detector variation
    ±14%
    ±14%
    tot-factor for d > 200 m
    -
    +10%
    ice properties
    ±20%
    ±20%
    charm BG
    negl.
    -
    tot-corr. and N2-laser cal.
    ±100%
    ±10%
    neutrino cross section [20]
    -
    ±4%
    LPM effect
    -
    −3%
    total (added in quadrature)
    ±131%
    ±32%
    TABLE II
    SUMMARY OF MODEL PREDICTIONS TESTED BY THIS ANALYSIS.
    MODELS WITH A MRF < 1 ARE EXCLUDED AT 90%, WHILE
    MODELS WITH A MRF > 1 ARE CONSISTENT WITH THESE
    RESULTS.
    Model
    ν
    all
    MRF
    Reference
    AGN RL A-jet
    1.10
    3.05
    Mannheim 95 PG [1]
    AGN RL B-jet
    17.8
    0.19
    Mannheim 95 RL [1]
    AGN-jet
    14.6
    0.23
    MPR 00 [2]
    AGN-core
    3.12
    1.07
    Stecker 05 [3]
    Waxman-Bahcall
    4.04
    0.83
    WB 99 [4]
    GZK mono-energetic
    5.50
    0.61
    KKSS 02 [21]
    GZK index α=2
    4.68
    0.72
    KKSS 02 [21]
    GZK full evol.
    0.28
    12.0
    ESS 01 [22]
    ing proton and iron CR primaries, and interaction models
    by using two different hadronic models (QGSJET and
    SIBYLL) were used to estimate variations in background
    event rate; Uncertainties in detector sensitivity which
    mostly depend on the absolute sensitivity of the PMT’s,
    were also included; Variations have been estimated due
    to the difference in detector response observed for the
    three years studied in the analysis; Variations in the
    spread of the time-over-threshold for distances d > 200
    m were evaluated for the impact on signal efficiency;
    Studying ice properties with two different models gave
    a max variation of 16% in signal sensitivity; The impact
    on systematic uncertainties due to ice properties has
    been further studied by varying the length of photon
    propagation in ice for distances characteristic of high
    energy signal, and by incorporating this variation into
    the effect of detector sensitivity to estimate the impact
    on signal sensitivity; Background from charm production
    has been estimated to be negligible in this analysis;

    4
    A. SILVESTRI et al. SEARCH FOR UHE NEUTRINOS
    sin(δ)
    -1
    -0.8 -0.6 -0.4 -0.2
    0
    0.2
    0.4
    0.6
    0.8
    1
    )
    -1
    s
    -2
    GeV cm
    -6
    dN/dE (10
    2
    E
    10
    -2
    10
    -1
    1
    10
    AMANDA (this work)
    preliminary
    MACRO [30]
    Super-K [31]
    AMANDA [32]
    Fig. 3.
    Point flux limits as a function of declination sin(δ) for
    the Southern Sky averaged over azimuth, solid line. Also included are
    limits from other experiments [30], [31] and for the Northern Sky [32].
    Variations in the N
    2
    -laser calibration for the spread of
    the time-over-threshold were estimated for background
    event rate and signal efficiency; Uncertainties in the
    neutrino cross section [20] for energies relevant for this
    analysis were incorporated, and impact due to LPM
    effect for signal above 10
    8
    GeV were also included. The
    estimated systematic uncertainties have been added in
    quadrature and incorporated in the final results of the
    analysis.
    The diffuse limit has been used to test a number
    of model predictions different from the E
    −2
    spectrum.
    Model predictions with a ratio ?µ
    90
    (n|b)?/n
    sig
    < 1 are
    excluded by this analysis. The models tested and the
    corresponding MRF have been summarized in Tab. II. A
    class of AGN predictions based on jet-models scenario,
    such as [1] (RL B) and [2] have been excluded,
    while prediction [1] (RL A) is not, and AGN prediction
    based on core-models scenario [3] is almost excluded.
    From the class of models excluded by this analysis we
    can conclude jet-models normalized to diffuse x-ray or
    GeV/TeV emission from individual sources are generally
    disfavored. These limits are consistent, and below the
    maximum upper bound to neutrino flux predicted by [4],
    [6], and also below the maximum neutrino flux due to
    possible extra-galactic component of low-energy protons
    of 10
    17
    eV [5]. These results are also consistent and
    below the bounds on neutrino fluxes presented by [2],
    computed by assuming optically thin (thick) sources
    to pion photo-production processes. Models on GZK
    neutrino spectrum were also tested, predictions [21] are
    excluded, while prediction [22] is still compatible with
    these results. The limits from this work to an E
    −2
    neutrino flux as a function of energy are shown in Fig. 2,
    thick solid line. Model predictions are represented by
    dotted curves, and solid lines show limits presented by
    other experiments [23], [24], [25], [26], [27], [28], [29].
    At UHE energies this analysis is sensitive to search
    for point source of neutrinos in the Southern Sky.
    Simulation shows that muons are reconstructed with
    angular resolution of ∼ 7
    o
    over the entire Southern
    hemisphere. Except for a small band near the horizon,
    signal originating from the Southern Sky will be ob-
    served in the Southern Sky. The sensitivity only depends
    on zenith angle and is roughly independent of azimuth,
    and maximum sensitivity peaks at the horizon [7]. Since,
    no excess of events were observed, a flux limit as a
    function of declination is derived, Fig. 3, and fitted
    with a function of δ, as E
    2
    dN/dE
    ν
    (sin δ) ≤ [1.3 ×
    e
    −(2 sin δ)
    ]×10
    −7
    GeV cm
    −2
    s
    −1
    with 10% accuracy,
    valid for −0.98 < sin(δ) < 0, and for 10
    15.2
    eV <
    E
    ν
    < 10
    18.8
    eV. These point flux limits are valid for
    energies above PeV, and are compatible with results from
    other experiments, which cover lower energy intervals
    between 10 GeV - 100 TeV [30], and between 10 GeV
    - 100 GeV [31].
    To summarize, we have presented in this paper the
    most stringent limits to date for neutrino energies above
    1 PeV. These experimental limits begin to restrict the
    largest possible fluxes of the WB upper bound [4], [5],
    [6].
    IV. ACKNOWLEDGMENT
    The author acknowledges support from U.S. National
    Science Foundation-Physics Division, and the NSF-
    supported TeraGrid system at the San Diego Supercom-
    puter Center (SDSC).
    REFERENCES
    [1] K. Mannheim, Astropart. Phys. 3, 295 (1995).
    [2] K. Mannheim et al., Phys. Rev. D 63, 023003 (2001).
    [3] F. Stecker, Phys. Rev. D 72, 107301 (2005).
    [4] E. Waxman and J. Bahcall, Phys. Rev. D 59, 023002 (1999).
    [5] J. Bahcall and E. Waxman, Phys. Rev. D 64, 023002 (2002).
    [6] E. Waxman, Phil.Trans.Roy.Soc.Lond. A 365, 1323 (2007).
    [7] A. Silvestri, Ph.D. Thesis, University of California, Irvine (2008).
    [8] E. Andres et al., Nature 410, 441 (2001).
    [9] W. Wagner et al., Proc. 28th ICRC, 2, 1365 (2003).
    [10] A. Silvestri et al., Proc. 29th ICRC, 5, 431 (2005).
    [11] A. Silvestri et al., Mod. Phys. Lett. A 22, 1769 (2007).
    [12] D. Heck, DESY-PROC-1999-01, 227 (1999).
    [13] D. Chirkin and W. Rhode, hep-ph/0407075 (2004).
    [14] S. Hundertmark,Proc.1st νTelescopes Workshop, Zeuthen (1998).
    [15] M. Ackermann et al., J. Geophys. Res. 111, D13203 (2006)
    [16] M. Kowalski and A. Gazizov, Comput. Phys. Comm. 172, 203
    (2005).
    [17] G. Hill and K. Rawlins, Astropart. Phys. 19, 393 (2003).
    [18] G. J. Feldman and R. D. Cousins, Phys. Rev. D 57, 3873 (1998).
    [19] G. Hill, Phys. Rev. D 67, 118101 (2003).
    [20] A. Cooper-Sarkar and S. Sarkar, JHEP 75, 801 (2008).
    [21] O. E. Kalashev et al., Phys. Rev. D 66, 063004 (2002).
    [22] R. Engel et al., Phys. Rev. D 64, 093010 (2001).
    [23] I. Kravchenko et al., Phys. Rev. D 73, 082002 (2006).
    [24] A. Achterberg et al., Phys. Rev. D 76, 042008 (2007).
    [25] K. Martens et al., astro-ph/0707.4417, (2007).
    [26] M. Ackermann et al., Astrophys. J. 675, 1014 (2008).
    [27] P. Gorham et al., astro-ph/0812.2715, (2008).
    [28] J. Abraham et al., astro-ph/0903.3385, (2009).
    [29] K. Mase, A. Ishihara and S. Yoshida et al., Proc. 31th ICRC,
    these Proceedings (2009).
    [30] M. Ambrosio et al., Astrophys. J. 546, 1038 (2001).
    [31] S. Desai et al., Astropart. Phys. 29, 42 (2008).
    [32] R. Abbasi et al., Phys. Rev. D 79, 062001 (2009).

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁOD´ Z´ 2009
    1
    Selection of High Energy Tau Neutrinos in IceCube
    Seon-Hee Seo
    and P. A. Toale
    for the IceCube Collaboration
    Oskar Klein Centre and Dept. of Physics, Stockholm University, SE-10691 Stockholm, Sweden
    Dept. of Physics, Pennsylvania State University, University Park, PA 16802, USA
    See the special section of these proceedings
    Abstract. Astrophysical neutrino sources are ex-
    pected to produce electron and muon flavor neutrinos
    via charged pion decay. Over cosmological distances,
    standard neutrino oscillations will change the flavor
    content to include equal fluxes of all three flavors. Tau
    neutrinos with energies above a few PeV will produce
    characteristic signatures known as double-bangs and
    lollipops. In contrast to searches for cosmological
    electron and muon neutrinos, which must contend
    with backgrounds from atmospheric neutrinos, tau
    neutrinos are expected to be background-free. Thus
    far no searches for tau neutrino events with these
    characteristic signatures have been performed be-
    cause their detection requires a kilometer-scale de-
    tector. In this talk, we will present current results
    from several methods for searching for high energy
    tau neutrinos in IceCube.
    Keywords: Tau neutrinos, Double-bangs, IceCube
    I. INTRODUCTION
    One of the main research topics in neutrino tele-
    scopes such as ANTARES and IceCube is to search for
    neutrinos of astrophysical origin. Astrophysical particle
    accelerators like AGNs and GRBs may produce high
    energy neutrinos [1], [2]. As daughters of charged pion
    decay, the emerging neutrinos are expected to have the
    flavor flux ratio of 1:2:0 (ν
    e
    µ
    τ
    ). Due to neutrino
    oscillations, this neutrino flux is expected to be observed
    in the flavor ratio of approximately 1:1:1 on Earth. There
    are also models which predict different ratios of neutrino
    flux observed on the Earth but they all lead to non-zero
    flux of ν
    τ
    [3], [4].
    Here we discuss aspects of a search for high energy
    (greater than a few PeV) ν
    τ
    ’s with the IceCube 22-
    string array (“IC-22”). High energy ν
    τ
    ’s can leave very
    distinctive signatures in the IceCube detector owing to
    the very short life time and numerous decay channels of
    tau leptons. We denote these signatures “lollipops,” “in-
    verted lollipops” and “double-bangs” [5], [6]. Although
    high energy ν
    τ
    ’s can traverse the Earth through the
    “regeneration” process [7], they typically emerge with
    energies too low to create any of the signatures under
    study here. The low energy (below PeV) ν
    τ
    ’s can be
    detected in 4π in IceCube but are seen as ”cascade-like”
    events, which is described elsewhere.
    The lollipop and inverted lollipop topologies are char-
    acterized by having either the production or decay vertex
    Fig. 1.
    A simulated double-bang event produced from a primary
    neutrino energy of 47 PeV that enters the IC-22 detector with 35
    o
    zenith angle. The two showers are separated by a tau track of 332 m
    long. The colors (online version only) represent the relative hit times,
    with red for the earliest hits, blue for the latest hits, and other times
    in between according to the colors of the rainbow.
    of the tau lepton well outside the detector fiducial vol-
    ume, respectively. In these topologies we expect to see
    a track due to the tau lepton and a single shower at the
    contained vertex. The double-bang topology as shown
    in Fig. 1 is characterized by having both production
    and decay vertices contained within the detector fiducial
    volume, and the tau track long enough to clearly separate
    the two showers from one another.
    These astrophysical high energy ν
    τ
    events are con-
    taminated much less by atmospheric background from
    cosmic ray interactions compared to ν
    µ
    and ν
    e
    [8]. This
    is because the conventional atmospheric ν
    τ
    flux is nearly
    zero, and the prompt ν
    τ
    flux produced from charmed
    meson decay in the atmosphere is also expected to be
    very small [9], [10].
    II. SIGNATURE BASED SEARCH METHOD
    In IC-22, the search for high energy ν
    τ
    does not incor-
    porate full event reconstruction [11], but instead relies
    on a simpler approach that exploits the unique signatures

    2
    S. H. SEO et al. TAU NEUTRINOS IN ICECUBE
    time (ns)
    9500 10000 10500 11000 11500 12000 12500 13000 13500 14000 14500
    Q per DOM
    0
    200
    400
    600
    800
    1000
    1200
    1400
    1600
    1800
    sliding time window
    Fig. 2. Charge (number of photo-electrons) per DOM distribution as
    a function of light arrival time (ns) for a simulated inverted lollipop
    event produced from a primary neutrino energy of 64 PeV. The initial
    peak corresponds to a shower from ν
    τ
    charged-current interaction,
    followed by a tau track.
    of these events. An example of a simple criterion is
    given in Fig. 2 which shows the distribution of detected
    charge (proportional to the amount of Cherenkov light)
    per digital optical module (DOM) as a function of the
    time at which the light arrived, for a simulated inverted
    lollipop event.
    As shown in the figure the inverted lollipop event
    produces a unique topology consisting of a shower
    followed by a track inside the detector compared to
    typical muons that produce simple track-like signatures
    with smooth light deposition along its track length. A
    set of simple variables based on the differences in the
    topologies can select (inverted-)lollipop and double-bang
    events while removing track-like muon backgrounds.
    One of the variables invented for this purpose is maxi-
    mum “current ratio,” I
    R,max
    . This variable is defined as
    a ratio of the two currents, I
    R
    , themselves defined as
    the amount of charge per unit time, inside and outside
    a sliding time window, as shown in Fig. 2. When the
    sliding time window passes through the event’s time-
    ordered hits, I
    R
    is calculated, and its maximum value
    I
    R,max
    is used as a cut variable. For signal events,
    I
    R,max
    is expected to be greater than 1 but for simple
    track-like muon backgrounds it is expected to be closer
    to 1.
    However, energetic muons can leave a big shower
    from bremsstrahlung during their passage through the
    fiducial volume so that these events could survive the
    I
    R,max
    cut due to the similarity of the event topology.
    To remove these energetic muon events another variable
    called the “local current,” I
    L
    , is used. The I
    L
    is defined
    as the current calculated in three equally-spaced time
    regions of an event’s time-ordered hits. Of the first and
    last third, we choose the one with the largest IL as the
    selection criterion. We intentionally ignore the middle
    third to help reject energetic muons that have an accom-
    panying bremsstrahlung somewhere in the middle of its
    track length. This variable showed good discrimination
    power between signal and background events.
    III. CUTS AND EFFICIENCIES
    So far the cuts have been developed in six distinct
    levels after application of a trigger and online filter.
    The trigger, denoted “SMT8,” applies a simple majority
    condition of 8 hits within 5 µs to the data as it is
    acquired. The online filter is a logical OR of IceCube’s
    cascade and Extremely High Energy (EHE) filters. The
    cascade filter is designed to select events which satisfy
    minimum condition of “cascade-like” events [12]. For
    the EHE filter, a minimum of 80 hits were required.
    The level 0 and 1 (L0, L1) cuts are designed to
    remove track-like muon backgrounds. The L2 cuts are
    designed mainly to remove energetic muons accompa-
    nying bremsstrahlung in the middle of their passage.
    The L3 cuts are designed to remove downwards-going
    events, and the L4 cuts are designed to select events
    that look more “cascade-like” than “track-like” using
    different variables from those used in L0 and L1. The
    L5 cuts are designed to remove events which are not
    sufficiently contained inside the detector. Fig. 3 shows
    the relative efficiencies at each cut level for signal and
    background events.
    0
    1
    2
    3
    4
    5
    6
    7
    log10(Efficiencies w.r.t. SMT8)
    -8
    -7
    -6
    -5
    -4
    -3
    -2
    -1
    0
    lolliPop
    doubleBang
    nuTau E-2
    nuMu E-2
    nuE E-2
    prompt atm nu
    bartol atm nu
    Single mu
    Double mu
    Tau Cut Efficiencies
    Fig. 3.
    High energy ν
    τ
    selection cut efficiencies w.r.t. SMT8 for
    lollipop, double-bang, astrophysical ν
    τ
    , ν
    µ
    and ν
    e
    , and atmospheric
    background events for IC-22. The numbers from 0 to 8 on the x-axis
    represent SMT8, the online filter, and L0, L1, L2, L3, L4 and L5 cuts,
    respectively.
    As shown in Fig. 3, lollipop and double-bang events
    keep the highest efficiencies because they are specially
    selected from all generated ν
    τ
    ’s so that they are well
    contained within the IC-22 detector (“golden events”).
    Next highest efficiency group is astrophysical neutrinos
    of all flavors. Note that, for the astrophysical ν
    τ
    ’s, they
    are unbiased data samples including all generated ν
    τ
    events unlike the “golden events”. Atmospheric neutrino
    backgrounds, prompt and conventional, come next to as-
    trophysical neutrinos. Atmospheric muon backgrounds,
    single and coincident, show the lowest efficiencies even

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁOD´ Z´ 2009
    3
    though they run out of statistics from L3 cut which need
    statistical improvement in near future.
    It is good that the cuts developed so far segregate
    well between astrophysical signal and atmospheric back-
    ground events. Within astrophysical neutrinos, however,
    the cuts are almost equally sensitive to all flavors so that
    we lose the discrimination power for the specific flavor
    under study, ν
    τ
    . This is due to the fact that the cuts are
    still quite general. To better distinguish astrophysical ν
    τ
    from astrophysical ν
    µ
    and ν
    e
    , which can more easily
    mimic lollipop than double-bang signatures, the future
    direction this analysis will take is to focus exclusively
    on the double-bang topology.
    IV. DOUBLE-BANG SEARCH
    Fig. 4 shows charge per DOM distribution as a
    function of DOM hit time for a simulated double-bang
    event produced from a primary neutrino energy of 50
    PeV. As shown in the figure the double-bang event has
    two showers separated by a track (403 m long).
    time (ns)
    10400 10600 10800 11000 11200 11400 11600 11800 12000 12200
    Q per DOM
    0
    200
    400
    600
    800
    1000
    1200
    1400
    1600
    1800
    2000
    Fig. 4. Charge (number of photo-electrons) per DOM distribution as
    a function of DOM hit time (ns) for a simulated double-bang event.
    The first peak corresponds to a shower from ν
    τ
    CC interaction and the
    second peak from the tau decay. Only hits arriving within 900 ns of
    residual time were used. (Residual time is the time difference between
    expected and actual photon arrival time.)
    Using the local current variable described above,
    requiring a large I
    L
    in both the first and last parts
    of the event, double-bang events can be selected.
    However, very energetic muons which produce two
    bremsstrahlungs in sequence could survive this cut.
    Further cuts are still being developed and evaluated.
    V. CONCLUSION
    Nature produces high energy neutrinos and they can
    be observed in all flavors. We try to detect especially
    high energy ν
    τ
    ’s which can leave unique signatures
    inside the IceCube detector. So far our approach is rather
    simple but we will continue investigate the IceCube
    potential especially for double-bang type events.
    REFERENCES
    [1] F. Halzen and D. Hooper, Rept. Prog. Phys. 65, 1025 (2002).
    [2] A. Neronov and M. Ribordy, arXiv:0905.0509 (2009).
    [3] S. Pakvasa, Nucl. Phys. B137 295 (2004).
    [4] D. Meloni and T. Ohlsson, Phys. Rev. D75, 125017 (2007).
    [5] J.G. Learned and S. Pakvasa, Astropart. Phys. 3, 267 (1995).
    [6] D. F. Cowen, J. Phys. 60, 227 (2007).
    [7] E. Bugaev et al., Astropart. Phys. 21, 491 (2004).
    [8] T. DeYoung, S. Razzaque and D. F. Cowen, J. of Phys. 60, 231
    (2007).
    [9] L. Pasquali and M. H. Reno, Phys. Rev. D59, 093003 (1999).
    [10] R. Enberg et al., Phys. Rev. D78, 043005 (2008).
    [11] M. Ribordy, Nucl. Inst. Meth. A574, 137 (2006).
    [12] J. Kiryluk, “First search for extraterrestrial neutrino-induced
    cascades with IceCube” in these proceedings.

    .

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    1
    Search for quantum gravity with IceCube and high energy
    atmospheric neutrinos
    Warren Huelsnitz
    and John Kelley
    for the IceCube Collaboration
    Department of Physics, University of Maryland, College Park, MD 20742, USA
    Department of Physics, University of Wisconsin, Madison, WI 53706, USA
    See the special section of these proceedings
    Abstract
    . We present the expected sensitivity of
    an analysis that will use data from the IceCube
    Neutrino Observatory to search for distortions in
    the energy or directional dependence of atmospheric
    neutrinos. Deviations in the energy and zenith angle
    distributions of atmospheric neutrinos due to Lorentz
    invariance violation or quantum decoherence could
    be a signature of quantum gravity in the neutrino
    sector. Additionally, a periodic variation as a func-
    tion of right ascension is a possible consequence
    of a Lorentz-violating preferred frame. We use a
    likelihood method to constrain deviations in the
    energy and zenith angle distributions and a discrete
    Fourier transform method to constrain a directional
    asymmetry in right ascension. In the absence of new
    physics, the likelihood method can also constrain
    conventional and prompt atmospheric neutrino flux
    models. Results from a similar analysis using data
    from the AMANDA-II detector are also discussed.
    Keywords
    : quantum gravity, Lorentz violation, at-
    mospheric neutrinos
    I. INTRODUCTION
    Physicists have so far been unable to reconcile quan-
    tum field theory and general relativity into a coherent
    theory of quantum gravity (QG). Numerous approaches
    are in development, and common to many is the possibil-
    ity that Lorentz invariance is violated at extremely small
    distance scales (high energy scales), due to a discrete
    structure of spacetime or an invariant minimum length
    scale. Interactions with a spacetime foam, or virtual
    black holes, may also induce quantum decoherence in
    which pure quantum states evolve into mixed states [1].
    Neutrinos, lacking any gauge interactions other than
    weak, and having extremely high Lorentz factors, are
    sensitive probes of these effects. Violation of Lorentz
    invariance (VLI) can induce a number of flavor-changing
    signatures in neutrinos, including oscillations with
    unique energy dependencies or directional asymmetries
    due to a Lorentz-violating preferred frame. Quantum
    decoherence (QD) can also result in flavor-changing
    effects that depend upon the neutrino energy.
    Atmospheric neutrinos are produced in the decay
    chains of particles resulting from the interaction of cos-
    mic rays with the earth’s atmosphere [2,3]. The IceCube
    neutrino telescope [4], currently under construction in
    the glacial ice at the South Pole, detects the Cherenkov
    radiation emitted by charged particles produced by neu-
    trino interactions in the ice or rock. IceCube has already
    collected a large sample of atmospheric muon neutrinos
    in the energy range of 100 GeV to a few tens of TeV,
    and can search for deficits caused by possible QG effects
    such as VLI or QD.
    We will first review the phenomenological models of
    QG to be tested. Then, we will discuss event selection
    and the observables used for the analysis, followed
    by a discussion of the likelihood and discrete Fourier
    transform (DFT) methods we use. Finally, results from
    the AMANDA-II detector, and expected sensitivity of
    the 40-string configuration of IceCube will be discussed.
    II. PHENOMENOLOGY
    A. Violation of Lorentz Invariance
    For VLI models, we consider the case of a flavor-
    dependent dispersion relation, or, equivalently, flavor-
    dependent limiting velocities that differ from the speed
    of light [5, 6]. Further, we make the simplifying as-
    sumption of a two neutrino model in which the new
    eigenstates are characterized by a mixing angle
    ξ
    and
    a phase
    η
    . This leads to a muon neutrino survival
    probability of the form:
    P
    ν
    µ
    →ν
    µ
    = 1  sin
    2
    2Θ sin
    2
    ?
    ∆m
    2
    L
    4E
    ?
    .
    E
    is the neutrino energy and
    L
    the propagation distance
    for the atmospheric neutrino, which is a function of
    zenith angle.
    Θ
    is the effective mixing angle, given by
    sin
    2
    2Θ =
    sin
    2
    2θ + R
    2
    sin
    2
    + 2R sin 2θ sin 2ξ cos η)
    ?
    2
    .
    The effective oscillation wavelength is
    ℜ =
    1 + R
    2
    + 2R [cos2θ cos2ξ
    + sin 2θ sin 2ξ cos η])
     1/2
    .
    R
    is the ratio between the VLI oscillation wavelength
    and the mass-induced oscillation wavelength:
    R =
    ∆c
    c
    E
    2
    4E
    ∆m
    2
    .
    ∆c/c
    is the velocity splitting between eigenstates. The
    VLI oscillation length can be generalized to integral

    2
    W. HUELSNITZ
    et al.
    SEARCH FOR QUANTUM GRAVITY
    powers of neutrino energy:
    ∆c
    c
    LE
    2
    → ∆δ
    LE
    n
    2
    .
    Since mass-induced oscillations are suppressed in the
    energy range for this analysis, we can make a simplify-
    ing assumption and set
    η = π/2
    so that
    cos η = 0
    . We
    then have two physics parameters:
    ∆c/c
    and
    sin
    2
    .
    B. Decoherence
    For quantum decoherence, we use a full three-neutrino
    model, in which the muon neutrino survival probability
    can be written [7, 8]:
    P
    ν
    µ
    →ν
    µ
    =
    1
    3
    +
    1
    2
    ?
    e
     γ
    3
    L
    cos
    4
    θ
    23
    +
    1
    12
    e
     γ
    8
    L
    (1  3 cos 2θ
    23
    )
    2
    + 4e
     (γ
    6+
    γ
    7)
    L
    2
    cos
    2
    θ
    23
    sin
    2
    θ
    23
    ×
    ?
    cos
    L
    m
    ?
    2
    ?
    + sin
    L
    m
    ?
    2
    ?
    6
     γ
    7
    )
    ?
    m
    ??
    ,
    (1)
    with m ≡
    ?
    ?
    ?
    6
     γ
    7
    )
    2
    ∆m
    2
    23
    ?
    E
    ?
    2
    ?
    ?
    ?
    .
    To limit the number of physics parameters, we assume
    that
    γ
    3
    = γ
    8
    and
    γ
    6
    = γ
    7
    . The
    γ
    i
    can be generalized to
    integral powers of neutrino energy:
    γ
    i
    → γ
    i
    ?
    E
    GeV
    ?
    n
    GeV.
    The units of
    γ
    i
    are then GeV
     n+1
    .
    C. Directional Asymmetry
    The location of IceCube at the South Pole is ideally
    suited to search for a sidereal variation in the flux of
    atmospheric neutrinos. Right ascension (RA) is synony-
    mous with sidereal phase, and azimuthal asymmetries
    in the detector average out over a year. We use a
    two-neutrino model derived from the Standard Model
    Extension (SME), known as the vector model [9]. This
    model predicts a survival probability that depends on the
    direction of neutrino propagation:
    P
    ν
    µ
    →ν
    µ
    = 1  sin
    2
    ?
    L
    ?
    (A
    s
    )
    µτ
    sin (α + ϕ
    0
    )
    + (A
    c
    )
    µτ
    cos (α + ϕ
    0
    )
    ??
    .
    α
    is the RA of the neutrino and
    ϕ
    0
    is the offset between
    the origin of our coordinate system and a ’preferred’
    direction.
    A
    s
    and
    A
    c
    are functions of neutrino energy,
    E
    ,
    neutrino direction unit vectors,
    N
    ˆ
    , and four coefficients
    from the SME, the
    (a
    L
    )
    µ
    and
    (c
    L
    )
    µν
    :
    (A
    s
    )
    µτ
    = N
    ˆ
    Y
    a
    X
    L
     2Ec
    TX
    L
    ?
     N
    ˆ
    X
    a
    Y
    L
     2Ec
    TY
    L
    ?
    ,
    (A
    c
    )
    µτ
    =  N
    ˆ
    X
    a
    X
    L
     2Ec
    TX
    L
    ?
     N
    ˆ
    Y
    a
    Y
    L
     2Ec
    TY
    L
    ?
    .
    Typically, we assume
    a
    X
    L
    = a
    Y
    L
    and
    c
    T X
    L
    = c
    TY
    L
    in
    the analysis. Additionally, while constraining the
    a
    L
    coefficients, the
    c
    L
    are set to 0, and when constraining
    the
    c
    L
    coefficients, the
    a
    L
    are set to 0.
    III. EVENT SELECTION
    We are interested in upgoing atmospheric
    ν
    µ
    events,
    and the main background is cosmic-ray muons. Even
    after an initial event selection based on zenith angle,
    the event sample is dominated, by several orders of
    magnitude, by misreconstructed cosmic-ray muons. This
    background is further reduced by event selection cuts
    that are based on track quality parameters and on fits
    to alternative track hypotheses. Alternative track hy-
    potheses include downgoing versus upgoing tracks, and
    coincident muon events.
    Since the remaining background contamination is dif-
    ficult to model with simulation, we require an essentially
    pure neutrino sample. Final event selection to achieve
    this level of purity is done using a Boosted Decision
    Tree (BDT) [10]. The event sample for one year of data
    from 40-string IceCube is expected to be about 20,000
    upgoing neutrinos, with zenith angles between 90 and
    180 degrees, and neutrino energies from 100 GeV to
    about 30 TeV.
    IV. FLUX MODELING AND BINNING
    Simulated events are weighted by their contribution
    to conventional [2, 3] and prompt [11–13] atmospheric
    neutrino flux models. These weights are then multiplied
    by the applicable oscillation or decoherence survival
    probability. Nuisance parameters are used in the likeli-
    hood analysis to account for the more significant theoret-
    ical and experimental uncertainties in flux normalization,
    spectral index, and zenith angle distribution. Individual
    events are thus weighted as follows:
    w = A {Bw
    conv
    + Cw
    prompt
    } P
    ν
    µ
    →ν
    µ
    ,
    where
    A = ε
    ?
    E
    1 TeV
    ?∆
    γ
    ?
    1+2α
    ?
    cos θ
    Z
    +
    1/2
    ??
    ,
    B =
    ?
    1 + 2α
    c
    ?
    cos θ
    Z
    +
    1/2
    ??
    , and
    C = A
    p
    ?
    E
    5 TeV
    ?∆
    γ
    p
    .
    ε
    accounts for theoretical and experimental uncer-
    tainties in the overall flux normalization, such as ice
    model uncertainties, optical module (OM) sensitivity
    uncertainty, interaction rate uncertainties, reconstruction
    errors, etc.
    ∆γ
    accounts for the uncertainty in the
    primary cosmic ray slope as well as the impact of OM
    and ice model uncertainties on the observed spectral
    index.
    α
    accounts for the impact of OM and ice model
    uncertainties on the zenith angle tilt of the observed flux.
    α
    c
    accounts for theoretical uncertainty in the zenith-
    angle tilt of the conventional atmospheric neutrino flux,
    primarily due to uncertainty in the pion to kaon ratio.
    A
    p
    and
    ∆γ
    p
    account for theoretical uncertainty in the
    magnitude and spectral index of the prompt atmospheric
    neutrino flux, primarily due to uncertainties in charm
    production cross sections and fragmentation functions.

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    3
    θ
    Z
    is the zenith angle of the neutrino, and
    P
    ν
    µ
    →ν
    µ
    is
    the oscillation or decoherence survival probability. The
    particular forms of the spectral tilt and zenith angle tilt
    equations were chosen to minimize the impact on overall
    normalization as the correction factors are varied.
    Events are binned in
    log
    10
    (dE/dX)
    and
    cos(θ
    Z
    )
    for tests of VLI, decoherence, and atmospheric flux
    models, and in RA for the vector model.
    dE/dX
    , with
    units of GeV m
     1
    , is the average energy loss per unit
    propagation length of a muon that would produce the
    detected amount of light, and serves as an estimator
    for the original neutrino energy. The energy resolution
    is about 0.3 on a log scale, reducing sensitivity to
    VLI effects by a factor of two as compared to perfect
    energy resolution. Histograms for
    log
    10
    (dE/dX)
    and
    cos(θ
    Z
    )
    are 10 x 10, and range from -1.9 to 1.1 for
    log
    10
    (dE/dX)
    and -1 to 0 for
    cos(θ
    Z
    )
    . 32 bins, from
    0 to 360
    , are used for RA, and include events in the
    declination band 0 to -30
    (zenith band 90 to 120
    ).
    Since some of the
    ν
    µ
    are assumed to oscillate to
    ν
    τ
    ,
    ν
    τ
    -induced muons are included in the simulation chain.
    Finally, bin counts for toy Monte Carlo (MC) histograms
    are varied according to Poisson distributions.
    V. LIKELIHOOD RATIO TEST
    To determine the compatibility of various new physics
    hypotheses with the data and identify acceptance re-
    gions, we use a likelihood-ratio test and the ordering
    principle of Feldman and Cousins [14]. The signal we
    are looking for is a distortion, or a warping, of the
    event counts in the energy-zenith plane. A likelihood
    analysis takes advantage of this shape of the distribution
    and provides a convenient way to include systematic
    uncertainties in the overall normalization and shape of
    the atmospheric neutrino flux. Systematic uncertainties
    are included using the nuisance parameters discussed
    above, and the profile construction method [15,16]. The
    likelihood function is:
    L ({n
    ij
    }|{µ
    ij
    r
    , θ
    s
    )}) =
    ?
    i,j
    µ
    n
    ij
    ij
    n
    ij
    !
    e
     µ
    ij
    .
    n
    is the binned toy MC or real data and
    µ
    is the
    prediction.
    θ
    r
    represents the physics parameters and
    θ
    s
    the nuisance parameters. In practice, this function is
    maximized by finding the minimum of the negative log
    of the likelihood, using the Minuit2 package in ROOT
    [17]. The test statistic is the likelihood ratio,
    R =  2ln
    L
    0
    L
    ˆ
    .
    where
    L
    0
    is the maximum likelihood, i.e., the best fit
    to the data or the toy MC histogram, with physics
    parameters held fixed and nuisance parmaters allowed
    to vary over the ranges of their uncertainties.
    L
    ˆ
    is the
    maximum likelihood when physics as well as nuisance
    parameters are allowed to vary.
    In the absence of new physics effects, the likelihood
    method will be used to evaluate theoretical uncertainties
    in conventional and prompt atmospheric neutrino flux
    models. For these analyses, experimental and theoretical
    uncertainties are split into separate nuisance parameters,
    and those associated with theoretical uncertainties in
    the conventional and/or prompt neutrino flux become
    physics parameters.
    VI. DFT ANALYSIS
    Neutrino oscillations in the vector model depend on
    the x and y components of the neutrino propagation di-
    rection. Hence, a phase angle specifying the offset from
    a preferred direction is required. To conduct a model-
    independent search for a sidereal signal independent of
    an arbitrary assumption about this phase angle, we use a
    DFT analysis. This analysis is done in two stages and has
    been adapted from a similar analysis performed with the
    MINOS detector to search for a directional dependence
    [18]. In the first stage, the data is checked for consistency
    with the hypothesis of no sidereal signal. In the second
    stage, constraints are placed on the SME coefficients of
    the vector model.
    First, a large number of toy experiments are performed
    in which the right ascensions of all events in the data
    are randomly redistributed. The power spectral densities
    (PSDs) in the
    n = 1
    to
    n = 4
    components of a DFT are
    computed for each of these ’noise-only’ toy experiments.
    The corresponding frequencies are
    n/T
    , where
    T
    is
    a sidereal day. The PSDs of the true data histogram are
    then computed and compared to the range of PSDs from
    the toy experiments. This indicates whether the data is
    consistent with the hypothesis of no sidereal signal.
    In the vector model, muon neutrino survivial probabil-
    ity varies with RA with a modulation frequency of
    4/T
    .
    To constrain the vector model, we look for an excess of
    power in the
    n = 4
    harmonic. The energy and zenith
    angle distributions are modeled using simulated events
    and the best-fit nuisance parameter values from the data.
    A large number of toy MC experiments are created,
    using these best-fit values. The physics parameters of
    the vector model are then increased, and the simulated
    events reweighted accordingly, until a PSD greater than
    the 99th percentile of the PSDs from the noise-only toy
    experiments is obtained. The values found in each of
    these trials are then averaged to find the sensitivity of
    this analysis given the data and the absence of a signal.
    VII. RESULTS AND EXPECTATIONS
    In a previous analysis of atmospheric muon neutrino
    events collected from 2000 to 2006 with the AMANDA-
    II detector [19], the data were consistent with the Stan-
    dard Model, and upper limits on QG parameters were
    set. A VLI upper limit at the 90% CL was found of
    ∆c/c < 2.8 × 10
     27
    for VLI oscillations proportional to the neutrino energy.
    A QD upper limit at the 90% CL was found of
    γ
    < 1.3 × 10
     31
    GeV
     1

    4
    W. HUELSNITZ
    et al.
    SEARCH FOR QUANTUM GRAVITY
    sin
    2
    (2ξ)
    0
    0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
    1
    c/c)
    Δ
    (
    10
    log
    -28
    -27.5
    -27
    -26.5
    -26
    -25.5
    -25
    -24.5
    -24
    Fig. 1: VLI Model, 90% CL curves. Dashed: AMANDA-
    II [19]. Dotted: SuperK + K2K [5]. Solid with hash
    marks: expected 80-string IceCube sensitivity [6]. Solid:
    expected 40-string IceCube sensitivity (preliminary).
    )
    7
    ,
    γ
    6
    10
    log
    -34 -33.5 -33 -32.5 -32 -31.5 -31 -30.5 -30
    )
    8
    γ
    ,
    3
    γ
    (
    10
    log
    -33
    -32
    -31
    -30
    -29
    -28
    -27
    -26
    -25
    Fig. 2: Decoherence model, 90% CL curves. Dashed:
    AMANDA-II [19]. Solid: expected 40-string IceCube
    sensitivity (preliminary). The black box indicates region
    scanned in AMANDA-II analysis.
    for decoherence effects proportional to
    E
    2
    and with all
    γ
    i
    assumed equal. For one year of data from 40-string
    IceCube, we expect about a factor of three improvement:
    ∆c/c < 9.0 × 10
     28
    and
    γ
    < 2.5 × 10
     32
    GeV
     1
    .
    Figure (1) shows 90% CL curves for the
    n = 1
    VLI
    model. Included are the AMANDA-II analysis, SuperK
    and K2K [5], and expected sensitivity for ten years of
    data from the full, 80-string, IceCube detector [6]. Also
    included is the 90% CL curve expected for the 40-string
    IceCube detector, based on a preliminary treatment of
    nuisance parameters. Figure (2) shows 90% CL curves
    for the
    n = 2
    decoherence model from the AMANDA-
    II analysis, and the expected sensitivity for 40-string
    IceCube (also preliminary).
    For the vector model, the sensitivity of the 40-string
    IceCube detector, at the 99% CL, is expected to be:
    a
    X
    L
    = a
    Y
    L
    < 2.0 × 10
     23
    GeV,
    c
    T X
    L
    = c
    TY
    L
    < 6.6 × 10
     27
    .
    These limits are three orders of magnitude lower for the
    a
    L
    terms and four orders of magnitude lower for the
    c
    L
    terms than the limits reported in [18]. This is due to
    the longer baseline of atmospheric neutrinos, and higher
    energy reach of IceCube.
    Data from AMANDA-II were used to find that the
    best-fit flux
    Φ
    for conventional atmospheric neutrinos,
    starting with the flux
    Barr
    )
    of reference [2] is:
    Φ = (1.1 ± 0.1)
    ?
    E
    640 GeV
    ?0
    .056
    Φ
    Barr
    .
    This likelihood methodology will also be used with
    IceCube data to constrain conventional and prompt at-
    mospheric neutrino flux models.
    VIII. CONCLUSIONS
    Data from IceCube’s 40-string configuration will im-
    prove constraints on VLI and decoherence models be-
    yond that achieved with AMANDA-II. Additionally, it
    will significantly improve constraints on a certain class
    of direction-dependent oscillation models.
    The IceCube detector will be able to provide improved
    constraints on various models of quantum gravity and
    atmospheric neutrino flux models as the detector grows
    to its final design configuration and as data collection
    continues in the following years. The likelihood method
    and the flux weighting discussed above provide flexi-
    bility to adjust nuisance parameter ranges as IceCube
    systematic uncertainties become better constrained.
    REFERENCES
    [1] S.W. Hawking,
    Commun. Math. Phys.
    , 87:395 (1982).
    [2] G.D. Barr
    et al.
    ,
    Phys. Rev. D.
    , 70:023006 (2004).
    [3] M. Honda
    et al.
    ,
    Phys. Rev. D.
    , 75:043006 (2007).
    [4] A. Karle,
    arXiv:0812.3981v1
    .
    [5] M.C. Gonzalez-Garcia and M. Maltoni,
    Phys. Rev. D.
    , 70:033010
    (2004).
    [6] M.C. Gonzalez-Garcia, F. Halzen and M. Maltoni,
    Phys. Rev. D.
    ,
    71:093010 (2005).
    [7] A.M. Gago
    et al.
    ,
    arXiv:hep-ph/0208166v1
    .
    [8] L. Anchordoqui and F. Halzen,
    Annals Phys.
    , 321:2660 (2006).
    [9] V.A. Kostelecky and M. Mewes,
    Phys. Rev. D.
    , 70:076002
    (2004).
    [10] A. Hocker
    et al.
    ,
    arXiv:physics/0703039v4
    .
    [11] R. Enberg, M.H. Reno and I. Sarcevic,
    Phys. Rev. D.
    , 78:043005
    (2008).
    [12] A.D. Martin, M.G. Ryskin and A.M. Stasto,
    Acta Phys. Polon.
    B
    , 34:3273 (2003).
    [13] G. Fiorentini, A. Naumov and F.L. Villante,
    Phys. Lett. B
    ,
    510:173 (2001).
    [14] G.J. Feldman and R.D. Cousins,
    Phys. Rev. D.
    , 57:3873 (1998).
    [15] G.J. Feldman, ”Multiple measurements and parameters in the
    unified approach,” Workshop on Confidence Limits, Fermilab
    (2000).
    [16] K. Cranmer, in
    Statistical Problems in Particle Physics,
    Astrophysics and Cosmology: Proceedings of PHYSTAT05
    ,
    L. Lyons and M.
    U¨nel
    , eds., (Univ. of Oxford, U.K.,
    2005);
    arXiv:physics/0511028v2
    .
    [17] R. Brun and F. Rademakers,
    Phys. Res. A
    , 389:81 (1997).
    [18] P. Adamson
    et al.
    ,
    Phys. Rev. Lett.
    , 101:151601 (2008).
    [19] R. Abbasi
    et al.
    ,
    arXiv:0902.0675v1
    .

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    1
    A First All-Particle Cosmic Ray Energy Spectrum From IceTop
    Fabian Kislat
    , Stefan Klepser
    , Hermann Kolanoski
    and Tilo Waldenmaier
    for the IceCube Collaboration
    DESY, D-15738 Zeuthen, Germany
    IFAE Edifici Cn., Campus UAB, E-08193 Bellaterra, Spain
    Institut fu¨r Physik, Humboldt-Universita¨t zu Berlin, D-12489 Berlin, Germany
    See special section of these proceedings
    Abstract
    . The IceTop air shower array is presently
    under construction at the geographic South Pole as
    part of the IceCube Observatory. It will consist of
    80 stations which are pairs of Ice-Cherenkov tanks
    covering an area of
    1 km
    2
    . In this paper a first
    analysis of the cosmic ray energy spectrum in the
    range
    2 · 10
    15
    eV
    to
    10
    17
    eV
    is presented using data
    taken in 2007 with 26 IceTop stations. The all-particle
    spectrum has been derived by unfolding the raw
    spectrum using response matrices for different mass
    compositions of the primaries. Exploiting the zenith
    angle dependence of the air shower development we
    have been able to constrain the range of possible
    composition models.
    Keywords
    : IceTop - Energy - Spectrum
    I. INTRODUCTION
    The IceTop air shower array is currently under con-
    struction as part of the IceCube Observatory at the
    geographic South Pole [1], [2]. Its 80 detector stations
    will cover an area of about
    1km
    2
    at an atmospheric
    depth of
    680 g/cm
    2
    . Each station consists of two ice
    filled
    1.8 m
    diameter tanks each equipped with two
    Digital Optical Modules (DOMs) [3] as photon sensors
    (the same as used by IceCube in the deep ice). The
    photomultipliers inside the two DOMs are operated at
    different gains to increase the dynamic range.
    Air showers are detected via the Cherenkov light
    emitted by charged particles inside the ice tanks. The
    light intensity is recorded by an ‘Analog Transient
    Waveform Digitizer’ (ATWD) at a
    300 MSPS
    sampling
    rate. When the signal inside a DOM crosses a threshold a
    ‘local coincidence’ signal is sent to neigbouring DOMs.
    Data taking is started when a local coincidence signal is
    received from the high gain DOM in the other tank of
    a station within
    250 ns
    . This ensures that the two tanks
    of a station always trigger together. This analysis only
    uses events where at least five stations have triggered.
    The analysis uses the signal sizes, obtained by inte-
    grating the waveforms, and the arrival times, determined
    from the leading edge of a waveform. Because of its high
    altitude of
    2835 m
    , IceTop is located close to the shower
    maximum for showers of energies between
    10
    15
    eV
    and
    10
    17
    eV
    (about
    550 g/cm
    2
    to
    720 g/cm
    2
    ). Therefore,
    the signals measured by IceTop are dominated by the
    electromagnetic component of the air showers.
    Fig. 1. An example of a lateral signal size distribution from a single
    event in IceTop. Each data point corresponds to the signal in terms of
    Vertical Equivalent Muons measured in an IceTop tank. The curve is
    a fit of the lateral distribution function (1).
    II. ENERGY RECONSTRUCTION
    Generally, the shower direction can be reconstructed
    from the arrival times of signals, while the core position
    and the primary energy of the air shower are inferred
    from the lateral distribution of signal sizes. These signal
    sizes are calibrated with signals from vertical muons to
    eliminate differences between the tanks. Signals given in
    units of Vertical Equivalent Muons (VEM) are a detector
    independent measure of the intensity of an air shower at
    the position of a tank.
    A first estimate of the shower core position is obtained
    by finding the signal center-of-gravity which is defined
    as the average of the tank positions weighted by the
    square-root of the signal size. The shower direction is
    obtained by fitting a plane to the measured signal times.
    These two first guesses are used as an input to a more
    accurate iterative maximum likelihood fit.
    In the latter procedure a lateral distribution function
    is fitted to the measured signal sizes and an arrival
    time distribution is fitted to the signal times. The lateral
    distribution function [4] has been obtained from air
    shower and detector simulations and corresponds to a
    second order polynomial on a double logarithmic scale:
    S (r) = S
    ref
    ?
    r
    R
    ref
    ?
     β
    ref
     κ log
    10
    (r/R
    ref
    )
    (1)

    2
    FABIAN KISLAT
    et al.
    ICETOP ENERGY SPECTRUM
    S
    ref
    is the signal expectation value at the reference ra-
    dius
    R
    ref
    from the shower axis,
    β
    ref
    is a slope parameter
    related to the shower age and
    κ
    is a curvature parameter
    of the lateral signal distribution function. Based on a
    study of the stability of fit results the reference radius
    has been fixed to
    R
    ref
    = 125 m
    which also corresponds
    to the IceTop grid spacing. An example is shown in
    Figure 1. The likelihood function used in the fitting
    procedure is based on a study of signal fluctuations and
    also takes into account stations that do not trigger [4].
    Taking into account the spatial curvature of the shower
    front, which is assumed to have a fixed profile [5], the
    arrival time distribution
    t(r)
    is given by:
    t(r) = 19.41 ns
    e
    (
    r
    118.1m
    )
    2
     1
    ?
     4.823 · 10
     4
    ns
    m
    2
    r
    2
    (2)
    This
    t(r)
    indicates the time delay of a signal at distance
    r
    from the shower axis with respect to a planar shower
    front through the shower core and perpendicular to the
    shower axis. Using this shower front parametrisation
    improves the direction resolution compared to the plane
    fit first guess result.
    The complete log-likelihood function, therefore, has
    three terms,
    L = L
    hit
    + L
    nohit
    + L
    time
    (3)
    L
    hit
    is based on the log-normal distribution of signal
    sizes obtained from the study of signal size fluctuations,
    L
    time
    is based on an assumed Gaussian distribution of
    arrival times and
    L
    nohit
    is defined by
    L
    nohit
    =
    ?
    log
    1  P
    A
    hit
    P
    B
    hit
    ?
    (4)
    P
    i
    hit
    is the probability that tank
    i
    of a station triggers
    given the signal expectation of the fit at that iteration
    step, given the shower core position and direction. The
    likelihood accounts for the ‘local coincidence’ condition
    which requires both tanks of a station to trigger before
    signals are transmitted. The sum in (4) runs over all
    stations that did not trigger.
    The fit procedure is divided into several steps to
    improve the stability. At first the direction is fixed
    to the initial first guess value and only the lateral
    signal distribution (1) is fitted. In a second iteration
    the direction is also varied but the parameters of the
    lateral fit are limited to a
    ±3σ
    range around the values
    obtained in the first step. Finally a last iteration with
    fixed direction is performed. The slope parameter of the
    lateral distribution function is limited throughout this
    procedure to
    1.5 ≤ β
    125
    ≤ 5
    .
    Using the results of the fit a first-guess energy is
    determined as a function of shower size and the zenith
    angle. As a measure of the shower size, the signal
    expectation value
    S
    ?log r?
    , at the distance
    ?log r?
    from
    the core is used, where
    ?log r?
    is the average of the
    logarithmic distance of tanks to the shower axis. The size
    parameter
    S
    ?log r?
    has been found to be least correlated
    to the other fit parameters (in general
    S
    125
    and the slope
    10
    (E/PeV)
    log
    -1
    -0.5
    0
    0.5
    1
    1.5
    2
    -1
    sr
    -1
    s
    -2
    dI/dlg(E) / m
    -10
    10
    -9
    10
    -8
    10
    -7
    10
    -6
    10
    -5
    10
    0°<θ<30°
    30°<θ<40°
    40°<θ<46°
    preliminary
    Fig. 2.
    Raw IceTop energy spectrum divided into three different
    zenith angle ranges (see text for details).
    β
    125
    are correlated). From air shower simulations an
    energy estimator
    E
    rec
    (S
    ?log r?
    , θ)
    for any given
    ?log r?
    based on a proton hypothesis has been derived.
    To ensure the quality of the reconstructed data several
    quality criteria are applied:
    Only showers with a zenith angle
    θ < 46
    are con-
    sidered restricting the analysis to a well understood
    zenith angle range;
    The value of the slope parameter must be
    1.55 ≤
    β
    125
    < 4.95
    removing all showers for which the
    likelihood fit ran into the limits on this parameter;
    The uncertainty on the shower core position must
    be less than
    20m
    ;
    The core position from the likelihood fit and the
    center-of-gravity first guess must be inside the
    IceTop array,
    50m
    away from the array border.
    Furthermore, the station with the largest signal must
    not be on the border of the IceTop array.
    The last item is to exclude showers with a core outside
    the array which have a high probability to be mis-
    reconstructed. Using the likelihood fit result alone is
    not sufficient because the fit tends to reconstruct these
    cores inside the array. These strict requirements on the
    core reconstruction are necessary because the position
    of the core directly influences the interpretation of the
    fitted lateral distribution. Data was taken between June
    1st 2007 and October 31st 2007 with the 26 IceTop
    stations operated at that time. The filter level data sample
    contained 11 262 511 events. After all of the above cuts,
    4 131 343 events remained.
    The raw energy spectrum, that is the distribution of the
    reconstructed energies
    E
    rec
    without any further correc-
    tions or unfolding applied, but after the abovementioned
    cuts, is shown in Figure 2. The data is split into three
    zenith angle ranges,
    0 ≤ θ < 30
    ,
    30
    ≤ θ < 40
    and
    40
    ≤ 46
    .
    III. AIR SHOWER SIMULATIONS
    The relation between the measured signals and the
    properties of the primary particle can only be obtained

