Observation of high energy atmospheric neutrinos with the Antarctic muon
    and neutrino detector array
    J. Ahrens,
    9
    E. Andre
    ´
    s,
    14
    X. Bai,
    1
    G. Barouch,
    11
    S. W. Barwick,
    8
    R. C. Bay,
    7
    T. Becka,
    9
    K.­H. Becker,
    2
    D. Bertrand,
    3
    F. Binon,
    3
    A. Biron,
    4
    J. Booth,
    8
    O. Botner,
    13
    A. Bouchta,
    4,
    *
    O. Bouhali,
    3
    M. M. Boyce,
    11
    S. Carius,
    5
    A. Chen,
    11
    D. Chirkin,
    7
    J. Conrad,
    13
    J. Cooley,
    11
    C. G. S. Costa,
    3
    D. F. Cowen,
    10
    E. Dalberg,
    14,†
    C. De Clercq,
    15
    T. DeYoung,
    11,‡
    P. Desiati,
    11
    J.­P. Dewulf,
    3
    P. Doksus,
    11
    J. Edsjo
    ¨
    ,
    14
    P. Ekstro
    ¨
    m,
    14
    T. Feser,
    9
    J.­M. Fre
    `
    re,
    3
    T. K. Gaisser,
    1
    M. Gaug,
    4,§
    A. Goldschmidt,
    6
    A. Hallgren,
    13
    F. Halzen,
    11
    K. Hanson,
    10
    R. Hardtke,
    11
    T. Hauschildt,
    4
    M. Hellwig,
    9
    H. Heukenkamp,
    4
    G. C. Hill,
    11
    P. O. Hulth,
    14
    S. Hundertmark,
    8
    J. Jacobsen,
    6
    A. Karle,
    11
    J. Kim,
    8
    B. Koci,
    11
    L. Ko
    ¨
    pke,
    9
    M. Kowalski,
    4
    J. I. Lamoureux,
    6
    H. Leich,
    4
    M. Leuthold,
    4
    P. Lindahl,
    5
    I. Liubarsky,
    11
    P. Loaiza,
    13
    D. M. Lowder,
    7,
    i
    J. Madsen,
    12
    P. Marciniewski,
    13,¶
    H. S. Matis,
    6
    C. P. McParland,
    6
    T. C. Miller,
    1,
    **
    , Y. Minaeva,
    14
    P. Mioc
    ˇ
    inovic
    ´
    ,
    7
    P. C. Mock,
    8,††
    R. Morse,
    11
    T. Neunho
    ¨
    ffer,
    9
    P. Niessen,
    4,15
    D. R. Nygren,
    6
    H. O
    ¨
    gelman,
    11
    Ph. Olbrechts,
    15
    C. Pe
    ´
    rez de los Heros,
    13
    A. C. Pohl,
    5
    R. Porrata,
    8,‡‡
    P. B. Price,
    7
    G. T. Przybylski,
    6
    K. Rawlins,
    11
    C. Reed,
    8,§§
    W. Rhode,
    2
    M. Ribordy,
    4
    S. Richter,
    11
    J. Rodrı
    ´
    guez Martino,
    14
    P. Romenesko,
    11
    D. Ross,
    8
    H.­G. Sander,
    9
    T. Schmidt,
    4
    D. Schneider,
    11
    R. Schwarz,
    11
    A. Silvestri,
    2,4
    M. Solarz,
    7
    G. M. Spiczak,
    12
    C. Spiering,
    4
    N. Starinsky,
    11,
    ii
    D. Steele,
    11
    P. Steffen,
    4
    R. G. Stokstad,
    6
    O. Streicher,
    4
    P. Sudhoff,
    4
    K.­H. Sulanke,
    4
    I. Taboada,
    10
    L. Thollander,
    14
    T. Thon,
    4
    S. Tilav,
    1
    M. Vander Donckt,
    3
    C. Walck,
    14
    C. Weinheimer,
    9
    C. H. Wiebusch,
    4,
    *
    C. Wiedeman,
    14
    R. Wischnewski,
    4
    H. Wissing,
    4
    K. Woschnagg,
    7
    W. Wu,
    8
    G. Yodh,
    8
    and S. Young
    8
    ~
    AMANDA Collaboration
    !