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    3
    true
    /PeV)
    10
    (E
    log
    -1
    -0.5
    0
    0.5
    1
    1.5
    2
    )
    true
    (E
    10
    log
    -
    )
    reco
    (E
    10
    log
    -0.4
    -0.3
    -0.2
    -0.1
    0
    0.1
    0.2
    0.3
    0.4
    Proton
    °
    ≤θ<
    30
    °
    0
    °
    ≤θ<
    40
    °
    30
    °
    ≤θ<
    46
    °
    40
    (a) Primary protons
    true
    /PeV)
    10
    (E
    log
    -0.5
    0
    0.5
    1
    1.5
    2
    )
    true
    (E
    10
    log
    -
    )
    reco
    (E
    10
    log
    -0.4
    -0.3
    -0.2
    -0.1
    0
    0.1
    0.2
    0.3
    0.4
    Iron
    °
    ≤θ<
    30
    °
    0
    °
    ≤θ<
    40
    °
    30
    °
    ≤θ<
    46
    °
    40
    (b) Primary iron
    Fig. 3. Systematic energy shift for (a) protons and (b) iron when using the energy estimator
    E
    rec
    as a function of the true energy of a particle.
    This energy mis-reconstruction is one of the parameters describing the energy response of the IceTop detector. The strong increase towards low
    primary energies is a threshold effect as explained in the text. Since
    E
    rec
    is based on the proton assumption the energy mis-reconstruction for
    protons is small while that for primary iron larger and shows a clear zenith angle dependance.
    from simulations of air showers and the detector. There-
    fore, a Monte Carlo shower library containing 98 760
    proton and iron showers with primary energies between
    100 TeV
    and
    100 PeV
    has been generated with COR-
    SIKA [6]. The hadronic interaction models SIBYLL [7]
    and Fluka 2006.3b [8] and the CORSIKA atmosphere
    model No. 14 (South Pole, Dec. 31, 1997, MSIS-90-E)
    were used.
    IV. UNFOLDING THE SPECTRUM
    The energy estimator
    E
    rec
    does not fully account
    for the angular and energy dependence of the detec-
    tor acceptance and smearing. This biases the result
    especially in the threshold region where the detection
    efficiency increases strongly with energy. In addition,
    the calculation of
    E
    rec
    assumes proton primaries and
    one specific atmosphere profile (CORSIKA atmosphere
    12, South Pole July 01, 1997). A different atmosphere
    profile or a different composition of primary particles
    can result in a different energy spectrum. All these
    effects can be taken into account by unfolding the raw
    energy spectrum.
    The unfolding algorithm uses a response matrix which
    contains the probability that an incident primary particle
    with true energy
    E
    true
    will be assigned the energy
    E
    rec
    .
    Response matrices are generated from CORSIKA air
    shower simulations for proton and iron primaries in three
    different zenith angle ranges. Response matrices for
    mixed compositions are obtained as linear combinations
    of these single primary particle responses.
    To calculate the response matrices the simulated
    data are subdivided into 30 bins of true primary en-
    ergies, equidistant in
    log
    10
    (E
    true
    )
    . For each bin the
    energy response is obtained from a distribution of
    log
    10
    (E
    rec
    /E
    true
    )
    of all reconstructed events which is
    approximately normally distributed. Non-zero values of
    this quantity indicate a wrong energy reconstruction. As
    such the mean of this distribution is the ‘energy mis-
    reconstruction’, the width is the energy resolution. The
    reconstruction efficiency is the ratio between the number
    of reconstructed and generated events.
    Depending on the zenith angle, the IceTop detector
    gets fully efficient at about
    2 PeV
    with an effective area
    of
    0.096 km
    2
    . The energy resolution improves with in-
    creasing energy and reaches a value of
    0.05
    in logarithm
    of energy at
    10 PeV
    primary energy which corresponds
    to a statistical uncertainty of roughly
    12 %
    .
    An important property of the detector response is
    the ‘energy mis-reconstruction’ shown in Figure 3. It
    is the systematic difference between the true primary
    energy and the reconstructed energy. The energy mis-
    reconstruction shows a strong dependence on the pri-
    mary particle mass and thus on the assumed primary
    composition. The attenuation of iron induced showers
    with increasing slant depth is greater than for proton
    showers. This leads to an underestimation of the primary
    energy for inclined iron showers when using the proton-
    based energy estimator
    E
    rec
    . This is nicely visible in
    Figure 3b.
    V. PRELIMINARY ENERGY SPECTRUM FROM ICETOP
    Based on these response matrices the three raw energy
    spectra of Figure 2 are unfolded using an iterative
    unfolding method [9]. An unfolding based on three
    different composition assumptions has been made: pure
    proton, pure iron and the two-component model consist-
    ing of proton and iron primaries as motivated in [10].
    The results are shown in Figure 4. Above
    4 PeV
    the spectra have a spectral index ranging from
    2.93
    to
    3.19
    , depending on the composition assumption and the
    zenith angle band. The absolute normalization varies
    between
    3.77 · 10
     15
    m
     2
    s
     1
    sr
     1
    GeV
     1
    and
    7.93 ·
    10
     15
    m
     2
    s
     1
    sr
     1
    GeV
     1
    at an energy of
    10 PeV
    .
    The abovementioned mis-reconstruction of the energy
    of inclined iron showers leads to a relative shift of the

    4
    FABIAN KISLAT
    et al.
    ICETOP ENERGY SPECTRUM
    PROTON
    10
    (E/PeV)
    log
    0
    0.2
    0.4
    0.6
    0.8
    1
    1.2
    1.4
    1.6
    1.8
    2
    )
    -1
    sr
    -1
    s
    -2
    m
    1.5
    / (PeV
    1.5
    E
    ×
    dI/dlg(E)
    -6
    10
    -5
    10
    0°<
    θ<
    30°
    30°<
    θ
    < 40°
    40°<
    θ
    < 46°
    preliminary
    IRON
    10
    (E/PeV)
    log
    0
    0.2
    0.4
    0.6
    0.8
    1
    1.2
    1.4
    1.6
    1.8
    2
    )
    -1
    sr
    -1
    s
    -2
    m
    1.5
    / (PeV
    1.5
    E
    ×
    dI/dlg(E)
    -6
    10
    -5
    10
    0°<
    θ<
    30°
    30°<
    θ
    < 40°
    40°<
    θ
    < 46°
    preliminary
    TWO-COMPONENT
    10
    (E/PeV)
    log
    0
    0.2
    0.4
    0.6
    0.8
    1
    1.2
    1.4
    1.6
    1.8
    2
    )
    -1
    sr
    -1
    s
    -2
    m
    1.5
    / (PeV
    1.5
    E
    ×
    dI/dlg(E)
    -6
    10
    -5
    10
    0°<
    θ<
    30°
    30°<
    θ
    < 40°
    40°<
    θ
    < 46°
    preliminary
    Fig. 4. Unfolding result for three different assumptions on the primary
    compositions: pure protons, pure iron, two-component model [10].
    spectra from different zenith bands in the unfolding. In
    case of proton primaries the spectrum from the most
    vertical zenith bin shows the largest flux and the spec-
    trum from the
    40
    to
    46
    zenith band shows the lowest
    flux. This order is reversed if the primary particles are
    assumed to be purely iron. Since there are no indications
    of an anisotropy of the arrival directions of cosmic rays
    in the energy range under consideration, isotropy has to
    be assumed. An isotropic flux, however, means that the
    unfolded spectra obtained from different zenith bands
    must agree.
    VI. CONCLUSIONS AND OUTLOOK
    A first analysis of the energy spectrum of cosmic
    rays in the range between
    2 PeV
    and
    100 PeV
    has been
    presented. This analysis uses an unfolding procedure to
    account for detector and atmospheric influences as well
    as differences in the shower development of different
    primary particles. Energy spectra have been obtained for
    three different compositions and three different zenith
    angle ranges. In case of a pure proton or iron hypothesis
    the three different inclination spectra do not agree which
    is in conflict with the assumption of an isotropic flux.
    Therefore those two pure composition assumptions can
    be excluded.
    Several issues have not yet been addressed in this
    early study. Most importantly, the systematic uncertain-
    ties due to the interaction models used in the simula-
    tions must be studied. Also the influence of different
    atmosphere parametrisations in the simulation must be
    analysed and understood. This becomes even more im-
    portant when analysing data from a longer period of
    time. Currently, in 2009, the detector is running with
    nearly three quarters of the full detector.
    REFERENCES
    [1] T.K. Gaisser
    et al. IceTop: The Surface Component of IceCube
    ,
    in
    Proc. 28th ICRC
    , Tsukuba, Japan, 2003.
    [2] T.K. Gaisser
    et al. Performance of the IceTop Array
    , in
    Proc.
    30th ICRC
    , Me´rida, Mexico, 2007; arXiv:0711.0353v1.
    [3] R. Abbasi
    et al.
    “The IceCube Data Acquisition System: Signal
    Capture, Digitization, and Timestamping,” Nucl. Instrum. Meth.
    A
    601
    , 294 (2009).
    [4] S. Klepser
    et al. Lateral Distribution of Air Shower Signals and
    Initial Energy Spectrum above 1 PeV from IceTop
    , in
    Proc. 30th
    ICRC
    , Me´rida, Mexico, 2007; arXiv:0711.0353v1.
    [5] S. Klepser
    et al. First Results from the IceTop Air Shower Array
    ,
    in
    Proc. 21st ECRS
    , Kosˇice, Slovakia, 2008; arXiv:0811.1671v1.
    [6] D. Heck
    et al.
    , Report
    FZKA 6019
    (1998), Forschungszentrum
    Karlsruhe;
    http://www-ik.fzk.de/corsika/physics
    description/
    corsika phys.html
    [7] R. Engel, T.K. Gaisser, P. Lipari, and T. Stanev,
    Air shower
    calculations with the new version of SIBYLL,
    in
    Proc. 26th ICRC
    ,
    Salt Lake City, USA, 1999.
    [8] A. Ferrari, P.R. Sala, A. Fasso`, and J. Ranft, “FLUKA: a multi-
    particle transport code”, CERN-2005-10 (2005), INFN/TC 05/11,
    SLAC-R-773.
    [9] G. D’Agostini, “A multidimensional unfolding method based on
    Bayes’ theorem,” Nucl. Instrum. Meth. A
    362
    , 487 (1995)
    [10] R. Glasstetter
    et al. Analysis of electron and muon size spectra
    of EAS
    , in
    Proc. 26th ICRC
    , Salt Lake City, USA, 1999.
    [11] S. Klepser, “Reconstruction of Extensive Air Showers and Mea-
    surement of the Cosmic Ray Energy Spectrum in the Range
    of
    1  80 PeV
    at the South Pole”, Dissertation, Humboldt-
    Universita¨t zu Berlin, 2008.

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁOD´ Z´ 2009
    1
    Reconstruction of IceCube coincident events and study of
    composition-sensitive observables using both the surface and deep
    detector
    Tom Feusels
    , Jonathan Eisch
    and Chen Xu
    , for the IceCube Collaboration
    §
    Dept. of Subatomic and Radiation Physics, University of Gent, B-9000 Gent, Belgium
    Dept. of Physics, University of Wisconsin, Madison, WI 53706, USA
    Bartol Research Institute, University of Delaware, Newark, DE 19716, USA
    §
    See the special section in these proceedings
    Abstract. The combined information from cosmic
    ray air showers that trigger both the surface and
    underground parts of the IceCube Neutrino Observa-
    tory allows the reconstruction of both the energy and
    mass of the primary particle through the knee region
    of the energy spectrum and above. The properties
    of high-energy muon bundles, created early in the
    formation of extensive air showers and capable of
    penetrating deep into the ice, are related to the
    primary energy and composition.
    New methods for reconstructing the direction and
    composition-sensitive properties of muon bundles are
    shown. Based on a likelihood minimization pro-
    cedure using IceCube signals, and accounting for
    photon propagation, ice properties, and the energy
    loss processes of muons in ice, the muon bundle
    energy loss is reconstructed. The results of the high-
    energy muon bundle reconstruction in the deep ice
    and the reconstruction of the lateral distribution of
    low energy particles in the surface detector can be
    combined to study primary composition and energy.
    The performance and composition sensitivity for both
    simulated and experimental data are discussed.
    Keywords:
    Cosmic
    ray
    composition,
    IceTop/IceCube, high-energy muon bundles
    I. INTRODUCTION
    The cosmic ray spectrum covers many orders of mag-
    nitude in both energy and flux. In the energy range ac-
    cessible to the IceCube Neutrino Observatory (∼0.3 PeV
    to 1 EeV) the slope of the spectrum remains mostly
    constant, except for a feature at around 3 PeV where
    the spectrum steepens. This feature is called the knee
    of the cosmic ray spectrum, and its origin is unknown.
    Proposed explanations include changes in acceleration
    mechanisms or cosmic rays leaking from the galactic
    magnetic field starting at this energy. Measurement of
    mass composition in this range could give clues to
    the origin of these cosmic rays. The IceCube detector,
    together with the IceTop air shower detector, provides
    an opportunity to measure the composition of cosmic
    ray particles in the region of the knee and beyond.
    The IceTop detector, high on the Antarctic Plain at
    an average atmospheric depth of 680 g/cm
    2
    , consists of
    a hexagonal grid of detector stations 125 m apart. Each
    station consists of two ice tanks that act as Cherenkov
    media for measuring mainly the electromagnetic compo-
    nent of cosmic ray air showers. In each tank, two Digital
    Optical Modules (DOMs) are deployed, which contain a
    10 inch PMT and digital readout and control electronics.
    The primary energy can be reconstructed by the IceTop
    surface array [1].
    Deep below each IceTop station is a string of the
    IceCube detector with 60 DOMs evenly spaced between
    1.5 and 2.5 km in the ice. Combined, the IceTop and
    IceCube arrays can reconstruct the air shower core
    position and direction while measuring the shower signal
    strength at the surface and the energy deposition of
    the high-energy muon bundle in the deep ice. With
    these measurements, the energy and mass of the primary
    cosmic rays can be reconstructed.
    A characteristic difference between showers induced
    by light and heavy nuclei for a fixed primary energy
    is their number of muons, with higher mass primaries
    producing more muon-rich showers. However, the muon
    multiplicity is not directly measured by either the IceTop
    or IceCube detectors. The deep IceCube detector is
    sensitive to Cherenkov light coming mainly from energy
    loss processes of high-energy muon bundles. This energy
    loss is a convolution of the muon multiplicity of the
    shower, the muon energy distribution and the energy
    loss of a single muon. If the energy loss behavior of
    muon bundles can be reconstructed accurately, it can be
    used as a primary mass indicator [2]. In Section IV, the
    reconstructed muon bundle track described in Section III
    will be used as a seed to reconstruct the muon bundle
    energy loss. Simulated data is compared to experimental
    data in Section V to examine the detector performance
    and its sensitivity to composition.
    II. DATA SAMPLE AND SIMULATION
    Our experimental data sample was taken from the
    month of September, 2008. At that time, the detector
    consisted of 40 IceTop stations and 40 IceCube strings.
    A total livetime of 28.47 days was obtained by selecting
    only runs where both detectors were stable. The events
    were processed at the South Pole with a filter which
    required at least three triggered IceTop stations and at

    2
    T. FEUSELS et al. RECONSTRUCTION OF ICECUBE COINCIDENT EVENTS
    least 8 triggered IceCube DOMs. After this filter about
    3.31·10
    6
    coincident events remained.
    To study the direction and energy reconstruction of
    muon bundles in the ice, a large number of proton
    and iron showers between 10 TeV and 46.5 PeV were
    simulated with the CORSIKA [3] package. SIBYLL
    2.1 [4], [5] was used as the high-energy (>80 TeV)
    hadronic interaction model, while FLUKA08 [6], [7]
    was used as the low energy hadronic interaction model.
    For this study, CORSIKA was configured to use a model
    of the South Pole atmosphere typical for the month
    of July [8]. The showers were simulated according to
    an E
    −1
    spectrum and then reweighted according to an
    E
    −2.7
    spectrum before the knee (at 3 PeV) and an E
    −3.0
    spectrum after the knee.
    The IceCube software environment was used to re-
    sample each CORSIKA shower 500 times on and around
    the detector, to propagate the high-energy muons through
    the ice, and to simulate the detector response and trigger.
    The simulation was filtered the same way as data and
    yielded about 9.0·10
    4
    proton and 9.0·10
    4
    iron events.
    III. DIRECTION RECONSTRUCTION
    The direction reconstruction by IceTop will be de-
    scribed first because it will be used later as a seed for
    an IceCube muon bundle reconstruction algorithm.
    An initial shower core position and direction is de-
    termined using the extracted times and charges of the
    recorded pulses from the IceTop tanks. The first guess
    core position is the calculated center of gravity of tank
    signals, while the initial direction reconstruction assumes
    a flat shower front. This shower core and direction are
    then used as a seed to fit the lateral distribution of pulses
    with the double logarithmic parabola (DLP) function
    described in [9]. Because the reconstruction of the core
    and direction are highly correlated, the resolution is
    improved by fitting the shower core position and the
    shower direction together, with a curved shower front
    and the DLP lateral particle distribution. This fit, and
    some geometrical quality cuts discussed later, gives an
    angular resolution
    1
    of 1.0
    for iron showers (see the dot-
    dot-dashed line in Fig. 1) and a core resolution of 15.0 m
    (see dot-dashed line in Fig. 2).
    The resolution depends strongly on where the shower
    core lands with respect to the IceTop array. This analysis
    uses only events with a shower core reconstructed within
    the geometrical area of the IceTop detector. The recon-
    structed muon bundle track was also required to pass
    within the instrumented volume of the IceCube detector.
    The direction that was already reconstructed by the
    IceTop algorithms alone can serve as a seed for an algo-
    rithm more specialized in reconstructing muon bundles
    in the ice. This algorithm uses only the charges measured
    R
    x
    1
    The resolution of observable Y is defined according to
    0
    p(∆Y ) d∆Y = 0.683, where x is the resolution and p(∆Y )
    the frequency of distances between the true and the reconstructed
    observable.
    Δ Ψ
    (degrees)
    0
    0.5
    1
    1.5
    2
    2.5
    3
    Fraction
    0
    0.02
    0.04
    0.06
    0.08
    0.1
    Δ
    Δ ΨΨ
    distributiondistribution
    Fig. 1.
    The distribution of the angles between the true direction
    and the reconstructed direction for simulations of iron showers is
    shown for different algorithms. The dot-dot-dashed curve shows the
    reconstruction which uses IceTop information alone. For the dot-
    dashed line, the reconstructed core position by IceTop is fixed and the
    zenith and azimuth are determined by using a muon bundle algorithm
    seeded with the track determined by IceTop. The solid line illustrates
    the ideal limit of this reconstruction method by using the true shower
    core position instead of the reconstructed one. The dashed curve
    is obtained when a second iteration between IceTop and IceCube
    algorithms is used.
    by the IceCube DOMs and takes into account the range-
    out of the muons in a bundle [10]. By keeping the
    reconstructed core position on the surface fixed, a large
    lever arm of at least 1500 m is obtained. This limits
    the track parameters (zenith and azimuth) during the
    minimization procedure and reduces the number of free
    track parameters from 5 to 2.
    In Fig. 1, it can be seen that this method (dot-dashed
    line) improves on the angular resolution determined by
    IceTop alone. If the core position can be determined
    more accurately, the direction reconstruction will be
    even better. The solid line on Fig. 1 represents the ideal
    limit, obtained using the true core position. Therefore,
    to improve the core resolution the new direction is kept
    fixed and used to seed the IceTop lateral distribution
    function which then only fits the core position (dashed
    line on Fig. 2). Iterating over both the surface and the
    deep detector reconstructions with progressively better
    core position and direction seeds, leads to the optimal
    resolution. The ideal limit for the core resolution is
    acquired by seeding the IceTop algorithm with the true
    direction and is shown on Fig. 2 (solid line). After the
    second iteration this limit is already obtained and gives
    a core resolution of 14.0 m and an angular resolution of
    0.9
    (see dashed lines in Figures 1 and 2).
    Using this combined method for muon bundle di-
    rection reconstruction, an almost energy independent
    angular resolution of 0.8
    and a core resolution of
    12.5 m is obtained for proton induced showers. An im-
    provement of the core resolution and direction resolution
    also improves the resolution of the reconstructed shower
    size and shower age.

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁOD´ Z´ 2009
    3
    Δ
    R (m)
    0
    5
    10
    15
    20
    25
    30
    Fraction
    0
    0.01
    0.02
    0.03
    0.04
    Δ
    Δ
    R
    R distributiondistribution
    Fig. 2. The distribution of distances between the true shower core
    position and the reconstructed core position at the surface for iron
    showers is plotted for different reconstructions. For the dash-dotted
    line the IceTop DLP function that determines the core position used a
    first guess direction. The dashed line, where the IceTop algorithm was
    seeded with a better direction reconstruction from the deep ice, gives
    a slightly better core reconstruction. When the true direction is used
    as a seed, the ideal limit represented by the solid line is obtained.
    IV. MUON BUNDLE ENERGY RECONSTRUCTION
    An energy reconstruction algorithm for single muons
    was previously reported in Ref. [12]. Using IceCube
    signals and lookup tables together with a previously
    reconstructed track a constant energy loss is fit with a
    likelihood function. The lookup tables model the South
    Pole ice properties and the propagation of Cherenkov
    photons through the ice. A single high-energy muon
    above 730 GeV loses energy mainly by radiative pro-
    cesses like Bremsstrahlung and pair production, which
    produce secondary electromagnetic cascades in the ice
    along the muon bundle track. Therefore, an infinite light
    source with mono-energetic cascades every meter is used
    as a model for a single muon.
    This light model can also be used for muon bundles.
    The main difference is that a slant depth dependent
    energy loss will be needed because of the range-out of
    muons. Here, the slant depth is defined as the distance
    along the muon bundle track between the point where the
    muon enters the ice and the point where the Cherenkov
    light is emitted.
    The equation for the muon bundle energy loss is:
    μ
    dE
    µ
    dX
    Bundle
    (X) =
    Z
    E
    max(surf)
    E
    min(surf)
    dN
    µ
    dE
    µ
    dE
    µ
    dX
    dE
    µ
    (surf) ,
    (1)
    where
    dN
    µ
    dE
    µ
    is the energy distribution of the muons and
    dE
    µ
    dX
    is the energy loss of a single muon. E
    min(surf)
    =
    a
    b
    ¡
    e
    bX
    − 1
    ¢
    is the minimum energy that a muon needs
    to get to depth X. E
    max(surf)
    E
    0
    A
    is the maximum
    energy a muon from a shower induced by a particle with
    A nucleons and primary energy E
    0
    can have.
    Using a simple power-law as an approximation for the
    Elbert formula [13], which describes the multiplicity of
    Slant Depth (m)
    1400 1600 1800 2000 2200 2400 2600 2800
    Muon bundle energy loss (GeV/m)
    0
    20
    40
    60
    80
    100
    120
    140
    Reconstructed
    Reconstructed slant
    slant depth
    depth behaviorbehavior
    Fig. 3. An example of reconstructed muon bundle energy loss for a
    single 17 PeV iron and proton shower. The reconstructed slant depth
    behavior follows the true energy loss reasonably well. The triangles
    (squares) are the muon bundle energy loss processes for a certain
    proton (iron) 17 PeV shower with a zenith angle of 28.4
    (12.6
    ).
    The spread of the points illustrates the stochastic nature. The solid
    (dashed) line is the reconstructed muon bundle energy loss function,
    described in the text, for the proton (iron) shower.
    high-energy muons in air showers, the differential muon
    energy distribution becomes:
    dN
    µ
    dE
    µ
    = γ
    µ
    κ(A)
    μ
    E
    0
    A
    γ
    µ
    −1
    E
    −γ
    µ
    −1
    µ
    ,
    (2)
    where γ
    µ
    = 1.757 is the muon integral spectral index
    and κ is a normalization that depends on the shower
    properties.
    With the solution of the single muon energy loss
    equation, E
    µ
    (X) =
    ¡
    E
    µ
    (surf) +
    a
    b
    ¢
    e
    −bX
    a
    b
    , the
    average energy loss formula can be expressed as a
    function of the muon energy at the surface:
    dE
    µ
    dX
    (X) = −a − bE
    µ
    (X)
    = −b
    ³
    E
    µ
    (surf) +
    a
    b
    ´
    e
    −bX
    ,
    (3)
    with a = 0.260 GeV/m, the ionization energy loss
    constant and b = 0.000357 m
    −1
    , the stochastic energy
    loss constant from [11].
    The average muon bundle energy loss function is then
    obtained by integrating Eq. (1) using Eqs. (2) and (3).
    This energy loss fit function, with κ and E
    0
    /A as free
    parameters, will be used to scale the expected charges
    in a DOM from the likelihood formula in [12] instead
    of scaling it with a constant energy loss.
    In Fig. 3, the curves show the reconstructed muon
    bundle energy loss functions for 17 PeV primaries. The
    data points are the energy losses calculated from simula-
    tions. It can be clearly seen that the muon bundle energy
    loss function obtained by minimizing the likelihood
    function describes the depth behavior for these two

    4
    T. FEUSELS et al. RECONSTRUCTION OF ICECUBE COINCIDENT EVENTS
    (Reconstructed Primary Energy/GeV)
    10
    log
    6
    6.1
    6.2
    6.3
    6.4
    6.5
    6.6
    6.7
    6.8
    6.9
    7
    (Muon Bundle Energy Loss/(GeV/m))
    10
    log
    0.2
    0.4
    0.6
    0.8
    1
    1.2
    1.4
    1.6
    1.8
    Comparison of Simulation and Data
    Proton MC median
    Proton MC center 68%
    Iron MC median
    Iron MC center 68%
    Data center 68%
    IceCube Preliminary
    Fig. 4.
    The reconstructed muon bundle energy loss evaluated at a
    slant depth of 1650 m versus the reconstructed shower primary energy.
    Shown are the center 68% of the proton simulation (right diagonal
    hashes), the iron simulation (left diagonal hashes) and the center 68%
    of the data (enclosed in the rectangles). The median of the distribution
    is also shown for the simulated data sets. The events were filtered as
    described in Section II with the additional quality requirement that the
    muon bundle reconstruction algorithm used signals from at least 50
    DOMs, to remove events with poorly fit energy loss. This quality cut
    removes a higher portion of muon bundles with low energy loss and
    will be accounted for in a full composition analysis.
    showers better than a constant energy loss function.
    It has been shown in [2] that proton and iron show-
    ers are separated better by the muon bundle energy
    loss at smaller slant depths. The reconstructed IceCube
    composition-sensitive parameter which will be used fur-
    ther on in the coincidence analysis, is the energy loss
    at the top of the IceCube detector, at a slant depth of
    1650 m. At this slant depth, Cherenkov light from show-
    ers with a zenith angle up to 30
    can still be detected
    by the upper DOMs, making the energy reconstruction
    more accurate over the entire zenith range.
    V. COMBINED PRIMARY MASS AND ENERGY
    RECONSTRUCTION
    Fig. 4 shows a comparison of proton and iron primary
    simulation and experimental data described in Section
    II using the reconstruction methods described in the
    sections III and IV. While this plot has only rough
    quality cuts, it can be seen that the spread in the
    simulation is similar to the spread in the data. The
    median muon bundle energy loss from an iron primary
    shower is approximately a factor of 2 higher in the
    ice than the median muon bundle energy loss from
    a proton primary shower for the same reconstructed
    primary energy. There is a large overlap area, where
    shower to shower fluctuations overcome the effect of the
    primary mass. It is difficult to reconstruct the primary
    mass of a single shower with any certainty due to these
    large fluctuations.
    VI. CONCLUSION AND OUTLOOK
    The IceTop and IceCube detectors of the IceCube
    Neutrino Observatory can be used together for an im-
    proved air shower core location and direction reconstruc-
    tion. The promising method of reconstructing the muon
    bundle energy loss behavior will be further developed
    to be used in a measurement of cosmic ray mass and
    energy. This study used only a subsample of the available
    data; with more statistics and an enlarged detector these
    methods can be extended up to 1 EeV.
    VII. ACKNOWLEDGEMENTS
    This work is supported by the Office of Polar Pro-
    grams of the National Science Foundation and by FWO-
    Flanders, Belgium.
    REFERENCES
    [1] F. Kislat et al., A First All-Particle Cosmic Ray Energy Spectrum
    From IceTop, These Proceedings.
    [2] X. Bai et al., Muon Bundle Energy Loss in Deep Underground
    Detector, These Proceedings.
    [3] D. Heck et al., CORSIKA FZKA 6019, Forschungszentrum
    Karlsruhe, 1998.
    [4] R. S. Fletcher et al., SIBYLL: An event generator for simulation
    of high energy cosmic ray cascades, Phys. Rev. D, 50 5710,
    1994.
    [5] R. Engel et al., Air shower calculations with the new version of
    SIBYLL, In Proc. 26th ICRC, Salt Lake City, 1999.
    [6] A. Fasso` et al., FLUKA: a multi-particle transport code , CERN-
    2005-10 (2005), INFN/TC 05/11, SLAC-R-773.
    [7] G. Battistoni et al., The FLUKA code: Description and bench-
    marking , Proceedings of the Hadronic Shower Simulation Work-
    shop 2006, Fermilab 6–8 September 2006, M. Albrow, R. Raja
    eds., AIP Conference Proceeding 896, 31-49, (2007).
    [8] D. Chirkin, Parameterization based on the MSIS-90-E model,
    1997, Private communication.
    [9] S. Klepser et al., Lateral Distribution of Air Shower Signals and
    Initial Energy Spectrum above 1 PeV from IceTop, In Proc. 30th
    ICRC, Merida, Mexico, 2007.
    [10] K. Rawlins, Ph.D. Dissertation, UW-Madison (2001).
    [11] P. Miocinoˇ
    vic,´ Ph.D. Dissertation, UC Berkeley (2001).
    [12] S. Grullon et al., Reconstruction of high-energy muon events in
    IceCube using waveforms, In Proc. 30th ICRC, Merida, Mexico,
    2007.
    [13] J.W. Elbert, In Proc. DUMAND Summer Workshop (ed. A.
    Roberts), 1978, vol 2, p.101.

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    1
    Small air showers in IceTop
    Bakhtiyar Ruzybayev
    , Shahid Hussain
    , Chen Xu
    and Thomas Gaisser
    for the IceCube Collaboration
    Bartol Research Institute, Department of Physics and Astronomy, University of Delaware, Newark, DE 19716, U.S.A.
    See the special section of these proceedings.
    Abstract
    . IceTop is an air shower array that is
    part of the IceCube Observatory currently under
    construction at the geographic South Pole [1]. When
    completed, it will consist of 80 stations covering an
    area of 1 km
    2
    . Previous analyzes done with IceTop
    studied the events that triggered five or more stations,
    leading to an effective energy threshold of about
    0.5PeV [2]. The goal of this study is to push this
    threshold lower, into the region where it will overlap
    with direct measurements of cosmic rays which
    currently have an upper limit around 300 TeV [3]. We
    select showers that trigger exactly three or exactly
    four adjacent surface stations that are not on the
    periphery of the detector (contained events). This
    extends the energy threshold down to 150 TeV.
    Keywords
    : IceTop, Air showers, Cosmic rays
    around the “knee”.
    I. INTRODUCTION
    During 2008, IceCube ran with forty IceTop stations
    and forty IceCube strings in a triangular grid with a mean
    separation of 125 m. In the 2008–2009 season, additional
    38 IceTop tanks and 18 standard IceCube strings were
    deployed as shown in Fig.1. When completed, IceCube
    will consist of eighty surface stations, eighty standard
    strings and six special strings in the ”DeepCore” sub-
    array [4]. Each IceTop station consists of two ice filled
    tanks separated by 10 m, each equipped with two Digital
    Optical Modules (DOMs) [5]. The photo multipliers
    inside the two DOMs are operated at different gains
    to increase the dynamic range of the response of a
    tank. The DOMs detect the Cherenkov light emitted
    by charged shower particles inside the ice tanks. Data
    recording starts when local coincidence condition is
    satisfied, that is when both tanks are hit within a
    250 nanoseconds interval. In this paper we used the
    experimental data taken with the forty station array and
    compared to simulations of this detector configuration.
    Here we describe the response of IceTop in its threshold
    region.
    II. ANALYSIS
    The main difference between this study and analyzes
    done with five or more stations triggering is the accep-
    tance criterion. In previous analyzes, we accepted events
    with five or more hit stations and with reconstructed
    shower core location within the predefined containment
    area (shaded area in Fig.1). In addition, the station with
    X (m)
    -800
    -600
    -400
    -200
    0
    200
    400
    600
    Y (m)
    -800
    -600
    -400
    -200
    0
    200
    400
    600
    IceCube strings
    IceTop tanks
    2009 IceCube strings
    2009 IceTop tanks
    Fig. 1: Surface map of IceCube in 2009. New stations are
    unfilled markers. The shaded area (200795 m
    2
    ) contains
    stations that are defined as inner stations.
    the biggest signal in the event must also be located
    within the containment area.
    In the present analysis we used events that triggered
    only three or four stations, thus complementing analyzes
    with five or more stations. Selection of the events was
    based solely on the stations that were triggered. The
    criteria are:
    1) Triggered stations must be close to each other
    (neighboring stations). For three station events,
    stations form almost an equilateral triangle. For
    four station events, stations form a diamond shape.
    2) Triggered stations must be located inside the ar-
    ray (shaded region in Fig.1). Events that trigger
    stations on the periphery are discarded.
    Since we are using stations on the periphery as a veto,
    we ensure that our selected events will have shower cores
    contained within the boundary of the array. In addition,
    these events will have a narrow energy distribution. We
    analyzed events in four solid angle bins with zenith
    angles
    θ
    : 0
    –26
    , 26
    –37
    , 37
    –45
    , 45
    –53
    . Results
    for the first bin,
    θ = 0
    26
    , are emphasized in this
    paper. This near-vertical sample will include most of
    the events with muons seen in coincidence with the deep
    part of IceCube.

    2
    B. RUZYBAYEV
    et al.
    SMALL AIR SHOWERS
    Fig. 2: The all-particle spectrum from air shower measurements as summarized in Figure 24.9 of Review of Particle
    Physics [3]. The shaded area indicates the range of direct measurements. The thick black line shows the flux model
    used for this analysis and the vertical lines indicate the energy range responsible for 95% of the 3 and 4 station
    events.
    A. Experimental data and simulations
    The experimental data used in this analysis were
    taken during an eight hour run on September 1st, 2008.
    Two sets of air shower simulations were produced:
    pure proton primaries in the energy range of 21.4 TeV–
    10.0 PeV, and pure iron primaries in the energy range
    of 45.7 TeV–10.0 PeV. All air showers were produced in
    zenith angle range:
    0
    ≤ θ ≤ 65
    .
    Our simulation used the following flux model:
    dF
    dE
    = Φ
    0
    ?
    E
    E
    0
    ?
     γ
    (1)
    Φ
    0
    = 2.6 · 10
     4
    GeV
     1
    s
     1
    sr
     1
    m
     2
    E
    0
    = 1 TeV
    γ = 2.7
    for both proton and iron primaries. The normalizations
    were chosen such that the fluxes will fit the all particle
    cosmic ray spectrum as shown in Fig. 2. Simulated
    showers were dropped randomly in a circular area,
    around the center of the 40 station array (
    X = 100 m
    ,
    Y = 250 m
    ) with a radius of 600 m.
    B. Shower Reconstruction
    Since the showers that trigger only three or four
    stations are relatively small, we use a plane shower
    front approximation and the arrival times to reconstruct
    the direction. The shower core location is estimated by
    calculating the center of gravity of the square root of the
    charges in the stations. For the energy reconstruction
    we use the lateral fit method [6] that IceTop uses to
    reconstruct events with five or more stations triggered.
    This method uses shower sizes at the detector level
    to estimate the energy of the primary particle. Heavier
    primary nuclei produce showers that do not penetrate as
    deeply into the atmosphere as the proton primaries of the
    same energy. As a result, iron primary showers will have
    a smaller size at the detector level than proton showers
    of the same energy. We define a reconstructed energy
    based on simulations of primary protons and fitted to the
    lateral distribution and size of proton showers. Therefore
    the parameter for reconstructed energy underestimates
    the energy when applied to showers generated by heavy
    primaries. We observe a linear correlation between true
    and reconstructed energies in this narrow energy range
    and use this to correct the reconstructed energies. We
    reconstruct the experimental data assuming pure proton
    or pure iron primaries.
    C. Effective Area
    We use the simulations to determine the effective area
    as a function of energy. Effective area is defined as
    A
    eff
    =
    Rate[E
    min
    , E
    max
    ]
    ∆Ω · F
    sum
    (2)
    F
    sum
    =
    E
    ?
    max
    E
    min
    Φ
    0
    ?
    E
    E
    0
    ?
     γ
    dE
    (3)
    where Rate[E
    min
    ,E
    max
    ] is total rate for a given energy
    bin,
    ∆Ω
    is the solid angle of the bin and
    F
    sum
    is the

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    3
    true
    /GeV)
    10
    (E
    Log
    4
    4.5
    5
    5.5
    6
    6.5
    7
    )
    2
    Effective Area (m
    10
    2
    3
    10
    10
    4
    5
    10
    3 or more stations
    5 or more stations
    3 stations only
    4 stations only
    preliminary
    Fig. 3: Effective areas for different triggers in the most vertical zenith angle range:
    0
    ≤ θ ≤ 26
    , derived using
    true quantities from simulations.
    reco
    / GeV)
    10
    (E
    Log
    4.6
    4.8
    5
    5.2
    5.4
    5.6
    5.8
    6
    6.2
    6.4
    Rate / Hz
    0
    0.01
    0.02
    0.03
    0.04
    0.05
    0.06
    0.07
    θ
    : 0
    °
    - 26
    °
    θ
    : 26
    °
    - 37
    °
    θ
    : 37
    °
    - 45
    °
    θ
    : 45
    °
    - 53
    °
    preliminary
    (a) Proton simulation
    reco
    / GeV)
    10
    (E
    Log
    4.6
    4.8
    5
    5.2
    5.4
    5.6
    5.8
    6
    6.2
    6.4
    Rate / Hz
    0
    0.002
    0.004
    0.006
    0.008
    0.01
    0.012
    0.014
    0.016
    0.018
    0.02
    0.022
    0.024
    θ
    : 0
    °
    - 26
    °
    θ
    : 26
    °
    - 37
    °
    θ
    : 37
    °
    - 45
    °
    θ
    : 45
    °
    - 53
    °
    preliminary
    (b) Iron simulation
    Fig. 4: Reconstructed energy distributions for proton and iron simulations for 3 station events in four zenith bins.
    total flux in the given energy bin. Figure 3 shows the
    calculated effective areas, using the true values of energy
    and direction, for different trigger combinations, in the
    most vertical bin (
    θ = 0
    26
    ).
    III. RESULTS
    We summarize the results of simulations and compar-
    ison to data in Figures 4–6.
    In Fig.4, we see that the energy distribution of the
    event rates depends on the zenith angle and the primary
    type. As expected, the peak of the energy distribution
    moves to higher energies for larger zenith angles and
    heavier primaries; these features of the distributions will
    be very helpful in unfolding the cosmic ray spectrum and
    composition.
    Figure 5 shows the energy distributions in the most
    vertical zenith bin (
    θ = 0
    26
    ). Experimental data is
    reconstructed twice, first with a pure proton assumption,
    then with a pure iron assumption. For three stations
    triggered (Fig. 5a), the energy distribution for pure
    proton simulation with the flux model as defined in (1)
    has a better agreement to the experimental data than iron
    simulation. For four stations triggered (Fig. 5b), we have
    a similar picture but the peaks of the distributions are
    shifted to the right since on average we need a higher

    4
    B. RUZYBAYEV
    et al.
    SMALL AIR SHOWERS
    reco
    / GeV)
    10
    (E
    Log
    4.5
    5
    5.5
    6
    6.5
    Rate / Hz
    0
    0.01
    0.02
    0.03
    0.04
    0.05
    0.06
    0.07
    0.08
    0.09
    0.1
    Exp.data, proton assumption
    Proton simulation
    Exp.data, iron assumption
    Iron simulation
    preliminary
    (a) 3 stations
    reco
    / GeV)
    10
    (E
    Log
    4.5
    5
    5.5
    6
    6.5
    Rate / Hz
    0
    0.005
    0.01
    0.015
    0.02
    0.025
    0.03
    Exp.data, proton assumption
    Proton simulation
    Exp.data, iron assumption
    Iron simulation
    preliminary
    (b) 4 stations
    Fig. 5: Reconstructed energy distributions for 3 and 4 station events with zenith angles
    0
    26
    . Experimental data
    is reconstructed twice: assuming pure proton and pure iron primary.
    cos(
    θ
    )
    0.4
    0.5
    0.6
    0.7
    0.8
    0.9
    1
    Rate / Hz
    0
    0.02
    0.04
    0.06
    0.08
    0.1
    0.12
    0.14
    0.16
    0.18
    Experimental Data
    Proton simulation
    Iron Simulation
    preliminary
    Fig. 6: Reconstructed zenith distributions for 3 station
    events.
    energy primary to trigger four stations. By including
    three station events we can lower the threshold down
    to 130 TeV.
    Figure 6 shows the zenith distributions of the events.
    Distribution for pure iron simulation is lower than for
    proton simulation since fewer iron primaries reach the
    detector level at lower energies. The deficiency of sim-
    ulated events in the most vertical bin may be due to the
    fact that we used a constant
    γ
    of 2.7 for all energies
    and at these energies
    γ
    is most probably changing
    continuously. In the most vertical bin showers must have
    a lower energy than showers at greater zenith angle.
    Starting from a lower
    γ
    and gradually increasing it for
    higher energies will increase events in vertical bin and
    decrease them at higher energies, thus improving the
    zenith angle distribution. It is possible to further improve
    the fit of the proton simulation to the experimental data
    by adjusting the parameters
    γ
    and
    Φ
    0
    of the model.
    IV. CONCLUSION
    We have demonstrated the possibility of extending
    the IceTop analysis down to energies of 130 TeV, low
    enough to overlap the direct measurements of cosmic
    rays. Compared to IceTop effective area for five and
    more station hits, our results show a significant increase
    in effective area for energies between 100–300 TeV (Fig.
    3). We plan to include three and four station events in
    the analysis of coincident events to determine primary
    composition, along the lines described in [7]. Overall
    results of this analysis encourage us to continue and
    improve our analysis of small showers.
    REFERENCES
    [1] T.K. Gaisser et al., “Performance of the IceTop array”, in Proc.
    30th ICRC, Me´rida, Mexico, 2007.
    [2] F. Kislat et al., “A first all-particle cosmic ray energy spectrum
    from IceTop”, this conference.
    [3] C. Amsler et al. (Particle Data Group), Physics Letters
    B667
    , 1
    (2008).
    [4] A. Karle et al., arXiv:0812.3981, “IceCube: Construction Status
    and First Results”.
    [5] R. Abbasi et al., “The IceCube Data Acquisition System: Signal
    Capture, Digitization, and Timestamping”, Nucl. Instrum. Meth.
    A
    601
    , 294 (2009).
    [6] S. Klepser, PhD Thesis, Humboldt-Universita¨t zu Berlin (2008).
    [7] T. Feusels et al., “Reconstruction of IceCube coincident events
    and study of composition-sensitive observables using both the
    surface and deep detector”, this conference.

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    1
    Cosmic Ray Composition using SPASE-2 and AMANDA-II
    K. Andeen
    and K. Rawlins
    For the IceCube Collaboration
    University of Wisconsin-Madison, 1150 University Ave, Madison, WI
    University of Alaska Anchorage, 3211 Providence Dr, Anchorage, AK
    See special section of these proceedings
    Abstract
    . The precise measurement of cosmic ray
    mass composition in the region of the knee (3 PeV)
    is critical to understanding the origin of high energy
    cosmic rays. Therefore, air showers have been ob-
    served at the South Pole using the SPASE-2 surface
    array and the AMANDA-II neutrino telescope which
    simultaneously measure the electronic air shower
    component at the surface and the muonic air shower
    component in deep ice, respectively. These two com-
    ponents, together with a Monte Carlo simulation and
    a well-understood analysis method will soon yield the
    relative cosmic ray composition in the knee region.
    We report on the capabilities of this analysis.
    Keywords
    :
    composition,
    cosmic-ray,
    neural-
    network
    I. INTRODUCTION
    The mass composition of high-energy cosmic rays
    around and above the knee in the energy spectrum
    (∼3 PeV) is dependent upon the mechanisms of cosmic
    ray production, acceleration, and propagation. Therefore,
    the study of mass composition is critical to understand-
    ing the origins of cosmic rays in this energy region. At
    energies up to 10
    14
    eV mass composition can be mea-
    sured directly using balloon and satellite experiments;
    however, due to the low flux, composition above 10
    14
    eV
    must be obtained from indirect measurements on the
    ground. Indirect measurements of composition involve
    a close examination of the air shower produced as a
    cosmic ray primary smashes into Earth’s atmosphere. By
    utilizing information from more than one component of
    the shower, such as the electronic and muonic compo-
    nents, the energy and relative mass can be obtained from
    primary particles with much higher energies than those
    currently measurable by direct detection techniques.
    II. DATA AND RECONSTRUCTION
    One such indirect measurement is possible using the
    South Pole Air Shower Experiment (SPASE-2) in coin-
    cidence with the Antarctic Muon And Neutrino Detector
    Array (AMANDA-II). The SPASE-2 detector is situated
    on the surface of the South Pole at an atmospheric depth
    of
    ∼685
    g cm
     2
    and is composed of 30 stations in a
    30 m triangular grid. Each station contains four 0.2 m
    2
    scintillators. The AMANDA-II detector lies deep in the
    ice such that the center-to-center separation between
    the deep ice and the surface arrays is
    ∼1730
    m with
    an angular offset of 12
    . AMANDA-II consists of 677
    optical modules (OMs) deployed on 19 detector strings
    spanning depths from 1500-2000 m below the surface of
    the ice. Each OM contains a photomultiplier tube (PMT)
    which is optimized for detection of the Cherenkov light
    emitted by particles—namely muon bundles—passing
    through the ice. In addition to the composition analysis,
    this coincident configuration allows for calibration as
    well as measurement of the angular resolution of the
    AMANDA-II detector. [1]
    For this analysis, coincident data from the years 2003-
    2005 are used, for a total livetime of around 600 days.
    For comparison with the data, Monte Carlo simulated
    proton, helium, oxygen, silicon and iron air showers
    with energies between 100 TeV and 100 PeV have been
    produced using the CORSIKA air shower generator with
    the SIBYLL/GEISHA hadronic interaction models [2].
    At the surface the air showers are injected into GEANT4,
    which simulates the SPASE-II detector response [3]. The
    showers are then propagated through the ice and the
    response of AMANDA-II detector is simulated using the
    standard software package of the AMANDA collabora-
    tion. An E
     1
    spectrum is used for generation, but for
    analysis the events are re-weighted to the cosmic ray
    energy spectrum of E
     2.7
    at energies below the knee
    at 3 PeV and E
     3.2
    above. Both the experimental data
    and the Monte Carlo simulated data are then put through
    identical reconstruction chains.
    The first step in the reconstruction uses information
    from SPASE-2 only. The goal of this first reconstruction
    is to find the shower direction, shower size, and core
    position of the incoming air shower. The direction can
    be computed from the arrival times of the charged
    particles in the SPASE-2 scintillators, while the shower
    core position and shower size are acquired by fitting the
    lateral distribution of particle density to the Nishimura-
    Kamata-Greisen (NKG) function. Evaluating the fit at a
    fixed distance from the center of the shower (in this case
    30 m) [4] gives a parameter called S30, which has units
    of particles/m
    2
    and will be used throughout this paper
    as a measure of the electronic part of the air shower.
    The next step in the reconstruction provides a measure
    of the muon component of the air shower from a
    combined reconstruction which uses both the surface
    and deep ice detectors. The core position of the shower
    from the SPASE-only fit is held fixed while
    θ
    and
    φ
    are
    varied in the ice to find the best fit of the muon track
    in the AMANDA-II detector. Holding the core fixed at

    2
    K. ANDEEN
    et al.
    SPASE/AMANDA COMPOSITION
    the surface allows for a lever arm of about 1730 m
    when calculating directionality, providing a very tight
    angular resolution for the track. The expected lateral
    distribution function (LDF) of the photons resulting
    from the muon bundle in AMANDA-II is computed
    and corrected for both the ranging out of muons as
    they progress downward through the detector, as well
    as the changing scattering length as a function of depth
    in the ice caused by dust layers. The LDF is then fit to
    the hit optical modules and evaluated at a perpendicular
    distance of 50 m from the center of the shower [5]. This
    parameter, called K50, has units of photoelectrons/OM
    and will be used throughout the rest of this paper as the
    measure of the muon component of the air shower.
    III. ANALYSIS DETAILS
    Once reconstruction has been completed it is impor-
    tant to find and eliminate poorly reconstructed events.
    Thus events have been discarded which:
    have a reconstructed shower core outside the area
    of SPASE-2 or a reconstructed muon track passing
    outside the volume of AMANDA-II,
    have an unreasonable number of hits in the ice
    given S30 at the surface (these events represent
    large showers which landed outside of SPASE-
    2 and were misreconstructed within the array as
    having a small S30),
    have an unphysical reconstructed attenuation length
    of light in the ice (an unphysical reconstruction of
    attenuation length will lead to a misreconstructed
    value for K50), or
    are reconstructed independently in SPASE-2 and
    AMANDA-II as coming from significantly different
    locations in the sky.
    After these cuts have been made, it can be seen in Fig. 1
    that our two main observables, S30 and K50, form a
    parameter space in which primary energy and primary
    mass separate. This is expected, since the showers asso-
    ciated with the heavier primaries develop earlier in the
    atmosphere and hence have more muons per electron
    by the time they reach the surface than the showers
    associated with lighter primaries [6]. This means that
    K50, which is proportional to the number of muons in
    the ice, will be higher for heavier primaries than for
    lighter primaries of the same S30, as is observed.
    In the three-year data set used for this analysis, more
    than 100,000 events survive all quality cuts. It is inter-
    esting to notice that in the previous analysis, using the
    SPASE-2/AMANDA-B10 detector [5], the final number
    of events for one year was 5,655. Furthermore, the
    larger detector used here is sensitive to higher energy
    events. The significant increases in both statistics and
    sensitivity, along with a new detector simulation and
    revised reconstruction algorithm for the SPASE-2 array,
    are the basis for performing a new analysis.
    (S30)
    10
    log
    -0.5
    0
    0.5
    1
    1.5
    2
    2.5
    3
    (K50)
    10
    log
    -2.5
    -2
    -1.5
    -1
    -0.5
    0
    0.5
    1
    1.5
    Proton Showers
    GeV
    5.6
    E
    true
    = 10
    (S30)
    10
    log
    -0.5
    0
    0.5
    1
    1.5
    2
    2.5
    3
    (K50)
    10
    log
    -2.5
    -2
    -1.5
    -1
    -0.5
    0
    0.5
    1
    1.5
    Iron Showers
    GeV
    5.6
    E
    true
    = 10
    Fig. 1.
    The two main observables, log
    10
    (K50) vs log
    10
    (S30),
    in the Monte Carlo simulation with protons above (red) and iron
    below (blue). The black contours depict lines of constant energy from
    5.4
    >
    log
    10
    (Etrue/GeV)
    >
    6.8 marked every log
    10
    (Etrue/GeV) = 0.2.
    The black line along the energy gradient approximates a division
    between proton and iron showers and is included merely as a reference
    between the two plots. It is clear that mass and energy are on a roughly
    linear axis.
    A. Calibration
    To accurately measure the composition using both
    electron and muon information, the Monte Carlo sim-
    ulations must provide an accurate representation of the
    overall amplitude of light in ice (measured here as K50).
    However, due to the model-dependencies displayed by
    air shower simulations, the overall light amplitude is
    subject to systematic errors. Therefore, it is important to
    calibrate the composition measurements at low energies
    where balloon experiments have provided direct mea-
    surements of cosmic ray composition. In light of this, a
    vertical “slice” in S30 is selected for calibration which
    corresponds with the highest energies measured directly.
    At these energies, the direct measurements indicate that
    <lnA>≈
    2, or
    50%
    protons and
    50%
    iron [7]. The K50
    values of the data in this S30 “slice” are thus adjusted by

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    3
    )
    true
    (E
    10
    )-log
    NN
    (E
    10
    log
    -0.4
    -0.2
    0
    0.2
    0.4
    Fraction of Events
    0
    0.05
    0.1
    0.15
    0.2
    0.25
    0.3
    /GeV) < 5.8
    True
    (E
    10
    5.6 < log
    )
    true
    (E
    10
    )-log
    NN
    (E
    10
    log
    -0.4
    -0.2
    0
    0.2
    0.4
    Fraction of Events
    0
    0.05
    0.1
    0.15
    0.2
    0.25
    0.3
    /GeV) < 6.0
    True
    (E
    10
    5.8 < log
    )
    true
    (E
    10
    )-log
    NN
    (E
    10
    log
    -0.4
    -0.2
    0
    0.2
    0.4
    Fraction of Events
    0
    0.1
    0.2
    0.3
    0.4
    /GeV) < 6.2
    True
    (E
    10
    6.0 < log
    )
    true
    (E
    10
    )-log
    NN
    (E
    10
    log
    -0.4
    -0.2
    0
    0.2
    0.4
    Fraction of Events
    0
    0.1
    0.2
    0.3
    0.4
    /GeV) < 6.4
    True
    (E
    10
    6.2 < log
    )
    true
    (E
    10
    )-log
    NN
    (E
    10
    log
    -0.4
    -0.2
    0
    0.2
    0.4
    Fraction of Events
    0
    0.1
    0.2
    0.3
    0.4
    0.5
    /GeV) < 6.6
    True
    (E
    10
    6.4 < log
    )
    true
    (E
    10
    )-log
    NN
    (E
    10
    log
    -0.4
    -0.2
    0
    0.2
    0.4
    Fraction of Events
    0
    0.1
    0.2
    0.3
    0.4
    0.5
    0.6
    0.7
    /GeV) < 6.8
    True
    (E
    10
    6.6 < log
    Fig. 2. Energy resolution (the difference between the true primary energy and the energy reconstructed by the neural network) is shown in bins
    of true energy for iron (solid blue), proton (dashed red) and oxygen (shaded green) primaries. Each energy bin is bounded by two consecutive
    contours from Fig. 1 (where the indicated energy contour is the lower bound of the first energy bin above). For easier comparison, a Gaussian
    distribution has been fit to these energy resolution histograms and the mean and sigma of each Gaussian can be found in Table 1.
    TABLE I
    MEAN AND SIGMA OF A GAUSSIAN DISTRIBUTION FIT TO EACH ENERGY RESOLUTION HISTOGRAM IN FIG. 2.
    Shower Type
    Gaussian Statistic
    True Energy Bins (log
    10
    (E) / GeV)
    5.6 - 5.8
    5.8 - 6.0
    6.0 - 6.2
    6.2 - 6.4
    6.4 - 6.6
    6.6 - 6.8
    Proton
    Mean
    0.03
    0.02
    0.00
    0.02
    0.01
    0.00
    σ
    0.12
    0.11
    0.08
    0.06
    0.05
    0.04
    Oxygen
    Mean
    -0.02
    -0.01
    -0.03
    0.00
    0.01
    0.02
    σ
    0.12
    0.10
    0.07
    0.07
    0.05
    0.09
    Iron
    Mean
    -0.12
    -0.05
    -0.04
    -0.02
    -0.01
    0.05
    σ
    0.15
    0.09
    0.08
    0.08
    0.06
    0.15
    an offset to match the distribution of K50 corresponding
    to a
    50%  50%
    proton and iron sample.
    B. Neural Network Reconstruction of Energy and Mass
    Similar past investigations utilized the quasi-linear re-
    lationship between K50/S30 and mass/energy, as seen in
    Fig. 1, to employ an analysis wherein, after calibration,
    the axis is rotated to correspond to the mass/energy
    coordinate plane [5] [8]. The rotation analysis works
    quite well up to energies slightly above the knee;
    however, beyond the knee the relationship between
    K50/S30 and mass/energy becomes increasingly non-
    linear as the air showers approach the energy where
    the shower maximum occurs at the atmospheric depth
    of the SPASE-2/AMANDA-II detectors. As the data
    set used here has significantly more statistics at these
    higher energies than previous studies, it was important
    to find a new procedure for extracting the composition. A
    neural network technique has therefore been developed
    to resolve the mean logarithmic mass at all energies [9].
    The main change to the neural network technique
    since its development has been to distinguish between
    the calculation for energy and the calculation for mass
    by using two separate networks. The first neural network
    (NN1) is trained to find the primary energy by using
    log
    10
    (K50) and log
    10
    (S30) as input parameters, followed
    by one hidden layer. The second network (NN2) is
    trained to find the primary mass of the air shower. NN2
    also takes as input log
    10
    (K50) and log
    10
    (S30) and also
    has a single hidden layer of neurons. In both cases
    the network is trained through a number of “epochs”,
    or training cycles, on half of the simulated proton and
    iron showers and tested on the other half of the proton
    and iron showers. The results of testing determine the

    4
    K. ANDEEN
    et al.
    SPASE/AMANDA COMPOSITION
    Mass Output
    0
    0.2
    0.4
    0.6
    0.8
    1
    Fraction
    0.1
    of Events
    0.2
    NN Trained on All Types
    Mass Output
    0
    0.2
    0.4
    0.6
    0.8
    1
    Fraction of Events
    0.1
    0.2
    0.3
    NN Trained on Protons and Iron Only
    Fig. 3. Mass output of the neural network from the bin 6.2
    <
    log
    10
    (Ereco/GeV)
    <
    6.4 for three primary types: iron (solid blue), oxygen
    (shaded green), and protons (dashed red). On the left is the output from a network trained on all particle types. On the right is the output from
    a network trained only on protons and iron. As expected, when the neural network is trained on more particle types it begins to reconstruct
    the intermediate primaries in their proper location as opposed to attempting to classify them as proton or iron.
    number of “epochs” each network can be trained through
    without overtraining. (The intermediate primaries are
    currently used only for checking the mass reconstruction
    of the network. It is hoped that they will soon be
    plentiful enough to use as inputs for training as well.)
    As NN1 is given a full spectrum of energies on which
    to train, it very successfully reconstructs the energy of
    each shower. Plots of energy resolution separated into
    bins of true energy can be found in Fig. 2. The energy
    resolution is not only very good but also composition
    independent. This can be seen more clearly in Table I,
    which shows the Gaussian mean and sigma of each
    energy bin from Fig. 2. (Currently only six bins in
    energy are shown: higher energies are being generated
    and will be available before the conference.)
    The mass network outputs a reconstructed mass for
    each particle in terms of the primaries on which it has
    been trained. As the neural network is trained only on
    proton and iron showers, it reconstructs each shower
    as some combination of proton and iron. Currently a
    minimization technique is used to find the mean log
    mass for each energy bin. This minimization technique
    has been tested on intermediate primaries and proves to
    reconstruct them reasonably well for a network trained
    only on protons and iron. However, as the number
    of intermediate primaries generated is increasing, it is
    hoped that the mass network can soon be trained on a
    wider spectrum of primaries. A test of this has been run
    and a comparison between the output from a network
    trained on all particle types and a network trained only
    on protons and iron is shown in Fig. 3. As expected, it is
    evident that when the network is trained on intermediate
    primaries oxygen is reconstructed in its own location
    between protons and iron and no longer as a fraction of
    one or the other. This method appears very promising
    and, as more intermediate primaries are simulated, this
    is the direction in which the analysis will proceed.
    IV. DISCUSSION AND OUTLOOK
    The SPASE-2/AMANDA-II cosmic ray composition
    analysis has lately acquired new Monte Carlo simulation,
    a new detector simulation of the surface array and a
    revised surface-array reconstruction algorithm. Aided by
    these three new features a great deal of progress has
    been made. It can clearly be seen that the use of a
    modified version of the neural network technique seen in
    the ICRC proceedings of 2007 [9] can very accurately
    reconstruct the energy of cosmic ray primaries in the
    region of the knee in the cosmic ray spectrum. The inclu-
    sion of a larger variety of primary particles for training
    the neural network is seen to be very promising and,
    with increased statistics in the Monte Carlo simulation,
    a composition result will soon follow.
    V. ACKNOWLEDGMENTS
    The authors would like to acknowledge support from
    the Office of Polar Programs of the United States Na-
    tional Science Foundation as well as the Arctic Region
    Supercomputing Center.
    REFERENCES
    [1] J. Ahrens et al.
    Nuclear Instruments and Methods
    , 552:347-359,
    2004.
    [2] D. Heck and T. Pierog,
    Extensive Air Shower Simulations
    with CORSIKA: A User’s Guide
    http://www-ik.fzk.de/corsika/
    usersguide/corsika
    tech.html
    [3] GEANT4 Collaboration
    Nuclear Instruments and Methods, A
    ,
    506:250-303, 2003.
    [4] J.E. Dickinson et al.
    Nuclear Instruments and Methods A
    ,
    440:95-113, 2000.
    [5] K. Rawlins. PhD thesis, University of Wisconsin-Madison,
    (2001)
    [6] T. Gaisser.
    Cosmic Rays and Particle Physics
    . Cambridge Uni-
    versity Press, 1988.
    [7] J.R. Ho¨randel.
    International Journal of Modern Physics A
    ,
    20(29):6753-6764, 2005.
    [8] J. Ahrens et al.
    Astroparticle Physics
    , 21: 565-581, 2004.
    [9] K. Andeen et al,
    Proceedings of the 30th International Cosmic
    Ray Conference
    2007.

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    1
    Study of High
    p
    T
    Muons in IceCube
    Lisa Gerhardt
    ∗†
    and Spencer Klein
    ∗†
    for the IceCube Collaboration
    §
    Lawrence Berkeley National Laboratory, Berkeley, California 94720
    University of California, Berkeley, Berkeley, California 94720
    §
    See the special section of these proceedings
    Abstract
    . Muons with a high transverse momentum
    (
    p
    T
    ) are produced in cosmic ray air showers via
    semileptonic decay of heavy quarks and the decay
    of high
    p
    T
    kaons and pions. These high
    p
    T
    muons
    have a large lateral separation from the shower core
    muon bundle. IceCube is well suited for the detection
    of high
    p
    T
    muons. The surface shower array can
    determine the energy, core location and direction
    of the cosmic ray air shower while the in-ice array
    can reconstruct the energy and direction of the high
    p
    T
    muon. This makes it possible to measure the
    decoherence function (lateral separation spectrum) at
    distances greater than 150 meters. The muon
    p
    T
    can
    be determined from the muon energy (measured by
    dE/dx) and the lateral separation. The high
    p
    T
    muon
    spectrum may also be calculated in a perturbative
    QCD framework; this spectrum is sensitive to the
    cosmic-ray composition.
    Keywords
    : high transverse momentum muons, cos-
    mic ray composition, IceCube
    I. INTRODUCTION
    The composition of cosmic rays with energies above
    10
    6
    GeV is not well known. At these energies, the
    flux of cosmic rays is so low that direct detection
    with balloon or satellite-borne experiments is no longer
    feasible and indirect measurement with larger ground
    arrays must be used. These arrays measure the electronic
    and hadronic components of a cosmic ray air shower,
    and must rely on phenomemological interaction models
    to relate observables like muon and electron density
    to primary composition. These interaction models are
    based on measurements made at accelerators that reach
    a maximum energy roughly equivalent to a 10
    6
    GeV
    proton cosmic ray [1]. Extrapolation to the energy of
    the detected cosmic rays leads to uncertainties in the
    composition of cosmic rays at high energies. An alter-
    nate method of determining the composition is to use
    muons with a high transverse momentum (
    p
    T
    ) [2].
    At transverse momenta on the order of a few GeV/c,
    the muon
    p
    T
    spectrum can be calculated using per-
    turbative QCD (pQCD). Such calculations have been
    made for RHIC and the Tevatron, and the data is in
    quite good agreement with modern fixed order plus
    next-to-leading log (FONLL) calculations [3]. These
    experimental studies give us some confidence in pQCD
    calculations for air showers.
    Most of the high-energy muons that are visible in
    IceCube are produced from collisions where a high-
    energy (large Bjorken
    x
    ) parton interacts with a low
    Bjorken
    x
    parton in a nitrogen target. These collisions
    will produce heavy (charmed and bottom) quarks and
    also jets from high
    p
    T
    partons; the jets will fragment
    into pions and kaons. Pions and kaons produce “con-
    ventional” muons that have a soft spectrum roughly
    proportional to E
     3.7
    [4] and typicallly have a low
    p
    T
    . In
    contrast, charm and bottom quarks are produced early in
    the shower. The resulting muons (“prompt” muons) have
    a harder spectrum, a higher
    p
    T
    , and are the dominant
    source of atmospheric muons above 10
    5
    GeV [5].
    If these particles are produced in the forward direction
    (where they will be seen by ground based detectors),
    they can carry a significant fraction of the energy of
    the incident nucleon. The muon energy and
    p
    T
    can be
    related to the energy of the partons that make up the
    incident cosmic ray. For example, a
    10
    16
    eV proton
    has a maximum parton energy of
    10
    16
    eV, while the
    maximum parton energy for a
    10
    16
    eV iron nucleus
    (with A=56) is
    1.8 × 10
    14
    eV. These two cases have
    very different kinematic limits for high
    p
    T
    , high-energy
    muon production, so measurements of high
    p
    T
    muon
    spectra are sensitive to the cosmic-ray composition.
    High
    p
    T
    muons constitute a small fraction of the
    muons visible in IceCube; the typical muon
    p
    T
    is of
    order a few hundred MeV/c. In this domain of “soft
    physics,” the coupling constants are very large, and
    pQCD calculations are no longer reliable. One is forced
    to rely on phenomenological models; different hadronic
    interaction models give rather different results on com-
    position [6].
    IceCube, a kilometer-scale neutrino telescope, is well
    suited for the detection of muons with
    p
    T
    s above a few
    GeV/c [2]. When completed in 2011, it will consist of a
    1 km
    3
    array of optical sensors (digital optical modules,
    or DOMs) buried deep in the ice of the South Pole and a
    1 km
    2
    surface air shower array called IceTop. IceTop has
    an energy threshold of 300 TeV and can reconstruct the
    direction of showers with energies above 1 PeV within
    1.5
    and locate the shower core with an accuracy of
    9 m [7]. The in-ice DOMs (here referred to as IceCube)
    are buried in the ice 1450 m under IceTop in kilometer-
    long strings of 60 DOMs with an intra-DOM spacing of
    17 m. IceCube can reconstruct high energy muon tracks
    with sub-degree accuracy. The IceTop measurements
    can be used to extrapolate the interaction height of the

    2
    GERHARDT
    et al.
    HIGH
    P
    T
    MUONS IN ICECUBE
    shower and IceCube can measure the energy, position
    and direction of the muons. These values, can be used
    to calculate the
    p
    T
    :
    p
    T
    =
    dE
    µ
    hc
    (1)
    where
    E
    µ
    is the energy of the high
    p
    T
    muon,
    d
    is its
    lateral separation and
    h
    is the interaction height of the
    shower, here taken as an average value of 25 km. The
    interaction height loosely depends on the composition
    and a full treatment of this is planned in the future. Tak-
    ing 150 m, 25 meters more than the separation between
    strings of DOMs in IceCube, as a rough threshold for
    the two-track resolution distance of the high
    p
    T
    muon
    from the shower core gives a minimum resolvable
    p
    T
    of
    6 GeV/c for a 1 TeV muon.
    Multiple scattering and magnetic fields can deflect the
    muons as they travel to the IceCube detector, but this
    is only equivalent to a few hundred MeV worth of
    p
    T
    and is not a strong effect in this analysis. The high
    p
    T
    muon events are near-vertical which gives them a short
    slant depth. In order for muons to reach the IceCube
    detector they must have an energy of at about 500 GeV
    at the surface of the Earth. These higher energy muons
    are deflected less by multiple scattering and the Earth’s
    magnetic field.
    The combined acceptance for cosmic ray air show-
    ers which pass through both IceTop and IceCube is
    0.3 km
    2
    sr for the full 80-string IceCube array [8]. By
    the end of the austral summer of 2006/2007, 22 IceCube
    strings and 26 IceTop tanks had been deployed. The
    combined acceptance for showers that pass through both
    IceTop-26 and IceCube-22 is 0.09 km
    2
    sr. While this ac-
    ceptance is too small to expect enough events to generate
    a
    p
    T
    spectrum, it offers an excellent opportunity to test
    reconstruction and background rejection techniques.
    II. PREVIOUS MEASUREMENT OF HIGH
    p
    T
    MUONS
    MACRO, an underground muon detector, has previ-
    ously measured the separation between muons in air
    showers with energies ranging roughly from 10
    4
    GeV to
    10
    6
    GeV [9]. MACRO measured muon separations out
    to a distance of about 65 meters. The average separation
    between muons was on the order of 10 m, with 90% of
    the muons found with a separation of less than 20 m [9].
    MACRO simulated air showers and studied the muon
    pair separations as a function of the
    p
    T
    of their parent
    mesons. They verified the linear relationship between
    p
    T
    and separation shown in Eq. 1 (with a small offset due
    to multiple scattering of the muons) out to momenta up
    to 1.2 GeV/c.
    III. HIGH
    p
    T
    RATE ESTIMATIONS
    The decay of charm will produce roughly 10
    6
    prompt
    muons/year with energies in excess of 1 TeV inside the
    0.3 km
    2
    sr combined acceptance of the full IceCube array
    [10]. Different calculations agree well in overall rate in
    IceCube, but, at high (1 PeV) energies, differences in
    TABLE I: Estimated number of events from charm above
    different
    p
    T
    thresholds in 1 year with the 80-string
    IceCube array
    p
    T
    [GeV/c]
    Separation [m]
    Number
    6
    150
    500
    8
    200
    200
    16
    400
    5
    parton distribution functions lead to rates that differ by
    a factor of 3 [11].
    We can roughly estimate the fraction of muons that
    are produced at high
    p
    T
    using PYTHIA
    pp
    simulations
    conducted for the ALICE muon spectrometer at a center
    of mass energy of 14 TeV, which indicate that roughly
    0.6% of these events will have a
    p
    T
    of at least 6 GeV/c
    [12]. These simulations are done at a higher center of
    mass energy than the bulk of the IceCube events, and
    also were for muons produced at mid-rapidity (and,
    therefore, higher projectile
    x
    values and lower target
    x
    values than in IceCube). However, they should be ade-
    quate for rough estimates. These estimated rates of high
    p
    T
    muons are further reduced by approximately a factor
    of 10 by requiring that the high
    p
    T
    muon be produced in
    coincidence with a shower that triggers IceTop, leaving
    a rough expectation of
    ∼500
    events per year with a
    p
    T
    greater than 6 GeV/c in the 80-string configuration
    of IceCube. The estimated number of events above a
    given
    p
    T
    from the decay of charm is given in Table I
    neglecting the uncertainties mentioned above. At higher
    p
    T
    , bottom production may also contribute significantly
    to muon production.
    The rate of 1 TeV muons from conventional flux is
    expected to be about 10
    9
    events/year in the full IceCube
    configuration. The vast majority of these muons will
    have a low
    p
    T
    . Based on the
    p
    T
    spectrum measured
    from pions produced in 200 GeV center of mass energy
    pp
    collisions by the PHENIX collaboration, the number
    of events expected with
    p
    T
    >
    6 GeV/c is estimated to
    be 1 in 6
    ×
    10
    6
    [13]. Requiring a
    p
    T
    >
    6 GeV/c and
    an IceTop trigger leaves roughly 20 events/year in the
    full IceCube configuration from the conventional flux of
    muons.
    IV. RECONSTRUCTION METHODS
    High
    p
    T
    muons will appear as a separate track coinci-
    dent in time and parallel with the track from the central
    core of low
    p
    T
    muons. Generally the bundle of low
    p
    T
    muons leaves more light in the detector than the high
    p
    T
    muon.
    Current reconstruction algorithms in IceCube are de-
    signed to reconstruct single tracks. In order to recon-
    struct these double-track events the activated DOMs
    are split into groups using a k-means clustering algo-
    rithm [14]. The first group is the muon bundle and
    the second group is the high
    p
    T
    muon. Each group is
    then reconstructed with a maximum likelihood method
    that takes into account the scattering of light in ice.
    After this initial reconstruction, the groups are resplit

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    3
    Bundle True
    Θ
    - Reco
    Θ
    -30
    -20
    -10
    0
    10
    20
    30
    0
    10
    20
    30
    40
    50
    True Split
    σ
    =2.2
    Muon Bundle
    Reco Split
    σ
    =2.7
    Muon True
    Θ
    - Reco
    Θ
    -30
    -20
    -10
    0
    10
    20
    30
    0
    5
    10
    15
    20
    25
    30
    35
    40
    T
    Muon
    True Split
    σ
    =3.0
    High p
    Reco Split
    σ
    =4.1
    Fig. 1: Zenith angle resolution of the high
    p
    T
    reconstruc-
    tion algorithm. The sigma are the results of Gaussian fits
    to the distributions.
    according to their time residual relative to the muon
    bundle reconstruction and re-fit. The first splitting forces
    the activated DOMs into two groups, which is not correct
    for the background events which don’t have a high
    p
    T
    muon and generate only a single shower in the array.
    Splitting the activated DOMs a second time according
    to their time residual allows for the possibility for all
    the activated DOMs to end up in a single group. Figure
    1 shows the performance of this clustering algorithm.
    The zenith angle resolution for groups determined using
    the true simulation information (black, solid lines) is
    compared to the resolution for groups determined using
    the clustering algorithm (red, dashed lines) for the muon
    bundle (top) and high
    p
    T
    muon (bottom). Roughly 20%
    of the events fail to reconstruct because there are not
    enough DOMs in one of the groups.
    These reconstruction algorithms achieve a zenith an-
    gle resolution of 2.7
    for the muon bundle and 4.1
    for
    a high
    p
    T
    muon separated by 400 m. The resolution is
    worse for the high
    p
    T
    muon because fewer DOMs are
    activated. While high
    p
    T
    muons with a greater separation
    are much easier to resolve with the two track algorithm,
    they also tend to be lower energy (see Eq. 1) and activate
    fewer DOMs. The average number of DOMs activated
    by the high
    p
    T
    muon is 12, compared to 50 for the muon
    bundle. Additionally, because high
    p
    T
    muons have less
    activated DOMs, a DOM activated by the muon bundle
    that is incorrectly placed in the high
    p
    T
    group has a
    much larger effect on the reconstruction of the high
    p
    T
    track. These factors lead to a poorer resolution in the
    high
    p
    T
    track direction.
    V. SIGNAL AND BACKGROUND SEPARATION
    Many of the processes that can generate a high
    p
    T
    muon are not properly modeled or included in COR-
    SIKA [15] with the exception of the modified DPMJET
    model discussed in [16]. Also, since high
    p
    T
    muons
    will occur in only a fraction of simulated showers
    making simulation very time intensive, a toy model
    based on CORSIKA proton showers was used to model
    the signal. A single muon is inserted into an existing
    CORSIKA event containing a muon bundle from an air
    shower. This modified shower is then run through the
    standard IceCube propagation, detector simulation, and
    reconstruction routines. Simulations insert a muon with
    energy of 1 TeV separated 100, 150, 200, and 400 m
    from the shower core. This gave a clean sample ideal for
    the development of reconstruction routines optimized to
    identify simultaneous parallel tracks inside the detector.
    Cosmic rays air showers that do not generate a high
    p
    T
    muon are considered a background to this search.
    Since they generate only a single shower in the array,
    these events are mostly eliminated by requiring there be
    two well-reconstructed tracks in the IceCube detector.
    These single showers are well-reconstructed by a single
    track hypothesis, while the high
    p
    T
    muon events are not.
    Figure 2 shows the negative log of the reduced likelihood
    of a single track reconstruction for single showers, and
    showers with an inserted 4, 8, and 16 GeV/c
    p
    T
    , 1 TeV
    muon (separation of 100 m, 200 m, and 400 m from
    the shower core, assuming an average interaction height
    of 25 km). Well reconstructed events have a lower value
    on this plot. For large separations, this variable separates
    single showers from showers which contain a high
    p
    T
    muon. When the separation between the high
    p
    T
    muon
    and the shower core drops below the interstring distance
    (the blue, dot-dashed line in Figure 2), it is no longer
    possible to cleanly resolve the high
    p
    T
    muon from the
    shower core and the event looks very similar to a single
    shower.
    -Log (Single Track Reduced Likelihood)
    6
    6.5
    7
    7.5
    8
    8.5
    9
    9.5
    10
    Arbitrary Units
    0
    0.05
    0.1
    0.15
    0.2
    0.25
    4 GeV/c Muon
    8 GeV/c Muon
    16 GeV/c Muon
    Single CR Shower
    Fig. 2: Negative log of the reduced likelihood for the
    single track reconstruction
    The IceCube 22-string configuration is large enough
    that the rate of simultaneous events from cosmic rays
    is significant. Muon bundles from two (or more) air
    showers can strike the array within the 10
    µs
    event
    window, producing two separated tracks. These so-called
    double-coincident events are the dominant background
    for air showers with high
    p
    T
    muons. Since these double-
    coincident events are uncorrelated in direction and time,
    requiring that both reconstructed tracks be parallel and
    occur within 1
    µs
    can eliminate most of these events.

    4
    GERHARDT
    et al.
    HIGH
    P
    T
    MUONS IN ICECUBE
    Due to misreconstructions some events will survive
    this selection, and criteria to eliminate these misre-
    constructed events are being developed. However, an
    irreducible background remains from double-coincident
    events that happen to come from roughly the same
    direction and time.
    Fig. 3: Candidate shower with a high
    p
    T
    muon. The
    colored circles show DOMs that are activated by light.
    The color and size of the circle corresponds to the
    time (red being earliest) and magnitude of the signal,
    respectively and the white dots indicate DOMs that
    are not activated in the event. The red lines show the
    reconstructed tracks.
    After applying selection criteria to reduce the cosmic
    ray background, a 10% sample of the IceCube data was
    scanned by eye for high
    p
    T
    muon candidates. Several
    events were found and Figure 3 shows a representative
    event. The two tracks occur within 600 ns of each other
    and the spaceangle between the two reconstructions is
    5.6
    .
    VI. HORIZONTAL EVENTS
    One possible method to avoid the backgrounds from
    single and double-coincident cosmic ray air showers is to
    search for events that come from the horizontal direction.
    At a zenith angle of 60
    the slant depth through the
    ice is about 3 km. Only muons with energies in excess
    of several TeV will reach IceCube [17] and the event
    multiplicity is greatly reduced. Because the intra-DOM
    spacing is 17 m, compared to the intra-string spacing
    of 125 m relevant for vertical events, the two-track
    resolution should improve. A disadvantage to searching
    in the horizontal direction is the increase in the deflection
    from the magnetic field and multiple scattering with the
    slant depth, which could lead to a greater spread in the
    muon bundle and create event topologies which mimic
    a high
    p
    T
    muon event. Nevertheless, the advantage of
    background suppression is strong. In addition to high
    p
    T
    muons in cosmic rays, there are a number of other
    processes (such as the decay of the supersymmetric
    stau [18]) that can produced horizontal parallel tracks,
    making the horizon an interesting direction.
    A search was conducted in 10% of the IceCube data
    from 2007 for these horizontal events using simple topo-
    logical selection criteria. Events which reconstructed
    Fig. 4: Candidate double track near-horizontal event
    within 30
    of horizontal were selected and searched for
    a double track topology. Several interesting candidate
    events were found, one is shown in Figure 4. The two
    tracks occur within 1 ns of each other and the spaceangle
    between the two reconstructions is 33
    .
    VII. CONCLUSIONS
    IceCube is large enough that the study of high
    p
    T
    muons has become viable. Resolution of tracks separated
    by at least 150 meters from the shower core will allow
    identification of muons with
    p
    T
    s of at least 6 GeV/c. A
    two-track reconstruction algorithm has been developed
    and selection criteria for identification of air showers
    with high
    p
    T
    muons is under development. A search for
    horizontal double tracks is also under way. The rate of
    high
    p
    T
    muon production is sensitive to the composition
    of the cosmic rays and offers an alternative to existing
    composition studies.
    This work is made possible by support from the NSF
    and the DOE.
    REFERENCES
    [1] R. Engel, Nuclear Physics B Proc. Suppl.,
    151
    :437 (2006).
    [2] S. Klein and D. Chirkin, 30th ICRC (2007), arXiv:0711.0353v1.
    [3] R. Vogt, M. Cacciari and P. Nason, Nucl. Phys.
    A774
    , 661
    (2006).
    [4] T.K. Gaisser, “Cosmic Ray and Particle Physics,” Cambridge
    University Press, 1990.
    [5] L. Pasquali, M. Reno, and I. Sarcevic, Phys. Rev. D
    59
    :034020
    (1998).
    [6] W. Apel
    et al.
    , Astropart. Phys.
    29
    :412 (2008); J. Bluemer et
    al,arXiv:0904.0725.
    [7] S. Klepser, 21st European Cosmic Ray Symposium (2008),
    arXiv:0811.1671.
    [8] X. Bai, 30th ICRC (2007), arXiv:0711.0353v1.
    [9] Ambrosio
    et al.
    , Phys. Rev. D
    60
    :032001 (1999).
    [10] M. Thunman, P. Gondolo, and G. Ingelman, Astropar. Phys.
    5
    :309 (1996).
    [11] R. Enberg, M. Reno and I. Sarcevic, Phys. Rev. D
    78
    :043005
    (2008).
    [12] Z. Conesa del Valle, The European Physical Journal C,
    49
    :149
    (2007).
    [13] S. Adler
    et al.
    , Phys. Rev. Lett.
    91
    :241803 (2003).
    [14] D. MacKay “Information Theory, Inference and Learning Algo-
    rithms,” Cambridge University Press (2003).
    [15] D. Heck, Forschungszentrum Karlsruhe Report FZKA 6019
    (1998).
    [16] P. Berghaus, T. Montaruli, and J. Ranft, Journal of Cosmology
    and Astropar. Phys.
    6
    :003 (2008).
    [17] P. Berghaus, these proceedings.
    [18] A. Olivas, these proceedings.

    1
    Large Scale Cosmic Rays Anisotropy With IceCube
    Rasha U Abbasi
    , Paolo Desiati
    and Juan Carlos Velez
    for the IceCube Collaboration
    University of Wisconsin, IceCube Neutrino Observatory, Madison, WI 53703, USA
    http://icecube.wisc.edu/
    Abstract
    . We report on a study of the anisotropy
    in the arrival direction of cosmic rays with a median
    energy per Cosmic Ray (CR) particle of
    ∼ 14
    TeV
    using data from the IceCube detector. IceCube is
    a neutrino observatory at the geographical South
    Pole, when completed it will comprise 80 strings
    plus 6 additional strings for the low energy array
    Deep Core. The strings are deployed in the deep ice
    between 1,450 and 2,450 meters depth, each string
    containing 60 optical sensors. The data used in this
    analysis are the data collected from April 2007 to
    March 2008 with 22 deployed strings. The data
    contain
    ∼ 4.3
    billion downward going muon events.
    A two-dimensional skymap is presented with an
    evidence of
    0.06%
    large scale anisotropy. The energy
    dependence of this anisotropy at median energies per
    CR particle of 12 TeV and 126 TeV is also presented
    in this work. This anisotropy could arise from a
    number of possible effects; it could further enhance
    the understanding of the structure of the galactic
    magnetic field and possible cosmic ray sources.
    Keywords
    : IceCube, Cosmic rays, Anisotropy.
    I. I
    NTRODUCTION
    The intensity of Galactic Cosmic Rays (GCRs) have
    been observed to show sidereal anisotropic variation on
    the order of 10
    −4
    at energies in the range of 1-100
    TeV ([1], [2] and [3]). This anisotropy could arise from
    number of different combination of causes. One possible
    cause could be the Compton Getting (CG) effect. This
    effect was proposed in 1935 [4] predicting that CR
    anisotropy could arise from the movement of the solar
    system around the galactic center with the velocity of
    ∼ 220 kms
    −1
    such that an excess of CR would be
    present in the direction of motion of the solar system
    while a deficit would appear in the opposite direction.
    Another possible effect (proposed by Nagashima
    et.
    al.
    [1]) causing the excess in the anisotropy, which was
    referred to as ”tail-in”, originates from close to the tail
    of the hemisphere. While the deficit in the anisotropy,
    which was referred to as ”loss-cone”, originates from a
    magnetic cone shaped structure of the galactic field in
    the vicinity region.
    In this paper we present results on the observation of
    large scale cosmic ray anisotropy by IceCube. Previous
    experiments have published a 2-dimensional skymap of
    the northern hemisphere sky ([2]- [5]). This measure-
    ment presents the first 2-dimensional skymap for the
    southern hemisphere sky. In addition, we present the
    energy dependence of this anisotropy at median energies
    per CR particle of 12 TeV and 126 TeV.
    The outline of the paper is: the second section will
    describe the data used in this analysis, the analysis
    method and the challenges. The third section will discuss
    the results and the stability checks applied to the data.
    The fourth section will discuss the anisotropy energy
    dependence and the last section is the conclusion.
    II. D
    ATA ANALYSIS
    The data used in this analysis are the downward going
    muons collected by the IceCube neutrino observatory
    comprising 22 strings. The data were collected from
    June 2007 to March 2008. The events used in this
    analysis are those reconstructed by an online one iter-
    ation Likelihood (LLH) based reconstruction algorithm.
    The events selected online require at least ten triggered
    optical sensors on at least three strings. The average rate
    of these events is ∼ 240 Hz (approximately 40 % of the
    events at triggering level). Further selection criteria are
    applied to the data to ensure good quality and stable
    runs. The final data set consists of 4.3 × 10
    9
    events
    with a median angular resolution (angle between the
    reconstructed muon and the primary particle) of 3
    and
    a median energy per CR particle of 14 TeV as simulated
    according to (CORSIKA [6], SIBYLL [7], Ho¨randel [8]
    ).
    In this analysis we are searching for a high precision
    anisotropy. The sidereal variation of the CR intensity
    is induced by the anisotropy in their arrival direction.
    However, it can also be caused by the detector exposure
    asymmetries, non-uniform time coverage, diurnal and
    seasonal variation of the atmospheric temperature. Apart
    from these effects the remaining variations can only be
    of galactic origin.
    Due to the unique location of IceCube at the South
    Pole the detector observes the sky uniformly. This is not
    the case for all the previous experiments searching for
    large scale cosmic ray anisotropy (e.g. [2], [3], [5]).
    Due to their locations they need a whole solar day to
    scan the entire sky. As a result they need to eliminate the
    diurnal and seasonal variations using various approaches.
    For IceCube the diurnal variation does not effect the
    sidereal variation because the whole sky is fully visible
    to the detector at any given time and because there
    is only one day and one night per year. In addition,
    although the seasonal variation is on the order of 20%
    the variation is slow and does not affect the daily muon
    intensity significantly.

    2
    The remaining challenge for this analysis is account-
    ing for the detector asymmetry, and unequal time cover-
    age in the data due to the detector run selection. To illus-
    trate the detector asymmetry Figure 1 shows the IceCube
    22 string geographical configuration. This geographical
    asymmetry results in a preferred reconstructed muon
    direction since the muons would pass by more strings
    in one direction in the detector compared to another.
    The combination of detector event asymmetry with a
    non-uniform time coverage would induce an azimuthal
    asymmetry and consequently artificial anisotropy of the
    arrival direction of cosmic rays. This asymmetry is
    corrected for by normalizing the azimuthal distribution.
    Figure 2 shows the azimuthal distribution for the
    whole data set. It displays the number of events vs.
    the azimuth of the arrival direction of the primary
    CR particle. Note that the asymmetry in the azimuthal
    distribution due to detector geometry is modeled well by
    simulation.
    To correct for the detector azimuthal asymmetry we
    apply an azimuthal normalization. The azimuthal distri-
    bution is parameterized by N, n
    i
    , and nˉ, where N is the
    total number of bins, n
    i
    is the number of events per bin
    and nˉ is the average number of events, nˉ =
    1
    N
    ?
    N
    i=1
    n
    i
    .
    nˉ is denoted by the horizontal red line in Figure 2. The
    azimuthal normalization is applied by weighting each
    eventby
    n
    i
    for that event.
    In addition to the azimuthal asymmetry we also
    observe a zenith angle asymmetry (more events arrive
    from the zenith than from the horizon). Due to this
    declination dependence, the sky is divided into four
    declination bands such that the data is approximately
    equally distributed among the declination bands. For
    each band the azimuthal distribution is normalized for
    the whole year. The relative intensity for each bin in the
    2-dimensional skymap is then calculated by dividing the
    number of events per bin by the total number of events
    per declination.
    III. R
    ESULTS:
    Figure 3 shows the southern hemisphere skymap
    for well reconstructed downward going muons for the
    IceCube 22-Strings data set. The skymap is plotted in
    equatorial coordinates. The color scale represents the
    relative intensity of the rate for each bin per declination
    band where each bin rate is calculated by dividing
    the number of events for that bin over the average
    number of events for that bin’s declination band. The
    plot shows a large scale anisotropy in the arrival di-
    rection of cosmic rays. The amplitude and the phase
    of this anisotropy is determined by projecting the 2-
    dimensional skymap in Right Ascension (RA) as shown
    in Figure 4. Figure 4 shows the relative intensity vs.
    the RA. The data are shown in points with their error
    bars. The fit is the second-order harmonic function in
    the form of
    ?
    n=2
    i=1
    (A
    i
    × cos(i ((RA) − φ
    i
    ))) + B where
    A
    i
    is the amplitude and φ
    i
    is the phase and B is a
    constant. The fit A
    i
    , φ
    i
    and χ
    2
    /ndof for the second
    Fig. 1: The IceCube detector configuration. The filled
    green circles are the positions of IceCube strings and
    the filled blue circles display the position of the IceTop
    tanks.
    Azimuth
    0
    50
    100
    150
    200
    250
    300
    350
    Number of Events
    70
    80
    90
    100
    110
    120
    130
    ×
    10
    6
    Fig. 2: The azimuthal distribution for the whole data set.
    This plot shows the number of events vs. the azimuth
    of the arrival direction of the primary CR particle. The
    horizontal red line in the average number of events for
    the distribution.
    harmonic fit are listed in table I. The significance of
    the 2-dimensional skymap is shown in Figure 5. The
    significance is calculated for each bin from the average
    number of events for that bin’s declination band. Note
    that the significance of several bins in the excess region
    is greater than 4σ and in some bins in the deficit region
    is smaller the −4σ.
    A
    1
    .
    (10
    −4
    )
    φ
    1
    A
    2
    .
    (10
    −4
    )
    φ
    2
    χ
    2
    /ndof
    6.4±0.2 66.4
    ±2.6
    2.1±0.3 −65.6
    ±4
    22/19
    TABLE I: The second harmonic fit amplitude, phase,
    and χ
    2
    /ndof.
    To check for the stability of the measured large scale

    3
    Fig. 3: The IceCube skymap in equatorial coordinates
    (Declination (Dec) vs. Right Ascension (RA)). The color
    scale is the relative intensity.
    Right Ascension
    0
    50
    100
    150
    200
    250
    300
    350
    Relative Intensity
    0.9985
    0.999
    0.9995
    1
    1.0005
    1.001
    1.0015
    Fig. 4: The 1-dimensional projection of the IceCube 2-
    dimensional skymap. The line is the second harmonic
    function fit.
    Fig. 5: The IceCube significance skymap in equatorial
    coordinates (Dec vs. RA). The color scale is the signif-
    icance.
    anisotropy we performed a number of checks with the
    data set. One check was applied by dividing the data into
    two sets where one set contains sub-runs with an even
    index number and the other set contains sub-runs with
    an odd index number (a sub-run on average contains
    events collected for ∼ 20 minutes at a time). Another
    check is applied by dividing the data into two sets each
    set contains half the sub-runs selected randomly. The
    results for both tests were consistent.
    In addition, stability checks are applied to test for
    daily and seasonal variation effects. To test for the daily
    variation effect the data were divided in two sets: The
    first set contains sub-runs with rate values greater than
    the mean rate value for that sub-run’s day. The second
    set contains sub-runs with rate values less than the
    mean rate value for that sub-run’s day. Furthermore,
    to test for the seasonal variation effect the data set is
    divided in two sets: The first set holds the winter month’s
    sub-runs (June-Oct.). The second set holds the summer
    month’s sub-runs (Nov.-March). For both tests we see
    no significant changes in the value of the anisotropy of
    the two data sets.
    IV. E
    NERGY DEPENDENCE:
    In order to better understand the possible nature of
    the anisotropy we have searched for energy dependent
    effects using our data. To determine the energy de-
    pendence for the signal we divided the data into two
    energy bins. To accomplish that and to ensure constant
    energy distribution along our sky both the number of
    sensors triggered by the event and the zenith angle of
    the event are used for the energy bands selection. The
    first energy bin contains 3.8 × 10
    9
    events with a median
    per CR particle of 12.6 TeV and 90% of the events
    between 2 and 158 TeV. The second energy bin contains
    9.6 × 10
    8
    events with a median energy per CR particle
    of 126 TeV and 90% of the events between 10 TeV
    and 1 PeV. Each 2-dimensional skymap is projected
    to 1-dimensional variations in RA. In comparison to
    previous experiments the 1-dimensional RA distribution
    is fitted to a first harmonic fit. The first harmonic
    fit amplitude and phase for the first energy band are
    A
    1
    = (7.3 ± 0.3) × 10
    −4
    and φ
    1
    = 63.4
    ± 2.6
    , while
    the amplitude and phase for the second energy band are
    A
    1
    = (2.9±0.6)×10
    −4
    and φ
    1
    = 93.2
    ±12
    . Figure 6
    shows the amplitude of this analysis (in filled circles) in
    comparison to previous experiments (In empty squares).
    Note that the amplitude in this analysis shows a decrease
    of the harmonic amplitude value at the higher energies
    for the energy ranges of 10-100 TeV.
    10
    12
    10
    13
    10
    14
    10
    -5
    10
    -4
    10
    -3
    Fig. 6: The filled circles markers are the
    result of this analysis and the empty square
    markers are the result from previous experiments
    ([3], [9], [10], [11], [12], [13], [14], [15], [16], [17], [18],
    [19], [20], [21], [22], [23], [24], [25]).

    4
    V. CONCLUSION
    In this analysis we present the first 2-dimensional
    skymap for the southern hemisphere of 4.3 billion cos-
    mic rays with a median angular resolution of 3
    and a
    median energy per CR particle of 14 TeV as observed by
    IceCube. A large cosmic ray anisotropy with a second
    harmonic vector amplitude of A
    1
    = (6.4 ± 0.2) × 10
    −4
    and a phase of φ
    1
    = 66.4
    ± 2.6
    is observed. The
    significance of some bins in the excess and the deficit
    regions were found to be > |4σ|. This anisotropy is an
    extension of previously measured large scale anisotropy
    at the northern hemisphere reported by multiple experi-
    ments ([2]- [5]).
    In addition, we report on the anisotropy energy de-
    pendence. We report the amplitude of the first harmonic
    vector of the anisotropy for the two energy bands.
    The first energy band with a median energy per CR
    particle of 12.6 TeV the amplitude is found to be
    A
    1
    = (7.3±0.3)×10
    −4
    . The second energy band with a
    median energy per CR particle of 126 TeV the amplitude
    is found to be A
    1
    = (2.9 ± 0.6) × 10
    −4
    . The amplitude
    energy dependence is found to follow a decreasing trend
    with energy.
    VI. A
    CKNOWLEDGMENTS
    We acknowledge the support from the following agen-
    cies: U.S. National Science Foundation-Oce of Polar
    Program, U.S. National Science Foundation-Physics Di-
    vision, University of Wisconsin Alumni Research Foun-
    dation, U.S. Department of Energy, and National En-
    ergy Research Scientic Computing Center, the Louisiana
    Optical Network Initiative (LONI) grid computing re-
    sources; Swedish Research Council, Swedish Polar Re-
    search Secretariat, and Knut and Alice Wallenberg
    Foundation, Sweden; German Ministry for Education
    and Research (BMBF), Deutsche Forschungsgemein-
    schaft (DFG), Germany; Fund for Scientic Research
    (FNRS-FWO), Flanders Institute to encourage scientic
    and technological research in industry (IWT), Belgian
    Federal Science Policy Oce (Belspo); the Netherlands
    Organisation for Scientic Research (NWO); M. Ribordy
    acknowledges the support of the SNF (Switzerland); A.
    Kappes and A. Groß acknowledge support by the EU
    Marie Curie OIF Program.
    R
    EFERENCES
    [1] K. Nagashima
    et. al Journal of Geophysical Research
    , vol. 103,
    pp. 17429–17440, Aug. 1998.
    [2] M. Amenomori
    et. al Science
    , vol. 314, pp. 439–443, Oct. 2006.
    [3] G. Guillian and for the Super-Kamiokande Collaboration
    Physi-
    cal Review D
    , vol. 75, p. 062003, 2007.
    [4] A. H. Compton and I. A. Getting
    Physical Review
    , vol. 47,
    pp. 817–821, June 1935.
    [5] A. Abdo
    et. al. ArXiv:astro-ph/0806.2293
    , 2008.
    [6] CORSIKA
    http://www-ik.fzk.de/corsika/
    .
    [7] R. Engel vol. 1 of
    International Cosmic Ray Conference
    , p. 415,
    1999.
    [8] J. R. Horandel¨
    Astroparticle Physics
    , vol. 19, pp. 193–220, May
    2003.
    [9] K. Munakata
    et. al. Physical Review D
    , vol. 56, p. 23, 1997.
    [10] M. Ambrosio
    et. al. Physical Review D
    , vol. 67, p. 042002, 2003.
    [11] D. Swinson
    et. al. Planet. Space Sci.
    , vol. 33, p. 1069, 1985.
    [12] K. Nagashima
    et. al. Planet. Space Sci.
    , vol. 33, p. 395, 1985.
    [13] H. Ueno
    et. al.
    vol. 6 of
    International Cosmic Ray Conference
    ,
    p. 361, 1990.
    [14] T. Thambyahpillai vol. 3 of
    International Cosmic Ray Confer-
    ence
    , p. 383, Aug. 1983.
    [15] K. Munakata
    et. al.
    vol. 4 of
    International Cosmic Ray Confer-
    ence
    , p. 639, 1995.
    [16] S. Mori
    et. al.
    vol. 4 of
    International Cosmic Ray Conference
    ,
    p. 648, 1995.
    [17] M. Bercovitch
    et. al.
    vol. 10 of
    International Cosmic Ray
    Conference
    , p. 246, 1981.
    [18] B. K. Fenton
    et. al.
    vol. 4 of
    International Cosmic Ray Confer-
    ence
    , p. 635, 1995.
    [19] Y. W. Lee
    et. al.
    vol. 2 of
    International Cosmic Ray Conference
    ,
    p. 18, 1987.
    [20] D. J. Cutler
    et. al. Astrophys. J.
    , vol. 376, pp. 322–334, July
    1991.
    [21] Y. M. Andreyev
    et. al.
    vol. 2 of
    International Cosmic Ray
    Conference
    , p. 22, 1987.
    [22] K. Munakata vol. 7 of
    International Cosmic Ray Conference
    ,
    p. 293, 1999.
    [23] V. V. Alexeyenko
    et. al.
    vol. 2 of
    International Cosmic Ray
    Conference
    , p. 146, 1981.
    [24] M. Aglietta
    et. al.
    vol. 2 of
    International Cosmic Ray Conference
    ,
    p. 800, 1995.
    [25] T. Gombosi
    et. al.
    vol. 2 of
    International Cosmic Ray Confer-
    ence
    , pp. 586–591, 1975.

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁOD´ Z´ 2009
    1
    Atmospheric Variations as observed by IceCube
    Serap Tilav
    , Paolo Desiati
    , Takao Kuwabara
    , Dominick Rocco
    ,
    Florian Rothmaier
    , Matt Simmons
    , Henrike Wissing
    §¶
    for the IceCube Collaboration
    ?
    Bartol Research Institute and Dept. of Physics and Astronomy, University of Delaware, Newark, DE 19716, USA.
    Dept. of Physics, University of Wisconsin, Madison, WI 53706, USA.
    Institute of Physics, University of Mainz, Staudinger Weg 7, D-55099 Mainz, Germany.
    §
    III Physikalisches Institut, RWTH Aachen University, D-52056 Aachen, Germany.
    Dept. of Physics, University of Maryland, College Park, MD 20742, USA.
    ?
    See the special section of these proceedings
    Abstract
    . We have measured the correlation of rates
    in IceCube with long and short term variations in
    the South Pole atmosphere. The yearly temperature
    variation in the middle stratosphere (30-60 hPa) is
    highly correlated with the high energy muon rate
    observed deep in the ice, and causes a
    ±
    10% seasonal
    modulation in the event rate. The counting rates of
    the surface detectors, which are due to secondary
    particles of relatively low energy (muons, electrons
    and photons), have a negative correlation with tem-
    peratures in the lower layers of the stratosphere
    (40-80 hPa), and are modulated at a level of
    ±
    5%.
    The region of the atmosphere between pressure levels
    20-120 hPa, where the first cosmic ray interactions
    occur and the produced pions/kaons interact or decay
    to muons, is the Antarctic ozone layer. The anti-
    correlation between surface and deep ice trigger
    rates reflects the properties of pion/kaon decay and
    interaction as the density of the stratospheric ozone
    layer changes. Therefore, IceCube closely probes the
    ozone hole dynamics, and the temporal behavior of
    the stratospheric temperatures.
    Keywords
    : IceCube, IceTop, South Pole
    I. INTRODUCTION
    The IceCube Neutrino Observatory, located at the
    geographical South Pole (altitude 2835m), has been
    growing incrementally in size since 2005, surrounding
    its predecessor AMANDA[1]. As of March 2009, Ice-
    Cube consists of 59 strings in the Antarctic ice, and
    59 stations of the IceTop cosmic ray air shower array
    on the surface. Each IceCube string consists of 60
    Digital Optical Modules (DOMs) deployed at depths of
    1450-2450m, and each IceTop station comprises 2 ice
    Cherenkov tanks with 2 DOMs in each tank.
    IceCube records the rate of atmospheric muons with
    E
    µ
    400 GeV. Muon events pass the IceCube Simple
    Majority Trigger (InIce SMT) if 8 or more DOMs are
    triggered in 5
    µ
    sec. The IceTop array records air showers
    with primary cosmic ray energies above 300 TeV. Air
    showers pass the IceTop Simple Majority Trigger (Ice-
    Top SMT) if 6 or more surface DOMs trigger within
    5
    µ
    sec.
    As the Antarctic atmosphere goes through seasonal
    changes, the characteristics of the cosmic ray interac-
    tions in the atmosphere follow these variations. When
    the primary cosmic ray interacts with atmospheric nu-
    clei, the pions and kaons produced at the early interac-
    tions mainly determine the nature of the air shower. For
    the cosmic ray energies discussed here, these interactions
    happen in the ozone layer (20-120 hPa pressure layer
    at 26 down to 14 km) of the South Pole stratosphere.
    During the austral winter, when the stratosphere is cold
    and dense, the charged mesons are more likely to interact
    and produce secondary low energy particles. During the
    austral summer when the warm atmosphere expands and
    becomes less dense, the mesons more often decay rather
    than interact.
    Figure 1 demonstrates the modulation of rates in
    relation to the temporal changes of the South Pole strato-
    sphere. The scaler rate of a single IceTop DOM on the
    surface is mostly due to low energy secondary particles
    (MeV electrons and gammas,
    ∼1
    GeV muons)[2] pro-
    duced throughout the atmosphere, and therefore highly
    modulated by the atmospheric pressure. However, after
    correcting for the barometric effect, the IceTop DOM
    counting rate also reflects the initial pion interactions in
    the middle stratosphere. The high energy muon rate in
    the deep ice, on the other hand, directly traces the decay
    characteristics of high energy pions in higher layers
    of the stratosphere. The IceCube muon rate reaches its
    maximum at the end of January when the stratosphere
    is warmest and most tenuous. Around the same time
    Icetop DOMs measure the lowest rates on the surface
    as the pion interaction probability reaches its minimum.
    The high energy muon rate starts to decline as the
    stratosphere gets colder and denser. The pions will
    interact more often than decay, yielding the maximum
    rate in IceTop and the minimum muon rate in deep ice
    at the end of July.
    II. ATMOSPHERIC EFFECTS ON THE ICECUBE RATES
    The Antarctic atmosphere is closely monitored by the
    NOAA Polar Orbiting Environmental Satellites (POES)
    and by the radiosonde balloon launches of the South
    Pole Meteorology Office. However, stratospheric data

    2
    ATMOSPHERICVARIATION
    180
    200
    220
    240
    260
    T
    p
    [K]
    (a)
    p=20-30hPa
    30-40hPa
    40-50hPa
    50-60hPa
    60-70hPa
    70-80hPa
    80-100hPa
    1200
    1300
    1400
    1500
    1600
    1700
    640
    660
    680
    700
    720
    740
    IceTop DOM count. [Hz]
    Pressure [hPa]
    (b)
    Observed
    Pressure corrected
    Pressure
    900
    1000
    1100
    1200
    05/01
    2007
    07/01
    2007
    09/01
    2007
    11/01
    2007
    01/01
    2008
    03/01
    2008
    05/01
    2008
    07/01
    2008
    09/01
    2008
    11/01
    2008
    01/01
    2009
    03/01
    2009
    180
    200
    220
    240
    IceCube muon rate [Hz]
    T
    eff
    [K]
    (c)
    Observed
    T
    eff
    Fig. 1. The temporal behavior of the South Pole stratosphere from May 2007 to April 2009 is compared to IceTop DOM counting rate and
    the high energy muon rate in the deep ice. (a) The temperature profiles of the stratosphere at pressure layers from 20 hPa to 100 hPa where
    the first cosmic ray interactions happen. (b) The IceTop DOM counting rate (black -observed, blue -after barometric correction) and the surface
    pressure (orange). (c) The IceCube muon trigger rate and the calculated effective temperature (red).
    is sparse during the winter when the balloons do not
    reach high altitudes, and satellite based soundings fail
    to return reliable data. For such periods NOAA derives
    temperatures from their models. We utilize both the
    ground-based data and satellite measurements/models
    for our analysis.
    A. Barometric effect
    In first order approximation, the simple correlation
    between log of rate change
    ∆{lnR}
    and the surface
    pressure change
    ∆P
    is
    ∆{lnR} = β · ∆P
    (1)
    where
    β
    is the barometric coefficient.
    As shown by the black line in the Figure 1b, the
    observed IceTop DOM counting rate varies by
    ±10%
    in
    anti-correlation with surface pressure, and the barometric
    coefficient is determined to be
    β = −0.42
    %/hPa. Using
    this value, the pressure corrected scaler rate is plotted
    as the smoother line (blue) in Figure 1b. The cosmic
    ray shower rate detected by the IceTop array also varies
    by
    ±17%
    in anti-correlation with surface pressure, and
    can be corrected with a
    β
    value of
    −0.77
    %/hPa. As
    expected [3], the IceCube muon rate shown in Figure
    1c is not correlated with surface pressure. However,
    during exceptional stratospheric temperature changes,
    the second order temperature effect on pressure becomes
    large enough to cause anti-correlation of the high energy
    muon rate with the barometric pressure. During such
    events the effect directly reflects sudden stratospheric
    density changes, specifically in the ozone layer.
    B. Seasonal Temperature Modulation
    Figure 1 clearly demonstrates the seasonal temper-
    ature effect on the rates. The IceTop DOM counting
    rate, after barometric correction, shows
    ±5%
    negative
    temperature correlation. On the other hand, the IceCube
    muon rate is positively correlated with
    ±10%
    seasonal
    variation.
    From the phenomenological studies [4][5], it is known
    that correlation between temperature and muon intensity
    can be described by the effective temperature
    T
    eff
    ,
    defined by the weighted average of temperatures from
    the surface to the top of the atmosphere.
    T
    eff
    approxi-
    mates the atmosphere as an isothermal body, weighting
    each pressure layer according to its relevance to muon
    production in atmosphere [5][6].
    The variation of muon rate
    ∆R
    µ
    / < R
    µ
    >
    is related
    to the effective temperature as
    ∆R
    µ
    <R
    µ
    >
    = α
    T
    ∆T
    eff
    <T
    eff
    >
    ,
    (2)
    where
    α
    T
    is the atmospheric temperature coefficient.
    Using balloon and satellite data for the South Pole
    atmosphere, we calculated the effective temperature as
    the red line in Figure 1c. We see that it traces the
    IceCube muon rate remarkably well. The calculated
    temperature coefficient
    α
    T
    = 0.9
    for the IceCube muon