    1
    Bartol Research Institute, University of Delaware, Newark, Delaware 19716
    2
    Fachbereich 8 Physik, BUGH Wuppertal, D­42097 Wuppertal, Germany
    3
    Universite
    ´
    Libre de Bruxelles, Science Faculty CP230, Boulevard du Triomphe, B­1050 Brussels, Belgium
    4
    DESY­Zeuthen, D­15735 Zeuthen, Germany
    5
    Department of Technology, Kalmar University, S­39182 Kalmar, Sweden
    6
    Lawrence Berkeley National Laboratory, Berkeley, California 94720
    7
    Department of Physics, University of California, Berkeley, California 94720
    8
    Department of Physics and Astronomy, University of California, Irvine, California 92697
    9
    Institute of Physics, University of Mainz, Staudinger Weg 7, D­55099 Mainz, Germany
    10
    Department of Physics and Astronomy, University of Pennsylvania, Philadelphia, Pennsylvania 19104
    11
    Department of Physics, University of Wisconsin, Madison, Wisconsin 53706
    12
    Physics Department, University of Wisconsin, River Falls, Wisconsin 54022
    13
    Division of High Energy Physics, Uppsala University, S­75121 Uppsala, Sweden
    14
    Department of Physics, Stockholm University, SCFAB, SE­10691 Stockholm, Sweden
    15
    Vrije Universiteit Brussel, Dienst ELEM, B­1050 Brussel, Belgium
    ~
    Received 1 February 2002; published 31 July 2002
    !
    The Antarctic muon and neutrino detector array
    ~
    AMANDA
    !
    began collecting data with ten strings in 1997.
    Results from the first year of operation are presented. Neutrinos coming through the Earth from the Northern
    Hemisphere are identified by secondary muons moving upward through the array. Cosmic rays in the atmo­
    sphere generate a background of downward moving muons, which are about 10
    6
    times more abundant than the
    upward moving muons. Over 130 days of exposure, we observed a total of about 300 neutrino events. In the
    same period, a background of 1.05
    3
    10
    9
    cosmic ray muon events was recorded. The observed neutrino flux is
    consistent with atmospheric neutrino predictions. Monte Carlo simulations indicate that 90% of these events lie
    in the energy range 66 GeV to 3.4 TeV. The observation of atmospheric neutrinos consistent with expectations
    establishes AMANDA­B10 as a working neutrino telescope.
    DOI: 10.1103/PhysRevD.66.012005 PACS number
    ~
    s
    !
    : 95.85.Ry, 95.55.Vj, 96.40.Tv
    *
    Now at CERN, CH­1211, Gene
    `
    ve 23, Switzerland.
    Now at Defense Research Establishment
    ~
    FOA
    !
    , S­17290 Stockholm, Sweden.
    Now at Santa Cruz Institute for Particle Physics, University of California—Santa Cruz, Santa Cruz, CA 95064.
    §
    Now at IFAE, 08193 Barcelona, Spain.
    i
    Now at MontaVista Software, 1237 E. Arques Ave., Sunnyvale, CA 94085.
    Now at The Svedberg Laboratory, S­75121 Uppsala, Sweden.
    **
    Now at Johns Hopkins University, Applied Physics Laboratory, Laurel, MD 20723.
    ††
    Now at Optical Networks Research, JDS Uniphase, 100 Willowbrook Rd., Freehold, NJ 07728­2879.
    ‡‡
    Now at L­174, Lawrence Livermore National Laboratory, 7000 East Ave., Livermore, CA 94550.
    §§
    Now at Dept. of Physics, Massachussetts Institute of Technology, Cambridge, MA.
    ii
    Now at SNO Institute, Lively, ON, Canada P3Y 1M3.
    PHYSICAL REVIEW D
    66
    , 012005
    ~
    2002
    !
    0556­2821/2002/66
    ~
    1
    !
    /012005
    ~
    20
    !
    /$20.00 ©2002 The American Physical Society
    66
    012005­1

    I. INTRODUCTION
    Energetic cosmic ray particles entering the Earth’s atmo­
    sphere generate a steady flux of secondary particles such as
    electrons, muons and neutrinos. The electronic component of
    cosmic rays is quickly absorbed. High energy muons pen­
    etrate the Earth’s surface for several kilometers, while atmo­
    spheric neutrinos can easily pass the Earth up to very high
    energies. Interactions of hadronic particles, similar to the
    ones that create the atmospheric neutrino flux, will generate
    neutrinos at sites where cosmic rays are generated and where
    they interact as they travel through the Universe. The goal of
    observing neutrinos of astrophysical origin determines the
    design and the size of neutrino telescopes.
    The primary channel through which neutrino telescopes
    detect neutrinos above energies of a few tens of GeV is by
    observing the Cherenkov light from secondary muons pro­
    duced in
    n
    m
    ­nucleon interactions in or near the telescope. To
    ensure that the observed muons are produced by neutrinos,
    the Earth is used as a filter and only upward moving muons
    are selected. A neutrino telescope consists of an array of
    photosensors embedded deeply in a transparent medium. The
    tracks of high energy muons—which can travel many hun­
    dreds of meters, or even kilometers, through water or ice—
    can be reconstructed with reasonable precision even with a
    coarsely instrumented detector, provided the medium is suf­
    ficiently transparent. A location deep below the surface
    serves to minimize the flux of cosmic­ray muons.
    In this paper we demonstrate the observation of atmo­
    spheric muon neutrinos with the Antarctic muon and neu­
    trino detector array
    ~
    AMANDA
    !