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁOD´ Z´ 2009
    3
    TABLE I
    Temperature and correlation coefficients of rates for different
    stratospheric layers of 20-100 hPa and
    T
    eff
    .
    IceCube Muon
    IceTop Count.
    P
    <T
    p
    >
    α
    p
    T
    γ
    α
    p
    T
    γ
    (hP a)
    (K)
    20-30
    214.0
    0.512
    0.953
    -0.194
    -0.834
    30-40
    208.7
    0.550
    0.986
    -0.216
    -0.906
    40-50
    207.3
    0.591
    0.993
    -0.240
    -0.946
    50-60
    206.6
    0.627
    0.985
    -0.261
    -0.968
    60-70
    206.3
    0.656
    0.971
    -0.278
    -0.975
    70-80
    206.3
    0.679
    0.954
    -0.292
    -0.975
    80-100
    206.5
    0.708
    0.927
    -0.310
    -0.971
    < T
    eff
    >
    α
    T
    γ
    α
    T
    γ
    211.3
    0.901
    0.990
    -0.360
    -0.969
    rate agrees well with the expectations of models as well
    as with other experimental measurements[7][8].
    In this paper, we also study in detail the relation be-
    tween rates and stratospheric temperatures for different
    pressure layers from 20 hPa to 100 hPa as
    ∆R
    <R>
    = α
    p
    T
    ∆T
    p
    <T
    p
    >
    .
    (3)
    The temperature coefficients for each pressure layer
    α
    p
    T
    and the correlation coefficient
    γ
    are determined from
    regression analysis. Pressure-corrected IceTop DOM
    counting rate, and IceCube muon rate are sorted in
    bins of
    10 days, and deviations
    ∆R/ < R >
    from
    the average values are compared with the deviation of
    temperatures at different depths
    ∆T
    p
    / < T
    p
    >
    .
    We list the values of
    α
    p
    T
    and
    γ
    for IceCube muon
    rate and IceTop DOM counting rate in Table 1. We
    find that the IceCube muon rate correlates best with
    the temperatures of 30-60 hPa pressure layers, while the
    IceTop DOM counting rate shows the best correlation
    with 60-80 hPa layers. In Figure 2. we plot the rate and
    temperature correlation for layers which yield the best
    correlation.
    III. EXCEPTIONAL STRATOSPHERIC EVENTS AND
    THE MUON RATES
    The South Pole atmosphere is unique because of
    the polar vortex. In winter a large-scale counter clock-
    wise flowing cyclone forms over the entire continent
    of Antarctica, isolating the Antarctic atmosphere from
    higher latitudes. Stable heat loss due to radiative cooling
    continues until August without much disruption, and the
    powerful Antarctic vortex persists until the sunrise in
    September. As warm air rushes in, the vortex loses its
    strength, shrinks in size, and sometimes completely dis-
    appears in austral summer. The density profile inside the
    vortex changes abruptly during the sudden stratospheric
    warming events, which eventually may cause the vortex
    collapse. The ozone depleted layer at 14-21 km altitude
    (ozone hole), observed in September/October period, is
    usually replaced with ozone rich layer at 18-30 km soon
    after the vortex breaks up.
    -15
    -10
    -5
    0
    5
    10
    15
    -20 -15 -10 -5 0 5 10 15 20
    Δ
    R / < R > [%]
    Δ
    T
    p
    / < T
    p
    > [%]
    IceCube muon rate vs. T
    p
    (p=40-50hPa)
    α
    = 0.591
    γ
    = 0.993
    (a)
    -15
    -10
    -5
    0
    5
    10
    15
    -20 -15 -10 -5 0 5 10 15 20
    Δ
    R / < R > [%]
    Δ
    T
    eff
    / < T
    eff
    > [%]
    IceCube muon rate vs. T
    eff
    α
    = 0.901
    γ
    = 0.990
    (b)
    -15
    -10
    -5
    0
    5
    10
    15
    -20 -15 -10 -5 0 5 10 15 20
    Δ
    R / < R > [%]
    Δ
    T
    p
    / < T
    p
    > [%]
    IceTop DOM count. vs. T
    p
    (p=60-70hPa)
    α
    = -0.278
    γ
    = -0.975
    (c)
    -15
    -10
    -5
    0
    5
    10
    15
    -20 -15 -10 -5 0 5 10 15 20
    Δ
    R / < R > [%]
    Δ
    T
    eff
    / < T
    eff
    > [%]
    IceTop DOM count. vs. T
    eff
    α
    = -0.360
    γ
    = -0.969
    (d)
    Fig. 2. Correlation of IceCube muon and IceTop DOM counting rates
    with stratospheric temperatures and
    T
    eff
    . (a) IceCube muon rate vs.
    temperature at 40-50 hPa pressure layer. (b) IceCube muon rate vs.
    effective temperature. (c) IceTop DOM counting rate vs. temperature at
    70-80 hPa pressure layer. (d) IceTop DOM counting rate vs. effective
    temperature.
    Apart from the slow seasonal temperature variations,
    IceCube also probes the atmospheric density changes
    due to the polar vortex dynamics and vigorous strato-
    spheric temperature changes on time scales as short as
    days or even hours, which are of great meteorological
    interest.
    An exceptional and so far unique stratospheric event
    has already been observed in muon data taken with
    IceCube’s predecessor AMANDA-II.
    A. 2002 Antarctic ozone hole split detected by AMANDA
    In late September 2002 the Antarctic stratosphere
    underwent its first recorded major Sudden Stratospheric
    Warming (SSW), during which the atmospheric temper-
    atures increased by 40 to 60 K in less than a week.
    This unprecedented event caused the polar vortex and
    the ozone hole, normally centered above the South Pole,
    to split into two smaller, separate off-center parts (Figure
    3) [9].
    Figure 4 shows the stratospheric temperatures between
    September and October 2002 along with the AMANDA-
    II muon rate. The muon rate traces temperature varia-
    tions in the atmosphere in great detail, with the strongest
    correlation observed for the 40-50 hPa layer.
    B. South Pole atmosphere 2007-2008
    Unlike in 2002, the stratospheric conditions over
    Antarctica were closer to average in 2007 and 2008.
    In 2007 the polar vortex was off-center from the South
    Pole during most of September and October, resulting in
    greater heat flux into the vortex, which decreased rapidly
    in size. When it moved back over the colder Pole region
    in early November it gained strength and persisted until
    the beginning of December. The 2008 polar vortex was

    4
    ATMOSPHERICVARIATION
    Fig. 3.
    Ozone concentration over the southern hemisphere on
    September 20th 2002 (left) and September 25th 2002 (right) [10].
    Date
    09/09 23/09 07/10 21/10
    R
    a
    t
    e
    [
    H
    z
    ]
    64
    66
    68
    70
    72
    74
    T
    [
    K
    ]
    180
    190
    200
    210
    220
    230
    240
    250
    20 - 30 hPa
    30 - 40 hPa
    40 - 50 hPa
    50 - 60 hPa
    60 - 70 hPa
    70 - 80 hPa
    80 -100 hPa
    Fig. 4. Average temperatures in various atmospheric layers over the
    South Pole (top) and deep ice muon rate recorded with the AMANDA-
    II detector (bottom) during the Antarctic ozone hole split of September-
    October 2002.
    one of the largest and strongest observed in the last 10
    years over the South Pole. Because of this, the heat flux
    entering the area was delayed by 20 days. The 2008
    vortex broke the record in longevity by persisting well
    into mid-December.
    In Figure 5 we overlay the IceCube muon rate over
    the temperature profiles of the Antarctic atmosphere pro-
    duced by the NOAA Stratospheric analysis team[11]. We
    note that the anomalous muon rates (see, for example,
    the sudden increase by 3% on 6 August 2008) observed
    by IceCube are in striking correlation with the middle
    and lower stratospheric temperature anomalies.
    We are establishing automated detection methods for
    anomalous events in the South Pole atmosphere as well
    as a detailed understanding and better modeling of
    cosmic ray interactions during such stratospheric events.
    IV. ACKNOWLEDGEMENTS
    We are grateful to the South Pole Meteorology Office
    and the Antarctic Meteorological Research Center of
    the University of Wisconsin-Madison for providing the
    Fig. 5.
    The temperature time series of the Antarctic atmosphere
    produced by NOAA[11] are shown for 2007 and 2008. The pattern
    observed in the deep ice muon rate (black line) is superposed onto
    the plot to display the striking correlation with the stratospheric
    temperature anomalies.
    meteorological data. This work is supported by the
    National Science Foundation.
    REFERENCES
    [1] Karle, A. for the IceCube Collaboration (2008), IceCube: Con-
    struction Status and First Results, arXiv:0812.3981v1.
    [2] Clem, J. et al. (2008), Response of IceTop tanks to low-energy
    particles, Proceedings of the 30th ICRC, Vol. 1 (SH), p.237-240.
    [3] Dorman, L. (2004), Cosmic Rays in the Earth’s Atmosphere and
    Underground, Springer Verlag.
    [4] Barrett P. et al. (1952), Interpretation of cosmic-ray measurements
    far underground, Reviews of Modern Physics, Vol. 24, Issue 3,
    p.133-178.
    [5] Ambrosio, M. et al. (1997), Seasonal variations in the underground
    muon intensity as seen by MACRO, Astroparticle Physics, Vol. 7,
    Issue 1-2, p.109-124.
    [6] Gaisser, T. (1990), Cosmic rays and particle physics. Cambridge,
    UK: Univ. Pr.
    [7] Grashorn, E. et al., Observation of the seasonal variation in
    underground muon intensity, Proceedings of the 30th ICRC, Vol.
    5 (HE part 2), p.1233-1236.
    [8] Osprey, S., et al. (2009), Sudden stratospheric warmings seen in
    MINOS deep underground muon data, Geophys. Res. Lett., 36,
    L05809, doi:10.1029/2008GL036359.
    [9] Varotsos, C., et al. (2002), Environ. Sci. & Pollut. Res., p.375.
    [10] http://ozonewatch.gsfc.nasa.gov
    [11] http://www.cpc.noaa.gov/products/stratosphere/strat-trop/

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    1
    Supernova Search with the AMANDA / IceCube Detectors
    Thomas Kowarik
    , Timo Griesel
    , Alexander Pie´gsa
    for the IceCube Collaboration
    Institute of Physics, University of Mainz, Staudinger Weg 7, D-55099 Mainz, Germany
    see the special section of these proceedings
    Abstract
    . Since 1997 the neutrino telescope
    AMANDA at the geographic South Pole has been
    monitoring our Galaxy for neutrino bursts from su-
    pernovae. Triggers were introduced in 2004 to submit
    burst candidates to the Supernova Early Warning
    System SNEWS. From 2007 the burst search was
    extended to the much larger IceCube telescope, which
    now supersedes AMANDA. By exploiting the low
    photomultiplier noise in the antarctic ice (on average
    280 Hz for IceCube), neutrino bursts from nearby
    supernovae can be identified by the induced collective
    rise in the pulse rates. Although only a counting
    experiment, IceCube will provide the world’s most
    precise measurement of the time profile of a neutrino
    burst near the galactic center. The sensitivity to
    neutrino properties such as the
    θ
    13
    mixing angle and
    the neutrino hierarchy are discussed as well as the
    possibility to detect the deleptonization burst.
    Keywords
    : supernova neutrino IceCube
    I. INTRODUCTION
    Up to now, the only detected extra-terrestrial sources
    of neutrinos are the sun and supernova SN1987A. To
    extend the search to TeV energies and above, neutrino
    telescopes such as AMANDA and IceCube [1] have been
    built. It turns out that the noise rates of the light sensors
    (OMs) in the antarctic ice are very low (
    ∼ 700 Hz
    for AMANDA,
    ∼ 280 Hz
    for IceCube) opening up
    the possibility to detect MeV electron anti-neutrinos
    from close supernovae by an increase in the collective
    rate of all light sensors. The possibility to monitor the
    galaxy for supernova with neutrino telescopes such as
    AMANDA has first been proposed in [2] and a first
    search has been performed using data from the years
    1997 and 1998 [3].
    The version of the the supernova data acquisition
    (SNDAq) covered in this paper has been introduced for
    AMANDA in the beginning of the year 2000 and was
    extended to IceCube in 2007. The AMANDA SNDAq
    has been switched off in February 2009. We will in-
    vestigate data recorded by both telescopes concentrating
    on the 9 years of AMANDA measurements and make
    predictions for the expected sensitivity of IceCube.
    II. DETECTORS AND DATA ACQUISITION
    In AMANDA, the pulses of the 677 OMs are collected
    in a VME/Linux based data acquisition system which
    operates independently of the main data acquisition
    aimed at high energy neutrinos. It counts pulses from
    Rate / Hz
    100 150 200 250 300 350 400 450
    Number of Entries
    1
    10
    10
    2
    10
    3
    10
    4
    10
    5
    Fig. 1. Rate distribution of a typical IceCube module
    The recorded rate distribution (filled area) encompasses about 44
    days and has been fitted with a Gaussian (solid line,
    χ
    2
    /N
    dof
    =
    56.5
    ) defined by
    µ = 271 Hz
    and
    σ = 21 Hz
    . However, a
    lognormal function (dotted line,
    χ
    2
    /N
    dof
    = 0.66
    ) with geomet-
    ric mean
    µ
    geo
    = 6.4 ln(Hz)
    , geometric standard deviation
    σ
    geo
    =
    0.03 ln(Hz)
    and a shift of
    x
    0
    = 377 Hz
    is much better at describing
    the data.
    every connected optical module in a 20 bit counter in
    fixed
    10ms
    time intervals that are synchronized by a
    GPS-clock.
    In IceCube, PMT rates are recorded in a
    1.6384ms
    binning by scalers on each optical module. The infor-
    mation is locally buffered and read out by the IceCube
    data acquisition system. It then transfers this data to the
    SNDAq, which synchronizes and regroups the informa-
    tion in
    2ms
    bins.
    The software used for data acquisition and analysis
    is essentially the same for AMANDA and IceCube. The
    data is rebinned in
    500 ms
    intervals and subjected to an
    online analysis described later. In case of a significant
    rate increase (“supernova trigger”), an alarm is sent to
    the Supernova Early Warning System (SNEWS, [4]) via
    the Iridium satellite network and the data is saved in a
    fine time binning (
    10ms
    for AMANDA and
    2ms
    for
    IceCube).
    III. SENSOR RATES
    In
    500 ms
    time binning, the pulse distribution of
    the average AMANDA or IceCube OM conforms only
    approximately to a Gaussian. It can more accurately be
    described by a lognormal distribution (see figure 1).
    The pulse distributions exhibit Poissonian and cor-
    related afterpulse components contributing with similar

    2
    T.KOWARIK
    et al.
    SUPERNOVA SEARCH WITH ICECUBE
    strengths. The correlated component is anticorrelated
    with temperature and arises form Cherenkov light caused
    by
    40
    K
    decays and glass luminescence from radioactive
    decay chains. A cut at
    250 µs
    on the time difference
    of consecutive pulses effectively suppresses afterpulse
    trains, improves the significance of a simulated super-
    nova at
    7.5kpc
    by approximately
    20%
    and makes the
    pulse distribution more Poissonian in nature.
    IV. EFFECTIVE VOLUMES
    Supernovae radiate all neutrino flavors, but due to the
    relatively large inverse beta decay cross section, the main
    signal in IceCube is induced by electron anti-neutrinos
    (see [5]). The rate
    R
    per OM can be approximated
    weighing the energy dependent anti-electron neutrino
    flux at the earth
    Φ(E
    ν
    ¯
    e
    )
    (derived from the neutrino
    luminosity and spectra found in [6]) with the effective
    area for anti-electron neutrino detection
    A
    eff
    (E
    ν
    ¯
    e
    )
    and
    integrating over the whole energy range:
    R =
    ?
    0
    dE
    ν
    ¯
    e
    Φ(E
    ν
    ¯
    e
    )A
    eff
    (E
    ν
    ¯
    e
    ) , with
    A
    eff
    (E
    ν
    ¯
    e
    ) = n
    ?
    0
    dE
    e
    +
    dE
    e
    +
    (E
    ν
    ¯
    e
    , E
    e
    +
    ) V
    eff,e
    +
    (E
    e
    +
    ) .
    dE
    e
    +
    (E
    ν
    ¯
    e
    , E
    e
    +
    )
    is the inverse beta decay cross section,
    n
    the density of protons in the ice and
    V
    eff,e
    +
    (E
    e
    +
    )
    the
    effective volume for positron detection of a single OM.
    V
    eff,e
    +
    (E
    e
    +
    )
    can be calculated by multiplying the
    number of Cherenkov photons produced with the effec-
    tive volume for photon detection
    V
    eff,γ
    ch
    .
    By tracking Cherenkov photons in the antarctic ice
    around the IceCube light sensors [7] and simulating
    the module response one obtains
    V
    eff,γ
    ch
    = 0.104 m
    3
    for the most common AMANDA sensors and
    V
    eff,γ
    =
    0.182 m
    3
    for the IceCube sensors. With a
    GEANT-4
    simulation, the amount of photons produced by a
    positron of an energy
    E
    e
    +
    can is estimated to be
    N
    γ
    ch
    = 270 E
    e
    +
    / MeV
    . Consequently, the effective
    volumes for positrons as a function of their energies
    are
    V
    AMANDA
    eff,e
    +
    = 19.5 E
    e
    +
    m
    3
    MeV
    and
    V
    IceCube
    eff,e
    +
    =
    34.2 E
    e
    +
    m
    3
    MeV
    . Uncertainties in the effective volumes
    derive directly from the uncertainties of the ice models
    (
    ∼ 5%
    ) and in the OM sensitivities (
    ∼ 10%
    ).
    In this paper we assume a supernova neutrino pro-
    duction according to the
    Lawrence-Livermore
    model [8]
    as it is the only one that provides spectra for up to
    15 s
    . It gives a mean electron anti-neutrino energy of
    15 MeV
    corresponding to an average positron energy of
    (13.4 ± 0.5) MeV
    .
    V. ANALYSIS PROCEDURE
    A simple investigation of the rate sums would be
    very susceptible to fluctuations due to variations in the
    detector response or external influences such as the
    seasonal variation of muon rates. Medium and long term
    fluctuations are tracked by estimating the average count
    rate by a sliding time window. The rate deviation for
    a collective homogeneous neutrino induced ice illumi-
    nation is calculated by a likelihood technique. In time
    bins of
    0.5 s
    and longer, the pulse distributions can be
    approximated by Gaussian distributions. For a collective
    rate increase
    ∆µ
    , the expectation value of the average
    mean rate
    µ
    i
    of a light sensor
    i
    with relative sensitivity
    ǫ
    i
    increases to
    µ
    i
    + ǫ
    i
    ∆µ
    . The mean value
    µ
    i
    and
    its standard deviation
    σ
    i
    are averaged over a sliding
    window of
    10 min
    , excluding
    15 s
    before and after the
    0.5 s
    time frame
    r
    i
    studied. By taking the product of
    the corresponding Gaussian distributions the following
    likelihood for a rate deviation
    ∆µ
    is obtained:
    L =
    N
    ?
    OM
    i=1
    1
    2π σ
    i
    exp 
    (r
    i
     (µ
    i
    + ǫ
    i
    ∆µ))
    2
    2
    i
    .
    Minimization of
     ln L
    leads to:
    ∆µ =
    N
    ?
    OM
    i=1
    ǫ
    2
    i
    σ
    2
    i
    ?
     1
    ?
    ?? ?
    = σ
    2
    ∆µ
    N
    ?
    OM
    i=1
    ǫ
    i
    (r
    i
     µ
    i
    )
    σ
    2
    i
    .
    The data is analyzed in the three time binnings
    0.5 s
    ,
    4 s
    and
    10 s
    for the following reasons: First, the finest time
    binning accessible to the online analysis is
    0.5 s
    . Second,
    as argued in [9], the neutrino measurements of SN1987A
    are roughly compatible with an exponential decay of
    τ = 3s
    . The optimal time frame for the detection of
    a signal with such a signature is
    ≈ 3.8s
    . Last,
    10s
    are
    the approximate time frame where most of the neutrinos
    from SN1987A fell.
    To ensure data quality, the optical modules are sub-
    jected to careful quality checks and cleaning. Those
    modules with rates outside of a predefined range, a high
    dispersion w.r.t. the Poissonian expectation or a large
    skewness are disqualified in real time.
    Since SNEWS requests one alarm per 10 days, the
    supernova trigger is set to
    6.3 σ
    .
    To ensure that the observed rate deviation is homo-
    geneous and isotropic, the following
    χ
    2
    discriminant is
    examined:
    χ
    2
    (∆µ) =
    N
    ?
    OM
    i=1
    ?
    r
    i
     (µ
    i
    + ǫ
    i
    ∆µ)
    σ
    i
    ?
    2
    .
    We demand the data to conform to a
    χ
    2
    -confidence
    level of
    99.9 %
    . However, it was found that the
    χ
    2
    cannot clearly distinguish between isotropic rate changes
    and fluctuations of a significant number of OMs: If
    e.g.
    20%
    of the OMs record rate increases of about
    1 σ
    (
    ≈ 20 pulses/0.5 s
    ), the significance for isotropic
    illumination can rise above
    6 σ
    without being rejected
    by the
    χ
    2
    condition. Still, no method was found that
    performed better than
    χ
    2
    .
    VI. EXTERNAL PERTURBATIONS
    Figure 2 shows the distribution of the significance
    ξ =
    ∆µ
    σ
    ∆µ
    for AMANDA. From the central limit theorem, one
    would expect a Gaussian distribution with a width of

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    3
    Significance /
    σ
    −10
    −5
    0
    5
    10
    Number of Entries
    1
    10
    10
    2
    10
    3
    10
    4
    10
    5
    10
    6
    Fig. 2. Significances of AMANDA in
    0.5 s
    time frames
    The filled area shows the significance of the data taken in 2002 with
    a spread of
    σ = 1.13
    . The solid line denotes the initial background
    simulation (
    σ ≈ 1
    ) and the dotted line is a toy monte carlo taking
    into account the fluctuations due to muon rates (
    σ ≈ 1.11
    ).
    σ = 1
    . This expectation is supported by a background
    simulation using lognormal representations of individual
    light sensor pulse distributions. However, one finds that
    the observable is spread wider than expected and exhibits
    a minor shoulder at high significances.
    As this observable is central to the identification of
    supernovae, its detailed understanding is imperative and
    a close investigation of these effects is necessary.
    Faults in the software were ruled out by simulating
    data at the most basic level, before entering the SNDAq.
    Hardware faults are improbable, as the broadening of
    the rates is seen both for AMANDA (
    σ = 1.13
    ) and
    IceCube (
    σ = 1.27
    ) and the detectors use different and
    independent power and readout electronics.
    One can then check for external sources of rate
    changes as tracked by magnetometers, rio- and pho-
    tometers as well as seismometers at Pole. All have
    been synchronized with the rate measurements of the
    OMs. Only magnetic field variations show a slight,
    albeit insignificant, influence on the rate deviation of
     4·10
     5
    Hz
    nT
    for AMANDA. Due to a
    µ
    metal wire mesh
    shielding, the influence on IceCube sensors is smaller by
    a factor
    ∼ 30
    .
    It turns out that the main reason for the broadening
    are fluctuations of the atmospheric muon rates. During
    0.5 s
    , AMANDA detects between
    ≈ 3.0 · 10
    3
    (June) and
    ≈ 3.4·10
    3
    (December) sensor hits due to muons. Adding
    hits from atmospheric muons broadens and distorts the
    distribution derived from noise rates. A simulation taking
    into account the muon triggered PMT hits increases the
    width of the significance distribution to
    ≈ 1.11
    (see
    figure 2). Although the investigations are still ongoing
    for IceCube, we expect the larger
    σ
    can be ascribed to
    its lower noise rate leading to a higher resolution and
    thereby a stronger sensitivity to perturbations.
    A Fourier transformation was performed on the data
    stream. No evidence for periodically recurring events
    was found, but occasional correlations between subse-
    quent
    0.5 s
    time frames both in the summed noise rate
    and the rate deviation could be identified. We determined
    the mean significance before and after a rate increase
    exceeding a predefined level. One observes a symmetric
    correlation in the data mainly between
    ±10 s
    . The origin
    of the effect, which is present both in AMANDA and
    IceCube data, is under investigation.
    VII. EXPECTED SIGNAL AND VISIBILITY RANGE
    A preliminary analysis of AMANDA data from 2000
    to 2003 yielded a detection range of
    R
    4s
    = 14.5kpc
    at the optimal binning of
    4s
    at an efficiency of
    90%
    ,
    encompassing
    81%
    of the stars of our Galaxy. As signals
    with exponentially decreasing luminosity (at
    τ = 3 s
    )
    have been used as underlying models, the two other
    binnings were less efficient with
    R
    0.5 s
    = 10 kpc
    (
    56%
    )
    and
    R
    10 s
    = 13 kpc
    (
    75%
    ), respectively.
    The rates of the IceCube sensors are stable and
    uniform across the detector. Scaling the observed rates
    to the full 80 string detector, a summed rate of
    ?R
    IceCube
    ? = (1.3· 10
    6
    ± 1.8· 10
    2
    ) Hz
    is expected. While
    a single DOM would only see an average rate increase of
    13 Hz
    or
    0.65 σ
    , the signal in the whole IceCube detector
    would be
    6.1 · 10
    4
    Hz
    or
    34 σ
    for a supernova at
    7.5kpc
    distance. Using this simple counting method, IceCube
    would see a supernova in the Magellanic Cloud with
    5 σ
    significance.
    Figure 3 shows the expected signal in IceCube for
    a supernova at
    7.5 kPc
    conforming to the
    Lawrence-
    Livermore
    model with
    ≈ 10
    6
    registered neutrinos in
    15s
    and a statistical accuracy of
    0.1%
    in the first
    2s
    .
    Assuming
    2 · 10
    4
    events in Super-Kamiokande (scaled
    from [12]), one arrives at an accuracy of
    1%
    in the same
    time frame. While IceCube can neither determine the
    directions nor the energies of the neutrinos, it will pro-
    vide worlds best statistical accuracy to follow details of
    Time after Core Bounce / s
    −5
    0
    5
    10
    15
    20
    Summed Detector Hits
    700
    750
    800
    850
    900
    950
    ×10
    3
    Fig. 3. Expected signal of a supernova at
    7.5 kpc
    in IceCube.
    The solid line denotes the signal expectation and the bars a randomized
    simulation.

    4
    T.KOWARIK
    et al.
    SUPERNOVA SEARCH WITH ICECUBE
    Distance / kpc
    10
    20
    30
    40
    50
    60
    Significance
    5
    10
    15
    20
    25
    Fig. 4. Supernova detection ranges using
    0.5 s
    time frames
    Significances which supernovae conforming to the
    Lawrence-
    Livermore
    model would cause as function of their distance. IceCube
    performance is described by the solid line, AMANDA by the dotted
    line. Both are given for the
    0.5 s
    binning.
    the neutrino light curve. Its performance in this respect
    will be in the same order as proposed megaton proton
    decay and supernova search experiments. As the signal
    is seen on top of background noise, the measurement
    accuracy drops rapidly with distance. Figure 4 shows the
    significance at which IceCube and AMANDA would be
    able to detect supernovae.
    VIII. DELEPTONIZATION PEAK AND NEUTRINO
    OSCILLATIONS
    As mentioned before, the signal seen by AMANDA
    and IceCube is mostly dominated by electron anti-
    neutrinos with a small contribution by electron neutrino
    scattering, thereby leading to a sensitivity which is
    strongly dependent on the neutrino flavor and is thus
    sensitive to neutrino oscillations.
    As the neutrinos pass varying levels of density within
    the supernova, the flux of electron and electron anti-
    neutrinos
    Φ
    released is different from the initial flux
    Φ
    0
    produced during the collapse:
    Φ
    ν
    e
    = p Φ
    0
    ν
    e
    + (1  p)Φ
    0
    ν
    x
    ,
    Φ
    ν
    ¯
    e
    = p¯Φ
    0
    ν
    ¯
    e
    + (1  p¯) Φ
    0
    ν
    ¯
    x
    .
    The survival probability
    p/p¯
    for the
    ν
    e
    /ν¯
    e
    ’s varies with
    the mass hierarchy and the mixing angle
    θ
    13
    [10]:
    neutrino oscillation parameters
    p
    m
    2
    2
    <m
    2
    3
    , sin
    2
    θ
    13
    > 10
     3
    ≈ 0%
    69%
    m
    1
    2
    >m
    2
    3
    < 0 , sin
    2
    θ
    13
    > 10
     3
    31%
    ≈ 0%
    any hierarchy,
    sin
    2
    θ
    13
    < 10
     5
    31%
    69%
    At the onset of the supernova neutrino burst during the
    prompt shock, a
    ∼ 10 ms
    long burst of electron neutrinos
    gets emitted when the neutron star forms. As the shape
    and rate of this burst is roughly independent of the
    properties of the progenitor stars, it is considered as
    Time after Core Bounce / s
    −0.02 −0.01
    0
    0.01 0.02
    0.03 0.04
    0.05 0.06
    0.07
    Summed Detector Hits
    2400
    2600
    2800
    3000
    3200
    3400
    3600
    3800
    deleptonization peak
    Fig. 5. Neutrino signals of a supernova at
    7.5 kpc
    distance modified
    by oscillations as seen in IceCube.
    The line shows the expectation without neutrino oscillations, the
    squares show the simulated signal for the normal mass hierarchy
    with
    sin
    2
    θ
    13
    > 10
     3
    (
    sin
    2
    θ
    13
    < 10
     5
    lies nearly on the same
    line) and the triangles show the signal for inverted mass hierarchy at
    sin
    2
    θ
    13
    > 10
     3
    .
    a standard candle, allowing one to determine neutrino
    properties without knowing details of the core collapse.
    The first
    0.7 s
    of a supernova signal were modeled af-
    ter [11]. Figure 5 shows the expectation for a supernova
    at a distance of
    7.5 kpc
    in the
    2 ms
    binning of IceCube.
    Due to statistics and the rising background from the
    starting electron anti-neutrino signal, the identification
    of the neutronization burst is unlikely at this distance.
    However, with a supernova at
    7.5 kpc
    one should be able
    to draw conclusions for the mass hierarchy, depending
    on the reliability of the models.
    IX. CONCLUSIONS AND OUTLOOK
    With 59 strings and 3540 OMs installed, IceCube
    has reached
    86%
    of its final sensitivity for supernova
    detection. It now supersedes AMANDA in the SNEWS
    network. With the low energy extension DeepCore,
    IceCube gains 360 additional OMs with a
    ∼ 30%
    higher
    quantum efficiency (rates
    ∼ 380 Hz
    ). These modules
    have not been considered in this paper.
    REFERENCES
    [1] A. Karle et al.
    arXiv:0812.3981v1
    [2] F. Halzen, J. E. Jacobsen and E. Zas
    Phys. Rev.
    , D49:1758–1761,
    1994
    [3] J. Ahrens et al.
    Astropart. Phys.
    16:345-359, 2002
    [4] P. Antonioli et al.
    New J. Phys.
    6:114, 2004
    [5] W. C. Haxton.
    Phys. Rev.
    , D36:2283, 1987
    [6] M. Th. Keil, G. G. Raffelt and H.-T. Janka.
    Astrophys. J.
    ,
    590:971–991, 2003
    [7] M. Ackermann et al.
    J. Geophys. Res.
    , 111:D13203, 2006
    [8] T. Totani, K. Sato, H. E. Dalhed and J. R. Wilson
    Astrophys. J.
    ,
    496:216–225, 1998
    [9] J. F. Beacom and P. Vogel
    Phys. Rev.
    D60:033007, 1999
    [10] A. Dighe.
    Nucl. Phys. Proc. Suppl.
    , 143:449–456, 2005
    [11] F. S. Kitaura, Hans-Thomas Janka and W. Hillebrandt
    Astron.
    Astrophys.
    450:345-350, 2006
    [12] M. Ikeda et al.
    Astrophys. J.
    , 669:519-524, 2007

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁOD´ Z´ 2009
    1
    Physics Capabilities of the IceCube DeepCore Detector
    Christopher Wiebusch
    ¤
    for the IceCube Collaboration
    y
    ¤
    III.Physikalisches Institut, RWTH Aachen, University, Germany
    y
    See the special section of these proceedings
    Abstract.
    IceCube-DeepCore
    is
    a
    compact
    Cherenkov Detector located in the clear ice of the
    bottom center of the IceCube Neutrino Telescope.
    Its purpose is to enhance the sensitivity of IceCube
    for low neutrino energies (< 1 TeV) and to lower
    the detection threshold of IceCube by about an
    order of magnitude to below 10 GeV. The detector is
    formed by 6 additional strings of 360 high quantum
    efficiency phototubes together with the 7 central
    IceCube strings. The improved sensitivity will
    provide an enhanced sensitivity to probe a range
    of parameters of dark matter models not covered
    by direct experiments. It opens a new window for
    atmospheric neutrino oscillation measurements of
    º
    ¹
    disappearance or º
    ¿
    appearance in an energy
    region not well tested by previous experiments, and
    enlarges the field of view of IceCube to a full sky
    observation when searching for potential neutrino
    sources. The first string was succesfully installed in
    January 2009, commissioning of the full detector is
    planned early 2010.
    Keywords: Neutrino-astronomy, IceCube-DeepCore
    I. INTRODUCTION
    Main aim of the IceCube neutrino observatory [1] is
    the detection of high energy extraterrestrial neutrinos
    from cosmic sources, e.g. from active galactic nuclei.
    The detection of high energy neutrinos would help to
    resolve the question of the sources and the acceleration
    mechanisms of high energy cosmic rays.
    IceCube is located at the geographic South-Pole. The
    main instrument of IceCube will consist of 80 cable
    strings, each with 60 highly sensitive photo-detectors
    which are installed in the clear ice at depths between
    1450 m and 2450 m below the surface. Charged leptons
    with an energy above 100 GeV inside or close to the
    detector produce enough Cherenkov light to be de-
    tected and reconstructed using the timing information of
    the photoelectrons recorded with large area phototubes.
    While the primary goal is of highest scientific interest,
    the instrument can address a multitude of scientific
    questions, ranging from fundamental physics such as
    physics on energy-scales beyond the reach of current
    particle accelerators to multidisciplinary aspects e.g. the
    optical properties of the deep Antarctic ice which reflect
    climate changes on Earth.
    IceCube is complemented by other major detector
    components. The surface air-shower detector IceTop is
    used to study high energy cosmic rays and to calibrate
    ?
    ?
    ????????
    ????????????
    ????????????
    ????????
    ????????
    ???????????
    ???????????
    ??????????
    ??????? ??
    ??????
    ????????
    ??????????????????
    ??????
    ????????????????????????? ?
    ????????????
    Fig. 1.
    Geometry of the DeepCore Detector. The top part shows
    the surface projection of horizontal string positions and indicates the
    positions of AMANDA and DeepCore. The bottom part indicates the
    depth of sensor positions. At the left the depth-profile of the optical
    transparency of the ice is shown.
    IceCube. R&D studies are underway to supplement Ice-
    Cube with radio (AURA) and acoustic sensors (SPATS)
    in order to extend the energy range beyond EeV en-
    ergies. Six additional and more densely instrumented
    strings will be deployed in the bottom center of the Ice-
    Cube detector and form the here considered DeepCore
    detector.
    A first DeepCore string has been succesfully installed
    in January 2009 and is taking data since then. The
    DeepCore detector will be completed in 2010 and will
    replace the existing AMANDA-II detector, which has
    been decomissioned in May 2009. DeepCore will lower
    the detection threshold of IceCube by an order of mag-
    nitude to below 10 GeV and, due to its improved design,
    provide new capabilities compared to AMANDA. In
    this paper we describe the design of DeepCore and the
    enhanced physics capabilities which can be addressed.
    II. DEEPCORE DESIGN AND GEOMETRY
    The geometry of DeepCore is sketched in figure 1.
    The detector consists out of 6 additional strings of

    2
    CHRISTOPHER WIEBUSCH et al. CAPABILITIES OF ICECUBE DEEPCORE
    0
    10
    20
    30
    40
    0
    40
    80
    120
    Optical Efficiency (%)
    DOMs
    p
    y( )
    2
    High QE
    Standard IceCube
    Fig. 2. Results of the quantum efficiency calibration at ¸ = 405 nm
    of the DeepCore phototubes compared to standard IceCube phototubes.
    60 phototubes each together with the 7 central IceCube
    strings. The detector is divided into two components:
    Ten sensors of each new string are at shallow depths
    between 1750 m and 1850 m, above a major dust-layer
    of poorer optical transparency and will be used as a veto-
    detector for the deeper component. The deep component
    is formed by 50 sensors on each string and is installed
    in the clear ice at depths between 2100 m and 2450 m.
    It will form, together with the neighbouring IceCube
    sensors the main physics volume.
    The deep ice is on average twice as clear as the
    average ice above 2000 m [2]. The effective scattering
    length reaches 50 m and the absorption length 230 m.
    Compared to AMANDA a substantially larger number
    of unscattered photons will be recorded allowing for
    an improved pattern recognition and reconstruction of
    neutrino events in particular at lower energies.
    Another important aspect is a denser spacing of photo-
    sensors compared to IceCube: The horizontal inter-string
    spacing is 72 m (IceCube: 125 m). The vertical spacing
    of sensors along a string is only 7 m (IceCube: 17 m).
    The next major improvement with respect to IceCube
    and AMANDA is the usage of new phototubes (HAMA-
    MATSU R7081-MOD) of higher quantum efficiency.
    This hemispherical 10” photomultiplier is identical to
    the standard IceCube PMT [3], but employs a modified
    cathode material of higher quantum efficiency (typically
    33 % at ¸ = 390 nm). Calibrations of the phototubes for
    DeepCore confirm a sensitivity improvement of 30 %-
    40 % with respect to the standard IceCube PMT (figure
    2). Also regular IceCube strings will be equipped with
    these phototubes if within the DeepCore volume.
    The net effect of the denser instrumentation is a factor
    » 6 gain in sensitivity for photon detection and superior
    optical clarity of the ice. This is an imporant prerequisite
    for a substantially lower detection threshold.
    III. DEEPCORE PERFORMANCE
    The electronic hardware of the optical sensors is
    identical to the standard IceCube module [3] and this
    significantly reduces the efforts for maintenance and
    operations compared to AMANDA. The DeepCore de-
    tector is integrated into a homogeneous data aquisition
    model of IceCube which will be only supplemented
    by additional trigger. Initial comissioning data of the
    first installed DeepCore string verifies that the hardware
    works reliably and as expected.
    The IceCube detector is triggered if typically a mul-
    tiplicity of 8 sensors within » 5 ¹s observe a signal
    coincident with a hit in a neighbouring or next to
    neighbouring sensor. For each trigger, the signals of the
    full detector are transferred to the surface.
    For the sensors within the considered volume the
    data taking is supplemented with a reduced multiplicity
    requirement of typically 3¡4. As shown in figure 7, such
    a trigger is sufficient to trigger atmospheric neutrino
    events down to a threshold of 1 GeV, sufficiently below
    the anticipated physics threshold.
    The chosen location of DeepCore allows to utilize the
    outer IceCube detector as an active veto shield against
    the background of down-going atmospheric muons.
    These are detected at a » 10
    6
    higher rate than neutrino
    induced muons. The veto provides external informa-
    tion to suppress this background and standard up-going
    neutrino searches will strongly benefit from a larger
    signal efficiency and a lower detection threshold as the
    demands on the maturity of recorded signals decrease.
    Even more intriguing is the opportunity to identify
    down-going º induced ¹, which may, unlike cosmic ray
    induced atmosperic ¹, start inside the DeepCore detec-
    tor. Simulations [4] show that three rings of surrounding
    IceCube strings and the instrumentation in the upper
    part of IceCube are sufficient to achieve a rejection of
    atmospheric muons by a factor > 10
    6
    maintaining a
    large fraction of the triggered neutrino signals. A further
    interesting aspect is the proposal in [6] to veto also atmo-
    spheric º by the detection of a correlated atmospheric ¹.
    This could provide the opportunity to reject a substantial
    part of this usually irreducable background for extra-
    terrestrial neutrino searches.
    Triggered events which start inside the detector will
    be selected online and transmitted north by satellite.
    Already simple algorithms allow to suppress the back-
    ground rate by a factor > 10
    3
    and meet the bandwidth
    requirements while keeping 90% of the signal [4]. A
    typical strategy requires that the earliest hits are located
    inside DeepCore and allows for later hits in the veto-
    region only if the time is causal consistent with the
    hypothesis of a starting track. A filter which selects
    starting tracks in IceCube is active since 2008 and
    allowed to verify the performance of such filters with
    experimental data and to benchmark the subsequent
    physics analysis.
    The filtered events are analyzed offline with more
    sophisticated reconstruction algorithms. Here, the fo-
    cus is to improve the purity of the sample and to
    reconstruct direction, energy and the position of the
    interaction vertex. A particularily efficient likelihood
    algorithm (finiteReco [4]) capable of selecting starting
    muons evaluates the hit probabilities of photomultipliers
    with and without a signal in dependence of the distance
    to the track. It estimates the most probable position of

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁOD´ Z´ 2009
    3
    E
    ν
    [GeV]
    0
    20
    40
    60
    80
    100
    120
    140
    160
    180
    200
    220
    [m]
    Reco
    L
    0
    50
    100
    150
    200
    250
    300
    350
    400
    450
    Entries 642
    Mean x 57.25
    Mean y
    142.1
    RMS x 46.98
    RMS y
    101.9
    Entries 642
    Mean x 57.25
    Mean y
    142.1
    RMS x 46.98
    RMS y
    101.9
    preliminary
    Fig. 3. The reconstructed length of ¹ contained tracks in DeepCore,
    based on the reconstructed start- and stop- vertex with the finiteReco
    algorithm. The data are º induced ¹-tracks from the upper hemisphere,
    which are reconstructed to start within DeepCore.
    10
    (Primary Neutrino Energy - GeV)
    log
    1
    1.2
    1.4
    1.6
    1.8
    2
    2.2
    2.4
    2.6
    2.8
    3
    )
    2
    Effective Neutrino Area (m
    -6
    10
    -5
    10
    -4
    10
    -3
    10
    -2
    10
    -1
    10
    1
    preliminary
    Fig. 4. Effective neutrino detection area of IceCube (trigger level)
    versus the energy for up-going neutrinos. The squares are IceCube
    only. The circles represent the area if DeepCore is included.
    the start-vertex and provides the probablity that a track
    may have reached this point undetected by the veto.
    The reconstruction algorithms are still under develop-
    ment but initial results are promising. As an example,
    figure 3 shows the reconstructed length of ¹ tracks as
    function of the º energy. Already the currently achieved
    resolution of » 50 m results in a visible correlation with
    the neutrino energy in particular for energies . 100 GeV.
    Note, that the resolution is substantially better for verti-
    cal tracks.
    The effective detection area of IceCube for neutrinos
    for triggered events is shown in figure 4. Despite Deep-
    Core being much smaller than IceCube, a substantial
    gain of up to an order of magnitude is achieved by
    the additional events detected in DeepCore. Higher level
    event selections for specific physics analysis benefit
    strongly from the higher information content of events
    and the gain of DeepCore further improves.
    IV. PHYSICS POTENTIAL
    A. Galactic point sources of neutrinos
    The analysis of IceCube data greatly benefits from
    the location at the geographic South Pole because the
    celestial sphere fully rotates during one sideral day.
    Azimuthal detector effects are largly washed out because
    Fig. 5.
    Interesting celestrial object with known emission of TeV
    gamma rays.
    each portion of the sky is observed with the same expo-
    sure and same inclination. However, the aperture of the
    conventional up-going muon analysis is restricted to only
    the Northern hemisphere and leaves out a large fraction
    of the galactic plane and a number of interesting objects
    such as the galactic center (see figure 5). Extending the
    field of view of IceCube at low energies (. 1 TeV) to a
    full sky observation will greatly enlarge the number of
    interesting galactic sources in reach of IceCube
    1
    .
    The energy spectrum of gamma rays from supernova
    remnants show indications of a potential cut-off at a few
    TeV [7]. Under the assumption of a hadronic production
    mechanism for these gamma rays the corresponding
    neutrino fluxes would show a similar cut-off at typ-
    ically half of that cut-off value. The high sensitivity
    of DeepCore for neutrinos of TeV energies and below
    will complement the sensitivity of IceCube which is
    optimized for energies of typically 10 TeV and above.
    B. Indirect detection of dark matter
    The observation of an excess of high energy neutrinos
    from the direction of the Sun can be interpreted by
    means of annihilations of WIMP-dark matter in its
    center. The energy of such neutrinos is a fraction of the
    mass of the WIMP particles (expected on the TeV-scale)
    and it depends on the decay chains of the annihilation
    products. The large effective area of DeepCore and the
    possibility of a highly efficient signal selection greatly
    improves the sensitivity of IceCube. In particular it is
    possible to probe regions of the parameter space with
    soft decay chains and WIMP masses below » 200
    GeV and which are not disfavored by direct search
    experiments.
    An example of the sensitivity for the hard annihilation
    channel of supersymmetric neutralino dark matter is
    1
    Note, that at high energies > 1 PeV the background of atmospheric
    muons rapidly decreases and also here neutrinos from the Southern
    hemisphere can be detected by IceCube [8]. However, galactic sources
    are usually not expected to produce significant fluxes of neutrinos at
    energies around the cosmic ray knee and above.

    4
    CHRISTOPHER WIEBUSCH et al. CAPABILITIES OF ICECUBE DEEPCORE
    Fig. 6.
    The expected upper limit of IceCube DeepCore at 90%
    confidence level on the spin-dependent neutralino-proton cross section
    for the hard (W+W
    ¡
    ) annihilation channel as a function of the
    neutralino mass for IceCube including Deep Core (solid line). Also
    shown are limits from previous direct and indirect searches. The shaded
    areas represent MSSM models which are not disfavoured by direct
    searches, even if their sensitivity would be improved by a factor 1000.
    Primary Neutrino Energy - GeV
    10
    20
    30
    40
    50
    60
    70
    80
    90
    100
    0
    500
    1000
    1500
    2000
    2500
    3000
    3500
    4000
    Preliminary
    Fig. 7.
    Number of triggered vertical atmospheric neutrinos per
    year (per 3GeV) versus the neutrino energy. Events from 1:6¼ sr are
    accepted. Shown are the numbers without (squares) and with (circles)
    the inclusion of oscillations (¢m
    2
    atm
    = 0:0024 eV2, sin(2μ
    23
    ) = 1).
    shown in figure 6.
    C. Atmospheric neutrinos
    DeepCore will trigger on the order of 10
    5
    atmo-
    spheric neutrinos/year in the energy range from 1 GeV to
    100 GeV. Atmospheric neutrinos are largly unexplored
    in this energy range. Smaller experiments like Super-
    Kamiokande cannot efficiently measure the spectrum
    for energies above 10 GeV and measurements done by
    AMANDA only start at 1 TeV. In the range between
    30 ¡ 50 GeV decays of charged kaons become dominant
    over decays of charged pions [10] for the production of
    atmospheric neutrinos and the systematic error of flux
    calculations increases. A measurement of this transition
    could help to reduce systematic errors of the flux of
    atmospheric neutrinos at TeV energies.
    The first maximum of disappearence of atmospheric
    º
    ¹
    due to oscillations appears at an energy of about
    25 GeV for vertically up-going atmospheric neutrinos
    [5]. The energy threshold of about 10 GeV would allow
    to measure atmospheric neutrino oscillations by means
    of a direct observation of the oscillation pattern in
    this energy range. In addition, DeepCore would aim to
    observe the appearance of º
    ¿
    by the detection of small
    cascade-like events in the DeepCore volume at a rate
    which is anti-correlated with the disappearance of º
    ¹
    .
    Similar to º
    ¿
    , the signature of º
    e
    events are cascade-
    events with a large local light deposition without the
    signature of a track. The dominant background to these
    events are charged current º
    ¹
    interactions with a small
    momentum transfer to the ¹. Analyses like these will
    have to be performed considering all three flavors and
    their mixing. Note, that only for a further reduction of
    the energy threshold smaller than 10 GeV matter effects
    in the Earth’s core would become visible [5].
    D. Other physics aspects
    Two remaining items are only briefly mentioned here.
    Slowly moving magnetic monopoles, when catalizing
    proton decays, produce subsequent energy depositions
    of » 1 GeV along their path with time-scales of ¹s to
    ms. Initial studies are under-way to develop a dedicated
    trigger for this signature using delayed coincidences.
    DeepCore extends the possibility to search for neu-
    trino emission in coincidence with gamma ray bursts
    (GRB) to lower energies. According to [11] GRB may
    emit a burst of neutrinos. However, predicted energies
    are only a few GeV and the event numbers are small
    (» 10 yr
    ¡1
    km
    ¡2
    ). Additional studies are required to
    evaluate the sensitivity for such signals.
    V. SUMMARY AND OUTLOOK
    This paper summarizes the enhancement of the
    physics profile of IceCube by the DeepCore detector.
    The geometry of DeepCore has been optimized and
    construction has started. Detailed MC studies and ex-
    perimental analyses are currently under way to optimize
    and finalize the analysis procedures. First data from the
    full detector will be available in spring 2010, the veto
    will be fully completed latest 2011.
    ACKNOWLEDGEMENT
    This work is supported by the German Ministry for
    Education and Research (BMBF). For a full acknowl-
    edgement see [1].
    REFERENCES
    [1] J. Ahrens et al. (IceCube Collaboration), Astropart. Phys. 20
    (2004) 507-532, 2004
    [2] M. Ackermann et al. (IceCube Collaboration), J. of Geophys.
    Res. 111 (2006) D13203, July 2006
    [3] R. G. Stokstad et al. (IceCube Collaboration), Nucl. Phys. B
    (Proc. Suppl.) 118 (2003) 514.
    [4] O. Schulz et al. (IceCube Collaboration), these proceedings.
    [5] D. Grant et al. (IceCube Collaboration), these proceedings.
    [6] S. Schonert¨
    et al., Phys. Rev. D 79, 043009 (2009)
    [7] F. Aharonian et al. (HESS Collaboration), Astron. Astrophys. 464,
    235-243 (2007)
    [8] J. Dumm et al. (IceCube Collaboration), these proceedings.
    [9] M. Danninger, priv. Comm., see also R. Abbasi et al. (IceCube
    Collaboration), arXiv:0902.2460, accept. by Phys. Rev. Lett..
    [10] see e.g. Athar, PhysRev.D 71, 103008 (2005)
    [11] J. Bahcall,&,P. Meszaros, Phys. Rev. Lett. 85, 1362-1365 (2000).

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    1
    Fundamental Neutrino Measurements with IceCube DeepCore
    Darren Grant
    , D. Jason Koskinen
    , and Carsten Rott
    for the IceCube collaboration.
    Dept. of Physics, Pennsylvania State University, University Park, PA 16802, USA
    Dept. of Physics and Center for Cosmology and Astro-Particle Physics, Ohio State University, Columbus, OH 43210, USA
    1
    Abstract
    . The recent deployment of the first string
    2
    of DeepCore, a low-energy extension of the Ice-
    3
    Cube neutrino observatory, offers new opportunities
    4
    for fundamental neutrino physics using atmospheric
    5
    neutrinos. The energy reach of DeepCore, down to
    6
    ∼ 10
    GeV, will allow measurements of atmospheric
    7
    muon neutrino disappearance at a higher energy
    8
    regime than any past or current experiment. In
    9
    addition to a disappearance measurement, a flavor-
    10
    independent statistical analysis of cascade-like events
    11
    opens the door for the measurement of tau neutrino
    12
    appearance via a measurable excess of cascade-like
    13
    events. In the event of a relatively large value of
    14
    sin
    2
    13
    , a multi-year measurement of the suppres-
    15
    sion of muon neutrino disappearance due to earth
    16
    matter effects may show a measurable dependence on
    17
    the sign of the mass hierarchy (normal vs. inverted).
    18
    19
    Keywords
    : Oscillations, DeepCore, hierarchy
    20
    I. ICECUBE DEEPCORE
    21
    The IceCube neutrino telescope is a multipurpose
    22
    discovery detector under construction at the South Pole,
    23
    which is currently about three quarters completed [1].
    24
    After completion in 2011, IceCube will have instru-
    25
    mented a volume of approximately one cubic kilometer
    26
    utilizing 86 strings, each instrumented with 60 Digital
    27
    Optical Modules (DOMs) at a depth between 1450 m
    28
    and 2450 m. Eighty of these strings (the baseline design)
    29
    will be arranged in a hexagonal pattern with an inter-
    30
    string spacing of about 125 m and with 17 m verti-
    31
    cal separation between DOMs. This baseline design is
    32
    complemented by six more strings, that form a more
    33
    densely instrumented sub-array, located at the center of
    34
    IceCube. These strings will be spaced in between the
    35
    regular strings, so that an interstring-spacing of 72 m
    36
    is achieved. Together with the seven adjacent standard
    37
    IceCube strings these six strings form the DeepCore
    38
    array in the center of IceCube (shown in Figure 1).
    39
    DeepCore strings have a different distribution of the
    40
    60 DOMs on them. Fifty out of the 60 DOMs on a
    41
    DeepCore string will be installed in the deep clear ice,
    42
    below an ice-layer of short absorption length (1970 -
    43
    2100 m) also labeled ”dust-layer”. The top 10 DOMs on
    44
    the DeepCore strings will be deployed above this dust-
    45
    layer. Those DOMs add to the effective veto capability
    46
    of the surrounding IceCube strings against down-going
    47
    muons.
    Fig. 1: IceCube with the DeepCore sub-detector in the
    center deep clear ice. The illustration on the left shows
    the depth-profile of the optical transparency of the ice.
    48
    The DeepCore extension will significantly improve
    49
    IceCube’s low energy performance and allow neutrino
    50
    detection to approximately 10 GeV (see Figure 2). This
    51
    is accomplished by having DeepCore strings with a
    52
    dense vertical spacing of 7 m between DOMs, which are
    53
    deployed in the deepest ice where the scattering length is
    54
    approximately twice that compared to the upper part of
    55
    the IceCube detector [2]. Coupled with the spacing and
    56
    ice clarity the DOMs themselves are instrumented with
    57
    high quantum efficiency photomultiplier tubes (HQE
    58
    PMTs), that have a
    40%
    efficiency increase, wavelength
    59
    dependent, compared to regular IceCube PMTs [3]. The
    60
    aforementioned properties make the DeepCore an ideal
    61
    detector for low energy-low rate neutrino physics.

    2
    D. GRANT
    et al.
    FUNDAMENTAL NEUTRINO MEASUREMENTS
    (Primary Neutrino Energy - GeV)
    10
    log
    1
    1.2
    1.4
    1.6
    1.8
    2
    2.2
    2.4
    2.6
    2.8
    3
    )
    2
    Effective Neutrino Area (m
    -6
    10
    -5
    10
    -4
    10
    -3
    10
    -2
    10
    -1
    10
    1
    10
    2
    10
    Preliminary
    Fig. 2: Comparison of preliminary study of effective area
    A
    eff
    at trigger level for the 80 IceCube string array
    without DeepCore (squares) and in addition with the
    six DeepCore strings (open circles). The addition of
    DeepCore increases the effective area of the detector at
    low energies significantly.
    5
    10
    15
    20
    25
    30
    35
    40
    45
    50
    E
    ν
    [GeV]
    0
    0.2
    0.4
    0.6
    0.8
    1
    Oscillation probabilities
    ν
    μ
    − ν
    μ
    ν
    μ
    − ν
    τ
    Fig. 3:
    ν
    µ
    survival probability and
    ν
    µ
    →
    ν
    τ
    oscillation
    probability for vertically upward going neutrinos, where
    sin
    2
    2
    θ
    13
    =0.1 [6].
    62
    II. DEEPCORE NEUTRINO OSCILLATION PHYSICS
    63
    The lower energy reach achieved with the DeepCore
    64
    opens the possibility to investigate atmospheric neutrino
    65
    oscillations in the primarily unexplored energy regime
    66
    of a tens of GeV. In Figure 3 we show the expected
    67
    ν
    µ
    survival probability and
    ν
    µ
    →
    ν
    τ
    oscillation curves
    68
    for which the DeepCore will have sensitivity and re-
    69
    late directly to the potential measurements discussed in
    70
    the following subsections. In addition to the improved
    71
    energy reach, part of what make such measurements
    72
    possible is the innate background rejection built into
    73
    the DeepCore design: increased overburden reduces the
    74
    number of atmospheric muons and the surrounding
    75
    IceCube strings provide an in situ veto. Simple veto
    76
    methods have achieved background reductions of four
    77
    orders of magnitude with excellent signal retention and
    78
    have potential for greater than 6 orders of magnitude
    79
    rejection utilizing reconstruction veto methods [3].
    80
    A. Muon-Neutrino Disappearance
    81
    Previous measurements of neutrino oscillations at the
    82
    atmospheric-scale have been significantly decreased in
    83
    both energy reach and active volume size of the detectors
    84
    compared to DeepCore. With an approximate 13 MT
    85
    fiducial volume, DeepCore has the capacity to make a
    86
    precision measurement of atmospheric neutrino oscilla-
    87
    tions above 10 GeV [4]. The
    ν
    µ
    survival probability
    88
    curve shown in Figure 3 illustrates an expectation for
    89
    a significant deficit in neutrino flux, shown in Fig.
    90
    4 at a previously unexplored energy region. An issue
    91
    associated with such an analysis is the angular resolution
    92
    of neutrino induced muon tracks at these energies is
    93
    fundamentally limited by the kinematics of the neutrino-
    94
    nucleon interaction. Low energy
    ν
    µ
    interactions have
    95
    a much bigger opening angle between the incoming
    96
    neutrino and outgoing muon than high energy interac-
    97
    tions, which leads the muon to have a higher probability
    98
    of being noncollinear with the incoming neutrino. The
    99
    intrinsic uncertainty on the opening angle reduces any
    100
    experiments ability to identify perfectly upward going
    101
    neutrinos, where the uncertainty can be approximated,
    in the full data sample, by
    ∆φ ≃ 30
    ×
    ?
    102
    (GeV)/E
    ν
    µ
    .
    103
    However, oscillations can be observed with very high
    104
    significance with an inclusive measurement over the
    105
    zenith angle range
     1.0 < cos φ <  0.6
    , and in-
    106
    corporation of angular dependence will only improve
    107
    the result. Fig. 4 shows a simulation of this muon
    108
    disappearance effect, which would be an approximate 20
    109
    sigma effect with just one year of IceCube DeepCore
    110
    data. The current study only discusses the effect on
    111
    the signal, taking statistical uncertainties into account.
    112
    Systematic uncertainties remain to be studied, as well
    113
    as the background prediction to this measurement.
    114
    B. Tau-Neutrino Appearance
    115
    Returning to Figure 3 and looking at the
    ν
    µ
    →
    ν
    τ
    116
    oscillation curve, we expect a fraction of the incoming
    117
    atmospheric neutrino flux to have a
    ν
    τ
    component.
    118
    Given the higher parent
    ν
    µ
    flux and different decay
    119
    kinematics of tau events relative to that of
    ν
    e
    charge-
    120
    current (CC) and
    ν
    x
    neutral current (NC) (x=e,
    µ
    ,
    τ
    )
    121
    events, we should be able to detect
    ν
    τ
    both via the
    122
    excess of cascade (hadronic and electromagnetic shower)
    123
    events and possibly through the resulting spectral energy
    124
    distortion. This measurement would not only represent
    125
    the largest sample of tau neutrinos ever collected (albeit
    126
    inclusively), it may also be competitive with OPERA [5]
    127
    in making an appearance measurement of tau neutrinos
    128
    due to oscillations.
    129
    Future components of this analysis comprise devel-
    130
    oping a dedicated energy reconstruction for low energy
    131
    cascade events as well as examining the background
    132
    caused by short muon tracks that mimic cascades as well
    133
    as the impact of misidentification of up versus down
    134
    going neutrinos. Good energy resolution will increase
    135
    sensitivity to
    ν
    τ
    appearance over regions where neutrino
    136
    oscillations are maximal.

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    3
    NChannel
    0
    10
    20
    30
    40
    50
    60
    70
    80
    90
    100
    Events/NChannel/Year
    0
    200
    400
    600
    800
    1000
    1200
    1400
    1600
    1800
    2000
    Preliminary
    Fig. 4: Simulated
    ν
    µ
    disappearance with 1 year of
    DeepCore data. The lower curve (open circles) show the
    number of upward-going muons observed over the zenith
    angle cos
    φ <
    -0.6 with oscillations, while the upper
    curve (squares) is the corollary without oscillations. No
    systematic errors or background have been included.
    The effect is approximately 20 sigma based purely on
    statistical errors (approximately 28000 events per year
    without oscillations and 16000 events with oscillations
    assuming
    sin
    2
    2
    θ
    23
    =1.0 and
    m
    2
    23
    =0.0024). Note that
    NChannel is a crude energy estimator for the detector
    based on the number of hit DOMs.
    137
    C. Matter Effects and Neutrino Mass Hierarchy
    138
    Depending on detection efficiency and purity, a suf-
    139
    ficiently large value of
    sin
    2
    θ
    13
    [6], and control of
    140
    systematics the DeepCore detector may be used in an
    141
    ambitious multi-year effort to determine the sign of
    142
    the neutrino mass hierarchy. This may be accomplished
    143
    by measuring a small enhancement/supression from the
    144
    MSW effect [11] of the expected number of
    ν
    µ
    events.
    145
    The
    ν
    µ
    oscillation probability, shown in Figure 5, indi-
    146
    cates that the neutrino rate over the 8-25 GeV energy
    147
    region is enhanced for the normal hierarchy (NH),
    148
    and enhanced for anti-neutrinos for the inverted mass
    149
    hierarchy (IH). A complication of this measurement is
    150
    that DeepCore cannot distinguish between neutrino and
    151
    anti-neutrinos, however at the relevant energy range of
    152
    10 GeV
    < E
    ν
    <
    30 GeV the interation cross-section
    153
    between neutrinos and anti-neutrinos differs by a factor
    154
    of two;
    σ(ν
    x
    ) ≃ 2σ(ν¯
    x
    )
    . The difference in interaction
    155
    cross-section translates into a difference in the number
    156
    of observed muon neutrino candidate events. Based on
    157
    statistical discrimination only, it may be possible to
    158
    distinguish normal from inverted hierarchies. In Fig. 6
    159
    we show the expected results from 5 years of DeepCore
    160
    data, with a statistical separation between the normal
    161
    and inverted hierarchies of approximately 10 sigma. The
    162
    described effect is pending on a sufficiently large value
    163
    of
    θ
    13
    (that is expected to be measured by the time of the
    164
    data for this measurement has been obtained). Further
    165
    the signal systematic uncertainties need to be sufficiently
    166
    small and remaining background need to be effectively
    167
    removed.
    168
    III. CONCLUSIONS
    169
    We have completed a full Monte Carlo study of the
    170
    IceCube DeepCore detector that shows the potential
    171
    for measurements of fundamental neutrino properties.
    172
    We have discussed the expected effects on the signal
    173
    for
    ν
    µ
    disappearance at energies higher than previously
    174
    measured, a measurement of
    ν
    τ
    appearance as well as
    175
    resolution of the neutrino mass hierarchy.
    176
    REFERENCES
    177
    [1] A. Achterberg
    et al.
    , Astropart. Phys.
    26
    , 155 (2006).
    178
    [2] M. Ackermann
    et al.
    , J. Geophys. Res.
    111
    , 02201 (2006).
    179
    [3] D. Cowen, for the IceCube coll., NUTEL 09, forthcoming.
    180
    [4] Y. Fukuda
    et al.
    [Super-Kamiokande Collaboration], Phys. Rev.
    181
    Lett.
    81
    , 1562 (1998) [arXiv:hep-ex/9807003].
    182
    [5] R. Acquafredda
    et al.
    [OPERA Collaboration], New J. Phys.
    8
    ,
    183
    303 (2006) [arXiv:hep-ex/0611023].
    184
    [6] O. Mena, I. Mocioiu and S. Razzaque, Phys. Rev. D
    78
    , 093003
    185
    (2008)
    186
    [7] M. Sanchez [MINOS Collaboration], Moriond EW 2009, forth-
    187
    coming.
    188
    [8] H. L. Ge, C. Giunti and Q. Y. Liu, arXiv:0810.5443 [hep-ph].
    189
    [9] G. L. Fogli, E. Lisi, A. Marrone, A. Palazzo and A. M. Rotunno,
    190
    Phys. Rev. Lett.
    101
    , 141801 (2008) [arXiv:0806.2649 [hep-ph]].
    191
    [10] M. Apollonio
    et al.
    [CHOOZ Collaboration], Eur. Phys. J. C
    27
    ,
    192
    331 (2003) [arXiv:hep-ex/0301017].
    193
    [11] E. K. Akhmedov, M. Maltoni and A. Y. Smirnov, JHEP
    0705
    ,
    194
    077 (2007)
    195
    [12] C. Rott
    et al.
    for the IceCube coll., these proceedings.
    196
    [13] D. Grant
    et al.
    (DeepCore reconstruction) for the IceCube coll.,
    197
    these proceedings.
    198
    [14] C. Wiebusch
    et al.
    (DeepCore) for the IceCube coll., these
    199
    proceedings.
    200
    [15] R. Abbasi
    et al.
    [IceCube Collaboration], Nucl. Instrum. Meth.
    201
    A
    601
    , 294 (2009)
    202
    [16] E. Resconi for the IceCube coll., astro-ph/0807.3891.

    4
    D. GRANT
    et al.
    FUNDAMENTAL NEUTRINO MEASUREMENTS
    5
    10
    15
    20
    25
    E
    ν
    [GeV]
    0
    0.2
    0.4
    0.6
    0.8
    1
    P(
    ν
    μ
    − ν
    μ
    )
    sin
    2
    2
    θ
    13
    = 0.1 (ΝΗ)
    sin
    2
    13
    = 0.1 (ΙΗ)
    sin
    2
    13
    = 0.06 (ΝΗ)
    sin
    2
    13
    = 0.06 (ΙΗ)
    Fig. 5: Oscillation probabilities for
    ν
    µ
    →
    ν
    µ
    transitions for upward going neutrinos [6]. For a value of
    sin
    2
    13
    =0.1, the difference in the survival probability between the normal hierarchy (solid black line) and inverted
    hierarchy (dashed red line) is
    7%, and increases for higher values of
    sin
    2
    13
    . Recent measurements [7] as well
    as global fits [8], [9] prefer that
    sin
    2
    13
    is non-zero, while the value of 0.10 reflects the 90% Confidence Limit
    set by CHOOZ [10].
    Energy of detected muon - GeV
    5
    10
    15
    20
    25
    30
    35
    40
    45
    50
    Events/5 GeV
    0
    1000
    2000
    3000
    4000
    5000
    6000
    7000
    8000
    9000
    Preliminary
    Fig. 6: Predicted rate with 5 years of data for normal
    (squares) and inverted (open circles) hierarchy, for
    ν
    µ
    induced muon tracks within 45 degrees of vertical that
    start within the DeepCore fiducial volume. We find that
    in the first two bins the rate for the inverted hierarchy
    is above that for the normal hierarchy and that in the
    remaining bins the rates overlap. Systematic errors are
    not yet estimated. Statistical errors are too small to be
    visible. Note that
    sin
    2
    θ
    13
    =0.1 in the presented case.

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    1
    Implementation of an active veto against atmospheric muons in
    IceCube DeepCore
    Olaf Schulz
    , Sebastian Euler
    and Darren Grant
    for the IceCube Collaboration
    §
    Max-Planck Institut fu¨r Kernphysik, Saupfercheckweg 1, 69171 Heidelberg, Germany
    III. Physikalisches Institut, RWTH Aachen University, 52056 Aachen, Germany
    Dept. of Astronomy and Astrophysics, Pennsylvania State University, University Park, PA 16802, USA
    §
    See the special section of these proceedings
    Abstract
    . The IceCube DeepCore [1] has been
    designed to lower the energy threshold and broaden
    the physics capabilities of the IceCube Neutrino
    Observatory. A crucial part of the new opportunities
    provided by DeepCore is offered by the possibility to
    reject the background of atmospheric muons. This
    can be done by using the large instrumented volume
    of the standard IceCube configuration around Deep-
    Core as an active veto region. By thus restricting
    the expected signal to those neutrino events with an
    interaction vertex inside the central DeepCore region,
    it is possible to look for neutrinos from all directions,
    including the Southern Hemisphere that was previ-
    ously not accessible to IceCube. A reduction of the
    atmospheric muon background below the expected
    rate of neutrinos is provided by first vetoing events
    in DeepCore with causally related hits in the veto
    region. In a second step the potential starting vertex
    of a muon track is reconstructed and its credibility
    is estimated using a likelihood method. Events with
    vertex positions outside of DeepCore or with low
    starting probabilities are rejected. We present here
    these newly developed veto and vertex reconstruction
    techniques and present in detail their capabilities in
    background rejection and signal efficiency that have
    been obtained so far from full Monte Carlo studies.
    Keywords
    : high energy neutrino-astronomy, Ice-
    Cube, DeepCore
    I. INTRODUCTION
    The IceCube Neutrino Observatory [2] is currently
    being built at the geographic South Pole in Antarctica.
    After completion it will consist of
    4800 digital optical
    modules (DOMs) on 80 strings instrumenting one cubic
    kilometer of ice at a depth between 1450 m and 2450 m.
    Each DOM consists primarily of a photomultiplier tube
    and read-out electronics in a glass pressure vessel.
    IceCube is designed to detect highly energetic neutrino-
    induced muons as well as hadronic or electro-magnetic
    showers (cascades) that produce Cherenkov radiation
    in the medium. Significant backgrounds to the signal,
    caused by muons from atmospheric air showers above
    the detector, limit the field of view to the Northern
    hemisphere for many studies that use neutrino events in
    IceCube. In addition to its nominal layout, the DeepCore
    Fig. 1. Schematic view of the IceCube DeepCore
    extension to the observatory will lower the IceCube
    energy threshold from
    100 GeV down to neutrino
    energies as low as 10 GeV. This improvement in the
    detector energy response is achieved by including 6 extra
    strings, deployed in a denser spacing, around a standard
    central IceCube string. Each of these strings will be
    equipped with 60 DOMs, containing Hamamatsu high
    quantum efficiency photo multiplier tubes (HQE PMTs).
    50 of these DOMs will be placed in a dense spacing
    of
    7 m in the lowest part of the detector where the
    ice is clearest and scattering and absorption lengths are
    considerably longer [3]. The remaining 10 modules are
    to be placed in a 10 m spacing at a depth from 1760 m
    to 1850 m. This position has been chosen in order to
    improve IceCube’s capabilities to actively identify and
    reduce the atmospheric muon background to the central
    DeepCore volume, as described below. The HQE PMTs
    have a quantum efficiency that is up to 40% higher,
    depending on wavelength, compared to the standard
    IceCube PMTs, while their noise rate of
    380 Hz is

    2
    O. SCHULZ
    et al.
    DEEPCORE VETO IMPLEMENTATION
    on average increased by about 32%. Together with the 7
    neighboring IceCube strings DeepCore will consist of 13
    strings and be equipped with 440 optical modules instru-
    menting a volume of
    13 megatons water-equivalent.
    DeepCore will improve the IceCube sensitivity for
    many different astrophysics signals like the search for
    solar WIMP dark matter and for neutrinos from Gamma-
    Ray Bursts [4]. It also opens the possibility to investigate
    atmospheric neutrino oscillations in the energy range of
    a few tens of GeV [5]. An additional intriguing oppor-
    tunity offered by DeepCore is the possibility to identify
    neutrino signals from the southern hemisphere. Such a
    measurement requires a reduction of the atmospheric
    muon background by more than a factor
    10
    6
    in order to
    obtain a signal of atmospheric neutrinos to background
    rate better than one. A first step toward achieving this
    reduction is implicit in the design of DeepCore. Back-
    ground events which trigger DeepCore with a minimum
    number of hits in the DeepCore fiducial volume must
    pass through a larger overburden resulting in a order of
    magnitude decrease to the atmospheric muon rate, with
    respect to the whole IceCube detector. Two additional
    steps are then performed to attain the remaining 10
    5
    rejection factor. The first is a veto of DeepCore events
    with causally related hits in the surrounding IceCube
    volume, reducing the background rate by
    10
    2
    to
    10
    3
    .
    Then we apply a vertex reconstruction algorithm based
    on a maximum-likelihood method that determines the
    approximate neutrino interaction vertex. By rejecting
    events with a reconstructed vertex outside the central
    DeepCore volume a full 10
    6
    background reduction may
    be achieved.
    II. TRIGGERING DEEPCORE
    The first reduction of the atmospheric muon back-
    ground rate, with respect to IceCube, is achieved by
    applying a simple majority trigger (SMT) to the Deep-
    Core region, based on number of channels registering
    a hit in coincidence with a neighbor DOM. The trigger
    hit coincidence requirement, also known as hard local
    coincidence (HLC), is such that each channel is accom-
    panied by at least one more hit on one of the four closest
    neighboring modules within a time window of
    ±
    1000 ns.
    The rate of atmospheric muons triggering DeepCore is
    largely dependent on the multiplicity that is required. In
    this study we applied a trigger requirement of 6 HLC hits
    (SMT6), which translates to an average neutrino energy
    of approximately 10 GeV.
    Table I shows the approximate detector rates for the
    trigger level and after application of the veto algorithm.
    The final rates after application of the cuts on the
    reconstructed vertex are not given in the table, since they
    are analysis dependent and vary strongly with the cut
    strength. The background events, muons from cosmic
    ray air showers, are simulated using CORSIKA [6]. In
    this study only the most energetic muon is propagated
    through the full detector simulation. This is a conserva-
    tive approach since the other muons could only improve
    Fig. 2. Scheme of the veto principle: Illustration of the DeepCore
    hit center of gravity (COG), vertex time and particle speed per hit.
    the veto efficiency. The signal rate given here is the rate
    of neutrinos produced in atmospheric air showers and
    has been determined following the flux calculations of
    the Bartol group [7]. Neutrino oscillation effects have
    not been taken into account. Since the goal is to identify
    starting muon tracks specifically, the signal is restricted
    to those events with a simulated interaction vertex within
    the DeepCore volume. The efficiencies given in Table I
    relate to events that fulfill this requirement and have
    a DeepCore SMT6 trigger. The background rejection
    factors refer to the expected main IceCube trigger rate,
    build up from an IceCube only SMT8 trigger, a String
    Trigger which requires 5 out of 7 aligned modules on
    a string to be fired in a trigger window of 1500 ns, as
    well as the DeepCore SMT6 trigger itself. Applying the
    SMT6 trigger gives a background rate which exceeds
    the signal by a factor
    ∼ 10
    5
    . This sets the challenge for
    the performance of the veto algorithms to be applied.
    III. THE VETO ALGORITHM
    DeepCore is surrounded by more than 4500 DOMs
    that can be used as an active veto volume to reject
    atmospheric muons. If hits in the surrounding standard
    IceCube array are consistent with a particle moving
    downwards with v=c the event is rejected. For the veto
    algorithm, and also for any following reconstructions,
    it is essential to keep as many physics hits as possible.
    Therefore all hits in the detector are used here, including
    hits on DOMs without HLC (a mode called soft local
    coincidence, or SLC). To reduce the amount of dark
    noise hits, we reject any hits that are isolated from
    others by more than 150 m in distance or by more
    than 1000 ns in hit time. To determine whether or not
    to reject an event we initially compute the average hit
    PMT position and an approximate start time (vertex

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    3
    TABLE I
    BACKGROUND AND SIGNAL RATES AFTER DEEPCORE TRIGGER AND CAUSAL HIT VETO
    atm.
    µ
    (CORSIKA)
    rejection
    atm.
    ν
    µ
    (Bartol)
    eff.
    atm.
    ν
    µ
    upwards
    atm.
    ν
    µ
    down-
    wards
    main IceCube triggers
    2279 Hz
    -
    -
    -
    -
    -
    DeepCore event selection
    102 Hz
    4.5
    ·10
     2
    1.799
    ·10
     3
    Hz
    100%
    0.901
    ·10
     3
    Hz
    0.895
    ·10
     3
    Hz
    after Veto
    1.2 Hz
    5.4
    ·10
     4
    1.719
    ·10
     3
    Hz
    95.5%
    0.863
    ·10
     3
    Hz
    0.856
    ·10
     3
    Hz
    particle speed [m/ns]
    -1
    -0.8
    -0.6
    -0.4
    -0.2
    0
    0.2
    0.4
    0.6
    0.8
    1
    probability per event
    0
    0.1
    0.2
    0.3
    0.4
    0.5
    0.6
    corsika
    signal
    Fig. 3. Particle speed probabilities per event for atmospheric muons
    (dotted line) and muons induced by atmospheric neutrinos inside
    DeepCore (solid line).
    time) of the DeepCore fiducial volume hits (see Fig.
    2). The interaction position is determined by using the
    subset of those hit DOMs that have times within one
    standard deviation of the first guess vertex time. This has
    the benefit of reducing the contribution from PMT dark
    noise and, by weighting the DOMs by their individual
    charge deposition, a reasonable center of gravity (COG)
    for the event is computed. By making the assumption
    that roughly all light in the DeepCore volume originates
    from the COG a more thorough estimation of the vertex
    time is possible. For each individual hit the time light
    would have needed to travel from the COG to the hit
    module is calculated and subtracted from the original
    PMT hit time. The average of these corrected PMT hit
    times is then considered as the vertex time.
    Each hit in the veto region gets assigned a particle
    speed, defined as the spatial hit distance to the DeepCore
    COG divided by the time difference to the DeepCore
    vertex time. This speed is defined to be positive if the
    hit occurred before the vertex time and negative if it
    appeared after. Causally related hits in the veto region
    are generally expected to have a speed close to the
    speed of the muon, which is very close to the speed
    of light in vacuum (0.3 m/ns). Smaller speeds occur
    for hits that have been scattered and thus arrive late.
    Larger speeds are in principle acausal, but since the
    vertex time represents the start of a DeepCore event,
    whereas the COG defines its center, the particle speeds
    for early hits are slightly overestimated. Late hits on the
    other hand have typically lower speeds. Fig. 3 shows
    the probability of the occurrence of a particular particle
    speed per event. The dotted curve describes the simu-
    Fig. 4. Principle of the vertex reconstruction
    lated muon background from air-showers (CORSIKA)
    and the solid curve the atmospheric neutrino signal [7]
    with an interaction vertex inside DeepCore. The peak for
    the CORSIKA muons is slightly above +0.3 m/ns while
    muons induced by neutrinos in DeepCore mainly give
    hits with negative particle speeds. The peak at positive
    speeds close to zero is mainly due to early scattered
    light. By cutting out all events with more than one
    hit within a particle speed window between 0.25 and
    0.4 m/ns we achieve an overall background rejection on
    the order of
    5 · 10
     4
    (see Table I).
    IV. THE VERTEX RECONSTRUCTION
    To achieve the remaining background rejection, a
    second algorithm is used. It analyzes the pattern of
    hits in an event in conjunction with an input direction
    and position of a reconstructed track. From the track
    the algorithm estimates the neutrino interaction vertex
    and calculates a likelihood ratio which is used as a
    measurement for the degree of belief that the track is
    starting at the estimated position.
    As shown in Fig. 4, we trace back from each hit
    DOM to the reconstructed track using the Cherenkov
    angle of 41
    in ice. This projection is calculated for
    all DOMs within a cylindrical volume of radius 200 m
    around the track and the DOMs are ordered according
    to this position. (Note that 200 m is large enough to
    contain virtually all photons produced by the track.)
    The projection of the first hit DOM in the up-stream
    direction defines the neutrino interaction (reconstructed)
    vertex. A reconstructed vertex inside IceCube indicates
    a potential starting (neutrino-induced) track. Due to the
    large distance between neighboring strings, atmospheric
    muons may leak through the veto, producing their first
    hit deep inside the detector and thus mimicking the

    4
    O. SCHULZ
    et al.
    DEEPCORE VETO IMPLEMENTATION
    likelihood ratio
    -20 -18 -16 -14 -12 -10
    -8
    -6
    -4
    -2
    0
    # events [a.u.]
    -3
    10
    -2
    10
    10
    -1
    corsika
    signal
    reconstructed vertex r [m]
    0
    100 200 300 400 500 600 700 800 900 1000
    reconstructed vertex depth [m]
    -3000
    -2800
    -2600
    -2400
    -2200
    -2000
    -1800
    -1600
    -1400
    -1200
    -1000
    # events [a.u.]
    1
    10
    2
    10
    3
    10
    corsika
    reconstructed vertex r [m]
    0
    100 200 300 400 500 600 700 800 900 1000
    reconstructed vertex depth [m]
    -3000
    -2800
    -2600
    -2400
    -2200
    -2000
    -1800
    -1600
    -1400
    -1200
    -1000
    # events [a.u.]
    1
    10
    2
    10
    signal
    Fig. 5. Distributions of the cut parameters of the vertex reconstruction: Likelihood ratio (left) and position of the reconstructed vertex for
    atmospheric muons from CORSIKA (middle) and atmospheric neutrinos (right).
    signature of a starting track. Therefore it is necessary
    to quantify for each event the probability of actually
    starting at the reconstructed vertex. To determine this
    starting likelihood, one first selects all DOMs without
    a hit and with a projection on the assumed track up-
    stream of the first hit DOM. The probability that each
    of these DOMs did not receive a hit is calculated
    assuming two track hypotheses: a track starting at the
    reconstructed vertex and a track starting outside the
    detector volume. Under the assumption of an external
    track
    p(noHit|
    Track) is calculated. Here, for each DOM
    the probability of not being hit (in spite of the passing
    track) depends on track parameters (energy of the light
    emitting particle, position and direction of the track) and
    ice properties. The probability is calculated from the
    expected number of photoelectrons, taken from
    Photorec
    tables of the
    Photonics
    project [8], assuming Poisson
    statistics:
    p
    λ
    (noHit) = p
    λ
    (0) =
    λ
    0
    0!
    e
     λ
    = e
     λ
    .
    (1)
    λ
    is the expected number of photoelectrons. Under
    the assumption of a starting track
    p(noHit|
    noTrack) is
    calculated, which is equal to the probability of a noise
    hit and can therefore be calculated from measured noise
    rates.
    The likelihood for the observed pattern of hit DOMs
    may now be constructed as the product of the individual
    hit probabilities. A track is classified as starting in the
    detector according to the probability given by the ratio of
    the likelihoods. For a clearly starting track this ratio is a
    negative number, and the larger the value the higher the
    starting probability for the track. To select tracks starting
    inside the detector, cuts are applied on the position of
    the reconstructed vertex and on the likelihood ratio. The
    distributions of the cut parameters are shown in Fig. 5.
    Preliminary studies are up to now utilizing the true
    simulated track, since dedicated low energy track re-
    constructions are still under development. Even though
    idealized, these studies strongly indicate that an overall
    background rejection of
    > 10
    6
    can be achieved without
    having to extend the vertex cuts into the densely instru-
    mented DeepCore fiducial region and with keeping the
    majority of the signal events.
    V. SUMMARY AND OUTLOOK
    We have presented the methods developed thus far to
    reduce the rate of background muon events within the
    IceCube DeepCore detector. Utilizing the instrumented
    standard IceCube volume around DeepCore as an active
    veto to identify and reject atmospheric muon events
    improves the possibility of detecting neutrino induced
    muons and cascades independent of direction. The rate
    of atmospheric muons is mainly reduced in a two step
    process. First, a veto algorithm is applied against Deep-
    Core events with causally related hits in the surrounding
    IceCube region. Second, applied to the veto surviving
    events, a cut has been defined, using a likelihood ratio,
    to determine the probability that the event had a starting
    vertex within the fiducial region of the detector. Monte
    Carlo studies indicate that both methods together will
    be suitable to reduce the background muon rate by
    more than the factor of
    10
    6
    needed to obtain a signal
    (atmospheric neutrinos) to background ratio of
    > 1
    .
    IceCube and DeepCore are currently under construction
    and will be finished in 2011. The fully deployed Deep-
    Core detector will provide an effective volume of several
    megatons of water equivalent for neutrino events with an
    energy above 10 GeV and a starting vertex in DeepCore.
    The exact volume will depend on the required signal-to-
    noise ratio and the individual analysis strategies.
    REFERENCES
    [1] E. Resconi, et al., IceCube Collaboration, Proceedings to
    VLVnT08, (2008).
    [2] A. Achterberg, et al., IceCube Collaboration, Astroparticle
    Physics 26, 155 (2006).
    [3] M. Ackermann, et al., IceCube Collaboration, Journal of Geo-
    physical Research, Vol. 111, D13203 (2006).
    [4] J. N. Bahcall and P. Meszaros, Physics Review Letters, Vol. 85,
    p. 1362, (2000).
    [5] E. K. Akhmedov, M. Maltoni and A. Y. Smirnov, Journal of High
    Energy Physics 5, 77 (2007).
    [6] D. Heck et al, Report FZKA 6019 (1998).
    [7] G. D. Barr, et al., Physical Review D 70, 023006 (2004).
    [8] J. Lundberg et al., Nuclear Instruments and Methods, Vol. A581,
    pp. 619-631, (2007).

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    1
    Acoustic detection of high energy neutrinos in ice: Status and
    results from the South Pole Acoustic Test Setup
    Freija Descamps
    for the IceCube Collaboration
    Department of Subatomic and Radiation Physics, University of Ghent, 9000 Ghent, Belgium,
    See the special section of these proceedings.
    Abstract
    . The feasibility and specific design of an
    acoustic neutrino detection array at the South Pole
    depend on the acoustic properties of the ice. The
    South Pole Acoustic Test Setup (SPATS) has been
    built to evaluate the acoustic characteristics of the
    ice in the 1 to 100 kHz frequency range. The most
    recent results of SPATS are presented.
    Keywords
    : SPATS, acoustic neutrino detection,
    acoustic ice properties
    I. INTRODUCTION
    The predicted ultra-high energy (UHE) neutrino fluxes
    from both hadronic processes in cosmic sources and
    interaction of high energy cosmic rays with the cosmic
    microwave background radiation are very low. There-
    fore extremely large detector volumes, on the order of
    100 km
    3
    or more, are needed to detect a significant
    amount of these neutrinos. The idea of a large hybrid
    optical-radio-acoustic neutrino detector has been de-
    scribed and simulated in [1], [2]. The density of detectors
    in such a possible future UHE neutrino telescope is
    dictated by the signal to noise behaviour with travelled
    distance of the produced signals. The ice has optical
    attenuation lengths on the order of 100 m, enabling
    the construction and operation of the IceCube [3] de-
    tector. The relatively short optical attenuation length
    makes building a detector much larger than IceCube pro-
    hibitively expensive. In contrast, the attenuation lengths
    of both radio and acoustic waves are expected to be
    larger [4], [5] and may make construction of a larger
    detector feasible. The South Pole Acoustic Test Setup
    (SPATS) has been built and deployed to evaluate the
    acoustic attenuation length, background noise level, tran-
    sient rates and sound speed in the South Pole ice-
    cap in the 1 to 100 kHz region so that the feasibility
    and specific design of an acoustic neutrino detection
    array could be assessed. Acoustic waves are bent to-
    ward regions of lower propagation speed and the sound
    speed vertical profile dictates the refraction index and
    the resulting radius of curvature. An ultra-high energy
    neutrino interaction produces an acoustic emission disk
    that will be deformed more for larger sound speed
    gradients and so the direction reconstruction will be
    more difficult. A good event vertex reconstruction allows
    accurate rejection of transient background. The absolute
    level and spectral shape of the continuous background
    noise determine the threshold at which neutrino induced
    signals can be extracted from the background and thus
    set the lower energy threshold for a given detector
    configuration. Transient acoustic noise sources can be
    misidentified as possible neutrino candidates, therefore
    a study of transient sources and signal properties needs
    to be performed. The acoustic signal undergoes a geo-
    metric 1/r attenuation and on top of that scattering and
    absorption. The overall acoustic attenuation influences
    detector design and hence cost.
    II. INSTRUMENTATION
    A. The SPATS array
    The South Pole Acoustic Test Setup consists of four
    vertical strings that were deployed in the upper 500 me-
    ters of selected IceCube holes [3] to form a trapezoidal
    array, with inter-string distances from 125 to 543 m.
    Each string has 7 acoustic stages. Figure 1 shows a
    schematic of the SPATS array and its in-ice and on-
    ice components. It also shows a schematic drawing
    of an acoustic stage comprised of a transmitter and
    sensor module. The transmitter module consists of a
    steel pressure vessel that houses a high-voltage pulse
    generator board and a temperature or pressure sensor.
    Triggered HV pulses are sent to the transmitter, a ring-
    shaped piezo-ceramic element that is cast in epoxy for
    electrical insulation and positioned
    ∼13
    cm below the
    steel housing. Azimuthally isotropic emission is the
    motivation for the use of ring shaped piezo-ceramics.
    The actual emission directivity of such an element was
    measured in azimuthal and polar directions [6]. The
    sensor module has three channels, each 120
    apart in
    azimuth, to ensure good angular coverage.
    B. The retrievable pinger
    A retrievable pinger was deployed in 10 water-filled
    IceCube holes down to a depth of 500 m: 6 holes were
    pinged in December 2007-January 2008 and 4 more
    holes were pinged using an improved pinger design in
    December 2008-January 2009. The emitter
    1
    is a broad
    band omni-directional transmitter which has a transmis-
    sion power of 149 dB re (
    µ
    Pa/V) at 1 meter distance.
    Upon receiving a trigger signal, a short (∼
    50µs
    ) HV
    pulse is sent to the piezo-ceramic transmitter that is
    suspended about 2 m below the housing. A broadband
    acoustic pulse is then emitted. Both the SPATS array
    and the pinger are GPS synchronized so that the arrival
    times of the pinger-pulses can be determined. Data was
    1
    ITC-1001 from the International Transducer Company

    2
    FREIJA DESCAMPS
    et al.
    SPATS STATUS AND RESULTS
    Fig. 1. Schematic of the SPATS detector.
    collected for all SPATS-sensor levels at 80, 100, 140,
    190, 250, 320, 400, 430 and 500 m depth and the
    pinger was pulsed at repetition rates of 1, 8 or 10 Hz.
    At the SPATS instrumented depths, the pinger lowering
    was stopped for five minutes so that a signal could
    be recorded with every SPATS sensor at that pinger
    position. For the December 2008-January 2009 holes,
    no stop was made at 80, 100, 140 and 430 m depth.
    III. RESULTS
    A. Sound speed
    The sound speed analysis uses data from December
    2007-January 2008 geometries where the pinger and
    sensor were at the same depth and 125 m apart. The
    pinger emitter was situated in a column of water where
    no shear waves can propagate, nevertheless shear waves
    were generated at the water-ice boundary where mode
    conversion was expected. Therefore transit times can
    be extracted from the data for both pressure and shear
    waves for many instrumented SPATS levels. There is
    agreement with [7] for the pressure wave measurement
    in the not fully compacted ice of the upper region (firn).
    The extracted shear wave speed is about half of the
    pressure wave speed, as expected [8]. A linear fit was
    made to the data in the deep and fully compacted ice
    between 250 and 500 m depth. We find following results
    for the pressure and shear wave sound speeds (
    v
    p
    and
    v
    s
    ) and their variation with depth (gradient:
    g
    p
    and
    g
    s
    ):
    v
    p
    (375m) = (3878 ± 12)m/s,
    g
    p
    = (0.09 ± 0.13)(m/s)/m,
    v
    s
    (375m) = (1975.8 ± 8.0)m/s,
    g
    s
    = (0.067 ± 0.806)(m/s)/m.
    The gradient for both pressure and shear waves is
    consistent with zero. Both sound speed measurements
    are performed with a better than 1% precision, taking
    into account the errors on the horizontal distance, pinger
    and sensor depths, emission and arrival times. For more
    details on the SPATS sound speed analysis, see [9].
    B. Gaussian noise floor
    The noise is monitored in SPATS through a forced
    200 kHz read-out of all sensor channels for 0.5 s
    every hour. The distribution of ADC counts in each
    of the operational channels is Gaussian and stable dur-
    ing the present observation time of more than 1 year
    with a typical deviation of the mean noise level of
    σ
    RM S
    <RMS>
    < 10
     2
    . The SPATS sensors on strings
    A, B and C have been calibrated in liquid water at
    0
    C in the 10 to 80 kHz frequency range prior to
    deployment [6]. Lab measurements have shown [6],
    [10] that the sensitivity of the SPATS sensor in air at
    atmospheric pressure increases by a factor of 1.5
    ±
    0.2
    when cooled down from 0
    C to -50
    C. We use this
    value to estimate the sensitivity of the SPATS sensors
    after they were deployed in the cold Antarctic ice. A

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    3
    measurement of the sensitivity at room temperature as a
    function of static pressure was performed in a pressure
    vessel and the results indicate a change in sensitivity of
    < 30%
    in the 1 to 100 bar region [10]. The noise level
    and fluctuations are high for all 4 strings in the firn
    region, where the transition from a snow/air mixture to
    compact bulk ice takes place. This is consistent with the
    effect of a large sound speed gradient in that layer so
    that the surface noise is refracted back to the surface. In
    the fully compacted ice, below the firn, noise conditions
    are more stable and we can derive an average noise
    level below 10mPa in the relevant frequency range (10
    to 50kHz). For more details on the SPATS noise floor
    analysis, see [11].
    C. Transient events
    The SPATS detector has been operated in transient
    mode for 45 minutes of every hour since August 2008.
    If the number of ADC counts on any of the twelve
    monitored channels exceeds a certain level above noise,
    we record a 5ms window of data around the trigger on
    that channel. The resulting trigger rate is stable and on
    the order of a few triggers every minute for each of
    the twelve monitored channels. Most of these events
    are Gaussian noise events, where only one sample is
    outside the trigger boundaries. The transient events are
    processed off-line and analysed for time-coincidence
    clustering. Figure 2 shows the spatial distribution of a
    total of 4235 reconstructed transient events as detected
    by SPATS between 1 September 2008 and 23 April
    2009 (4422 hours integrated lifetime). The data shows
    clear and steady sources or ’hot spots’ that can be
    correlated with the refreezing process of the water in
    sub-surface caverns. Each IceCube drilling season, a
    Fig. 2.
    An overview of the spatial distribution of transient events
    as detected by SPATS between September 2008 and April 2009. The
    circles indicate the positions of the IceCube holes, the SPATS strings
    are indicated by their corresponding letter-ID. The “Rodriguez well”
    are represented by squares. The smearing-effect is an artefact of the
    event reconstruction algorithm due to not accounting for refraction in
    the firn.
    “Rodriguez wel” (RW) is used as a water reservoir.
    The main source of transients is the 07-08 RW. There
    is also steady detection of the 05-06/04-05 RW. The
    06/07 RW was a steady source until October 2008.
    The 08-09 RW has not yet been detected. Transient
    data-taking continued during the IceCube 08-09 drilling
    season and the refreezing of 12 holes nearest to the
    SPATS array are audible whereas 7 of the farthest are
    not. No vertices were reconstructed deeper than 400 m.
    D. Attenuation length
    We have performed 3 classes of attenuation analyses,
    all of which use the sensors on the frozen-in SPATS
    strings but with a different sound source: the retrievable
    pinger, the frozen-in SPATS transmitters and transient
    events. Table I gives an overview of the SPATS atten-
    uation length studies with their respective uncertainties.
    distance [m]
    250
    300
    350
    400
    450
    500
    550
    600
    ln (P R) a.u.
    11.0
    11.2
    11.4
    11.6
    11.8
    12.0
    12.2
    12.4
    12.6
    12.8
    λ
    = 280 +/− 13 m
    preliminary
    Fig. 3. An example of a (ln(amplitude
    ·
    distance) vs. distance)-fit for
    a single channel at 400 m depth.
    1) Pinger attenuation length:
    The retrievable pinger
    was recorded simultaneously by all SPATS sensors for
    each of the 4 IceCube holes in which it was deployed
    during the December 2008-January 2009 period. The
    pinger holes were almost perfectly aligned relative to
    the SPATS array, making the single-channel analysis
    independent of polar and azimuthal sensitivity variation.
    The energy of the signal, which dominates in the fre-
    quency range from 5 to 35 kHz, was extracted both
    from the waveforms in the time domain (4 pinger holes)
    and from the power spectrum (3 pinger holes). Figure 3
    shows an example of a single-level fit for the energy in
    the frequency domain (single channel at 400 m depth).
    For both analyses, the quoted value in Table I is the
    mean of all fits with the standard deviation as error.
    The result for the time domain analysis uses all levels
    as shown in Fig. 4 and points which have large error
    bar (
    λ
    σ
    λ
    < 3)
    have been excluded (4/47 combinations).
    There is no evidence for a depth-dependence of the
    attenuation length.
    2) Inter-string attenuation length:
    The complete
    SPATS inter-string set of April 2009 consists of all
    possible transmitter-sensor combinations excluding the
    shallowest transmitters. The transmitters were fired at
    25Hz repetition rate and for each combination a total
    of 500 pulses was recorded simultaneously by the three
    channels of the sensor module. An averaged waveform

    4
    FREIJA DESCAMPS
    et al.
    SPATS STATUS AND RESULTS
    TABLE I
    SPATS ATTENUATION LENGTH STUDIES.
    Attenuation analysis
    λ
    (m)
    uncertainty
    comment on uncertainty
    Pinger energy time domain
    320
    60m
    standard deviation of distribution
    Pinger energy frequency domain
    270
    90m
    standard deviation of distribution
    Interstring energy all levels
    320
    100m
    standard error of weighted mean of distribution
    Interstring energy 3-level ratio
    193
    -
    best fit
    Transient event analysis
    200
    -
    best fit
    0
    100
    200
    300
    400
    500
    600
    700
    800
    900
    1000
    module
    C190
    C250
    C320
    C400
    B190
    B250
    B320
    B400
    A190
    A250
    A320
    A400
    D190
    D250
    D320
    D400
    D500
    attenuation length [m]
    Preliminary
    Fig. 4. Pinger energy analysis result from time domain for all recorded
    levels.
    can therefore be obtained and the pressure wave energies
    extracted. An averaged noise-waveform allows us to
    subtract the noise contribution. Every transmitter in
    SPATS can be detected by 3 sensors at the same depth
    and different distances (strings), this means that only
    the unknown sensor sensitivities have to be included
    into the systematic error for a single-level attenuation
    length fit. We have fitted all data of the transmitters that
    are detected at three different distances. Each single fit
    does not constrain the attenuation length very well due
    to the large sensor-to-sensor variations in sensitivity. By
    combining all fits, we find a weighted mean and the
    standard deviation of the weighted mean as value for
    the attenuation length and its error (Table I). A way
    to work around the missing calibration of the sensors
    and transmitters is to build ratios of amplitudes using
    two transmitter-sensor pairs. For isotropic sensors and
    transmitters, one such measurement should yield the at-
    tenuation length. To minimize the remaining variation of
    the sensitivity due to the varying polar angle, we limited
    the amplitude ratios to the levels at 250, 320 and 400 m
    depth. This ratio-method yields an attenuation length
    with a large systematic error due to unknown azimuthal
    and remaining polar sensor orientations, therefore only
    the best fit is quoted in Table I.
    3) Transient attenuation length:
    If a transient source
    can be localized, its corresponding signals can be used
    to estimate the acoustic attenuation length. The strongest
    transient source (07/08 RW) is excluded since it is too
    loud so that some channels are saturated. Moreover, all
    four of the SPATS strings are at similar distances from
    07/08 RW. The transient events from the refreezing of
    hole 37 offer the advantage of mostly being inside the
    dynamical range of all sensors. In addition, hole 37 was
    pinged and all sensors recorded signals simultaneously
    from the pinger at a depth of 320 m. This pinger
    data can be used to extract relative sensitivity-correction
    factors for that precise direction for all sensor channels
    at 320 m depth on all 4 SPATS strings. The best fit of
    the sensitivity corrected single level transient data yields
    an attenuation length of 200 m.
    IV. CONCLUSIONS
    We have presented the most recent results from the
    SPATS setup. Both pressure and shear wave speeds have
    been mapped versus depth in firn and bulk ice. This is
    the first measurement of the pressure wave speed below
    190 m depth in the bulk ice, and the first measurement
    of shear wave speed in the South Pole ice. The resulting
    vertical sound speed gradient for both pressure and
    shear waves is consistent with no refraction between
    250 and 500 m depth. Extrapolating sensitivities from
    laboratory calibrations gives a first estimation of the
    absolute noise level at depths larger than 200 m and
    indicates values below 10 mPa integrated over the 10 to
    50 kHz frequency range. The in-ice transient rates are
    low, most of the transient events can be correlated with
    well-know anthropogenic sources. We have presented an
    overview of the different SPATS attenuation length stud-
    ies, the preliminary results show that the analyses favour
    attenuation lengths in the 200-350 m region. The current
    dataset does not allow a distinction to be made between
    absorption- or scatter-dominated attenuation length and
    dedicated measurements are under consideration.
    V. ACKNOWLEDGMENT
    We are grateful for the support of the U.S. National
    Science Foundation and the hospitality of the NSF
    Amundsen-Scott South Pole Station.
    REFERENCES
    [1] D. Besson
    et al.
    , Int. J. Mod. Phys. A
    21S1
    (2006) 259.
    [2] D. Besson
    et al.
    , Nucl. Instr. and Meth. A (2009),
    doi:10.1016/j.nima.2009.03.047.
    [3] A. Achterberg
    et al.
    , Astropart. Phys.
    26
    (2009) 155.
    [4] S. Barwick
    et al.
    , J. Glaciol.
    51
    (2005) 231.
    [5] P.B. Price, J. Geophys.Res.
    111
    (2006) B02201.
    [6] J.-H. Fisher, Diploma thesis, Humboldt-Universitat zu Berlin
    (2006).
    [7] J. G. Weihaupt, Geophys. 28
    (4)
    (1963) 582.
    [8] D.G. Albert, Geophys. Res. Let.
    25 No 23
    (1998) 4257.
    [9] F. Descamps
    et al.
    ,
    Nucl. Instr. and Meth.
    A(2009),
    doi:10.1016/j.nima.2009.03.062.
    [10] T. Karg
    et al.
    , these proceedings.
    [11] T. Karg
    et al.
    , Nucl. Instr. and Meth. A (2009),
    doi:10.1016/j.nima.2009.03.063.

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    1
    Sensor development and calibration for acoustic
    neutrino detection in ice
    Timo Karg
    , Martin Bissok
    , Karim Laihem
    , Benjamin Semburg
    , and Delia Tosi
    for the IceCube collaboration
    §
    Dept. of Physics, University of Wuppertal, D-42119 Wuppertal, Germany
    III Physikalisches Institut, RWTH Aachen University, D-52056 Aachen, Germany
    DESY, D-15735 Zeuthen, Germany
    §
    See the special section of these proceedings.
    Abstract
    . A promising approach to measure the
    expected low flux of cosmic neutrinos at the highest
    energies (E
    >
    1 EeV) is acoustic detection. There
    are different in-situ test installations worldwide in
    water and ice to measure the acoustic properties
    of the medium with regard to the feasibility of
    acoustic neutrino detection. The parameters of inter-
    est include attenuation length, sound speed profile,
    background noise level and transient backgrounds.
    The South Pole Acoustic Test Setup (SPATS) has
    been deployed in the upper 500 m of drill holes for
    the IceCube neutrino observatory at the geographic
    South Pole. In-situ calibration of sensors under the
    combined influence of low temperature, high ambient
    pressure, and ice-sensor acoustic coupling is difficult.
    We discuss laboratory calibrations in water and ice.
    Two new laboratory facilities, the Aachen Acoustic
    Laboratory (AAL) and the Wuppertal Water Tank
    Test Facility, have been set up. They offer large
    volumes of bubble free ice (3 m
    3
    ) and water (11 m
    3
    )
    for the development, testing, and calibration of acous-
    tic sensors. Furthermore, these facilities allow for
    verification of the thermoacoustic model of sound
    generation through energy deposition in the ice by a
    pulsed laser. Results from laboratory measurements
    to disentangle the effects of the different environmen-
    tal influences and to test the thermoacoustic model
    are presented.
    Keywords
    : acoustic neutrino detection, thermo-
    acoustic model, sensor calibration
    I. INTRODUCTION
    The detection and spectroscopy of extra-terrestrial
    ultra high energy neutrinos would allow us to gain new
    insights in the fields of astroparticle and particle physics.
    Apart from the possibility to study particle acceleration
    in cosmic sources, the measurement of the guaranteed
    flux of cosmogenic neutrinos [1] opens a new window
    to study cosmic source evolution and particle physics
    at unprecedented center of mass energies. However, the
    fluxes predicted for those neutrinos are very low [2], so
    detectors with large target masses are required for their
    detection. One possibility to instrument volumes of ice
    of the order of
    100km
    3
    with a reasonable number of
    sensor channels is to detect the acoustic signal emitted
    from the particle cascade at a neutrino interaction vertex
    [3].
    To study the properties of Antarctic ice relevant for
    acoustic neutrino detection the South Pole Acoustic Test
    Setup (SPATS) [4] has been frozen into the upper part of
    IceCube [5] boreholes. SPATS consists of four vertical
    strings reaching a depth of 500 m below the surface.
    The horizontal distances between strings cover the range
    from 125 m to 543 m. Each string is instrumented with
    seven acoustic sensors and seven transmitters. The ice
    parameters to be measured are the sound speed profile,
    the acoustic attenuation length, the background noise
    level, and transient noise events in the frequency range
    from 1 kHz to 100 kHz.
    For the design of a large scale acoustic neutrino
    detector it is crucial to fully understand the
    in-situ
    response of the sensors as well as the thermoacoustic
    sound generation mechanism.
    II. SENSOR CALIBRATION
    To study the acoustic properties of the Antarctic
    ice, like the absolute background noise level, and to
    deduce the arrival direction and energy of a neutrino
    in a future acoustic neutrino telescope it is essential to
    measure the sensitivity and directionality of the sensors
    used, i.e. the output voltage as function of the incident
    pressure, and its variation with the arrival direction of
    the incident acoustic wave relative to the sensor. These
    measurements can be carried out relatively easily in the
    laboratory in liquid water. The two calibration methods
    most commonly used are
    the comparison method, where an acoustic signal
    sent by a transmitter (with negligible angular vari-
    ation) is simultaneously recorded at equal distance
    with a pre-calibrated receiver and the sensor to be
    calibrated. A comparison of the signal amplitudes
    in the two receivers allows for the derivation of the
    desired sensitivity from the sensitivity of the pre-
    calibrated sensor.
    the reciprocity method, which makes use of the
    electroacoustic reciprocity principle to determine
    the sensitivity of an acoustic receiver without hav-
    ing to use a pre-calibrated receiver (see e.g. [7]).

    2
    KARG
    et al.
    ACOUSTIC SENSOR CALIBRATION
    All SPATS sensors have been calibrated in
    0
    C water
    with the comparison method [8]. However, both calibra-
    tion methods are not suitable for in-situ calibration of
    sensors in South Pole ice. There are no pre-calibrated
    sensors for ice available, and reciprocity calibration
    requires large setups which are not feasible for de-
    ployment in IceCube boreholes. Further, directionality
    studies require a change of relative positioning between
    emitter and receiver which is difficult to achieve in a
    frozen-in setup.
    It is not clear how results obtained in the laboratory
    in liquid water can be transferred to an in-situ situation
    where the sensors are frozen into Antarctic ice. We
    are studying the influence of the following three envi-
    ronmental parameters on the sensitivity separately: low
    temperature, increased ambient pressure, and different
    acoustic coupling to the sensor. We will assume that sen-
    sitivity variations due to these factors obtained separately
    can be combined to a total sensitivity change for frozen
    in transmitters. This assumption can then be checked
    further using the two different sensor types deployed
    with the SPATS setup. Apart from the standard SPATS
    sensors with steel housing two HADES type sensors
    [9] have been deployed with the fourth SPATS string.
    These contain a piezoceramic sensor cast in resin and
    are believed to have different systematics.
    A. Low temperatures
    The ice temperature in the upper few hundred meters
    of South Pole ice is
     50
    C [10]. It is not feasible to
    produce laboratory ice at this temperature in a large
    enough volume to carry out calibration studies. We
    study the dependence of the sensitivity on temperature
    in air. A signal sent by an emitter is recorded with a
    sensor at different temperatures. To prevent changes in
    the emissivity of the transmitter, the transmitter is kept
    at constant temperature outside the freezer, and only
    the sensor is cooled down. The recorded peak-to-peak
    amplitude is used as a measure of sensitivity. First results
    indicate a linear increase of sensitivity with decreasing
    temperature (cf. Fig. 1). The sensitivity of a SPATS
    sensor is increased by a factor of
    1.5 ± 0.2
    when the
    temperature is lowered from
    0
    C to
     50
    C (averaged
    over all three sensor channels).
    B. Static pressure
    Acoustic sensors in deep polar ice are exposed to
    increased ambient static pressure. During deployment
    this pressure is exerted by the water column in the bore-
    hole (max. 50 bar at 500 m depth). During re-freezing it
    increases since the hole freezes from the top, developing
    a confined water volume. The pressure is believed to
    decrease slowly as strain in the hole ice equilibrates to
    the bulk ice volume. The final static pressure on the
    sensor is unknown.
    A
    40.5
    cm inner diameter pressure vessel is available
    at Uppsala university that allows for studies of sensor
    −70
    −60
    −50
    −40
    −30
    −20
    −10
    0
    10
    1
    1.5
    2
    2.5
    3
    3.5
    4
    4.5
    5
    preliminary
    temperature [C]
    amplitude peak to peak [V]
    SignalVsTemp−Ch0
    fit: y=a x+b
    a = −0.021 +/ 0.001 T/
    o
    C
    b = 2.32 +/ 0.03
    Fig. 1.
    Measured peak-to-peak amplitude of one SPATS sensor
    channel in air at different temperatures and linear fit to the data.
    0
    20
    40
    60
    80
    100
    0
    1
    2
    3
    4
    5
    6
    preliminary
    pressure [bar]
    amplitude Vpp
    Feb09−POS3−20kHz−4V−Pscan−allCh
    0
    1
    2
    Fig. 2. Measured peak-to-peak amplitude of a SPATS sensor excited
    by a transmitter coupled from the outside to the pressure vessel. All
    three channels of the sensor are shown.
    sensitivity as function of ambient pressure. Static pres-
    sures between 0 and 800 bar can be reached. In this
    study the pressure is increased up to 100 bar. Acoustic
    emitters for calibration purposes can be placed inside
    the vessel or, free of pressure, outside of it. Cable feeds
    allow one to operate up to two sensors or transmitters
    inside the vessel. A sensor is placed in the center of
    the water filled vessel. The transmitter is coupled from
    the outside to the vessel. The recorded peak-to-peak
    amplitude is used as a measure of sensitivity while the
    pressure is increased. The sensor sensitivity is measured
    by transmitting single cycle gated sine wave signals with
    different central frequencies from 5 kHz to 100 kHz.
    Figure 2 shows the received signal amplitudes for the
    three sensor channels of a SPATS sensor as a function
    of ambient pressure. No systematic variation of the
    sensitivity with ambient pressure is observed. Combin-

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    3
    ing all available data we conclude that the variation
    of sensitivity with static pressure is less than 30% for
    pressures below 100 bar.
    C. Sensor-ice acoustic coupling
    The acoustic coupling, i.e. the fractions of signal en-
    ergy transmitted and reflected at the interface of medium
    and sensor, differs significantly between water and ice.
    It can be determined using the characteristic acoustic
    impedance of the medium and sensor, which is the
    product of density and sound velocity and is equivalent
    to the index of refraction in optics. Due to the different
    sound speeds the characteristic acoustic impedance of
    ice is about
    2.5
    times higher than in water.
    Its influence will be studied in the Aachen Acoustic
    Laboratory (Sec. III), where it will be possible to carry
    out reciprocal sensor calibrations in both water and ice,
    and also to use laser induced thermoacoustic signals as
    a calibrated sound source.
    III. NEW LABORATORY FACILITIES
    Two new laboratories have been made available to the
    IceCube Acoustic Neutrino Detection working group for
    signal generation studies and sensor development and
    calibration.
    a) Wuppertal Water Tank Test Facility:
    For rapid
    prototyping of sensors and calibration studies in water,
    the Wuppertal Water Tank Test Facility offers a cylin-
    drical water tank with a diameter of
    2.5
    m and depth
    of
    2.3
    m (
    11m
    3
    ). The tank is built up from stacked
    concrete rings and has a walkable platform on top. It
    is equipped with a positioning system for sensors and
    transmitters and a 16-channel PC based DAQ system
    (National Instruments USB-6251 BNC).
    The size of the water volume allows for the clean sep-
    aration of emitted acoustic signals and their reflections
    from the walls and surface. This makes it possible to
    install triangular reciprocity calibration setups with side
    lengths of up to 1 m. Further, installations to measure the
    polar and azimuthal sensitivity of sensors are possible.
    b) Aachen Acoustic Laboratory:
    The Aachen
    Acoustic Laboratory is dedicated to the study of thermo-
    acoustic sound generation in ice. A schematic overview
    of the setup can be seen in Fig. 3. The main part is a
    commercial cooling container (
    6 × 2.5 × 2.5 m
    3
    ), which
    can reach temperatures down to
     25
    C. An IceTop
    tank, an open cylindrical plastic tank with a diameter
    of 190 cm and a height of 100 cm [6], is located inside
    the container. The IceTop tank has a freeze control unit
    by means of which the production of bubble-free ice is
    possible. The freeze control unit mainly consists of a
    cylindrical semipermeable membrane at the bottom of
    the container, which is connected to a vacuum reservoir
    and a pressure regulation system. The membrane allows
    for degassing of the water. A total volume of
    ≈ 3 m
    3
    of
    bubble-free ice can be produced. A full freezing cycle
    takes approximately sixty days with the freezing going
    from top to bottom.
    Fig. 3. Overview of the AAL setup with zoom on the mirror holder
    (top), the sensor positioning system (bottom, left) and a sensor (bottom,
    right).
    On top of the container, a Nd:YAG Laser is installed in
    a light-tight box with an interlock connected to the laser
    control unit. The laser has a pulse repetition rate of up to
    20 Hz and a peak energy per pulse of 55 mJ at 1064 nm,
    30 mJ at 532 nm, and 7 mJ at 355 nm wavelength. The
    laser beam is guided into the container and deposited in
    variable positions on the ice surface by a set of mirrors
    with coatings for the above mentioned frequencies. The
    optical feed-through consists of a tilted quartz window
    to avoid damage of the laser cavity by reflected laser
    light. For the detection of thermoacoustic signals, 18
    sensors are mounted on a sensor positioning system.
    The positioning system has three levels, on each level 6
    sensors are placed in a hexagonal geometry. Along with
    the sensors, 18 sound emitters are deployed for calibra-
    tion and test purposes. The sensors will be calibrated
    reciprocally. The positioning system will also include
    a reciprocal calibration setup for HADES sensors and
    the ability to install a SPATS sensor for calibration pur-
    poses. In addition, two temperature sensors are deployed
    at each level. The acoustic sensors are pre-amplified
    low-cost piezo based ultrasound sensors, usually used
    for distance measurement. The sensors show a strong
    variation of signal strength with incident angle. This
    directionality has to be studied but is rather useful for
    the suppression of reflected signals. The sensors are read
    out continuously by a LabVIEW-based DAQ framework
    with a NI PCIe-6259M DAQ card. The framework
    includes a temperature and acoustic noise monitoring

    4
    KARG
    et al.
    ACOUSTIC SENSOR CALIBRATION
    system.
    IV. STUDIES OF THERMOACOUSTIC SIGNAL
    GENERATION
    A detailed understanding of thermoacoustic sound
    generation in ice is crucial for designing an acoustic
    extension to the IceCube detector. The dependence of the
    signal strength on the deposited energy as well as on the
    distance to the sensor is of great interest. Also the pulse
    shape and the frequency content have to be studied sys-
    tematically with respect to various cascade parameters.
    The spatial distribution of the acoustic signal has to be
    investigated, i.e. the acoustic disk and its dependence on
    the spatial and temporal energy deposition distribution.
    In addition, the AAL setup will be able to study the
    thermoacoustic effect in a wide temperature range from
    20
    C to
     25
    C and possible differences of the effect
    in ice and water.
    In the Aachen Acoustic Laboratory, the thermo-
    acoustic signal is generated by a Nd:YAG laser. A laser-
    induced thermoacoustic signal differs from a signal pro-
    duced in a hadronic cascade. While a cascade’s energy
    deposition profile can be described by a Gaisser-Hillas
    function, the laser intensity drops of exponentially. Also
    the lateral profile of a cascade follows a NKG function,
    where, assuming a TEM
    00
    mode in the far field region,
    the typical laser-beam profile is Gaussian. Knowing this,
    a recalculation of the signal properties from a laser-
    induced pulse to a cascade-generated pulse is possible.
    The frequency content of the signal is expected to vary
    with the beam diameter, while a too short penetration
    depth will result in an acoustic point source rather than
    a line source. The absorption coefficient of light in
    water or ice varies strongly with wavelength. The first
    wavelength of the laser (1064 nm) is absorbed after few
    centimeters, while the second harmonic (532 nm) has
    an absorption length of
    ≈ 60
    m. The third harmonic at
    355 nm with an absorption length of
    ≈ 1
    m is expected
    to be the most suitable wavelength to emulate a hadronic
    cascade with a typical length of 10 m and a diameter of
    10 cm. The diameter of the heated ice volume has to be
    controlled by optics inside the container that will widen
    the beam.
    With an array of 18 sensors, the Aachen Acoustic Lab-
    oratory allows the study of the spatial distribution of the
    generated sound field, as well as the frequency content
    with varying beam parameters. The first thermoacoustic
    signal has been generated and detected in a test setup
    with preliminary sensor electronics in order to determine
    a reasonable gain for the pre-amplifier while avoiding
    saturation. A small volume of bubble-free ice has been
    produced, containing a sensor and an emitter. Laser
    pulses (wavelength 1064 nm, 55 mJ per pulse) have been
    shot at the ice block. A zoom on the first waveform is
    presented in Fig. 4. The distance between laser spot and
    sensor is approximately
    15 cm
    and the sensor gain factor
    is 22. A Fourier transform of the pulse implies a pulse
    central frequency of
    ≈ 100 kHz
    , which is expected for
    t [ms]
    0
    0.5
    1
    1.5
    2
    2.5
    3
    3.5
    4
    U [V]
    −0.1
    −0.05
    0
    0.05
    0.1
    Fig. 4. Thermoacoustic pulse in ice, generated with a laser at 1064 nm
    wavelength and a beam diameter of
    ≈ 1 µm
    .
    such a small beam diameter. In order to see the expected
    bipolar pulse, further studies have to be performed to
    determine the transfer function of the sensors.
    V. CONCLUSIONS
    Detailed understanding of the thermoacoustic sound
    generation mechanism and the response of acoustic sen-
    sors in Antarctic ice is necessary to design an acoustic
    extension for the IceCube neutrino telescope. While in-
    situ calibrations in deep South Pole ice are inherently
    difficult, different environmental influences can be stud-
    ied separately in the laboratory. No change in sensor
    response with increasing ambient pressure was found; a
    linear increase in sensitivity with decreasing temperature
    was observed. Intense pulsed laser beams can be used
    to generate thermoacoustic signals in ice which can also
    be used as an in-ice calibration source.
    ACKNOWLEDGMENTS
    We are grateful for the support of the U.S. National
    Science Foundation and the hospitality of the NSF
    Amundsen-Scott South Pole Station.
    This work was supported by the German Ministry for
    Education and Research.
    REFERENCES
    [1] V. S. Berezinsky and G. T. Zatsepin, Phys. Lett. B
    28
    (1969)
    423.
    [2] R. Engel, D. Seckel and T. Stanev, Phys. Rev. D
    64
    (2001)
    093010,
    arXiv:astro-ph/0101216
    .
    [3] G. A. Askaryan, At. Energ.
    3
    (1957) 152.
    [4] J. Vandenbroucke et al., Nucl. Instr. and Meth. A (2009),
    doi:10.1016/j.nima.2009.03.064,
    arXiv:0811.1087
    [astro-ph]
    .
    [5] http://icecube.wisc.edu/
    [6] T. K. Gaisser, in Proceedings of the 29th International Cosmic
    Ray Conference (ICRC 2005),
    arXiv:astro-ph/0509330
    .
    [7] R. J. Urick,
    Principles of underwater sound
    , 3rd ed., Peninsula
    Publishing (1983).
    [8] J.-H. Fischer, Diploma Thesis, Humboldt-Universita¨t zu Berlin
    (2006).
    [9] B. Semburg et al., Nucl. Instr. and Meth. A (2009),
    doi:10.1016/j.nima.2009.03.069,
    arXiv:0811.1114
    [astro-ph]
    .
    [10] P. B. Price et al., Proc. Nat. Acad. Sci. U.S.A.
    99
    (2002) 7844.

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    1
    A new method for identifying neutrino events in IceCube data
    Dmitry Chirkin
    University of Wisconsin, Madison, U.S.A.
    Abstract
    . A novel approach for selecting high-
    quality muon neutrino events in IceCube data is
    presented. The rate of air shower events mis-
    reconstructed as signal is first reduced via the use of
    the geometrical (software) trigger. The final event se-
    lection is performed with a machine-learning method,
    designed specifically for IceCube data. It takes into
    account some generic properties of IceCube events,
    e.g., the fact that separation of signal from back-
    ground is more difficult (requiring tighter cuts on
    the quality parameters) for horizontal rather than
    vertically up-going tracks. The method compares
    favorably to other techniques in situations with both
    high and low simulation statistics.
    Keywords
    : neutrino search, machine learning, event
    selection
    I. INTRODUCTION
    An important task of a neutrino telescope like IceCube
    is identifying extra-terrestrial neutrinos that are inter-
    spersed between orders of magnitude higher background
    of particles originating in the showers produced by
    cosmic rays in the Earth’s atmosphere.
    As a first step a high purity atmospheric (plus possible
    extraterrestrial) neutrino event sample is selected, with
    only a small contaminating fraction of mis-reconstructed
    atmospheric muon events. Only neutrinos can cross the
    overburden of the Earth in the upward direction; how-
    ever, selecting events reconstructed as upward-moving
    leaves many mis-reconstructed atmospheric muons in the
    sample, improving the ratio of neutrino to contaminating
    muon events (initially at
    ∼ 10
     6
    ) by only a factor of
    ∼ 100
    .
    The problem is further exacerbated by a highly uneven
    contamination of the mis-reconstructed muons in several
    of the analysis variables, most importantly the zenith
    angle. This contamination is smaller for up-going direc-
    tions and increases for more horizontal tracks, growing
    rapidly near and above the horizon. It is therefore
    difficult to arrive at an event selection method that
    provides optimized cut surfaces simultaneously for all
    zenith angles. Splitting the cut optimization in different
    zenith bins leads to fluctuations of the cut parameters
    from one zenith angle bin to the next that are perceived
    as unphysical. In situations with limited simulated data,
    splitting it in several zenith angle bins is undesirable.
    This author has also performed an SVM-based event
    selection optimization and found that training the SVM
    gets more difficult for zenith angle ranges extending
    above the horizon.
    The above considerations led to the development of
    a new framework for selecting and applying cuts on
    quality parameters, that in the following is called “Subset
    Browsing Method”, or
    SBM
    for short.
    The quality parameters used with the event selection
    method of this paper build upon those discussed previ-
    ously [1].
    II. SIMPLE EXAMPLE
    First consider a simple example employing only two
    parameters that select events with lower background
    contamination for lower value of the quality parameter.
    These can be, e.g., zenith angle (0 degrees for up-
    going to 90 for horizontal tracks) and estimated angular
    resolution (e.g., describing the half-width of the like-
    lihood function at the minimum corresponding to the
    reconstructed track direction). Both of these can be used
    to remove the background of mis-reconstructed events,
    one through the basic reconstruction property, and the
    other though our prior knowledge that the contamination
    is higher for tracks near the horizon.
    The toy simulated events are divided into two groups
    (randomly): the training set that is used for the training
    of the machine, and the testing set, that is used to
    judge its performance. Both sets, while drawn from the
    same parent distribution, are statistically independent.
    The toy “data” events are also simulated and drawn from
    a somewhat wider signal distribution (to demonstrate the
    effect of cuts in the transitional region between signal
    and background). The 3 steps of the machine application
    are the steps
    1
    a
    ,
    1
    b
    , and 2 as shown on Figure 1.
    The training of the machine in this simple example
    is achieved by identifying the “outlying” background
    events (on the signal side of the distribution), and
    creating the “angle cuts” (shown with black straight
    lines) by cutting away everything on the rejected sides
    of such cuts (i.e., everything above and to the right of
    the background event, including that background event
    itself).
    The cuts so identified will obviously remove all back-
    ground events in the training dataset. As seen from the
    second row of Figure 1, these cuts do not remove all
    of the background events when applied to the testing
    dataset, so a further step, here called
    1
    b
    is necessary.
    Using the angle cuts derived in step
    1
    a
    a quality param-
    eter (
    SBM*
    ) is constructed, which is simply the count
    of “angle cuts” of step
    1
    a
    that fall into the bad quadrant
    (up and to the right) of a tested event, see Figure 2.
    This quality parameter could also be constructed as a

    2
    CHIRKIN
    et al.
    A METHOD FOR EVENT SELECTION IN ICECUBE
    training
    testing
    step 1
    a
    data
    step 1
    b
    step 2
    Fig. 1.
    Two event populations are shown: red is signal for the
    training and simulated testing event sets, and all data in the data set.
    Blue points (located up and to the right from red points) describe
    background events. Lower values on both x and y mean better quality.
    In steps
    1
    a
    and
    1
    b
    empty circles show background events removed by
    the
    skeleton cuts
    , aqua and pink points show background and signal
    (or data) events respectively that are removed by the machine quality
    parameter cut set at 1.5. In the third column events removed by step
    2 are shown as empty circles and events additionally removed by the
    quality parameter cut are shown in pink, same as before. The black
    lines show the
    skeleton cuts
    and go through the out-most background
    events of the training set.
    weighted sum, as described in the following section,
    shown for comparison as
    SBM
    in Figure 4.
    A map of the quality parameter (
    SBM*
    ) is shown
    in Figure 3. It is clear that through application of the
    quality parameter some space is inserted between the
    cut structure achieved in step
    1
    a
    and events with quality
    parameter greater than 0. Figure 4 shows that a value of
    SBM*
    =2.5 completely separates signal from background
    in the testing dataset of this simple example.
    III. UNSIMULATED EVENTS AND STEP 2
    After the application of steps
    1
    a
    and
    1
    b
    to the real
    IceCube data it became obvious that, although much of
    the background events like those present in our sim-
    ulation was removed, some “unsimulated” background
    remained in data. This affected agreement in parameter
    distributions and was particularly evident upon visual
    inspection. One class of such events appeared to contain
    two or more coincident but independent muon hit pat-
    terns that happened near each other with much overlap
    in time and failed to be split by the topological trigger.
    With more simulation we would most likely have been
    able to correctly identify these events.
    Another class of events appeared to contain a bright
    electromagnetic or hadron shower (
    cascade
    ), with rate of
    occurrence higher than that predicted by the simulation.
    While these events, when understood, could be very
    interesting, making a valuable contribution to the final
    event selection, it is unclear at this point whether they
    should be classified as signal or background. Moreover,
    since they are not simulated as either, the detector
    0
    1
    2
    3
    4
    5
    6
    7
    basic cut variable
    quality cut parameter
    -8
    -6
    -4
    -2
    0
    2
    4
    6
    8
    -8
    -6
    -4
    -2
    0
    2
    4
    6
    8
    Fig. 2. Shown are events remaining in the testing dataset after the
    application of step
    1
    a
    , signal in red, and background in blue. The
    legend indicates the value of the
    SBM*
    quality parameter for the shown
    events. For each value of the
    SBM*
    quality parameter the bad sides
    of a representative event are shown with straight lines. The
    SBM*
    quality parameter is simply the count of the “outlying” background
    events of the training dataset that gave rise to the “angle cuts” of step
    1
    a
    (indicated with black squares and black solid lines).
    effective area to these events cannot be estimated, so
    they do not contribute to any of the physics results.
    A common way to deal with this is by raising the
    quality parameter of an analysis past the point where all
    simulated background is removed and to the point where
    an agreement between data and signal simulation is
    achieved. This is possible if the quality parameter judges
    not only how far a given event is from the background
    region, but also how close it is to the signal region.
    In the method presented here the quality parameter
    is constructed using only the information about the
    outlying background events that form the “angle cuts”
    (after some initial amount of signal events is carved
    out by the step
    1
    a
    ). Thus, the quality parameter judges
    only the distance to the background region (contrary to
    other approaches). To achieve the “similarity with signal
    events” another step (called step 2) becomes necessary.
    This event selection step is achieved by removing all
    data events, to the bad sides of which there are no signal
    events of the training set (remaining after the application
    of steps
    1
    a
    and
    1
    b
    ). The effect of step 2 is demonstrated
    in the last column of Figure 1: all data events on the
    background side of the signal region are removed.
    IV. MULTI-DIMENSIONAL GENERALIZATION
    First we re-iterate that the SBM method relies on the
    important observation that most of the quality parame-
    ters used in the analysis of IceCube data have the follow-
    ing property: as the fits become less constrained at lower
    number of channels
    N
    ch
    or strings
    N
    str
    (that received
    hits), the cuts on the quality parameters necessary to
    reach a given signal purity need to be tightened. Alter-
    natively, the cuts applied to quality parameters of events
    with higher
    N
    ch
    or
    N
    str
    can be relaxed somewhat. A

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    3
    0
    1
    2
    3
    4
    5
    6
    7
    basic cut variable
    quality cut parameter
    -8
    -6
    -4
    -2
    0
    2
    4
    6
    8
    -8
    -6
    -4
    -2
    0
    2
    4
    6
    8
    Fig. 3. Map of the quality parameter (
    SBM*
    ) calculated according
    to prescription of Figure 2. Highest-quality region is shown in red.
    quality parameter SBM
    counts
    quality parameter SBM*
    counts
    1
    10
    10
    2
    -1
    0
    1
    2
    3
    4
    5
    6
    1
    10
    10
    2
    -1
    0
    1
    2
    3
    4
    5
    6
    7
    8
    Fig. 4. Quality parameters: simple sum over the “angle cuts” (
    SBM*
    ),
    and weighted sum (
    SBM
    ). Red solid and blue dotted lines show the
    distribution of signal and background events, respectively.
    similar behavior of cuts on quality parameters can be
    argued for their dependence on the reconstructed zenith
    angle
    θ
    : at angles closer to the horizon the number of
    background events seeping through is higher than for
    tracks going up closer to the vertical, so to reach the
    same purity the cuts on the quality parameters need to
    be tighter for events with higher reconstructed zenith
    angle
    θ
    . To summarize, we introduce the following
    Basic cut property
    : the cuts necessary to reach the
    same signal purity satisfy the following conditions:
    c(θ
    , N
    ch
    , N
    str
    ) ≤ c(θ
    0
    , N
    ch
    , N
    str
    )
    for
    θ
    ≥ θ
    0
    c(θ, N
    ch
    , N
    str
    ) ≤ c(θ, N
    0
    ch
    , N
    str
    )
    for
    N
    ch
    ≤ N
    0
    ch
    c(θ, N
    ch
    , N
    str
    ) ≤ c(θ, N
    ch
    , N
    0
    str
    )
    for
    N
    str
    ≤ N
    0
    str
    This relies on the
    assumption
    that lower cut values
    imply tighter cuts
    1
    . Parameters
    θ
    ,
    N
    ch
    , and
    N
    str
    that
    allow such a behavior of cuts are in the following
    1
    Some of the quality parameters may need to be taken with a minus
    sign or as one over their value to satisfy this assumption
    called
    basic cut variables
    . The following discussion is
    simplified with a
    Definition
    : a cut
    c
    defined for a set of events with
    θ
    ,
    N
    ch
    , and
    N
    str
    is said to be operating on a
    subset
    of events of a cut
    c
    0
    defined for a set of events with
    θ
    0
    ,
    N
    0
    ch
    , and
    N
    0
    str
    if
    θ
    ≥ θ
    0
    ,
    N
    ch
    ≤ N
    0
    ch
    , and
    N
    str
    N
    0
    str
    .
    Main cut property
    : a cut operating on a given set of
    events also operates on all its subsets.
    To rephrase, a cut
    c
    0
    defined for a set of events with
    θ
    0
    ,
    N
    0
    ch
    , and
    N
    0
    str
    also applies to any set of events with
    θ
    ,
    N
    ch
    , and
    N
    str
    (that has its own cut
    c
    ), if
    θ
    ≥ θ
    0
    ,
    N
    ch
    ≤ N
    0
    ch
    , and
    N
    str
    ≤ N
    0
    str
    . To prove we need to
    show that
    c
    0
    ≥ c
    . Using 2 intermediate sets of events,
    and the basic cut property introduced above
    c
    0
    = c(θ
    0
    , N
    0
    ch
    , N
    0
    str
    ) ≥ c(θ
    , N
    0
    ch
    , N
    0
    str
    ) ≥
    c(θ
    , N
    ch
    , N
    0
    str
    ) ≥ c(θ
    , N
    ch
    , N
    str
    ) = c
    This property allows us to consider all cuts as op-
    erating not only on events with
    θ = θ
    0
    ,
    N
    ch
    = N
    0
    ch
    ,
    and
    N
    str
    = N
    0
    str
    , but rather on all events with
    θ ≥ θ
    0
    ,
    N
    ch
    ≤ N
    0
    ch
    , and
    N
    str
    ≤ N
    0
    str
    .
    Definition
    : The cut
    c
    0
    associated with
    θ
    0
    ,
    N
    0
    ch
    , and
    N
    0
    str
    is considered
    redundant
    if there exists another cut
    c
    associated with some other
    θ
    ,
    N
    ch
    , and
    N
    such
    that
    c
    ≤ c
    0
    ,
    θ
    ≤ θ
    0
    ,
    N
    ch
    ≥ N
    0
    ch
    , and
    N
    str
    ≥ N
    0
    str
    .
    This is because the new cut
    c
    clearly implies
    c
    0
    by
    the main cut property.
    For each background event
    i
    b
    in the simulated training
    dataset its quality parameters are used to create
    n
    cuts
    associated with
    θ
    i
    b
    ,
    N
    i
    b
    ch
    , and
    N
    i
    b
    str
    of the event. The
    signal purity
    p
    i
    b
    = s
    i
    b
    /(s
    i
    b
    + b
    i
    b
    )
    of the events with
    θ ≥ θ
    i
    b
    ,
    N
    ch
    ≤ N
    i
    b
    ch
    , and
    N
    str
    ≤ N
    i
    b
    str
    is then calculated
    and used to find the
    i
    b
    that defines cuts in a region
    with the worst purity. Out of the
    n
    cuts associated
    with
    i
    b
    the cut that results in a smallest loss of signal
    events is then chosen and applied to the whole subset
    on which this cut operates. To accelerate this process if
    a cut in encountered that removes no signal events it is
    immediately used without taking into account the purity
    of the subset of events on which this cut operates.
    This procedure is then repeated until the background
    events in the simulated training dataset are exhausted. At
    that point all cuts of all background events are cycled
    through once again, and those that result in no further
    loss of remaining signal events (which are said to form
    a
    core
    or signal events) are saved into the
    trained cut
    set
    of the machine. One can further
    reduce
    this set by
    removing the
    redundant
    cuts from it, thus resulting in
    an
    irreducible trained cut set
    , which is the result of this
    machine training procedure.
    One can obviously remove all background events
    (i.e., reach a 100% signal purity) by applying all cuts
    from the
    irreducible trained cut set
    to the simulated
    training dataset. However, when the same is applied
    to the separately generated
    testing
    dataset a number of

    4
    CHIRKIN
    et al.
    A METHOD FOR EVENT SELECTION IN ICECUBE
    background events seep through and the signal purity
    never reaches 100%.
    This may happen, e.g., if we encounter a background
    event with, say,
    N
    ch
    higher than
    N
    i
    b
    ch
    of every back-
    ground event in the simulated training dataset, thus there
    are no cuts available that would remove such an event
    from the testing dataset.
    A way around this is to find at least one cut
    c
    with
    θ
    c
    ≥ θ
    ,
    N
    c
    ch
    ≤ N
    ch
    , and
    N
    c
    str
    ≤ N
    str
    such that
    c
    = q
    ≤ c
    (
    q
    being the quality parameter of the
    tested event). By the
    main cut property
    the cut that would
    achieve the same purity
    p
    c
    on the subset defined by
    θ
    ,
    N
    ch
    , and
    N
    str
    as the cut
    c
    on the subset defined by
    θ
    c
    ,
    N
    c
    ch
    , and
    N
    c
    str
    is necessarily no more tight as the cut
    c
    .
    That is, applying the cut
    c
    on the subset defined by
    θ
    ,
    N
    ch
    , and
    N
    str
    achieves at least the same or higher level
    of purity as
    p
    c
    . Now, if an event defined by its quality
    parameters
    c
    = q
    but at the
    basic cut variables
    θ
    c
    ,
    N
    c
    ch
    , and
    N
    c
    str
    of the cut
    c
    is passed by the
    trained cut
    set
    , cut
    c
    is called the
    purity cut
    defined for the original
    event (defined by its own
    c
    ,
    θ
    ,
    N
    ch
    , and
    N
    str
    ).
    Existence of at least one such cut
    c
    for each of the
    testing dataset events passed by the machine guarantees
    that the purity in the regions with extrapolated
    θ
    ,
    N
    ch
    ,
    and
    N
    str
    is at least as good or better than in the regions
    for which background events existed in the simulated
    training dataset. This is an important advantage of the
    discussed method compared to the other machine learn-
    ing techniques.
    Counting all
    purity cuts
    available for a given testing
    dataset event provides one with an important machine
    quality parameter which value is higher for events that
    are more likely to be signal and lower for events that are
    more likely to be background. It appears that a cut on
    this parameter improves the purity in all subsets by equal
    amount. In order to improve the purity in all subsets
    to the same final value one may weight the terms in
    the quality parameter sum with the initial purity of the
    simulated training dataset on the subsets of the cuts used
    in the sum (thus leading to the weighted sum definition
    of SBM, as shown in Figure 4).
    We call the machine learning method described here
    the
    subset browsing method
    because of the technique in
    which one has to
    browse
    through the
    subsets
    on which
    the cuts of the
    trained cut set
    are defined to calculate
    the quality parameter separating signal from background.
    The quality parameter itself is called the
    SBM quality
    parameter
    :
    S B M
    .
    The
    irreducible trained cut set
    forms a “skeleton” of
    cuts that are applied to the testing dataset achieving the
    initial SBM cut level
    :
    SBM = 0
    . The
    SBM
    quality
    parameter is usually normalized so that the highest value
    of
    SBM
    of a background event in a simulated testing
    dataset is 1 (e.g., in the plot of the
    SBM
    in Figure 5).
    V. CONCLUSIONS AND OUTLOOK
    We present a new framework for selecting and apply-
    ing cuts on the quality parameters, here called SBM. It is
    neutrino signal
    corsika single
    coincident showers
    data, SBM step 1
    data, SBM step 2
    log
    10
    (SBM)
    entries
    10
    -1
    1
    10
    -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1
    Fig. 5. Distribution of the quality parameter after step
    1
    a
    (for 1 year
    of IC-22 data). The
    >90%
    estimated purity is achieved at SBM=0.36.
    particularly well-suited for application to IceCube data
    analysis as it takes into account (both during training
    and for event classification) some obvious relationships
    between the quality parameters, which are hardwired
    into the algorithm.
    The SBM appears to separate signal and background
    well even if the number of simulated training dataset
    events is low. The SBM extrapolation behavior is very
    robust and the performance of extrapolation improves
    as the simulated testing dataset statistics increases, since
    more
    purity cuts
    become available to testing events in
    regions less populated by cuts of the
    trained cut set
    .
    The machine will not go around individual background
    events of the training dataset (a condition that may occur
    in other methods) because the cuts, by construction, are
    monotonous functions of the basic parameters.
    Additionally, the learning method itself is very simple
    and has virtually no parameters to set. Most of the
    machine is implemented with only an application of the
    operator (and a simple summation for estimating the
    quality parameter).
    The method splits the classification of the data events
    into two steps: dissimilarity with background, and sim-
    ilarity with signal, thus allowing one to investigate
    possible “unsimulated” classes of events.
    This method was used to identify atmospheric neu-
    trino events in IceCube data taken during the 2007
    operation season (see Figure 5 and reference [2]).
    As an outlook, this method may be well-suited to
    analyzes that depend on a veto region around the in-
    teresting events in the detector, as most of the cuts on
    the quality parameters can be relaxed as the veto region
    is expanded, thus satisfying the basic cut property if veto
    size is chosen as a basic cut variable.
    The method could be further improved in the future
    by implementing a technique similar to boosting of the
    BDT.
    REFERENCES
    [1] D. Chirkin, et al.,
    Effect of the improved data acquisition system
    of IceCube on its neutrino-detection capabilities
    , 30th ICRC,
    Merida, Mexico (arXiv:0711.0353)
    [2] D. Chirkin, et al.,
    Measurement of the atmospheric neutrino
    energy spectrum with IceCube
    , these proceedings

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    1
    Muon Production of Hadronic Particle Showers in Ice and Water
    Sebastian Panknin
    , Julien Bolmont
    , Marek Kowalski
    and Stephan Zimmer
    Humboldt-Universita¨t zu Berlin, Germany
    LPNHE, Universite Pierre et Marie Curie Paris 6, Universite Denis Diderot Paris 7, CNRS/IN2P3, France
    Humboldt-Universita¨t zu Berlin, Germany, now at Uppsala Universitet, Sweden
    Abstract
    . One of the neutrino signatures in
    Cerenkov neutrino detectors are isolated, particle
    showers induced by neutrinos of all flavors. Hadronic
    showers can produce muons during the shower
    development and the appearance of the showers can
    change significantly by such high-energy muons. We
    use a modified version of the air shower simula-
    tion program CORSIKA for the simulation of the
    generation of muons in salt water. We discuss how
    the results can be applied for ice. In addition, a
    simple analytical model is derived, that provides
    scaling relations for the muon energy spectrum and
    its dependence on the primary particle.
    Keywords
    : shower-simulation, muons, neutrino-
    detection
    I. INTRODUCTION
    Large neutrino telescopes in ice or water, like
    IceCube[1], Baikal[2] and ANTARES[3], are in opera-
    tion or construction. They detect Cerenkov light from
    particles created by neutrino interactions. The most
    prominent signature are up-going muons generated by
    charge-current interactions of muon neutrinos. Another
    possibility is the search for neutrino induced cascades
    [4]. Such a search is sensitive to electron and tau neu-
    trinos. In addition, unlike a search for neutrino induced
    muons, one is not restricted to up-going tracks, since
    the signature of isolated cascades allows in principle
    good separation from the down-going atmospheric muon
    background. A requirement for a search of neutrino
    induced events is their accurate simulation.
    In hadronic cascades also muons can be produced.
    Due to their track-like light signature they may influence
    the shape and thus should be simulated and parameter-
    ized.
    We develop a simple analytical calculation for the
    produced muon flux (section II). We show the results of
    a full shower simulation with a modified version of the
    well-known air-shower simulation program CORSIKA
    [5] (section III) and use this to parameterize the muon
    flux (section IV). We conclude in section V.
    II. ANALYTICAL MODEL FOR THE MUON FLUX
    The basic properties of electro-magnetic showers can
    be understood through the Heitler model [6]. After one
    interaction length a photon creates an electron-positron
    pair, and an electron/positron radiates a bremsstrahlung
    photon. This repeats every interaction length while the
    energy is distributed over the generated particles, so that
    more and more particles with lower and lower energies
    are created. Hadronic showers are more complicated,
    since in each interaction a wide range of hadronic
    particles can be created, which through decay increase
    the complexity further. However, expanding the simple
    Heitler model helps to get a basic understanding of the
    muon generation in hadronic cascades.
    We follow the extended Heitler model of [7] and
    consider a hadronic shower generated by a particle of
    primary energy
    E
    0
    . In each interaction
    N
    mul
    ≈ 10
    hadrons are produced. About one third of them will
    be neutral hadrons like
    π
    0
    , which will lead to an
    electromagnetic sub-shower.
    So in the generation
    n
    we have
    N
    H
    (n)
    hadrons with
    average energy
    E
    H
    :
    N
    H
    (n) =
    ?
    2
    3
    N
    mul
    ?
    n
    (1)
    E
    H
    (n) =
    E
    0
    N
    n
    mul
    (2)
    Combining this leads to the energy depending number
    of hadrons per generation
    ∆N
    H
    ∆ log
    N
    mul
    (E
    0
    /E
    H
    )
    (E
    H
    ) =
    ?
    E
    H
    E
    0
    ?
     κ
    with
    κ := 1 + log
    N
    mul
    2
    3
    (3)
    and to the hadronic flux
    dN
    H
    dE
    H
    (E
    H
    ) = 
    ln N
    mul
    E
    0
    ?
    E
    H
    E
    0
    ?
     (κ+1)
    .
    (4)
    Next we consider different types of hadrons
    h
    . In the
    following, we will keep using
    H
    if a variable is meant for
    all hadrons. We neglect the different reaction channels
    for different hadrons and assume a constant branching
    ratio
    B
    h
    for production of a hadron
    h
    , independent of
    the incident type. (We mainly focus on pions and kaons.)
    The muon flux can now be derived as the hadron
    flux multiplied with the decay probability
    P
    h→µ
    and
    folded by the energy distribution of the generated
    muon
    dn
    h
    dE
    µ
    (E
    µ
    , E
    h
    )
    :
    dN
    µ
    dE
    µ
    (E
    µ
    ) =
    ?
    h
    B
    h
    ?
    0
    dn
    h
    dE
    µ
    (E
    µ
    , E
    h
    ) P
    h→µ
    dN
    H
    dE
    H
    (E
    h
    ) dE
    h
    (5)

    2
    SEBASTIAN PANKNIN
    et al.
    MUONS FROM PARTICLE-SHOWERS IN ICE
    TABLE I
    USED NUMERICAL VALUES FOR PIONS AND KAONS. THEIR
    CONTRIBUTION
    A
    TO THE AMPLITUDE AS WELL AS THE FULL
    AMPLITUDE ARE PROVIDED.
    h
    B
    h
    b
    h→µ
    α
    r
    h
    A
    h
    [GeV
     1
    ]
    π
    0.9
    1.00
    67.1
    5.73 · 10
     1
    20.03 · 10
     3
    K
    0.1
    0.64
    9.03
    4.58 · 10
     2
    6.00 · 10
     3
    A =
    P
    h
    A
    h
    :
    26.30 · 10
     3
    The decay probability of a hadron with mass
    m
    h
    and
    lifetime
    τ
    h
    and a branching ratio
    b
    h→µ
    for the decay
    into muons is given by
    P
    h→µ
    = b
    h→µ
    Λ
    λ
    D
    =
    b
    h→µ
    1+ α
    h
    E
    h
    b
    h→µ
    α
    h
    E
    h
    with
    α
    h
    :=
    τ
    h
    m
    h
    λ
    I
    ,
    (6)
    where
    Λ :=
    1
    λ
    I
    +
    1
    λ
    D
    ,
    λ
    I
    is the interaction length of the
    hadron and
    λ
    D
    =
    E
    h
    τ
    h
    m
    h
    is the decay length, assuming
    E
    h
    ≫ m
    h
    . The approximation holds for
    α
    h
    E
    h
    ≫ 1
    .
    The probability to create a muon of energy
    E
    µ
    from
    the decay of a hadron with energy
    E
    h
    is given by
    the energy distribution
    dn
    h
    dE
    µ
    (E
    µ
    , E
    h
    )
    . Unpolarized me-
    son performing a two-body decay will generate mono-
    energetic muons isotropic distributed over the directions
    in their rest-frame (see [8]). This transforms in the
    laboratory system to a constant distribution between the
    minimal energy
    r
    h
    E
    h
    with
    r
    h
    =
    m
    2
    µ
    m
    2
    h
    and the maximal
    energy
    E
    h
    :
    dn
    h
    dE
    µ
    (E
    µ
    , E
    h
    ) =
    ?
    1
    (1 r
    h
    )E
    h
    r
    h
    E
    h
    ≤ E
    µ
    ≤ E
    h
    0
    else
    (7)
    Applying eq. (4), (6) and (7) on eq. (5) we obtain the
    muon flux
    dN
    µ
    dE
    µ
    (E
    µ
    ) = A
    ?
    E
    0
    GeV
    ?
    κ
    ?
    E
    µ
    GeV
    ?
     (κ+2)
    with
    A =
    ln N
    mul
    κ +2
    ?
    h
    B
    h
    b
    h→µ
    α
    h
    1  r
    κ+2
    h
    1  r
    h
    ?
    1
    GeV
    2
    .
    (8)
    This results in the numerical values for the amplitude
    A = 26.3 · 10
     3
    GeV
     1
    and for the exponent according
    to eq. (3)
    κ = 0.824
    . The values used for pions and
    kaons are summarized in Table I.
    III. CORSIKA SIMULATION
    For the simulation we used a modified version of
    CORSIKA based on the official version 6.2040 which
    enables shower simulation in salt water (see [9]). The
    used interaction models are Gheisha for low energies
    and QGSJet 01 for high energies. We simulated 1000
    showers at primary energy
    E
    0
    = 1 TeV
    , 1000 at
    10 TeV
    , 100 at
    100 TeV
    and ten at
    1 PeV
    with a proton
    as primary particle.
    The used configuration (Fig. 1) is a CORSIKA obser-
    vation level
    9 m
    behind the interaction point. Here the
    shower is expected to be fully developed, while only the
    very low energy muons will already be decayed.
    μ
    p
    salt water
    9m
    Observation Level
    Fig. 1. CORSIKA configuration: the incoming proton
    p
    interacts
    9 m
    above the observation level. The produced muon
    µ
    will roughly have
    the same direction as the incoming particle and are recorded, as they
    pass through the observation level.
    IV. DERIVED MUON FLUX
    With the simulation data we calculate the muon flux.
    To make this comparable to our model we need to
    calculate the muon energy at generation
    E
    µ
    from the
    muon energy at observation level
    E
    obs
    µ
    given in the
    simulation. This is a small correction, important only for
    low energetic muons. Muons that are most interesting
    for us are those with track-lengths above about
    10 m
    ,
    namely those with a range bigger than the typical shower
    size. These muons have energies above
    3 GeV
    , which is
    high enough, to reduce systematic error due to the energy
    correction.
    The energy loss can be approximated by:
    1
    ̺
    dE
    µ
    dx
    =  a  bE
    µ
    (9)
    where the medium density is
    ̺ = 1.02gcm
     3
    and the
    interaction constants are
    a = 2.68 MeVcm
    2
    g
     1
    and
    b =
    4.7 · 10
     6
    cm
    2
    g
     1
    (see [10]). Solving this provides the
    formula for the energy at generation:
    E
    µ
    =
    ?
    E
    obs
    µ
    +
    a
    b
    ?
    e
    ̺bx
    a
    b
    (10)
    Here we choose
    x = 7 m
    , the distance from first
    interaction point to the observation level reduced by
    about three radiation length [10].
    We show the normalized flux
    E
    2
    µ
    E
    κ
    0
    dN
    µ
    dE
    µ
    (E
    µ
    )
    which
    should follow according to eq. (8) a power law with
    small primary energie dependency (Fig. 2). The power
    law was fitted for the curves individually (Tab. II). This

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    3
    leads to the averaged parameterization:
    dN
    µ
    dE
    µ
    = A
    ?
    E
    0
    GeV
    ?
    κ
    ?
    E
    µ
    GeV
    ?
     (2+κ)
    A = (3.5 ± 0.5(
    stat
    ) ±
    6.5
    2.0
    (
    sys
    )) · 10
     3
    1
    GeV
    κ = 0.97 ± 0.07 (
    stat
    ) ± 0.12 (
    sys
    ).
    (11)
    The systematics were estimated by using different
    x
    -
    values for the energy correction.
    0.4 0.6 0.8 1.0 1.2
    1.4
    1.6 1.8 2.0
    log
    10
    (E
    /GeV)
    10
    -5
    10
    -4
    10
    -3
    10
    -2
    E
    2
    E
    0
    d
    N
    d
    E
    [
    1
    /
    GeV
    2
    ]
    1 TeV
    10 TeV
    100 TeV
    1 PeV
    Fit 1 TeV
    Fit 10 TeV
    Fit 100 TeV
    Fit 1 PeV
    Fig. 2. The muon flux
    d N
    µ
    dE
    µ
    is shown as function of muon energy
    E
    µ
    . We multiplied by
    E
    2
    µ
    E
    0
    to remove the primary energy dependency
    by some extend and improve readability. The results of the fit are given
    in Table II.
    Using the integral representation (Fig. 3), we can see
    that on average a hadronic cascade of
    100 TeV
    produces
    e.g. about one muon with an energy above
    10 GeV
    . Such
    a muon would have a track length of about
    36m
    .
    CORSIKA provides the information if a muon was
    generated from a pion. Using this information, Fig. 4
    shows the number of pion produced muons over all
    muons as a function of the energy. As expected, other
    0.4 0.6 0.8 1.0 1.2
    1.4
    1.6 1.8 2.0
    log
    10
    (E/GeV)
    10
    -3
    10
    -2
    10
    -1
    10
    0
    10
    1
    10
    2
    10
    3
    N
    (
    E
    >E
    )
    1 TeV
    10 TeV
    100 TeV
    1 PeV
    E
    0
    =1 TeV Fit
    E
    0
    =1 PeV
    Fig. 3. Integral muon flux
    N
    µ
    (E
    µ
    > E )
    . The fit is taken from the
    differential muon flux (Fig. 2 and Table II).
    TABLE II
    PARAMETERIZATION OF MUON FLUX: THE PARAMETERIZATION
    GIVEN IN (11) WAS FITTED TO THE MUON FLUX INDIVIDUALLY
    FOR THE DIFFERENT PRIMARY ENERGIES (S. FIG. 2).
    E
    0
    10
    3
    ·A [
    1
    GeV
    ]
    κ
    10TeV
    4.731
    ±0.995
    1.007
    ±0.104
    100TeV
    3.020
    ±0.703
    0.840
    ±0.113
    1TeV
    9.457
    ±6.792
    1.242
    ±0.371
    1PeV
    3.089
    ±0.965
    1.091
    ±0.156
    average
    3.490
    ±0.492
    0.971
    ±0.068
    production mechanisms (e.g. kaons) become more im-
    portant with higher energy. However the constant hadron
    fractions
    B
    h
    are a reasonable approximation. Our simple
    model with only pion and kaon (
    B
    K
    = 0.1
    ) predicts a
    constant
    N
    π
    µ
    N
    µ
    = 77%
    .
    0.0
    0.5
    1.0
    1.5
    2.0
    log
    10
    (E
    /GeV)
    0.0
    0.2
    0.4
    0.6
    0.8
    1.0
    π✆
    N
    /N
    1 TeV
    10 TeV
    100 TeV
    1 PeV
    model
    Fig. 4.
    Muon parent: the ratio of number of muons produced by
    pions
    N
    π
    µ
    over all number of muons
    N
    µ
    for each muon energy
    E
    µ
    is shown. A slight increase with energy of muons produced by other
    parent hadrons can be observed. However, the comparison to the model
    shows that the constant fraction is a reasonable approximation.
    V. DISCUSSION AND CONCLUSION
    We developed an analytical model describing the
    muon production in hadronic cascades, showing that it
    follows a power law with exponent
     (2 + κ)
    . The am-
    plitude scales linear with the medium density
    ̺
    and the
    primary energy
    E
    κ
    o
    . This is an effect of approximately
    10 %
    in amplitude for changing from salt water to ice.
    We compared this model to a full shower simulation
    with a modified CORSIKA version. The used setup
    forced us to correct for energy losses. To minimize the
    influence of the correction and because of the short track
    length of low energetic muons we focus on energies
    above
    3 GeV
    .
    We fitted a power law to the results (Fig 2 and Table
    II) and get the parameters with systematic errors due
    to the energy correction:
    A = (3.5 ± 0.5 (
    stat
    )  2.0 +
    6.5(
    sys
    )) · 10
     3
    GeV
     1
    and
    κ = 0.97 ± 0.07(
    stat
    ) ±
    0.12(
    sys
    ).
    Around an energy of
    E
    µ
    = 10GeV
    and a
    primary energy
    E
    0
    = 10 TeV
    , the value most important
    to us, the analytical model predicts a flux a factor three
    higher than the flux from the simulation results.

    4
    SEBASTIAN PANKNIN
    et al.
    MUONS FROM PARTICLE-SHOWERS IN ICE
    This study could be checked with a newer modified
    CORSIKA version [11], which would provide direct
    results for ice and the possibility to compare different
    hadronic interaction models and increase the statistic.
    The parameterization of muon production in hadronic
    cascades that we provide can be used to simulate more
    accurately neutrino induced cascades.
    VI. ACKNOWLEDGEMENTS
    We would like to thank Terry Sloan for providing
    us with his modified version of CORSIKA. Sebastian
    Panknin and Marek Kowalski acknowledge the support
    of the Deutsche Forschungsgemeinschaft (DFG).
    REFERENCES
    [1] THE ICECUBE COLLABORATION: J. AHRENS
    et al. Ice-
    Cube Preliminary Design Document
    http://www.icecube.wisc.
    edu/science/publications/pdd/pdd.pdf 2001
    [2] G.V. DOMOGATSKII (for THE BAIKAL COLLABORATION),
    Sta-
    tus of the BAIKAL neutrino project
    , arXiv:astro-ph/0211571v2,
    2002
    [3] Y. BECHERINI (for THE ANTARES COLLABORATION)
    Status
    report of the ANTARES experiment
    J.Phys.Conf.Ser.39:444-446,
    2006
    [4] AMANDA COLLABORATION: M. ACKERMANN,
    et al.
    Search for Neutrino-Induced Cascades with AMANDA
    Astropart.Phys.22:127-138,2004
    [5] D. HECK and T. PIEROG,
    Extensive Air Shower Simulation
    with CORSIKA: A Users Guide
    http://www-ik.fzk.de/corsika/
    usersguide/usersguide.pdf Version 6.9xx from April 3, 2009
    [6] W. HEITLER,
    The Quantum Theory of Radiation
    , third edition.
    Oxford University Press, London 1954
    [7] J. MATTHEWS,
    A Heitler model of extensive air showers
    , As-
    troparticle Physics 22 (2005), 387-397
    [8] T. K. GAISSER,
    Cosmic Rays and Particle Physics
    , Cambridge,
    1990
    [9] S. BEVAN
    et al.
    .,
    Astroparticle Physics
    , 28, 366, 2007
    [10] PARTICLE DATA GROUP,
    Particle Physics Booklet
    , Berkeley,
    2006
    [11] J.
    BOLMONT,
    CORSIKA-IW:
    a
    Modifed
    Version
    of
    CORSIKA
    for
    Cascade
    Simulations
    in
    Ice
    or
    Water
    ,
    http://nuastro-zeuthen.desy.de/e27/e711/e745/
    infoboxContent746/corsika-iw
    \
    julien
    \
    0812.pdf, 2008

    PROCEEDINGS OF THE 31
    st
    ICRC, ?OD´ Z´ 2009
    1
    Muon bundle energy loss in deep underground detector
    Xinhua Bai
    ?
    , Dmitry Chirkin
    y
    , Thomas Gaisser
    ?
    , Todor Stanev
    ?
    and David Seckel
    ?
    ?
    Bartol research Inst., Dept. of Physics and Astronomy, University of Delaware, Newark, DE 19716, U.S.A.
    y
    Dept. of Physics, University of Wisconsin, Madison, WI 53706, USA
    Abstract
    . High energy air showers contain bundles
    of muons that can penetrate deep underground.
    Study of these high energy muons can reveal the
    cosmic ray primary composition and some features
    of the hadronic interactions. In an underground neu-
    trino experiment like IceCube, high energy muons
    are also of interest because they are the dominant
    part of the neutrino background. We study muon
    bundle energy loss in deep ice by full Monte Carlo
    simulation to de?ne its ?uctuations and relation to
    the cosmic ray primary nuclei. An analytical formula
    of muon bundle mean energy loss is compared with
    the Monte Carlo result. We also use the simulation
    to set the background for muons with catastrophic
    energy loss much higher than those of normal muon
    bundles.
    Keywords
    : cosmic ray, muon bundle, energy loss.
    I. INTRODUCTION
    When high energy cosmic ray particles interact in the
    atmosphere and develop extensive air showers (EAS),
    muons are produced through the decay of mesons,
    such as pions and kaons. IceCube is primarily a high
    energy neutrino telescope but it can also study high
    energy muon bundles which, together with the EAS
    size determined by the surface array IceTop [1], will
    contribute to studies of the cosmic ray composition and
    hadronic interactions.
    A proton shower contains one muon above 500 GeV
    when the primary energy is about 100 TeV. Muons
    with that energy can reach the underground detector. At
    higher energies more high energy muons are produced
    in bundles. The muons above a several hundred GeV in
    the bundle are highly collimated and closer to each other
    in space [2], [3] than the distance between IceCube
    strings. This makes it very hard to count the number of
    individual muons in the bundle.
    In this work we study the energy loss of muon bundles
    produced by proton and iron showers at different primary
    energies by carrying out a Monte Carlo simulation. An
    analytical formula of muon bundle mean energy loss
    is compared with the Monte Carlo result in Section II.
    We discuss the ?uctuations in the energy loss and their
    association with cosmic ray primary particles in Section
    III. With the energy loss limits set by the simulation, we
    also discuss in Section IV the search for muons above
    100 TeV.
    We ?rst calculated proton and iron showers with
    CORSIKA (version 6.735, with SIBYLL 2.1 as high
    energy hadronic interaction model). This version of
    CORSIKA does not include charm production. All the
    muons produced are from pion or kaon decay. Two
    hundred showers were generated for eight energy points
    from 1 PeV to 1 EeV for proton and iron primaries. The
    production was set for the South Pole location (2835 me-
    ters above sea level). The atmospheric parametrization
    for July 01, 1997 was chosen. All muons above 200 GeV
    in each shower were extracted from the shower ground
    particle ?le. Each of these muons was then propagated
    through the ice to the depth of 2450 meters using the
    Muon Monte Carlo (MMC) simulation package [4]. All
    processes in which a muon loses more than 10
     3
    of
    its energy are treated stochastically. Energy losses due
    to ionization, bremsstrahlung, photo-nuclear interaction,
    pair production and their continuous components were
    kept for each of the ?ve-meter step size along the muon
    track.
    II. HIGH ENERGY MUONS IN AIR SHOWERS: ELBERT
    FORMULA AND CORSIKA SHOWERS
    The number of high energy muons in an EAS depends
    on the energy and mass of the primary cosmic ray
    particle. A well-known parametrization was given by
    Elbert [5] as follows:
    N
    ?;B
    (E
    ?
    > E
    ?
    (0))
    = A
    0:0145T eV
    E
    ?
    (0)cos?
    (
    E
    0
    AE
    ?
    (0)
    )
    p
    1
    ? (1 
    AE
    ?
    (0)
    E
    0
    )
    p
    2
    '
    0:0145T eV
    cos?
    E
    p
    1
    0
    A
    p
    1
     1
    E
    ?
    (0)
     p
    1
     1
    (1  p
    2
    A
    E
    0
    E
    ?
    (0))
    (1)
    in which A, E
    0
    and ? are the mass, total energy and
    zenith angle of the primary nucleus. p
    1
    = 0:757 and
    p
    2
    = 5:25. The approximation only keeps the ?rst two
    terms in the Taylor's expansion on (1 
    A
    E
    0
    E
    ?
    (0))
    p
    2
    . It
    will be used later in this paper to derive the mean energy
    loss of muon bundles in matter.
    Both Elbert parametrization and the approximation
    were compared with air shower Monte Carlo results for
    protons and irons from several hundred TeV up to 1
    EeV. Two examples are shown in Fig. 1. Except in the
    threshold region (where the Elbert parametrization may
    not be valid) the agreement over the whole energy region
    is remarkable. This is shown by the two peaks at zero
    in the two plots at the bottom in Fig. 1. One can also
    see that CORSIKA with SIBYLL 2.1 has slightly higher
    yield at higher muon energies. The excess increases with
    the energy of the primary particle.

    2
    X. BAI
    et al.
    MUON BUNDLE ENERGY LOSS ...
    E
    m
    (TeV)
    10
    -1
    1
    10
    10
    2
    10
    3
    )
    m
    (>E
    m
    N
    10
    -4
    10
    -3
    10
    -2
    10
    -1
    1
    10
    10
    2
    10
    3
    10
    4
    10
    5
    N
    CORSIKA
    (>E
    mu
    )-N
    Elbert
    (>E
    mu
    )
    -20
    -15
    -10
    -5
    0
    5
    10
    15
    20
    Number of Entries
    0
    100
    200
    300
    400
    500
    600
    N
    CORSIKA
    (>E
    mu
    )-N
    Elbert
    (>E
    mu
    )
    -20
    -15
    -10
    -5
    0
    5
    10
    15
    20
    Number of Entries
    0
    100
    200
    300
    400
    500
    600
    700
    Fig. 1. Number of muons in the bundles as a function of the muon
    energy. Top entry: 50 PeV proton at 30 degree zenith (lower black
    +) and 1 EeV iron at zero degree zenith (higher red ?). The open
    squares and circles are the averages over all 200 showers at each
    energy points. The curves represent Elbert formula. Dashed lines are
    the approximation to the Elbert formula, which correspond to the last
    line in Eq. ( 1). Bottom entries: The difference between the number
    of muons (for E
    ?
    ? 100 GeV and a increment step size of 100 GeV)
    in CORSIKA showers and that predicted by the Elbert formula. (left:
    50 PeV proton, right: 1 EeV iron)
    III. MUON BUNDLE ENERGY LOSS AND
    COMPOSITION SENSITIVITY
    A. Muon bundle energy loss in ice
    Since IceCube measures the energy loss instead of the
    number of muons in the bundle, we derive an analytical
    expression for the muon bundle mean energy loss to
    understand better the bundle energy loss. Starting from
    the approximation in Eq. ( 1) with muon energy loss
    dE
    ?
    (X)
    dX
    ˇ  a  b ? E
    ?
    (X);
    (2)
    the energy loss of a muon bundle at the slant depth X
    is an integral over the energy loss of the muons in the
    bundle:
    dE
    ?;B
    (X)
    dX
    =
    R
    E
    max
    ?
    (X)
    E
    min
    ?
    (X)
    dE
    ?
    (X)
    dX
    dN
    ?;B
    (X)
    dE
    ?
    (X)
    dE
    ?
    (X)
    = 
    R
    E
    max
    ?
    (X)
    E
    min
    ?
    (X)
    [a + b ? E
    ?
    (X)] ?
    dN
    ?;B
    (X)
    dE
    ?
    (X)
    dE
    ?
    (X):
    (3)
    Here, E
    min
    ?
    (X) and E
    max
    ?
    (X) represent the possible
    minimum and maximum energy of muons in the bundle
    at slant depth X. They can be written as follow [6]:
    E
    min
    ?
    (X) = maxf[(E
    ?
    (0) +
    a
    b
    )e
     bX
    a
    b
    ]
    min
    ; 0g
    =0
    E
    max
    ?
    (X) = (E
    max
    ?
    (0)+
    a
    b
    )e
     bX
    a
    b
    = (
    E
    0
    A
    +
    a
    b
    )e
     bX
    a
    b
    :
    (4)
    The approximate mean energy loss of a muon bundle
    can be obtained by doing the integration [6]. The ex-
    pression can be further simpli?ed by assuming that the
    high energy corrections can be ignored:
    dE
    ?;B
    (X)
    dX
    = ! ? b ? (p
    1
    +1)
    1
    V +1
    [
    1
    p
    1
    +1
    (
    a
    b
    )
     p
    1
    ?V
     p
    1
     1
    +
    1
    p
    1
    (
    a
    b
    )
     p
    1
    V
     p
    1
    1
    p
    1
    +1
    (
    a
    b
    )
    ?(
    E
    0
    A
    )
     p
    1
     1
    1
    p
    1
    (
    E
    0
    A
    )
     p
    1
    ]
    (5)
    in which ! =
    0:0145T eV
    cos?
    E
    p
    1
    0
    A
    p
    1
     1
    and V = (e
    bX
     1). For
    the ice at the South Pole, a = 0:26 GeV mwe
     1
    and
    b = 3:60 ? 10
     4
    mwe
     1
    [4] in units of meter water
    equivalent mwe. X is the slant depth in mwe along the
    muon bundle track.
    The comparison with the full Monte Carlo can be seen
    in Fig. 2. Despite the large ?uctuations in the energy loss
    mainly due to bremsstrahlung the mean energy loss in
    the full Monte Carlo (the blue and green dots) can be
    well described by the analytic approximation of Eq. ( 5).
    It needs to be pointed out, however, that for low energy
    showers, or for the energy loss at larger slant depth, Eq.
    ( 5) can have larger offset from the mean Monte Carlo
    values. This is due to the fact that the Elbert formula
    (Eq. ( 1)) is not exact at high E
    ?
    values or a and b can
    be different from the constant values being used here.
    Slant Depth (m)
    1000 1200 1400 1600 1800 2000 2200 2400 2600 2800
    /dX(GeV/m)
    tot.
    dE
    1
    10
    10
    2
    10
    3
    10
    4
    10
    5
    Fig. 2.
    Muon bundle energy loss as function as slant depth. Two
    examples are given in the ?gure: red ? for vertical 1 EeV iron showers
    and the black + for 30 degree 50 PeV proton showers. Open squares
    and circles are the mean value of the Monte Carlo results for iron and
    proton showers. Eq. ( 5) is represented by the two curves.
    B. Muon bundle energy loss and composition
    To study the muon bundle energy loss and composi-
    tion sensitivity, in the two component case of this work,

    PROCEEDINGS OF THE 31
    st
    ICRC, ?OD´ Z´ 2009
    3
    we ?rst de?ne the composition resolving parameters
    (AjY ) and (BjY ) based on an observable Y as follows:
    (AjY ) =
    P
    i=1
    N
    A
    i
    =(N
    A
    i
    + N
    B
    P
    i
    )
    i=1
    P
    A
    i
    (6)
    (BjY ) =
    P
    i=1
    N
    B
    i
    =(N
    A
    i
    + N
    B
    P
    i
    )
    i=1
    P
    B
    i
    (7)
    in which P
    A
    i
    = 1:0 (or = 0:0) when N
    A
    i
    6= 0:0 (or
    = 0:0), and N
    A
    i
    and N
    B
    i
    represent the number of proton
    and iron events in the i
    th
    bin on dN=dY distribution as
    shown in Fig. 3. (AjY ) and (BjY ) have values between
    0 and 1. When the two distributions are well separated
    from each other, both (AjY ) and (BjY ) are equal to
    1. When the two distributions are identical and fully
    overlapped, (AjY ) = (BjY ) = 0:5, which means the
    chance to assign a particle as proton or iron is 50%.
    N
    N
    1
    i
    N i
    N k
    1ki
    N
    k
    A
    A
    B
    B
    B
    dN/dY
    Y
    Y
    0
    5
    10
    15
    20
    25
    30
    35
    40
    dN/dY
    0
    500
    1000
    1500
    2000
    2500
    3000
    3500
    4000
    P: ( 1.0, 15.0, 4.0)
    (P|Y) =0.697
    Fe:( 1.0, 25.0, 4.0)
    (Fe|Y)=0.579
    Fig. 3.
    Left: De?nition of composition resolving parameters used
    in this work, (AjY ) and (BjY ): Two histograms represent the
    distribution of variable Y (e.g. muon bundle energy loss in deep ice
    detector) for two types of primary particles (say proton and iron).
    Right: Particles A and B are represented by two Gaussian shaped
    distributions with the amplitude, mean and ˙ in the parentheses.
    Particle A and B each has 320000 and 80000 samples in this test.
    The value of the parameter also depends on the
    frequency of the detectable signals of different particles.
    This can be seen in Fig. 3 (right-hand panel), in which
    (AjY ) = 0:697 and (BjY ) = 0:579. This de?nition
    can be easily extended to cases in which each event has
    multiple observable variables.
    IceCube is sensitive to the Cherenkov light emitted
    by high energy charged particles. Simulation shows
    the Cherenkov light yield of an in-ice event is nearly
    proportional to the particle total energy loss [7]. Since
    we are not doing the full experiment simulation, we
    use the total energy loss of muon bundles in small
    bins along the bundle track as a measure of the signal
    size. The distributions of muon bundle energy loss in
    ?ve-meter steps at the slant depths between 1958 m
    and 2010 m in the ice are shown in Fig. 4. Proton
    energy loss distribution has a signi?cant overlap with
    iron, much bigger than the overlap in the muon number
    distributions.
    To improve the separation between the energy loss
    signals from different nuclei we excluded the large
    energy loss events caused mostly by bremsstrahlung. The
    effect from eliminating all energy loss events larger than
    85% of the average (cut
    1
    , red) and 500% (cut
    2
    , blue)
    is shown in the bottom panel of Fig. 4.
    log
    10
    [E
    tot.loss
    /(GeV*5m)]
    1
    1.5
    2
    2.5
    3
    3.5
    4
    4.5
    5
    Number of Events
    1
    10
    10
    2
    P, N
    m
    Fe, N
    m
    P, E
    tot.loss
    Fe, E
    tot.loss
    log
    10
    (N
    m
    )
    1
    1.5
    2
    2.5
    3
    3.5
    4
    4.5
    5
    log
    10
    [E
    tot.loss
    /(GeV*5m)]
    1
    1.5
    2
    2.5
    3
    3.5
    4
    4.5
    5
    Number of Events
    1
    10
    10
    2
    P, E
    tot.loss
    , cut_1
    Fe, E
    tot.loss
    , cut_1
    P, E
    tot.loss
    , cut_2
    Fe, E
    tot.loss
    , cut_2
    Fig. 4.
    Top: Histograms of the number of muons and the muon
    bundle total energy loss in ?ve-meter steps at depths between 1958
    m and 2010 m in ice. The example is for 600 PeV proton and iron
    showers at zenith angle of 30 degrees. Bottom: Histograms of the
    muon bundle total energy loss for the same Monte Carlo data sample
    after two cuts were applied. See more details in the text.
    Fig. 5 summarizes how the composition resolving
    parameter varies at different slant depth after these two
    cuts in proton and iron showers. Several features can
    be seen in this ?gure, (1) the tighter cut (cut
    1
    ) gives
    a higher value of resolving parameter corresponding to
    less overlap between the two histograms shown at the
    bottom plot of Fig. 4, (2) when tighter cut is applied,
    the composition resolving power using the muon bundle
    energy loss can be close to that obtained by the number
    of muons in the bundle in the simulation, (3) using the
    same cut, the composition resolving power is slightly
    better at shallower depth. It would be very interesting to
    explore these features in real data analysis.
    IV. SIGNATURE OF HIGH ENERGY PROMPT MUONS IN
    MUON BUNDLE EVENTS
    Very high-energy muons in air showers are produced
    either in the decay of very short-lived particles, i.e.
    charm or from the ?rst interaction whether the parent
    is conventional (pion or kaon) or charm. The crossing
    from conventional to prompt muon ?uxes was estimated
    to happen between 40 TeV and 3 PeV [8]. Such muons
    may be used to study the composition of cosmic ray
    primaries, as well as heavy quark production in high
    energy p-N interactions. There are several ways to
    separate the prompt muons from the conventional ones in

    4
    X. BAI
    et al.
    MUON BUNDLE ENERGY LOSS ...
    Slant Depth (m)
    1200
    1400
    1600
    1800
    2000
    2200
    2400
    2600
    2800
    Resolving Probability
    0.4
    0.5
    0.6
    0.7
    0.8
    0.9
    1
    P, by N of muon
    P, by E
    tot.loss
    , cut_1
    P, by E
    tot.loss
    , cut_2
    P, by E
    tot.loss
    , no cut
    Slant Depth (m)
    1200
    1400
    1600
    1800
    2000
    2200
    2400
    2600
    2800
    Resolving Probability
    0.4
    0.5
    0.6
    0.7
    0.8
    0.9
    1
    Fe, by N of muon
    Fe, by E
    tot.loss
    , cut_1
    Fe, by E
    tot.loss
    , cut_2
    Fe, by E
    tot.loss
    , no cut
    Fig. 5. The values of composition resolving parameters for proton
    and iron under different cuts at different slant depths. The calculation
    was done with 200 proton showers and 200 iron showers with 600 PeV
    primary energy at zenith angle of 30 degrees. The two cuts correspond
    to the cuts used in Fig. 4
    underground experiments, such as using the difference
    in their zenith angle distributions, the different depth
    dependence at a given depth and zenith angle [9].
    Another technique that is explored here relies on recog-
    nizing catastrophic dE=dX signature from these leading
    muons as bursts of light on an otherwise smoother light
    deposition from a bundle of lower energy muons.
    The probability of ?nding a certain amount of energy
    loss in ?ve-meter steps from 1450 to 2450 meters under
    ice is shown in Fig. 6. The chance to have an energy
    loss of about 30 TeV (point A in Fig. 6) in a ?ve-meter
    step is much higher for a muon with energy of 100 TeV
    than conventional muon bundles from showers below
    100 PeV. If one sees a burst energy loss above 160 TeV,
    it is almost certain (P > 1? 10
     3
    versus P < 3? 10
     5
    )
    that the event consists of a single muon with energy
    above 1 PeV rather than showers below 1 EeV (point B
    in Fig. 6).
    Since the cosmic ray primary energy can be deter-
    mined by the surface array in IceCube, this method can
    be explored further with IceTop and in-ice coincidence
    data.
    V. SUMMARY
    In this work, we studied the muon bundle energy
    loss in ice and its association with cosmic ray primary
    composition. The analytic formula of the mean muon
    log
    10
    (E
    loss
    /(GeV*5m)
    234567
    Probability
    10
    -6
    10
    -5
    10
    -4
    10
    -3
    10
    -2
    10
    -1
    1
    muon bundle in H showers
    muon bundle in Fe showers
    muon, fixed energy
    AB
    Fig. 6. The probability of the energy loss (in a ?ve-meter step) of
    muon bundles in air showers (solid lines) and muons with ?xed energy
    (dashed lines). Vertical proton and iron showers with primary energy
    of 500 TeV, 10 PeV, 100 PeV, 600 PeV and 1EeV are plotted together
    with vertical muons with ?xed energy of 5 TeV, 100 TeV, 1 PeV, 6
    PeV, and 10 PeV on the surface. The probability increases at larger
    energy loss as the energy of primary or single muon goes higher. The
    energy loss sample for the probability calculation is taken from the
    depth of 1450. m to the bottom of the in-ice array.
    bundle energy loss given here has reasonably good
    agreement with the full Monte Carlo. It can be used
    in muon bundle event reconstruction in IceCube. The
    parameters (cosmic primary energy and mass) in the
    formula can be further explored in composition study
    using IceTop and in-ice coincidence data [10]. Using
    IceTop in-ice coincidence data, one can also look for
    signatures from very high energy muons from charm
    decay by recognizing large catastrophic dE=dX along
    the muon bundle track.
    Acknowledgment
    This work is supported in part by
    the NSF Of?ce of Polar Programs and by NSF grant
    ANT-0602679.
    REFERENCES
    [1] T.K. Gaisser for the IceCube Collaboration, Proc. 30th ICRC,
    Merida (Mexico, 2007), arXiv:0711.0353, p. 15-18.
    [2] T. Gaisser,
    Cosmic Rays and Particles Physics
    , p.208 (Cambridge
    University Press, 1990).
    [3] X. Bai, et al. Proc. 30th ICRC, Merida, Mexico, Vol.5, p.1209-
    1212, 2007.
    [4] D. Chirkin, W. Rhode, flPropagating leptons through matter with
    Muon Monte Carlo (MMC)fl, arXiv:hep-ph/0407075v2, 2008.
    [5] J.W. Elbert, In Proc. DUMAND Summer Workshop (ed. A.
    Roberts), 1978, vol. 2, p. 101.
    [6] X. Bai, flEnergy loss of muon bundles: An analytic approxima-
    tionfl, IceTop internal technical report. 2007.
    [7] Christopher Henrik V. Wiebusch, Ph.D. thesis, Physicalische
    Institute, RWTH Aachen, 1995.
    [8] Graciela Gelmini, Paolo Gondolo, and Gabriele Varieschi,
    arXiv:hep-ph/0209111v1.
    [9] E. V. Bugaev et al., Phys. Rev. D58(1998)054001
    [10] See for example the work by Tom Feusels, Jonathan Eisch and
    Chen Xu,
    Reconstruction of IceCube coincident events and study
    of composition-sensitive observables using both the surface and
    deep detector.
    These proceedings. More work on data is still
    needed.

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    1
    Constraints on Neutrino Interactions at energies beyond 100 PeV
    with Neutrino Telescopes
    Shigeru Yoshida
    Department of Physics, Faculty of Science, Chiba University, Chiba 263-8522, Japan
    Abstract
    . A search for extremely high energy cos-
    mic neutrinos has been carried out with the IceCube
    Neutrino Observatory. These event are neutrino-
    induced energetic charged leptons and their rate
    depends on the neutrino-nucleon cross-sections. The
    resultant event rate has implications for possible new
    physics beyond the standard model as it is predicted
    that the cross-sections can be much higher than
    the standard particle physics prediction if we live
    in more than four space-time dimensions. In this
    study we show the capability of neutrino telescopes
    such as IceCube to constrain neutrino cross-sections
    at energies beyond
    10
    7
    GeV. The constraints are
    obtained as a function of the extraterrestrial neutrino
    flux in the relevant energy range, which accounts for
    the astrophysical uncertainty of neutrino production
    models.
    Keywords
    : Neutrino, IceCube, cross-sections
    I. INTRODUCTION
    High energy cosmic neutrino observations provide a
    rare opportunity to explore the neutrino-nucleon (
    ν
    N)
    interaction behavior beyond energies accessible by the
    present accelerators. These neutrinos interact druing
    their propagation in the earth and produce energetic
    muons and taus. These secondary leptons reach un-
    derground neutrino detectors and leave the detectable
    signals. The detection rate is, therefore, sensitive to
    neutrino-nucleon interaction probability. The center-of-
    mass energy of the collision,
    s
    , is well above
    ∼ 10
    TeV for cosmic neutrino energies on the order of 1
    EeV (=
    10
    9
    GeV), a representative energy range for
    the bulk of the GZK cosmogenic neutrinos, generated
    by the interactions between the highest energy cosmic
    ray nucleons and the cosmic microwave background
    photons [1].
    The
    ν
    N collision cross-sections can be varied largely
    if non-standard particle physics beyond the Standard
    M
    odel (SM) are considered in the high energy regime of
    s ≫
    TeV. The extra-dimension scenarios, for example,
    have predicted such effects [2]. The cross-sections of
    black hole productions via
    ν
    N collisions can be larger
    than the SM prediction by more than two orders of
    magnitude [3]. The effect would be sizable enough to
    affect the expected annual event rate (
    O(0.1  1))
    of the
    GZK neutrinos by
    km
    3
    instrumentation volume of
    the underground neutrino telescope such as the IceCube
    observatory, thereby the search for the extremely-high
    energy (EHE) cosmic neutrinos leads to constraints on
    the non-standard particle physics.
    The IceCube neutrino observatory has already begun
    EHE neutrino hunting with the partially deployed under-
    ground optical sensor array. The 2007 partial IceCube
    detector realized a
    ∼ 0.5
    km
    2
    effective area for muons
    with
    10
    9
    GeV and recently placed a limit on the flux
    of EHE neutrinos at the level approximately an order of
    magnitude higher than the expected GZK cosmogenic
    neutrino intensities for 242 days of observation [4].
    Since new particle physics may vary the cross section by
    more than an order of magnitude as we noted above, this
    result should already imply a meaningful bound on the
    ν
    N cross-sections. In this paper, we study the constraint
    on the
    ν
    N cross-sections (
    σ
    ν N
    ) by the null detection of
    EHE neutrinos with the 2007 IceCube observation. A
    model-independent bound is derived by estimating the
    lepton intensity at the IceCube depth with the SM cross-
    sections scaled by a constant. The constraint is displayed
    in form of the excluded region on the plane of the cosmic
    neutrino flux and
    σ
    ν N
    . It is equivalent to upper-bound
    of
    σ
    ν N
    for a given flux of astrophysical EHE neutrinos.
    II. THE METHOD
    The neutrino and charged lepton fluxes at the IceCube
    depth originated in a given neutrino flux at are calculated
    by the coupled transportation equations:
    dJ
    ν
    dX
    =
     N
    A
    σ
    νN,CC +NC
    J
    ν
    +
    m
    l
    cρτ
    d
    l
    Z
    dE
    l
    1
    E
    l
    dn
    d
    l
    dE
    ν
    J
    l
    (E
    l
    )
    +N
    A
    Z
    dE
    ν
    ν N,N C
    dE
    ν
    J
    ν
    (E
    ν
    )
    +N
    A
    Z
    dE
    l
    lN,C C
    dE
    ν
    J
    l
    (E
    l
    )
    (1)
    dJ
    l
    dX
    =
     N
    A
    σ
    lN
    J
    l
    m
    l
    cρτ
    d
    l
    E
    l
    J
    l
    +N
    A
    Z
    dE
    ν
    ν N,C C
    dE
    l
    J
    ν
    (E
    ν
    )
    +N
    A
    Z
    dE
    l
    lN
    dE
    l
    J
    l
    (E
    l
    )
    +
    m
    l
    cρτ
    d
    l
    Z
    dE
    l
    1
    E
    l
    dn
    d
    l
    dE
    l
    J
    l
    (E
    l
    ),
    (2)
    where
    J
    l
    = dN
    l
    /dE
    l
    and
    J
    ν
    = dN
    ν
    /dE
    ν
    are differen-
    tial fluxes of charged leptons and neutrinos, respectively.
    X
    is the column depth,
    N
    A
    is the Avogadro’s number,
    ρ
    is the local density of the medium (rock/ice) in the
    propagation path,
    σ
    is the relevant interaction cross-
    sections,
    dn
    d
    l
    /dE
    is the energy distribution of the decay
    products which is derived from the decay rate per unit
    energy,
    c
    is the speed of light,
    m
    l
    and
    τ
    d
    l
    are the mass

    2
    S. YOSHIDA, CONSTRAINTS ON UHE
    ν
    INTERACTIONS
    GZK
    ν
    E > 10PeV
    x1
    x3
    x10
    cos (Zenith Angle)
    −18−2
    sec
    Flux [10 cm
    sr]
    −1
    −1
    −1
    −0.5
    0
    0.5
    1
    0
    1
    2
    3
    4
    Fig. 1.
    Integral fluxes of the muon and taus above 10 PeV
    (=
    10
    7
    GeV) at IceCube depth (∼ 1450m) of the GZK cosmogenic
    neutrinos [5]. The solid lines represent the muons while the dashed
    lines represent the taus. Numbers along each of the curves are the
    multiplication factors (N
    scale
    ) that enhance the standard
    νN
    cross-
    sections [6] in the relevant calculations.
    and the decay life time of the lepton
    l
    , respectively. In
    this paper we scale
    σ
    ν N
    to that of the SM prediction
    with the factor N
    scale
    ,
    i.e.
    ,
    σ
    ν N
    ≡ N
    scale
    σ
    SM
    ν N
    . It is an
    extremely intensive computational task to resolve the
    coupled questions above for every possible values of
    σ
    ν N
    . To avoid this difficulty, we introduce two assump-
    tions to decouple calculation of
    J
    ν
    from the charged
    lepton transportation equation. The first is that distortion
    of the neutrino spectrum by the neutral current reaction
    is small and the other is that the contribution of muon
    and tau decay to enhance the neutrino flux, which is
    represented by the second term on the right hand side of
    Eq. 1, is negligible. These are very good approximation
    in the energy region above
    10
    8
    GeV where even the
    tau is unlikely to decay before reaching the IceCube
    instrumentation volume. Then the neutrino flux is simply
    given by the beam dumping factor as
    J
    ν
    (E
    ν
    , X
    IC
    ) = J
    ν
    (E
    ν
    , 0)e
     N
    scale
    σ
    SM,CC
    ν N
    X
    IC
    ,
    (3)
    where
    X
    IC
    is column density of the propagation path
    from the earth surface to the IceCube depth. The charged
    lepton fluxes,
    J
    l=µ,τ
    (E
    l
    , X
    IC
    )
    , are obtained as
    J
    µ,τ
    (E
    µ,τ
    , X
    IC
    ) = N
    A
    Z
    X
    IC
    0
    dX
    Z
    dE
    µ,τ
    dN
    µ,τ
    dE
    µ,τ
    (E
    µ,τ
    → E
    µ,τ
    )
    Z
    dE
    ν
    N
    scale
    S M,C C
    νN
    dE
    µ,τ
    J
    ν
    (E
    ν
    , 0)e
     N
    scale
    σ
    SM,C C
    ν N
    X
    .
    (4)
    Here
    dN
    µ,τ
    /dE
    µ,τ
    (E
    µ,τ
    → E
    µ,τ
    )
    represents distribu-
    tions of muons and taus with energy of
    E
    µ,τ
    at
    X
    IC
    originated in those with energy of
    E
    µ,τ
    produced by
    νN
    collisions at depth
    X
    . This is calculated in the
    transportation equation, Eq. 2, with a replacement of
    J
    ν
    (E
    ν
    )
    by Eq. 3.
    Calculation of the neutrino and the charged lepton
    fluxes with this method is feasible for a wide range of
    N
    scale
    without any intensive computation. A comparison
    of the calculated fluxes with those obtained without the
    introduced simplification for a limited range of
    N
    scale
    indicates that the relative difference we found in the
    resultant
    J
    ν,µ,τ
    (X
    IC
    )
    is within 40%. Since this analysis
    is searching for at least an
    order
    of magnitude difference
    in
    σ
    ν N
    , the introdced simplifications provide sufficient
    accuracy for the present study.
    Fig. 1 shows the calculated intensities of the sec-
    ondary muons and taus for various
    N
    scale
    factors. One
    can see that the intensity is almost proportional to
    N
    scale
    as expected since the interaction probability to generate
    muons and taus linearly depends on
    σ
    ν N
    . It should
    be pointed out, however, that the dependence starts to
    deviate from the complete linearity when the propagation
    distance is comparable to the mean free path of neutri-
    nos, as one can find in the case of
    N
    scale
    = 10
    in the
    figure. This is because the neutrino beam dumping factor
    in Eq. 3 becomes significant under this circumstances.
    The flux yield of leptons,
    Y
    l
    ν
    (
    l = ν
    s,
    µ, τ
    ) from
    neutrinos with a monochromatic energy at earth surface,
    E
    s
    ν
    , is given by Eq. 4 with an insertion of
    J
    ν
    (E
    ν
    , 0) =
    δ(E
    ν
     E
    s
    ν
    )
    . The resultant event rate per
    neutrino energy
    decade
    is then obtained by,
    N
    ν
    (E
    s
    ν
    ) =
    X
    ν =ν
    e
    µ
    τ
    1
    3
    dJ
    ν
    e
    µ
    τ
    d log E
    ν
    (E
    s
    ν
    )
    Z
    dΩ
    X
    l=ν,µ,τ
    Z
    dE
    l
    A
    l
    (E
    l
    ) Y
    l
    ν
    (E
    s
    ν
    , E
    l
    , X
    IC
    (Ω), N
    scale
    )
    (5)
    where
    A
    l
    is the effective area of the IceCube to detect
    the lepton
    l
    . Note that the differential limit of the
    neutrino flux is given by Eq. 5 for
    N
    scale
    = 1
    with
    N
    ν
    = µ
    ¯
    90
    which corresponds to the 90 % confidence
    level average upper limit. This calculation is valid when
    the cosmic neutrino flux
    J
    ν
    and the cross section
    σ
    ν N
    do not rapidly change over a decade of neutrino energy
    around
    E
    s
    ν
    . Limiting
    σ
    ν N
    in the present analysis corre-
    sponds to an extraction of the relation between
    N
    scale
    and the (unknown) cosmic neutrino flux
    J
    ν
    e
    µ
    τ
    yielding
    N
    ν
    = µ
    ¯
    90
    . The obtained constraints on
    σ
    ν N
    is represented as a function of
    J
    ν
    e
    µ
    τ
    for a given
    energy of
    E
    s
    ν
    . It consequently accounts for astrophysical
    uncertainties on the cosmic neutrino flux.
    In scenarios with extra dimensions and strong grav-
    ity, Kaluza-Klein gravitons can change only the neu-
    tral current (NC) cross-sections because gravitons are
    electrically neutral. Any scenarios belonging to this
    category can be investigated by scaling only
    σ
    NC
    ν N
    in
    the present analysis. The event rate calculation by
    Eq. 5 is then performed for
    Y
    l
    ν
    (N
    scale
    = 1)
    with ef-
    fective area for
    ν
    ’s,
    A
    ν
    , enhanced by
    S M,C C
    ν N
    +
    N
    scale
    σ
    S M,N C
    ν N
    )/(σ
    S M,C C
    ν N
    + σ
    SM,NC
    ν N
    )
    since the rate of
    detectable events via the NC reaction by IceCube is
    proportional to
    σ
    N C
    ν N
    . We also show the constraint in
    this case.
    III. RESULTS
    In this analysis we use the IceCube observation results
    with 242 days data in 2007 to limit
    σ
    ν N
    using Eq. 5. No
    detection of signal candidates in the measurement has
    led to an upper limit of the neutrino flux at
    6 × 10
     7

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    3
    ]
    sec
    j(E) E[GeV cm
    −1
    sr
    −2
    −1
    2
    σ
    ν
    N
    [cm
    2
    ]
    1EeV
    10EeV
    10
    −33
    10
    −32
    10
    −31
    10
    −30
    10
    −29
    10
    −9
    10
    −8
    10
    −7
    10
    −6
    Fig. 2.
    Constraints on the all-flavor cosmic neutrino flux and the
    charged current
    νN
    cross-sections based on the null detection of
    neutrino signals by the IceCube 2007 observation. The right upper
    region is excluded by the present analysis. The cross points provide
    reference points where the standard cross section [6] and the expected
    GZK cosmogenic neutrino fluxes [7] is located.
    GeV cm
     2
    sec
     1
    sr
     1
    [4]. The effective area
    A
    l
    is 0.5
    km
    2
    for
    µ
    , 0.3 km
    2
    for
    τ
    , and
    3 × 10
     4
    km
    2
    for
    ν
    s.
    Constraints on
    σ
    ν N
    are then derived with Eq. 5. Here
    we assume that the effective area for
    ν
    s is proportional
    to
    N
    scale
    . The results for
    E
    s
    ν
    = 10
    9
    and
    10
    10
    GeV are
    shown in Fig. 2. Enhancing the charged current cross-
    sections by more than a factor of 30 for
    E
    ν
    = 1
    EeV
    (10
    9
    GeV) is disfavored if the astrophysical neutrino
    intensities are around
    ∼ 10
     7
    GeV cm
     2
    sec
     1
    sr
     1
    ,
    the expected range of the GZK cosmogenic neutrino
    bulk. Note that neutrino-nucleon collision with
    E
    ν
    = 1
    EeV corresponds to
    s ∼ 40
    TeV and the present
    limit on
    σ
    ν N
    would place a rather strong constraint
    on scenarios with extra dimensions and strong grav-
    ity, although more accurate estimation requires studies
    with a model-dependent approach which implements
    the cross-sections and the final-state particles from the
    collision predicted by a given particle physics model in
    the neutrino propagation calculation. Taking into account
    uncertainty on the astrophysical neutrino fluxes, any
    model that increases the neutrino-nucleon cross-section
    to produce charged leptons by more than two orders
    of magnitude at
    s ∼ 40
    TeV is disfavored by the
    IceCube observation. However, we should point out that
    the IceCube 2007 data could not constrain the charged
    current cross-sections if the intensity of cosmic neutrinos
    in the relevant energy region is fewer than
    ∼ 10
     8
    GeV
    cm
     2
    sec
     1
    sr
     1
    , approximately an order of magnitude
    lower than the predicted cosmogenic neutrino fluxes
    discussed in the literature. Absorption effects in the earth
    becomes sizable in this case, resulting in less sensitivity
    to the cross-section. This limitation will be improved for
    larger detection area of the full IceCube detector.
    Fig. 3 shows the constraint when only the NC cross
    section is varied. Enhancement of
    σ
    N C
    ν N
    by a factor
    beyond 100 at
    s ∼ 40
    TeV is disfavored, but this
    strongly depends on the cosmic neutrino flux one as-
    ]
    sec
    j(E) E[GeV cm
    −1
    sr
    −2
    −1
    2
    σ
    ν
    N
    [cm
    2
    ]
    1EeV
    10EeV
    10
    −33
    10
    −32
    10
    −31
    10
    −30
    10
    −29
    10
    −9
    10
    −8
    10
    −7
    10
    −6
    Fig. 3.
    Constraints on the all-flavor cosmic neutrino flux and the
    neutral current
    νN
    cross-sections for the scenario that only the neutral
    current reaction is enhanced by a new physics beyond the standard
    model. The right upper region is excluded by the present analysis.
    The crosses provide reference points for the standard cross section [6]
    and the expected GZK cosmogenic neutrino fluxes [7] is located.
    sumes. Because the NC interaction does not absorb
    neutrinos during their propagation though the earth, even
    the case when the neutrino flux is small could bound
    the cross-section, but the limit becomes rather weak;
    the allowed maximum enhancement factor is an order
    of
    ∼ 10
    3
    .
    IV. SUMMARY AND OUTLOOK
    The IceCube 2007 observation indicated that any
    scenario to enhance either the NC or both NC and CC
    equivalent cross-sections by more than 100 at
    s ∼ 40
    TeV is unlikely if the astrophysical neutrino fluxes are
    ∼ 10
     7
    GeV cm
     2
    sec
     1
    sr
     1
    in EeV region. A study
    of constraints on the model-dependent cross-sections
    predicted by the theories of the black holes creation with
    extra dimensions is underway with a dedicated treatment
    of final state particles produced from the black hole
    evaporation.
    ACKNOWLEDGMENTS
    This work was supported in part by the Grants-in-
    Aid in Scientific Research from the MEXT (Ministry of
    Education, Culture, Sports, Science, and Technology) in
    Japan.
    REFERENCES
    [1] V. S. Beresinsky, G. T. Zatsepin, Phys. Lett.
    28B
    , 423 (1969).
    [2] C. Tyler, A. V. Olinto, G. Sigl, Phys. Rev.
    D63
    05501 (2001).
    [3] J. Alvarez-Mun˜iz
    et. al.
    . Phys. Rev.
    D65
    124015 (2002).
    [4] K. Mase, A. Ishihara, S. Yoshida for the IceCube collaboration,
    this proceedings
    .
    [5] S. Yoshida, H. Dai, C.C.H. Jui, P. Sommers, Astrophys. J.
    479
    ,
    547 (1997).
    [6] R. Gandhi, C. Quigg, M. H. Reno, and I. Sarcevic, Astropart.
    Phys.
    5
    , 81 (1996); Phys. Rev. D
    58
    093009 (1998).
    [7] O. E. Kalashev
    et.al.
    , Phys. Rev. D
    66
    , 063004 (2002).

    .

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁOD´ Z´ 2009
    1
    Constraints on Extragalactic Point Source Flux from Diffuse
    Neutrino Limits
    Andrea Silvestri
    and Steven W. Barwick
    Department of Physics and Astronomy, University of California, Irvine, CA 92697, USA.
    Abstract. We constrain the maximum flux
    from extragalactic neutrino point sources by
    using diffuse neutrino flux limits. We show
    that the maximum flux from extragalactic
    point
    sources
    is
    E
    2
    (dN
    ν
    /dE) ≤ 5.1 × 10
    −9
    (L
    ν
    /10
    45
    erg/s)
    1/3
    GeV cm
    −2
    s
    −1
    from
    an
    ensemble of sources with average neutrino
    luminosity per decade, L
    ν
    . It depends only slightly
    on factors such as the inhomogeneous matter density
    distribution in the local universe, the luminosity
    distribution, and the assumed spectral index.
    Keywords: Extragalactic sources, diffuse and point
    sources, high energy neutrinos
    I. INTRODUCTION
    The origin of ultra high energy cosmic rays (UHECR),
    is still unknown. AGN, GRB’s, or processes beyond the
    standard model have been hypothesized to be the sources
    of UHECR’s, and may originate from regions of the sky
    correlated with AGN sources [1]. Therefore, if nearby
    AGN are the sources of the highest energy CR’s, and if
    AGN emit ν’s in addition to photons, protons and other
    charged particles, then the fluxes from individual AGN
    may be observable by current generation of neutrino
    detectors. Several models predict a diffuse neutrino flux
    from AGN, in particular ν-production has been predicted
    from the core of radio-quite AGN as presented in [2],
    [3], and from AGN jets and radio lobes as suggested
    in [4], [5], [6]. There are good but speculative reasons to
    expect a correlation between sources of cosmic rays and
    sources of neutrinos. Direct searches for diffuse [7] and
    point flux [8] by current telescopes have set the most
    stringent upper limits, but generally have not reached
    the sensitivity required, and the models suggest that
    challenges exist even for next generation telescopes.
    Of course, one of the primary motivations for the
    construction of ν-telescopes is to search for unexpected
    sources with no obvious connection to the power emitted
    in the electromagnetic (EM) band. However, we show
    in this paper that the ν-flux from EG point sources can
    be constrained by the measured diffuse ν-flux limits.
    We also test models from individual sources with the
    constraints.
    II. ANALYSIS
    If the diffuse ν-flux is generated by an ensemble
    of extragalactic (EG) sources, then only the nearest of
    the diffusely distributed sources would be detectable
    as point sources. Point sources of neutrinos are ob-
    served when several neutrinos originate from the same
    direction, and in the context of this study, only the
    very nearest of the uniformly distributed sources are
    detectable as point sources. The number of detectable
    (or resolvable) point sources, N
    s
    , first proposed in [9],
    is determined for a given diffuse ν-flux limit and point
    source sensitivity. The N
    s
    calculation is based on three
    general assumptions: (1) the sources are extragalactic
    and uniformly distributed in space; (2) the ν-luminosity
    follows a power law or broken power law distribution;
    (3) the sources are assumed to emit neutrinos with an
    E
    −2
    energy spectrum. Later, we discuss the robustness
    of the constraint by investigating the validity and caveats
    of the assumptions.
    The number of resolvable sources N
    s
    for a distribution
    of luminosities L
    ν
    per decade in energy is given by:
    N
    s
    ?
    3
    1
    r
    ln
    ³
    E
    max
    E
    min
    ´
    H
    0
    c
    K
    diff
    ν
    (C
    point
    )
    3/2
    ?L
    3/2
    ν
    ?
    ?L
    ν
    ?
    1
    ξ
    (1)
    where the parameter ξ depends on cosmology and source
    evolution as described in [9]. The ν-luminosity of the
    source, L
    ν
    has units of (erg/s), and (E
    min
    , E
    max
    )
    defines the energy range of the flux sensitivity, where
    E
    max
    = 10
    3
    E
    min
    for a typical experimental con-
    dition. For canonical energy spectrum proportional
    to E
    −2
    , we use the results for all-flavor diffuse
    flux limits presented in [7] to obtain the ν
    µ
    -diffuse
    flux: K
    diff
    ν
    ≡ E
    2
    Φ
    ν
    µ
    = (1/3) ∗ E
    2
    Φ
    ν
    all
    =
    (1/3) ∗ 8.4 × 10
    −8
    GeV cm
    −2
    s
    −1
    sr
    −1
    = 2.8 ×
    10
    −8
    GeV cm
    −2
    s
    −1
    sr
    −1
    valid for the energy interval
    of 1.6 PeV < E < 6.3 EeV. This is the energy interval of
    interest for CR interaction with energies above the ankle.
    Below PeV energies K
    diff
    ν
    can be obtained from [10],
    K
    diff
    ν
    < 7.4×10
    −8
    GeV cm
    −2
    s
    −1
    sr
    −1
    , valid between
    16 TeV to 2.5 PeV. So, similar diffuse flux limits,
    K
    diff
    ν
    , exist for the entire interval from TeV to EeV
    energies. C
    point
    is the experimental sensitivity to ν-
    fluxes from point sources for an E
    −2
    spectrum, and we
    used C
    point
    = E
    2
    (dN
    ν
    /dE) < 2.5 × 10
    −8
    GeV cm
    −2
    s
    −1
    [8].
    The diffuse flux K
    diff
    ν
    and the point flux sensitivity
    C
    point
    are linearly correlated by the following equation:
    4πK
    diff
    ν
    =
    .
    3
    μ
    c
    H
    0
    1
    r
    max
    N
    s
    ¸
    × C
    point
    (2)
    where (c/H
    0
    ) represents the Hubble distance given by
    c/H
    0
    = 3 × 10
    5
    (km s
    −1
    )/77 (km s
    −1
    Mpc
    −1
    ) ∼ 4

    2
    A. SILVESTRI AND S.W. BARWICK. FLUX CONSTRAINTS
    (GeV))
    ν
    μ
    10
    (E
    log
    3
    4
    5
    6
    7
    8
    9
    10
    11
    )
    -1
    s
    -2
    dN/dE (GeV cm
    2
    E
    10
    -10
    10
    -9
    10
    -8
    10
    -7
    UHE constraint (this work)
    45
    erg/s)
    ν
    = 10
    VHE constraint (L
    45
    erg/s)
    ν
    = 10
    (L
    40
    erg/s)
    ν
    = 10
    constraint (L
    40
    erg/s)
    ν
    = 10
    constraint (L
    AMANDA [8]
    IceCube 1-yr sensitivity [28]
    3C273 (SP92b)
    3C273 (N93)
    3C273 (M93)
    3C279 (AD04)
    NGC4151 (SP92)
    NGC4151 (SS96)
    Mkn 421 (SP92a)
    Mkn 501 (LM00)
    RQQ (AM04)
    GeV blazar (NS02)
    Cen A (AN04)
    Cen A (HO07)
    Cen A (CH08)
    M87 (AN04)
    Mkn 501 (MP01)
    3C273 (SS96)
    3C279 (SP92c)
    Coma (CB98)
    3C273 (SP92b)
    3C273 (N93)
    3C273 (M93)
    3C279 (AD04)
    NGC4151 (SP92)
    NGC4151 (SS96)
    Mkn 421 (SP92a)
    Mkn 501 (LM00)
    RQQ (AM04)
    GeV blazar (NS02)
    Cen A (AN04)
    Cen A (HO07)
    Cen A (CH08)
    M87 (AN04)
    Mkn 501 (MP01)
    3C273 (SS96)
    3C279 (SP92c)
    Coma (CB98)
    Fig. 1. Constraints on neutrino point fluxes derived from the UHE diffuse ν-flux limit [7], and from VHE limit [10], and assuming a range
    of neutrino luminosities L
    ν
    = (10
    40
    − 10
    45
    ) erg/s. Current AMANDA limit [8] and IceCube sensitivity [28] to ν-point fluxes are also shown
    (thin solid lines). Model predictions for ν
    µ
    -point flux from EG sources are displayed in thin dotted-dashed lines: emission from 3C273 predicted
    by [3C273 (SP 92)] [13], core emission due to pp interactions [3C273 (N93)] [14], including pp and pγ interactions [3C273 (M93)] [15];
    core emission due to pγ interaction [3C273 (SS96)] [24]; AGN jet, calculated for a 3C279 flare of 1 day period [3C279 (AD04)] [16] and
    continuous emission [3C279 (SP 92)] [13]; emission from NGC4151 by [NGC4151 (SP 92)] [13] and core emission from NGC4151 due
    to pγ interaction [NGC4151 (SS96)] [2]; Spectra predicted for Mkn 421 [Mkn 421 (SP 92)] [13], and for Mkn 501 during the outburst
    in 1997 [Mkn 501 (LM00)] [17] and blazar flaring Mkn 501 [Mkn 501 (MP 01)] [23]; radio-quiet AGN [RQQ (AM04)] [18] and
    GeV-loud blazars [GeV blazar (NS02)] [19]; emission from Cen A as described in [Cen A (AN04)] [20], [Cen A (HO07)] [21] and
    [Cen A (CH08)] [22]; emission from M87 [M87 (AN04)] [20], and emission from Coma galaxy cluster [Coma (CB98)] [25].
    Gpc, and r
    max
    defines the maximum observable distance
    for a point source of luminosity L
    ν
    , which is given by:
    r
    max
    =
    .
    L
    ν
    4π ln(E
    max
    /E
    min
    ) C
    point
    ¸
    1/2
    (3)
    The constraint also holds for time variable sources,
    since it depends only on the observed luminosity and
    is independent of the duration of the variability [11].
    Similarly, it holds for beamed sources, such as GRB’s.
    However for luminosities of the order of 10
    51
    erg/s
    typical of GRB emission, we found that a dedicated
    search for GRB’s leads to more restrictive limits [12].
    III. RESULTS
    We can now estimate a numerical value for N
    s
    by
    incorporating the ν-diffuse flux limit and the sensitivity
    to point sources in Eq. 1: N
    s
    ? (3.7 · 10
    −29
    cm
    −1
    ) ×
    (K
    diff
    ν
    ) × (C
    point
    )
    −3/2
    × (L
    45
    )
    1/2
    × 1/ξ ? 0.07
    computed assuming L
    45
    = 10
    45
    erg/s, and ξ = ξ
    AGN
    ?
    2.2 which defines the effects due to cosmology and
    source evolution that follows AGN [9]. The estimate for
    N
    s
    ? 0.07, which is compatible with the non-detection
    of any point sources.
    The constraint on ν-flux is determined by setting
    N
    s
    = 1 and inverting Eq. 1 to solve for C
    point
    :
    E
    2
    dN
    ν
    dE
    2
    66
    4
    3
    1
    r
    ln
    ³
    E
    max
    E
    min
    ´
    H
    0
    c
    · K
    diff
    ν
    p
    L
    ν
    ·
    1
    ξ
    3
    77
    5
    2/3
    E
    2
    dN
    ν
    dE
    ≤ 5.1 × 10
    −9
    μ
    L
    ν
    10
    45
    erg/s
    1/3
    μ
    GeV
    cm
    2
    s
    (4)
    valid for the same energy range 1.6 PeV < E < 6.3
    EeV of the diffuse flux limit K
    diff
    ν
    . This result defines a
    benchmark flux constraint Φ
    C
    ≡ E
    2
    (dN
    ν
    /dE) ≤ 5.1 ×
    10
    −9
    GeV cm
    −2
    s
    −1
    on neutrino fluxes from bright
    (L
    ν
    = 10
    45
    erg/s) extragalactic point sources, which
    is a factor five lower than present experimental limits
    from direct searches. Note from Eq. 2, that for the case
    of N
    s
    < 1 the distance ratio (c/H
    0
    )/r
    max
    > 1, which
    occurs for sources well within the Hubble distance.
    Fig. 1 shows these results represented by the con-
    straint derived from the Ultra High Energy (UHE) dif-
    fuse ν-flux limit for energies above PeV (thick solid
    line), and from the Very High Energy (VHE) limit in
    the TeV-PeV range (thick dotted line). Model predictions
    for ν
    µ
    -point flux from EG sources (dotted/dashed lines),

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁOD´ Z´ 2009
    3
    TABLE I
    SUMMARY OF MODELS FOR ν
    µ
    POINT FLUX FROM EXTRAGALACTIC SOURCES CONSTRAINED BY THE RESULTS FROM THIS WORK.
    MODELS ARE ORDERED ACCORDING TO THE TYPE OF THE NEUTRINO SOURCE. THE PARAMETER BAND
    γ
    REPRESENTS THE PHOTON
    ENERGY-BAND ASSUMED IN THE GIVEN MODEL. THE NEUTRINO FLUX PREDICTED BY A GIVEN MODEL FOR AN E
    −2
    SPECTRUM IS
    DENOTED BY Φ
    model
    ν
    AND NEUTRINO FLUXES FOR MODELS WHICH ARE ALMOST CONSTANT TO AN E
    −2
    SPECTRUM FOR A LARGE
    ENERGY RANGE. THE CORRESPONDING FLUX CONSTRAINED FOR AN E
    −2
    SPECTRUM IS DEFINED BY THE BENCHMARK FLUX Φ
    C
    . IF THE
    SOURCE IS COMMONLY REPLICATED IN THE UNIVERSE WITH OUR ASSUMPTIONS, THEN THE RATIO R
    flux
    = Φ
    C
    model
    ν
    < 1
    DETERMINES A MODEL CONSTRAINED BY THIS WORK.
    Model
    Band
    γ
    Φ
    model
    ν
    R
    flux
    Reference
    (GeV/cm
    2
    s)
    [3C273 (SP 92)]
    IR/x-ray
    1.0 × 10
    −8
    0.51
    [13]
    [3C273 (N93)]
    x-ray
    2.5 × 10
    −8
    0.20
    [14]
    [3C273 (M93)]
    γ-ray/IR
    1.0 × 10
    −8
    0.51
    [15]
    [3C279 (AD04)]
    GeV
    2.0 × 10
    −7
    0.03
    [16]
    [NGC4151 (SP 92)]
    IR/x-ray
    3.5 × 10
    −8
    0.14
    [13]
    [Mkn 421 (SP 92)]
    IR/x-ray
    9.0 × 10
    −9
    0.10
    [13]
    [Mkn 501 (LM00)]
    TeV
    2.5 × 10
    −8
    0.57
    [17]
    [RQQ (AM04)]
    x-ray/UV
    1.0 × 10
    −8
    0.51
    [18]
    [Cen A (AN04)]
    TeV
    1.5 × 10
    −8
    0.34
    [20]
    [Cen A (CH08)]
    TeV
    6.0 × 10
    −9
    0.85
    [22]
    have been tested by this analysis and are summarized in
    Tab. I.
    Tab. I summarizes the results from the constraint Φ
    C
    compared to a number of models of neutrino point fluxes
    from extragalactic sources. The fluxes Φ
    model
    ν
    predicted
    from these models can be directly compared to Φ
    C
    since either follow an E
    −2
    spectrum, or do cover a
    large energy range almost constant to an E
    −2
    spec-
    trum. These models are constrained since their predicted
    fluxes exceed the benchmark flux set by Φ
    C
    . If the
    source is commonly replicated in the universe with the
    assumptions defined in Sec II, then the ratio R
    flux
    =
    Φ
    C
    model
    ν
    < 1 determines a model constrained by this
    analysis.
    Models have also been presented which predict ν-
    fluxes from nearby AGNs [20], [21], [22], such as
    Centaurus A (Cen A) and M87 at a distance of 3.4 Mpc
    and 16 Mpc, respectively. We note these predictions lie
    below the upper value of the constraint Φ
    C
    , and are
    compatible with our results.
    A few other models, as shown in Fig. 1, present
    flux predictions which strongly deviate from an E
    −2
    spectrum and in this class of models a direct comparison
    with the benchmark flux Φ
    C
    is less straightforward.
    In these cases, the predicted energy spectra should
    be integrated over the energy interval that defines the
    constraint to obtain the total neutrino event rate, N
    model
    ν
    .
    This result should be compared to the integrated neutrino
    event rate N
    C
    determined by the constraint and by the
    given neutrino detector characteristics.
    IV. DISCUSSION
    The thick dark horizontal line in Fig. 1 indicates
    our primary constraint Φ
    C
    . It was derived for a mean
    neutrino luminosity that characterizes the brightest AGN
    in the EM band. The constraint is even stronger for
    less luminous classes of sources. In this section we
    address the robustness of the constraint by focusing the
    discussion on the three assumptions listed in Sec. II.
    A. Homogeneity of source distribution
    The matter distribution within 50 Mpc of the Milky
    Way is far from uniform, which suggests the possibility
    that the number density of neutrino sources, n
    s
    , may
    be higher than the universal average if n
    s
    is correlated
    with matter density. We argue that, in practice, the
    local inhomogeneity affects only the class of sources
    characterised by low luminosities. The bright sources are
    too rare to be affected by local matter density variation
    - the likelihood of finding a bright neutrino source
    within 50 Mpc is small to begin with (if EM luminosity
    and neutrino luminosity are comparable), and the local
    enhancements in matter density insufficient to change
    the probability of detection.
    On the other hand, low luminosity sources are more
    likely to be within 50 Mpc, and their density could be
    affected by fluctuations (e.g. by a factor of 15 [26] at 5
    Mpc) in the local matter density. In this case, the flux
    constraint (Eq. 4) should be adjusted to account for the
    higher density of local matter, Φ
    ?
    = Φ∗(n
    local
    /?n
    s
    ?)
    2/3
    (Tab II). However, the adjusted fluxes are below the
    benchmark flux constraint Φ
    C
    .
    To exceed Φ
    C
    a source of a given luminosity L
    ν
    must
    be within a distance d
    l
    = (4π/3)
    1/3
    ·r
    max
    ∗(Φ
    ?
    C
    )
    1/2
    .
    However, we found no counterparts in the EM band
    within a distance d
    l
    that would violate the constraint
    Φ
    C
    .
    TABLE II
    ADJUSTED Φ
    ?
    TO ACCOUNT FOR LOCAL n
    s
    ENHANCEMENT.
    L
    ν
    Φ
    n
    l
    /?n
    s
    ?
    Φ
    ?
    d
    l
    erg/s
    GeV/2s
    [26]
    GeV/cm2s
    Mpc
    8 × 10
    41
    0.5 × 10
    −9
    15
    2.8 × 10
    −9
    3.7
    1 × 10
    43
    1.1 × 10
    −9
    5
    3.1 × 10
    −9
    16
    1 × 10
    44
    2.4 × 10
    −9
    2.5
    4.3 × 10
    −9
    55
    B. Distribution function of ν-luminosity
    The number of detectable sources N
    s
    depends on
    ?L
    3/2
    ν
    ?/?L
    ν
    ?, but the luminosity distribution for neu-
    trino sources is not known. However, if the distribution

    4
    A. SILVESTRI AND S.W. BARWICK. FLUX CONSTRAINTS
    function follows a broken power law, which is measured
    for several class of energetic sources in various electro-
    magnetic bands, then the estimate for N
    s
    based on a
    full distribution agrees with an estimate using the mean
    luminosity of the distribution to within few percent, as
    shown in [11]. So, to an excellent approximation, the
    mean value of the luminosity distribution is sufficient
    to predict N
    s
    ∼ ?L
    ν
    ?
    1/2
    for power law or broken
    power law distributions. The most common distribution
    of luminosities can only be observed at relatively small
    distances, so source evolution and cosmological effects
    are negligible. Larger values of luminosity are too rare
    to contribute significantly.
    C. Energy spectrum of the source
    The constraint can be extended to energy spectra
    that differ from the assumed E
    −2
    dependence, but the
    constraint applies over a restricted energy interval that
    matches the energy interval of the diffuse neutrino limits.
    Experimental diffuse limits span two different energy
    regions, VHE and UHE, and either limit can be inserted
    into Eq. 4. The restriction in energy range is required to
    avoid extrapolating the energy spectrum to unphysical
    values. In other words, for power law indices far from
    2, the spectrum must cut-off at high energies for indices
    γ < 2, or at low energies for indices γ > 2. Subject
    to this restriction, we find that the constraint depends
    weakly on the assumed spectral index. For example,
    the constraints improve by a factor 2 for hard spectra
    (γ = 1) and weaken by roughly the same factor for soft
    spectra (γ = 3) [11].
    On the other hand, it could be argued that the energy
    spectrum dN
    ν
    /dE is completely unknown. In this case,
    instead of relying on the power law of neutrino luminos-
    ity L
    ν
    , one could derive the constraints by examining the
    measured number density n
    s
    , (n
    s
    ∝ 1/L
    ν
    ) for a given
    class of sources [27].
    V. CONCLUSION
    The constraint on neutrino fluxes from extra-
    galactic point sources is E
    2
    (dN
    ν
    /dE) ≤ 5.1 ×
    10
    −9
    (L
    ν
    /10
    45
    erg/s)
    1/3
    GeV cm
    −2
    s
    −1
    , which is a
    factor 5 below current experimental limits from direct
    searches if the average L
    ν
    distribution is comparable to
    the EM luminosity that characterizes the brightest AGN.
    We tested a number of model predictions for ν-point
    fluxes, and models which predict fluxes higher than the
    constraint have been restricted by this analysis.
    The constraint is strengthened for less luminous sources,
    and noncompetitive with direct searches for highly lu-
    minous explosive sources, such as GRB. We found
    that the constraint is robust when accounting for the
    non-uniform distribution of matter that surrounds our
    galaxy, or considering energy spectra that deviate from
    E
    −2
    , or various models of cosmological evolution. The
    constraint suggests that the observation of EG neutrino
    sources will be a challenge for kilometer scale detectors
    unless the source is much closer than the characteristic
    distance between sources, d
    l
    , after accounting for local
    enhancement of the matter density. Although the con-
    straint cannot rule out the existence of a unique, nearby
    EG neutrino sources, we note that we found no counter-
    parts in the EM band with the required luminosity and
    distance to violate the constraint, assuming L
    ν
    ∼ L
    γ
    .
    VI. ACKNOWLEDGMENT
    The authors acknowledge support from U.S. National
    Science Foundation-Physics Division, and the NSF-
    supported TeraGrid system at the San Diego Supercom-
    puter Center (SDSC).
    REFERENCES
    [1] J. Abraham et al., Science 318, 938 (2007).
    [2] F. Stecker et al., Phys. Rev. Lett. 66, 2697 (1991); erratum Phys.
    Rev. Lett. 69, 2738 (1992).
    [3] F. Stecker, Phys. Rev. D 72, 107301 (2005).
    [4] K. Mannheim, Astropart. Phys. 3, 295 (1995).
    [5] F. Halzen and E. Zas, Astrophys. J. 488, 669 (1997).
    [6] R. J. Protheroe, astro-ph/9607165 (1996).
    [7] A. Silvestri et al., Proc. 31th ICRC, these Proceedings (2009).
    [8] R. Abbasi et al., Phys. Rev. D 79, 062001 (2009).
    [9] P. Lipari, Nucl. Instrum. Meth. A 567, 405 (2006).
    [10] A. Achterberg et al., Phys. Rev. D 76, 042008 (2007).
    [11] A. Silvestri, Ph.D. Thesis, University of California, Irvine (2008).
    [12] A. Achterberg et al., Astrophys. J. 674, 357 (2008).
    [13] A. P. Szabo and R. J. Protheroe, in High Energy Neutrino
    Astrophysics, World Scientific (Singapore), 24 (1992).
    [14] L. Nellen et al., Phys. Rev. D 47, 5270 (1993).
    [15] K. Mannheim, Phys. Rev. D 48, 2408 (1993).
    [16] A. Atoyan and C. Dermer, New Astron. Rev. 48, 381 (2004).
    [17] J. G. Learned, and K. Mannheim, Annu. Rev. Nucl. Part. Sci. 50,
    679 (2000).
    [18] J. Alvarez-Muniz and P. Meszaros, Phys. Rev. D 70, 123001
    (2004).
    [19] A. Neronov and D. Semikoz, Phys. Rev. D 66, 123003 (2002).
    [20] L. A. Anchordoqui et al., Phys. Lett. B 600, 202 (2004).
    [21] F. Halzen and A. O’Murchadha, astro-ph/0802.0887 (2008).
    [22] A. Cuoco and S. Hannestad, Phys. Rev. D 78, 023007 (2008).
    [23] A. Mu¨cke and R. J. Protheroe, Astropart. Phys. 15, 121 (2001).
    [24] F. Stecker and M. H. Salamon, Space Sci. Rev. 75, 341 (1996).
    [25] S. Colafrancesco and P. Blasi, Astropart. Phys. 9, 227 (1998).
    [26] T. Jarrett, private communication.
    [27] M. Kowalski, private communication.
    [28] J. Ahrens et al., Astropart. Phys. 20, 507 (2004).

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁOD´ Z´ 2009
    1
    Study of electromagnetic backgrounds in the 25-300 MHz
    frequency band at the South Pole
    Jan Auffenberg
    ¤
    , Dave Besson
    y
    , Tom Gaisser
    z
    , Klaus Helbing
    ¤
    , Timo Karg
    ¤
    , Albrecht Karle
    x
    ,
    and Ilya Kravchenko
    {
    ¤
    Bergische Universitat¨ Wuppertal, Fachbereich C- Astroteilchenphysik, 42097 Wuppertal, Germany
    y
    Dept. of Physics & Astronomy, University of Kansas, Lawrence, KS 66045, USA
    z
    Bartol Research Institute, University of Delaware, Newark, DE 19716, USA
    x
    Dept. of Physics, University of Wisconsin, Madison, WI 53706, USA
    {
    Dept. of Physics & Astronomy, University of Nebraska, Lincoln, NE 68588, USA
    Abstract. Extensive air showers are detectable
    by radio signals with a radio surface detector. A
    promising theory of the dominant emission process
    is the coherent synchrotron radiation emitted by e+
    e- shower particles in the Earth’s magnetic field
    (geosynchrotron effect). A radio air shower detector
    can extend IceTop, the air shower detector on top
    of IceCube. This could increase the sensitivity of
    IceTop to higher shower energies and for inclined
    showers significantly. Muons from air showers are
    a major part of the background of the neutrino
    telescope IceCube. Thus a surface radio air shower
    detector could act as a veto detector for this muonic
    background. Initial radio background measurements
    with a single antenna in 2006 revealed a continuous
    electromagnetic background promising a low energy
    threshold of radio air shower detector. However,
    short pulsed radio interferences can mimic real sig-
    nals and have to be identified in the frequency range
    of interest. These properties of the electromagnetic
    background are being measured at the South Pole
    during the Antarctic winter 2009 with two different
    types of surface antennas. In total four antennas
    are placed at distances ranging up to 400m from
    each other. They are read out using the RICE DAQ
    with an amplitude threshold trigger and a minimum
    bias trigger. Results of the first three months of
    measurement are presented.
    Keywords: Radio air shower detection, EMI back-
    ground, South Pole
    I. INTRODUCTION
    The emission of coherent synchrotron radiation by
    e
    +
    e
    ¡
    shower particles in the Earth magnetic field
    provides a measurable broadband signal from 10 MHz-
    150 MHz on ground [1]. The South Pole site with
    its dedicated infrastructural environment and a limited
    number of radio sources is possibly one of the best
    places in the world for the detection of air showers by
    their low frequency radio emission. Another feature of
    the South Pole site in comparison to other radio quiet
    regions in the world is the possibility to make studies
    in coincidence with other astrophysical experiments like
    Fig. 1. Schematic view of a high energy radio air shower detector
    expanding the active area of IceTop. The distance between single
    antenna stations can be several hundred meters.
    the neutrino detector IceCube and the air shower detector
    IceTop. IceCube is a neutrino detector [2] embedded in
    the Antarctic ice. One of the main aims of IceCube is
    to measure neutrinos from cosmic sources. The strategy
    of IceCube is to measure up-going particles from the
    northern hemisphere. Only neutrinos or other weakly
    interacting particles are not absorbed by the Earth and
    are able to interact in the South Pole ice and produce
    measurable particles like muons. IceTop is built on the
    surface above IceCube (Fig. 1). It is designed to detect
    cosmic air showers from 10
    15
    eV up to 10
    18
    eV.
    This special environment leads to two different op-
    tions with different focus for a radio air shower detector
    on top of IceCube and its ambit [3].
    ² An infill detector built up of radio antennas on
    the same footprint as IceTop in similar distances
    but shifted with respect to the tank array. This
    provides an additional powerful observation tech-
    nique in cosmic ray research of air showers at
    the South Pole. It would be possible to study air
    showers by three independent detector systems,
    IceTop, IceCube and the radio detector.
    ² An areal expansion of IceTop with radio surface
    antennas is an extension of the air shower detector

    2
    JAN AUFFENBERG et al. ELECTROMAGNETIC BACKGROUND AT THE SOUTH POLE
    to higher energy primary particles and to higher
    inclination angles [6]. The idea is to build an
    antenna array in rings with increasing radius around
    the IceTop array. For ultra-high energetic neutrinos,
    E > 300 TeV, the neutrino nucleon cross section is
    large enough for absorption in the Earth to become
    increasingly important. For cosmogenic neutrinos,
    produced by the GZK mechanism, for example,
    most of the signal comes from near the horizon.
    Thus Muon bundles induced by air showers can be
    misinterpreted as a neutrino signal in the IceCube
    detector. The role of an antenna array field expan-
    sion of the IceTop detector is to detect these air
    showers with high inclination angle as a veto for
    the IceCube detector.
    II. ELECTROMAGNETIC BACKGROUND
    MEASUREMENTS AT THE SOUTH POLE WITH RICE
    Initial background measurements with a single an-
    tenna in 2006 indicated a continuous electromagnetic
    background promising a low detector threshold [3].
    Together with air shower simulations of the radio emis-
    sion the background measurements seem to allow the
    detection of air showers with a threshold lower than
    10 PeV in primary energy. The study presented here is
    aimed at long-term studies of the electromagnetic back-
    ground for several months to investigate the amount of
    pulsed radio frequency interference (RFI) and potential
    long-term variations in the continuous background. The
    data acquisition of the RICE experiment, constructed to
    investigate radio detection methods of high energy neu-
    trinos in ice [7], is suited to be extended by four surface
    antennas. The RICE DAQ consists of 6 oscilloscopes
    with 4 channels each. The sampling rate of each channel
    is 1GHz. The dynamic range of the scopes is §2 V with
    12 Bit digitization.
    Three different kinds of trigger are implemented in
    RICE:
    1) Unbiased events every 10 min which is a forced
    read out of all channels.
    2) The RICE simple multiplicity trigger. It is read
    out if the signal in four or more RICE antennas
    is above a threshold. The threshold is calculated
    at the beginning of every run to be 1.5 times
    above the RMS of unbiased events. These events
    should have no signal over threshold in the surface
    antennas.
    3) The RICE surface trigger. This is a RICE simple
    multiplicity trigger with one or more surface an-
    tennas above the threshold as part of the trigger.
    This includes RICE triggers where only surface
    antennas have a signal over threshold.
    The first and the third kind of trigger are of great
    interest for the surface radio background studies. The
    second trigger strategy is interesting to understand
    the in-ice RFI not reaching the surface. It is the most
    interesting event class for the RICE neutrino detection.
    Fig. 2.
    Top view of the IceCube footprint. Two Fat Wire Dipole
    antennas were deployed in 350 m distance from the MAPO building
    (crosses near the SPASE building). The signals are amplified with
    60dB MITEQ AU-4A-0150 low noise amplifiers and connected with
    RG59 cables to the MAPO building. Two four arm dipole antennas are
    located on the roof of the MAPO building and amplified with 39 dB
    MITEQ AU-1464-400.
    Fig. 3. Comparison of results from antenna simulations and measured
    properties (DATA). The data is a measurement of the voltage standing
    wave ratio (VSWR) of the Fat Wire Dipole on the South Pole snow
    at the final position of the antenna. The Simulation is made with
    EZNEC+ v. 5.0 without ground effects from the snow surface. The
    frequency response of the antenna is well described by the simulation.
    Including ground effects should even improve the agreement between
    simulation and data.
    In total four surface antennas were deployed in the
    South Pole season 2008/2009 on the IceCube footprint
    (Fig. 2). Two Fat Wire Dipole antennas (Fig. 4) are con-
    nected to RICE with RG59 signal cables (1505A Coax)
    of the decommissioned SPASE experiment (Fig. 2).
    These broad band antennas allow for measurements

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁOD´ Z´ 2009
    3
    TABLE I
    ANTENNA POSITION AND CABLE DELAYS TO THE RICE DAQ.
    Antenna
    x [m]
    y [m]
    z [m]
    cable delay [ns]
    MAPO1
    47
    -28.0
    18
    144
    MAPO2
    25
    -20
    18
    129
    SPASE1
    -135
    -366
    1
    2126
    SPASE2
    -148
    -348
    1
    2729
    Fig. 4. Conducting elements of the Fat Wire Dipole Antenna deployed
    near the SPASE building. It is 3 m long and 0.8 m in diameter. The
    wires are mounted to a wooden carcass.
    in the frequency range from 25-500 MHz. Figure 3
    shows measurements of the voltage standing wave ratio
    (VSWR) of a Fat Wire Dipole lying on the Antarctic
    snow in its final position compared to simulations of
    the antenna using the NEC2 software package [4].
    The frequency response is already well described by
    simulations without snow ground. The connection with
    the SPASE cables allows measurements over distances
    of several hundred meters from the MAPO building
    on the IceCube footprint, housing electronics, which
    is a potentially large RFI source, and the antennas
    on its roof. To compensate for the attenuation of the
    long signal cable (ca. 40 dB at 75 MHz) 60dB low
    noise preamplifiers (MITEQ AU-4A-0150) are imple-
    mented between antenna and signal cable. The power
    is transmitted through the same cable using bias tees.
    To avoid saturation of the preamplifiers by the input
    power, 25 MHz high pass filters and 300 MHz low
    pass filters were implemented between amplifiers and
    Fat Wire Dipole antenna. The antennas near the SPASE
    building are lying on the snow surface orthogonal to
    each other. Thus they measure orthogonal polarization
    of the signals. The other two antennas are deployed
    on the roof of the MAPO building. These four arm
    dipole antennas with an amplification of about -2 dB
    at >70 MHz are difficult to calibrate in the surrounding
    of the MAPO building and thus will only be used for
    event reconstruction (Fig. 6). The signal of the antennas
    on the roof of the MAPO building are amplified with
    39 dB preamplifiers (AU-1464-400). 300 MHz low pass
    Filters are used for these roof antennas, too. Table I
    shows the position of the surface antennas in AMANDA
    coordinates and their cable delays.
    Fig. 5. Picture of the four arm dipole antennas MAPO1 and MAPO2.
    Every arm has a length of 0.7 m. The wooden stand is 1.2 m high.
    The final position of the antennas is the roof of the MAPO building.
    III. ANALYSIS STRATEGIES
    From the technical point of view one can divide the
    analysis of the data into two parts.
    A. Event Reconstruction And Mapping
    The reconstruction of the origin of RFI events seen
    in more than two antennas will indicate possible noise
    sources at the South Pole e.g. the IceCube counting
    house or the South Pole Station. A map of these sources
    will help to improve air shower detection.
    A Â
    2
    minimization on time residuals is used to recon-
    struct the source location of single events. To test the
    event reconstruction algorithm, we use signals generated
    with a GHz horn antenna in front of the MAPO building.
    Considering the cable delays and the antenna positions
    (Table I) we are able to make a 3D and time reconstruc-
    tion of the events. Figure 6 shows the reconstruction
    with horn antenna signal data is working well. It is
    accurate within several meters and shows clearly the
    horn antenna lies in front of the MAPO building near
    the antenna MAPO1. The horn antenna data reconstructs
    to the actual position within 50 m with an RMS of
    2.3 m. Figure 7 shows a typical noise event triggered
    with the RICE surface trigger. The event can be nicely
    reconstructed to have its origin in the building of the
    10 m Telescope which is the topmost building in Fig. 2.
    B. Events and background in the frequency domain
    Another Topic is the rate and variation of the different
    RFI sources during a whole year of measurement. It
    is expected that RFI events have a typical signature in
    the frequency domain. This will help to find an ideal
    frequency region for a radio air shower detector. The
    continuous background is monitored over a whole year
    using the unbiased RICE events. One of the highest
    peaks on top of the continuous radio background is ex-
    pected to be the meteor radar at 46.6 MHz to 47.0 MHz
    and 49.6 MHz to 50.0 MHz [5]. It is clearly observable
    in the DFT of the recorded data together with a few
    other expected sources of filterable continuous narrow

    4
    JAN AUFFENBERG et al. ELECTROMAGNETIC BACKGROUND AT THE SOUTH POLE
    Fig. 6. Distribution of the reconstructed transmitter events in the xy-
    plane. The four crosses indicate the position of the antennas on the
    roof of the MAPO building (18m above the snow) and on the snow
    surface near the SPASE building. The dots are reconstructed positions
    of triggered signals from a GHz horn antenna, measured with all four
    antennas. The reconstructed events are in very good agreement to the
    transmitter position in front of the MAPO1 antenna and demonstrate
    the potential of the instrument.
    Fig. 7. Example of a triggerd event, seen by all four surface antennas
    in the time domain. The signal of the event reconstructs to be coming
    from the building of the 10m telescope.
    band RF signals.
    The DFT of the constant background measured with
    the fat wire dipoles is the basis to evaluate the limit
    of detectable signal strength. For this it is of great
    importance to correct the measured data for the antenna
    properties, the high- and low pass filtering, the amplifier
    response, and the attenuation of the signal cable. Another
    important topic will be to determine long term variations
    of the background during one year.
    RICE triggered surface events are studied in the fre-
    quency domain whether a discrimination of air shower
    radio signals from RFI noise is feasible. Most of the
    narrow band noise events could e.g. be filtered in a future
    air shower detector system.
    IV. CONCLUSION
    As a part of RICE the four antenna surface detection
    system for radio signals, is able to study the conditions
    of the radio background in the frequency range from
    25-150 MHz and higher at the South Pole. The thresh-
    old trigger strategy together with RICE allows for the
    estimation of the amount of RFI noise and its sources
    on the IceCube site. An analysis of the signals in the
    frequency domain shall be used to develop strategies
    to suppress the false trigger rate of a radio air shower
    detector. Measurements of the continuous background
    and its variations are the basis to estimate the energy
    threshold of a radio air shower detector in different
    frequency bands. The RFI measurements of the surface
    antennas will help to understand the signals measured
    with in ice radio detection systems.
    REFERENCES
    [1] T. Huege, H. Falcke, Astron. & Astrophys. 412, 1934, (2003).
    [2] http://www.icecube.wisc.edu.
    [3] J. Auffenberg et al. ICRC arXiv:0708.3331 (2007).
    [4] http://www.nec2.org.
    [5] E. M. Lau et al. Radio Sci. 41, RS4007, (2006).
    [6] J. Auffenberg et al. Arena, doi:10.1016/j.nima.2009.03.179
    (2008).
    [7] I. Kravchenko et al. Phys. Rev. D 73, 082002 (2006).

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    1
    Neutrino signal from
    γ
    -ray loud binaries powered by high-energy
    protons
    Andrii Neronov
    ∗†
    and Mathieu Ribordy
    INTEGRAL Science Data Centre, Ch. d’Ecogia 16, 1290, Versoix, Switzerland
    Geneva Observatory, University of Geneva, 51 ch. des Maillettes, 1290, Sauverny, Switzerland
    Ecole Politecnique Federale de Lausanne, 1015, Ecublens, Switzerland
    Abstract
    . We consider hardonic model of activity of
    Galactic
    γ
    -ray-loud binaries. We show that in such a
    model multi-TeV neutrino flux from the source can
    be much higher and/or harder than the detected TeV
    γ
    -ray flux. This is related to the fact that most of the
    neutrinos can be produced in
    pp
    interactions close to
    the bright massive star, in a region which is optically
    thick for the TeV
    γ
    -rays. Considering the example of
    LS I +61
    303, we show that the expected neutrino
    signal, detectable within
    ∼ 3
    years of exposure with
    ICECUBE, will be marginally sufficient to constrain
    the spectral characteristics of neutrino signal.
    Keywords
    : Neutrinos, gamma-ray astronomy.
    I. INTRODUCTION
    Recently discovered ”
    γ
    -ray-loud” binaries (GRLB)
    form a new sub-class of Galactic binary star systems
    which emit GeV-TeV
    γ
    -rays. These are high-mass X-
    ray binaries (HMXRB) composed of a compact object
    (a black hole or a neutron star) orbiting a massive star.
    The detection of
    γ
    -rays with energies up to 10 TeV
    from these systems shows that certain HMXRBs host
    powerful particle accelerators producing electrons and/or
    protons with energies above 10 TeV.
    Different theoretical models of
    γ
    -ray activity of
    HMXRBs (see [1] for a review) have at least one
    common point: they are all based on the assumption that
    the
    γ
    -ray emission is produced in result of interaction of
    a relativistic outflow from the compact object (jet from
    a black hole, or wide angle wind from a pulsar) with the
    wind and radiation emitted by the companion massive
    star. With the exception of Cyg X-1, all the known
    GRLBs have similar spectral energy distributions (SED),
    peaking in the MeV-GeV energy band. The physical
    mechanism of production of the MeV-GeV bump in the
    spectra is not clear. It can be the synchrotron emission
    from electrons with the energies much above TeV [2],
    [3]. Otherwise, it can be produced via inverse Compton
    (IC) scattering of the UV thermal emission from the
    massive star by electrons of the energies
    E ∼ 10
    MeV
    [4], [5].
    The available multi-wavelength data do not allow
    to constrain the composition of the relativistic outflow
    from the compact object. On one side, the multi-TeV
    or 10 MeV electrons, responsible for the production
    of the MeV-GeV bump in the SED, could be injected
    into the emission region from the
    e
    +
    e
    pairs loaded
    wind. Otherwise, these electrons can be secondary par-
    ticles produced in e.g. proton-proton collisions, if the
    relativistic wind is proton-loaded. The only direct way
    to test if relativistic protons are present in the
    γ
    -ray
    emission region would be detection of neutrinos of the
    muli-TeV energies. A general problem for the estimates
    of expected neutrino flux is the uncertainty of the atten-
    uation of the
    γ
    -ray flux in the TeV band, which makes
    the derivation of the estimate of the neutrino flux and
    spectral characteristics based on the observed TeV
    γ
    -ray
    emission highly uncertain (see [6] for the discussion of
    the particular case of LS I +61
    303).
    In the absence of direct relation between the character-
    istics of the observed TeV
    γ
    -ray and neutrino emission
    from a GRLB, the only way to constrain possible
    neutrino signal from the source is via detailed modelling
    of the broad band spectrum of the source within the
    hadronic model of activity. The idea is that the
    pp
    interactions, which result in the production of neutrinos,
    also result in production of the
    e
    +
    e
    pairs, which release
    their energy via synchrotron, IC and bremsstrahlung
    emission. The total power released in the
    pp
    interactions
    determines the overall luminosity of emission from the
    secondary pairs. The knowledge of the electromagnetic
    luminosity can be used to constrain the power released
    in
    pp
    interactions and hence the neutrino luminosity of
    the source.
    In this contribution we implement this method of
    estimation of neutrino flux to explore the possibility of
    detection of neutrino signal from GRLBs. We concen-
    trate mostly on the particular example of LS I +61
    303
    system, because it is the only known persitent GRLB
    in the Northern hemisphere, available for observations
    with ICECUBE neutrino telescope [7].
    II. HADRONIC MODEL OF
    γ
    -RAY ACTIVITY
    In the hadronic model, the primary source of the high-
    energy activity of the system are high-energy protons.
    The presence of the high-energy protons in relativistic
    outflows from compact objects (stellar mass and super-
    massive black holes, neutron stars) is usually difficult
    to detect, because of the very low energy loss rates of
    protons. GRLBs provide a unique possibility to ”trace”
    the presence of protons/ions in the relativistic outflows

    2
    NERONOV AND RIBORDY, NEUTRINO SIGNAL FROM BINARIES
    p
    p
    ν
    ν
    Be star
    Be star disk
    Observer
    Fig. 1. Mechanism of production of neutrinos in interactions of high-
    energy protons ejected by the compact object with the dense equatorial
    disk of Be star.
    generated by compact objects. The dense matter and ra-
    diation environment, created by the companion massive
    star provides abundant target material for the protons
    in the relativistic outflow. Interactions of high-energy
    protons with the ambient matter and radiation fields,
    created by the presence of a bright massive companion
    star, lead to the production and subsequent decays of
    pions. This results in emission of neutrinos and
    γ
    -
    rays from the source and to the deposition of
    e
    +
    e
    pairs throughout the proton interaction region. Radiative
    cooling of the secondary
    e
    +
    e
    pairs leads to the broad-
    band synchrotron and IC emission from the source.
    Interactions of the high-energy protons with the radia-
    tion field produced by the bright massive star in the sys-
    tem (e.g. a Be star with the temperature
    T
    ∼ 3 × 10
    4
    K
    in the case of LS I +61
    303 and PSR B1259-63)
    are efficient only for protons with the energies above
    E
    p
    ≥ [200
    MeV
    ]m
    p
    ≃ 2 × 10
    16
    /10
    eV] eV,
    where
    ǫ
    ≃ 3kT
    is the typical energy of photons of
    the stellar radiation. To the contrary, interactions of the
    high-energy protons with the protons from the dense
    stellar wind can be efficient for the protons of much
    lower energies.
    In the case when the massive star is a Be-type star,
    the rate of
    pp
    interactions can be highly enhanced if
    the high-energy protons are able to penetrate into the
    dense equatorial disk known to surround this type of
    stars. An obstacle for the penetration of the high-energy
    protons accelerated e.g. close to the compact object
    into the disk could be the presence of magnetic field,
    which would deviate proton trajectories away from the
    disk. However, the Larmor radius of the highest energy
    protons,
    R
    L
    ≃ 4 × 10
    12
    ?
    E
    p
    /10
    15
    eV
    ?
    [B/1
    G]
     1
    cm,
    where
    B
    is the magnetic field, could be comparable
    to the size of the system. Thus, if the magnetic field
    in the region of contact between the stellar wind and
    the relativistic outflow is not larger than several Gauss,
    the highest energy protons can freely penetrate into the
    dense stellar wind region.
    pp
    interactions result in the production of pions,
    which subsequently decay onto
    γ
    -rays, neutrinos and
    electrons/positrons. If the Larmor radius of the high-
    est energy protons is comparable to the size of the
    disk, most of the pions are produced by the protons
    propagating toward the companion star. In this case the
    neutrino and
    γ
    -ray emission from the pion decays is
    expected to be anisotropic, with most of the neutrino
    flux emitted toward the massive star, as it is shown in
    Fig. 1. To the contrary, the synchrotron and IC emission
    from the
    e
    +
    e
    pairs is, most probably, isotropized at
    the energies at which the radiative cooling time of
    electrons becomes longer than the period of giration in
    the magnetic field. Difference in the anisotropy patterns
    of neutrino emission and of the broad band emission
    from the secondary
    e
    +
    e
    pairs should, in principle,
    lead to significant difference in the expected orbital
    lightcurves of neutrino and electromagnetic emission
    from the source.
    The flux of
    γ
    -rays from the
    pp
    interaction region
    is absorbed due to the pair production on the ultra-
    violet photon background in the vicinity of Be star.
    Maximal optical depth with respect to the pair pro-
    duction is achieved at the energies
    E
    γ
    ≃ 4m
    2
    e
    0.2
    ?
    T
    /3 × 10
    4
    K
    ?
     1
    TeV, where the pair produciton
    cross section reaches the maximum
    σ
    γ γ
    ≃ 1.5 ×
    10
     25
    cm
    2
    . The optical depth for the
    γ
    -rays of this
    energy produced close to the masive star can be about
    τ
    γγ
    ≥ 10
    . At higher
    γ
    -ray energies,
    E
    γ
    T
    ≫ (m
    e
    c
    2
    )
    2
    ,
    the pair production cross-section and, respectively, the
    optical depth decrease as
    E
     1
    ln
    E
    . The attenuation of
    the
    γ
    -ray flux due to the pair produciton becomes small
    only at the energies above
    ∼ 10
    TeV.
    The power of the absorbed
    γ
    -rays is re-distributed
    to the secondary
    e
    +
    e
    pairs, which subsequently loose
    their energy onto the synchrotron and inverse Compton
    emission. Depending on the magnetic field strength in
    the pair production region, the bulk of electromag-
    netic emission from the secondary pairs of the energies
    10 GeV-10 TeV can be either re-emitted back in the
    GeV-TeV energy band, if the inverse Compton loss
    dominates, or in the X-ray band, in the case of the
    dominant synchrotron loss.
    The results of numerical modeling of the spectra
    of (isotropic) emission from the secondary
    e
    +
    e
    pairs
    produced in
    pp
    and
    γγ
    interactions are shown in Fig.
    2. Our numerical code follows evolution of the spectra
    of the secondary particles, produced in interactions of
    the high-energy protons with the stellar wind protons
    from the dense equatorial disk of Be star. We assume
    that the high-energy protons are initially injected close
    to the massive star and then escape, together with the
    secondary particles produced in
    pp
    interactions, toward
    larger distances. The primary proton injection spectrum
    is assumed to be hard, with
    p
    ≃ 0
    , so that most
    of the protons have the energy close to the cut-off
    energy assumed to be
    E
    cut
    = 10
    15
    eV. Such an almost
    monochromatic spectrum of protons injected into the
    stellar wind can be produced either if the high-energy
    protons originate from a ”cold” relativistic wind with the
    bulk Lorentz factor
    ∼ 10
    6
    , or because only the highest
    energy protons have large enough Larmor radii to be
    able to penetrate deep into the stellar wind. The magnetic

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    3
    XMM
    INTEGRAL
    BATSE
    OSSE
    COMPTEL
    EGRET
    MAGIC
    VERITAS
    Fig. 2. Broad band spectrum of emission from secondary
    e
    +
    e
    pairs
    produced in
    pp
    interactions close to the surface of Be star, calculated
    assuming almost monochromatic proton injection spectrum with
    p
    =
    0
    ,
    E
    p,cut
    = 10
    15
    eV. Upper panel shows the injection spectrum of
    e
    +
    e
    pairs (dashed line) and the spectrum formed in result of cooling
    via synchrotron, IC and bremsstrahlung emission as well as Coulomb
    energy loss. In the lower panel, thin red solid, dashed and dotted lines
    show, respectively, the synchrotron, IC and Bremsstrahlung emission
    from the pairs. The black thick solid line shows the overall broadband
    model spectrum.
    field is supposed to be characterized by the radial profile
    B = B
    0
    (D/R
    )
     α
    B
    with
    α
    B
    = 1
    and
    B
    0
    = 5
    G.
    Fig. 3 shows modifications of the spectrum introduced
    by (a) possible presence of an anisotropic
    γ
    -ray emission
    component produced by the
    π
    0
    decays and (b) by the
    strong anisotropic attenuation of the
    γ
    -ray flux by the
    pair production. Both effects are expected to be strongly
    variable with orbital phase. It is not possible to estimate
    the real value of average
    τ
    γγ
    as long as the details of
    the 3-dimenisonal geometry of the emission region and
    the mutual orientation of the extended emission region,
    of the Be star and of the observer are not known. Taking
    into account this uncertainty, we choose the minimal
    value of
    τ
    γγ
    at which the absorbed spectrum does not
    violate the upper bound on the flux at the energies above
    ∼ (several)
    TeV, found in VERITAS observations of
    the source.
    III. ESTIMATE OF THE NUMBER OF NEUTRINO
    EVENTS FOR ICECUBE
    Within the hadronic models of activity, the MeV-GeV
    bump in the spectral energy distribution of GRLBs is
    produced by the emission from the secondary
    e
    +
    e
    pairs
    from the
    pp
    interactions. This fact enables an estimate
    of the neutrino luminosity of the source, based on its
    MeV-GeV band luminosity. The only uncertainty of such
    an estimate is that the synchrotron emission from the
    secondary
    e
    +
    e
    pairs in the MeV-GeV energy band
    is, most probably, isotropic, while the neutrino and
    π
    0
    decay
    γ
    -ray emissions are not.
    XMM
    INTEGRAL
    BATSE
    OSSE
    COMPTEL
    EGRET
    MAGIC
    VERITAS
    Fig. 3.
    Broad band spectrum of emission from
    pp
    interactions,
    calculated assuming the same parameters as in Fig. 2, but considering
    the possibility of strong neutrino emission from the source. Notations
    for the
    γ
    -ray spectrum are the same as in the lower panel of Fig.
    2. Green thick solid line shows the spectrum of neutrinos.The black
    thick solid and dashed lines show the overall broadband model spectra,
    calculated Solid line shows the case of almost radial escape of the
    γ
    -
    rays, dashed line corresponds to the attenuation of the
    γ
    -rays emitted
    normally to the direction from the massive star. Dotted thick black line
    shows the modification of the broad band spectrum, if the emission
    from the tertiary
    e
    +
    e
    pairs produced via the
    γγ → e
    +
    e
    process
    is taken into account.
    Although the total power of neutrino emission can
    be estimated from the MeV-GeV luminosity of the
    γ
    -
    ray-loud binary, modelling of the broad band emission
    spectrum of the source gives only mild constraints on the
    neutrino emission spectrum: acceptable models of the
    broad band spectra can be found assuming the initial
    proton injection spectra ranging from
    E
     2
    powerlaw
    to almost monochromatic injection spectra (see [8] for
    details). The properties of the neutrino signal can be used
    to distinguish between different primary proton spectra,
    if one is able to derive the information on the energies
    of the detected neutrinos from the observational data.
    Extraction of information about the properties of the
    source neutrino spectrum from the data is complicated
    by the fact that (a) the neutrino telescopes measure only
    the energies of the detected muons, rather than that
    of the primary neutrinos, (b) the flux of the primary
    neutrinos and the energies of the secondary muons are
    attenuated by the effect of propagation through the Earth.
    The differential induced muon spectrum at the detector,
    which strongly depends on the source declination
    δ
    ,
    could be obtained after propagation of neutrinos up
    to the interaction point and further propagation of the
    muons to the detector. A semi-analytical method of
    calculation of the muon spectra, which is based on the
    use of the muon effective area and the knowledge of the
    details of the detector, namely the angular and energy
    resolution was developed in the Ref. [8].
    The results of such semi-analytical calculation of the
    atmospheric background-subtracted muon spectra, for
    the particular case of neutrinos from LS I +61
    303, are
    presented Fig. 4. The exposure time is taken to be 3 years
    of running the full ICECUBE array. For comparison,
    we also show by the solid thick line the level of the
    signal which is
    above the atmospheric background.
    Dotted line show the detected muon spectrum for the

    4
    NERONOV AND RIBORDY, NEUTRINO SIGNAL FROM BINARIES
    Fig. 4. Background-subtracted cumulative muon spectra
    N (E
    µ,thr
    )
    ,
    expected after the 3-year ICECUBE exposure for the three model
    neutrino spectra of LS I +61
    303, discussed in the previous sections
    (error bars of signal muon spectra are the sum in quadrature of
    statistical errors of signal + atmospheric neutrino background). Dotted
    line shows the spectrum for the proton injection spectrum with the
    spectral index
     = 2
    . Thin solid line corresponds to the initial proton
    injection spectrum with
     = 1
    , while the dashed line is for the
    proton injeciton spectrum with
     = 0
    . Thick solid line shows the
    5 σ
    excess above the atmospheric neutrino backgorund (the ”discovery
    threshold”). The bin radius is set to
    ψ = 1.3
    .
    model with the proton injection spectrum with spectral
    index
    p
    = 2
    , thin solid line corresponds to the proton
    injection spectrum with
    p
    = 1
    and cut-off at the same
    energy, while the dashed line corresponds to
    p
    = 0
    .
    In all three cases the cut-off energy is assumed to be
    E
    cut
    = 1
    PeV.
    Inspecting the muon spectra shown in the upper panel
    of Fig. 4, one can see that softed proton injection spec-
    trum results in a slight excess of muon events at lower
    energies. If the overall normalization of the neutrino
    flux would be known, measurement of the spectrum
    of muon events would allow to constrain the spectrum
    of the primary protons in the source. However, taking
    into account the uncertainty of the overall normalization
    of neutrino flux introduced by the uncertainty of the
    anisotropy pattern of neutrino emission, one can find
    that the statistics of the signal will not be enough for
    such a task. This is illustrated in the lower panel of Fig.
    4, where a comparison of the shapes of the muon spectra
    is shown. If one assumes the same total number of muon
    events, the difference in the spectra for the three models
    is always within
    ∼ 1σ
    errorbars, over the entire energy
    range.
    IV. CONCLUSIONS
    We have estimated the neutrino flux from GRLBs
    expected within the hadronic model of activity of these
    sources. Within such a model, the measured spectral
    characteristics of
    γ
    -ray emission from the source in
    the TeV energy band are not directly related to the
    spectral characteristics of the neutrino emission, because
    of absorption of the TeV
    γ
    -rays on the thermal photon
    background produced by the massive star in the system.
    The uncertainty of the calculation of the attenuation of
    the TeV
    γ
    -ray flux introduces a large uncertainty to the
    estimate of the neutrino flux based on the measured TeV
    γ
    -ray lfux.
    Taking into account this uncertainty, we have adopted
    a different approach for the estimate of the neutrino
    flux from a GRLB. Namely, we have noted that the
    energy output of proton-proton interactions, and hence
    the neutrino flux, can be instead constrained by the
    broad-band spectrum of the source.
    Although the observed bolometric luminosity of the
    source (i.e. the height of the MeV-GeV bump of the
    SED) constrains the overall neutrino luminosity, the
    shape of the neutrino spectrum and the overall normal-
    ization of the neutrino flux are only mildly constrained
    by the multi-wavelength data. Taking into account these
    uncertainties of the neutrino emission spectrum, we have
    estimated the expected number of neutrinos which will
    be detected by the ICECUBE, assuming that the neutrino
    flux saturates the upper bound imposed by the observed
    γ
    -ray flux in the MeV-GeV energy band. Considering
    the particular example of LS I +61
    303, we have found
    that if the spectrum of high-energy protons in the source
    extends to the PeV energies, the source could be readily
    detectable with 3 years of exposure with ICECUBE.
    We have also explored the potential of the full ICE-
    CUBE detector for the measurement of the spectral
    characteristics of the neutrino signal from LS I +61
    303.
    We find that in the case when the neutrino flux is at
    the level of the upper bound imposed by the observed
    MeV-GeV
    γ
    -ray flux, exposure time much longer than
    3 years is required to constrain the spectral index of the
    primary high-energy proton spectrum via observations
    of neutrino signal in ICECUBE.
    REFERENCES
    [1] I.F. Mirabel, Science, 312, 1759, (2006).
    [2] M. Tavani, J. Arons, Ap.J., 477, 439 (1997).
    [3] V. Bosch-Ramon et al., A&A, 447, 263 (2006).
    [4] L. Maraschi, A. Treves, MNRAS, 194, 1 (1981).
    [5] M. Chernyakova, A. Neronov, et al. MNRAS, 367, 1201 (2006).
    [6] D.F. Torres, F. Halzen, A.Ph., 27, 500, (2007).
    [7]
    http://icecube.wisc.edu/
    [8] A. Neronov, M. Ribordy, Phys. Rev. D79, 043013 (2009).

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    1
    Acoustic sensor development for ultra high energy neutrino
    detection
    Matt Podgorski
    and Mathieu Ribordy
    High Energy Physics Laboratory, EPFL, CH - 1015 Switzerland
    RWTH Aachen, visiting EPFL
    Abstract
    . The GZK neutrino flux characterization
    would give insights into cosmological source evo-
    lution, source spectra and composition at injection
    from the partial recovery of the degraded informa-
    tion carried by the ultra high energy cosmic rays. The
    flux is expected to be at levels necessitating a much
    larger instrumented volume (
    >
    100 km
    3
    ) than those
    currently operating. First suggested by Askaryan,
    both radio and acoustic detection techniques could
    render this quest possible thanks to longer wave
    attenuation lengths (predicted to exceed a kilometer)
    allowing for a much sparser instrumentation com-
    pared to optical detection technique.
    We present the current acoustic R&D activities at
    our lab developing adapted devices, report on the
    obtained sensitivies and triangulation capabilities we
    obtained, and define some of the requirements for
    the construction of a full scale detector.
    Keywords
    : Ultra high energy neutrinos. Acoustic
    detection techniques. Acoustic sensor studies.
    I. INTRODUCTION
    The IceCube detector [1] may well soon identify the
    first ultra high energy neutrino of cosmogenic origin,
    following interactions of ultra high energy cosmic rays
    with the cosmic microwave background [2]. Predic-
    tions for the cosmogenic neutrino flux, i.e. neutrinos
    from photo-disintegration, is at levels of the order of
    EdN/dE ∼ 10
     17
    s
     1
    cm
     3
    sr
     1
    at
    E = 10
    18
    eV,
    resulting in 0.01 - 1 event / year / km
    3
    in ice [3]. These
    predictions strongly depend on the primary cosmic ray
    composition [4]. Currently, the situation is uncertain:
    While the observed correlation of UHE CR sources with
    the AGN distribution by AUGER [5] hints toward a light
    composition (and in this case we lie close to the upper
    flux predictions), dedicated AUGER composition studies
    favor a composition turning heavier at UHE [6]. GZK
    neutrinos are astronomical messengers keeping track
    of the original CR direction, GZK interactions mostly
    occur close to the source. In case of the existence of a
    few UHE cosmic accelerators located close-by (Gpc),
    the detection of a substantial flux of GZK neutrinos
    from these directions using a multi-messenger approach
    would allow the possibility of pinpointing the nature of
    these CR accelerators.
    The characterization of the GZK neutrino flux spec-
    trum, and thus the recovery of the degraded informa-
    tion information carried by UHE CR, would allow the
    delineation of cosmological source evolution scenarios
    from source spectrum characteristics. To fulfill this goal,
    the event detection rate should be vastly increased.
    Therefore a much larger volume should be instrumented
    with an adequate technology for the detection of ultra
    high energy neutrino interactions. Two novel detection
    methods have been proposed, following signatures first
    discussed by Askaryan [7], [8]. An interacting neutrino
    emits a coherent Cherenkov pulse in the range of 0.1-
    1 GHz [9] close to the shower axis and a thin ther-
    moacoustic pancake normal to the shower axis. Both
    detection techniques are currently exploited by several
    detectors. In ice, both radio and acoustic emissions have
    rather large theoretical attenuation lengths [10]. While
    this has been convincingly demonstrated for the radio
    emission, it is still a work in progress for the acoustic
    emission and is one of the main goals for the South
    Pole Acoustic Test Setup (SPATS) [11]. With the data
    collected by the SPATS array, the sound speeds w.r.t. the
    depth have been determined with great accuracy, meeting
    theoretical expectations [12] and S-waves have been
    found as well. Unknown, however, remains the exact
    nature of the local source of noise and the exact value
    of the attenuation length. Newest experimental results
    hint toward a reduced pressure wave attenuation length
    on one hand and demonstrate favorable noise level below
    10 mPa on the other hand [13].
    Contrary to salt and water, ice is unique. It allows
    the detection of three distinct signatures accompanying
    a neutrino event: Optical Cherenkov light, coherent radio
    Cherenkov and thermoacoustic emissions, thus firmly
    establishing the event origin by a strong background re-
    duction. A possible layout for the hybrid instrumentation
    of a large volume of order of 100 km
    3
    at the South
    Pole would consist of strings deployed one kilometer
    apart down to a depth of 2 km (radio and acoustic
    attenuation lengths strongly vary with temperature and
    are decreasing with depth). Given the topologies for
    the radio & acoustic emissions, a string should be
    densely equipped with radio and acoustic devices with
    an option of supplementing it with PMT devices for
    optical detection.
    10  10
    3
    interacting GZK neutrinos in
    100 km
    3
    instrumented volume can be expected after 10
    years. The cosmogenic spectrum could be characterized
    (and consequently insights into the underlying physics),
    provided a high detection efficiency, a deep knowledge
    of the local source of noise and good energy resolution.
    Thermoacoustic models and Monte Carlo simulations

    2
    MATHIEU RIBORDY
    et al.
    ACOUSTIC SENSOR R&D
    predict that a signal from a neutrino with an energy
    E = 10
    18
    eV will typically have an amplitude of 10
    mPa at a distance of one kilometer [14] (to be rescaled
    for finite attenuation length). To keep a good S/N ratio,
    the sensitivity of the devices should be at the sub mPa
    level. Also, a good pointing resolution may serve the
    purpose of UHE point source search. Given the giant
    array layout introduced above, an acoustic signal will
    be recorded by a small number of acoustic devices. It
    is therefore desirable to design acoustic sensor devices
    with pointing capabilities of their own.
    In the next section, we present R&D activities which
    are taking place at our lab in regard of sensor design
    and construction and discuss its performances.
    II. R&D ACTIVITIES
    The design and construction of a multi-channel sensor
    was conducted at our lab, which use piezo transduc-
    ers (PZT) as sensitive elements. A noise level level
    S/N < 5
    mPa (
    S/N ≡ S
    RMS
    /N
    RMS
    ) and a good
    angular resolution were demonstrated, suggesting the
    possibility for excellent vertex localization combining
    the responses from all sensor hits. The design, which
    must still be improved to meet our design goals, could
    eventually allow for diffuse acoustic noise reduction
    through spectral shape analysis and accurate energy
    estimate of physical events.
    The setup for conducting the R&D activities consists
    of a bath, topped with a support structure for one
    absolutely calibrated hydrophone (Sensortech SQ03),
    one homemade sensor and one transmitter. A datataking
    LabView program interfaced to a National Instrument
    board is used for analog response digitization (12 bits
    ×
    12 channels, total 1.25 MHz), transmitter pulse gener-
    ation and
    relative orientation between the transmitter
    and the sensors for automatized sensor profiling.
    The experimental setup shown in Fig. 1 consists of the
    homemade hydrophone, a transmitter and the reference
    hydrophone. In the following, two different electric sig-
    nal shapes have been considered: a damped sin pulse and
    a gaussian pulse, resulting in a tripolar pressure pulse
    (the neutrino-induced thermoacoustic pulse is bipolar).
    SQ03 hydrophone
    homemade sensor
    large water tank
    transmitter
    Fig. 1. Experimental setup.
    A. Acoustic sensor design
    The sensor consists of an aluminium pressure vessel
    housing 4 channels to provide triangulation capabilities.
    The noise level at the amplifier input to 130 nV, reached
    frequency [Hz]
    10000 20000 30000 40000 50000 60000 70000 80000 90000
    pressure [mPa; S/N=1]
    1
    10
    2
    10
    piezo 1
    piezo 2
    piezo 3
    piezo 4
    RMS piezo noise: 3.2mV
    sensitivity SQ03: 7.94
    μ
    V/mPa
    Fig. 2. Pressure sensitivity as a function of frequency of damped sin
    transmitted pulses normalized to
    S/N = 1
    ratio.
    at the expense of some bandwidth reduction, peaking at
    22 kHz with
    ∼90
    dB amplification and sharply decreas-
    ing below 10 kHz and above 40 kHz. Whether that is
    optimal has to be studied further. It is manufacturable
    at relatively low cost. While aluminium is an adequate
    medium for use in a liquid water bath, it will be replaced
    by steel for application in ice (more adequate given both
    impedance and resistance to pressure).
    B. Sensitivity calibration
    Signals with peak frequencies in the range 10 - 90 kHz
    were recorded with a sampling rate of 330 kHz. A strong
    frequency correlation between the transmitted pulse and
    the sensor response was observed. Due to the finite size
    of the bath tub and given the sound speed velocity in
    water, only the first 150
    µ
    s following the pulse arrival
    time were analysed in what follows to avoid reflexion
    artefacts.
    With the collected data from the 4-channel sensor and
    from the commercial hydrophone, the absolute pressure
    sensitivity was calculated in the time domain using RMS
    values for signal and noise. Fig. 2 shows the absolute
    pressure sensitivity (defined as
    S/N = 1
    ) w.r.t. the
    dominant frequency of the sent signals.
    The measurements demonstrate the importance of the
    state of surface coupling the PZT to the housing: The
    polished surfaces for piezos 1 and 3 show a response
    ∼2
    times stronger than piezos 2 and 4 coupled to the
    housing through porous surfaces.
    C. Triangulation
    Time resolution is essential for triangulation and
    therefore a digitization frequency of 100/200 kHz
    is required in order to reach cm resolution in alu-
    minium/steel. This suggests that a sensor design should
    include digital electronics with at least 200 kHz sam-
    pling rate per channel
    1
    , in order to reach 0.5-5 ms
    1
    100 kHz (and therefore 200 kHz sampling rate) is by coincidence
    the value above which the ice attenuation length drops quickly and
    roughly the extension of the neutrino-induced thermoacoustic pulse
    spectrum.

    PROCEEDINGS OF THE 31
    st
    ICRC, ŁO´ DZ´ 2009
    3
    pulse start time resolution (depending on amplitude)
    which roughly corresponds to 2
    -20
    (
    4π/10
    2
    -
    4π/10
    4
    )
    angular resolution with the current multi-channel sensor.
    The design of a new digital (0.2 MHz/channel) 4-
    channel amplifier board has been started, with long range
    communication protocols. It does not yet include trigger
    logic. Digitization is necessary for a viable acoustic
    detector design in order to avoid losses in km long cables
    (of order of 3 dB/100 m in high quality cables) and thus
    keep both good sensitivity and time resolution. Once in
    operation, this will allow to define the requirements for
    future efficient trigger concept at the sensor level.
    degrees
    −200 −150 −100 −50
    0
    50
    100
    150
    200
    time [ms]
    0.62
    0.63
    0.64
    0.65
    0.66
    0.67
    −3
    ×
    10
    piezo 1
    piezo 2
    piezo 3
    piezo 4
    Graph
    Fig. 3.
    Time of first signal maximum as a function of the polar
    orientation of the sensor.
    Figure 3 demonstrates triangulation capabilities. Cou-
    pling between channels was found to provide in any
    sampled sensor positions 2 channels with signal within
    a factor 2 of the channel with the highest response. The
    resolution will nevertheless depend on the individual
    signal-to-noise ratios, but it shows potential for vertex
    reconstruction with a single sensor module.
    D. Outlook
    First positive results in the time domain were ob-
    tained. Absolute sensitivities are currently analyzed in
    the frequency domain. A second 4-channel sensor of
    similar design in a steel housing will be soon equipped
    with digital electronics. Further sensor tests are foreseen
    to happen next. Low temperature behavior test will be
    conducted in the laboratory and at large distances and
    depths in the lake Geneva (∼ 400 m depth) to avoid
    reflexions, assess the acoustic noise characteristics and
    probe the sensor design.
    III. CONCLUSIONS
    Acoustic neutrino detection techniques should be
    further developed, pushing the sensitivity at sub mPa
    levels, together with the characterization of noise sources
    which may impede the applicability of the technique.
    Noise measurement in an open media (water / ice) are
    required to characterize the noise rate and its spectral
    shape in order to investigate improved trigger schemes
    relying on signal processing within the sensors. These
    developments have been started with digitization board
    in the sensor, a necessary step for a viable acoustic array
    design, where signal attenuation along km transmission
    cable is excluded.
    While it seems clear that sub mPa sensors should
    be designed, the uncertain detection conditions at the
    South Pole make predictions concerning the detection
    efficiency difficult. The potential can be dangerously
    spoiled in the case the local source of acoustic noise
    mimick a neutrino event. Further dedicated studies are
    on-going to ensure that it could be possible to distinguish
    the event origin with high efficiency. The deployement
    of an acoustico-radio-optical hybrid detector would con-
    stitute a welcome option, allowing to reduce further
    possible background noises. Also, complementing the
    hybrid radio-acoustic strings with a few optical DOMs
    such as in IceCube would allow to unambiguously
    tag neutrino events (at these energies, given a
    >
    100m
    absorption length in the ice at these depths [15], photons
    may likely accompany a radio-acoustic signal in the case
    of a neutrino event).
    REFERENCES
    [1] R. Abbasi et al., Nucl. Instr. Meth.A601:294-316,2009.
    [2] V. Berezinsky and G. Zatsepin, Phys. Lett. B 28 (1969) 423.
    [3] R. Engel, D. Seckel & T. Stanev, Phys. Rev. D
    64
    (2001) 093010.
    [4] L. A. Anchordoqui, H. Goldberg, D. Hooper, S. Sarkar &
    A. M. Taylor, Phys. Rev. D
    76
    (2007) 123008.
    [5] J. Abraham et al. [Pierre Auger Collaboration], Astropart. Phys.
    29
    (2008) 188. [Erratum-ibid.
    30
    (2008) 45]
    [6] K. H. Kampert [Pierre Auger Collaboration], AIP Conf. Proc.
    1085
    (2009) 30.
    [7] G. Askaryan, JETP 14 (1962) 441; G. Askaryan, JETP 21 (1965)
    658.
    [8] G. Askaryan, Sov. J. Atom. Energy 3 (1957) 921.
    [9] E. Zas, F. Halzen & T. Stanev, Phys. Rev. D45 (1992) 362.
    [10] B. Price and L. Bergstrm, Appl. Opt. 36 (1997) 4181; L.
    Bergstrm, B. Price et al., Appl. Opt. 36, 4168 (1997); B. Price,
    J. Geophys. Res.111, B02201 (2006).
    [11] S. Boeser et al., arXiv:0807.4676 [astro-ph]; S. Boeser et al., Int.
    J. Mod. Phys. A21S1 (2006) 107.
    [12] J. Vandenbroucke et al., arXiv:0811.1087 [astro-ph].
    [13] F. Descamps et al. [IceCube coll.], in these proceedings.
    [14] D. Besson et al., Int. J. Mod. Phys. A21S1 (2006) 259.
    [15] M. Ackermann et al., J. Geophys. Res. 111, D13203 (2006).

    Back to top