    . These neutrinos constitute
    a convenient flux of fairly well known strength, angular dis­
    tribution, and energy spectrum, which can be used to verify
    the response of the detector. The paper will focus on the
    methods of data analysis and the comparison of observed
    data with simulations. After a brief description of the detec­
    tor, the data and the methods of simulation are introduced in
    Sec. III and the general methods of event reconstruction are
    described in Sec. IV. Two AMANDA working groups ana­
    lyzed the data in parallel. The methods and results of both
    analyses are described in Secs. V and VI. After a discussion
    of systematic uncertainties in Sec. VII we present the final
    results and conclusions.
    II. THE AMANDA DETECTOR
    The AMANDA detector uses the 2.8 km thick ice sheet at
    the South Pole as a neutrino target, Cherenkov medium and
    cosmic ray flux attenuator. The detector consists of vertical
    strings of optical modules
    ~
    OMs
    !
    —photomultiplier tubes
    sealed in glass pressure vessels—frozen into the ice at depths
    of 1500–2000 m below the surface. Figure 1 shows the cur­
    rent configuration of the AMANDA detector. The shallow
    array, AMANDA­A, was deployed at depths of 800 to 1000
    m in 1993–1994 in an exploratory phase of the project. Stud­
    ies of the optical properties of the ice carried out with
    AMANDA­A showed a high concentration of air bubbles at
    these depths, leading to strong scattering of light and making
    accurate track reconstruction impossible. Therefore, a deeper
    array of ten strings with 302 OMs was deployed in the aus­
    tral summers of 1995–1996 and 1996–1997 at depths of
    1500–2000 m. This detector is referred to as AMANDA­
    B10, and is shown in the center of Fig. 1. The detector was
    augmented by three additional strings in 1997–1998 and six
    in 1999–2000, forming the AMANDA­II array.
    In AMANDA B10, an optical module consists of a single
    8 in. Hamamatsu R5912­2 photomultiplier tube
    ~
    PMT
    !
    housed in a glass pressure vessel. The PMT is optically
    coupled to the glass housing by a transparent gel. Each mod­
    ule is connected to electronics on the surface by a dedicated
    electrical cable, which supplies high voltage and carries the
    anode signal of the PMT. For each event, the optical module
    is read out by a peak­sensing ADC and a TDC capable of
    registering up to eight separate pulses. The overall precision
    of measurement of photon arrival times is approximately
    5 ns. Details of deployment, electronics and data acquisition,
    calibration, and the measurements of geometry, timing reso­
    lution, and the optical properties of the ice can be found in
    @
    1,2
    #
    .
    The optical properties of the polar ice in which
    AMANDA is embedded have been studied in detail, using
    both light emitters located on the strings and the downgoing
    muon flux itself. These studies
    @
    3
    #
    have shown that the ice is
    not perfectly homogeneous, but rather that it can be divided
    into several horizontal layers which were laid down by vary­
    ing climatological conditions in the past
    @
    4
    #
    . Different con­
    centrations of dust in these layers lead to a modulation of the
    FIG. 1. The present AMANDA detector. This paper describes
    data taken with the ten inner strings shown in expanded view in the
    bottom center.
    J. AHRENS
    et al.
    PHYSICAL REVIEW D
    66
    , 012005
    ~
    2002
    !
    012005­2

    scattering and absorption lengths of light in the ice, as shown
    in Fig. 2. The average absorption length is about 110 m at a
    wavelength of 400 nm at the depth of the AMANDA­B10
    array, and the average effective scattering length is approxi­
    mately 20 m.
    III. DATA AND SIMULATION
    The data analyzed in this paper were recorded during the
    austral winter of 1997, from April to November. Subtracting
    downtime for detector maintenance, removing runs in which
    the detector behaved abnormally and correcting for deadtime
    in the data acquisition system, the effective livetime was
    130.1 days.
    Triggering was done via a majority logic system, which
    demanded that 16 or more OMs report signals within a slid­
    ing window of 2
    m
    s. When this condition was met, a trigger
    veto was imposed and the entire array read out. The raw
    trigger rate of the array was on average 75 Hz, producing a
    total data set of 1.05
    3
    10
    9
    events.
    Random noise was observed at a rate of 300 Hz for OMs
    on the inner four strings and 1.5 kHz for tubes on the outer
    six, the difference being due to different levels of concentra­
    tion of radioactive potassium in the pressure vessels
    ~
    details
    on noise rates can be found in Ref.
    @
    5
    #!
    . A typical event has
    a duration of 4.5
    m
    s, including the muon transit time and the
    light diffusion times, so random noise contributed on average
    one PMT signal per event.
    Almost all of the events recorded were produced by
    downgoing muons originating in cosmic ray showers. Trig­
    gers from atmospheric neutrinos contribute only a few tens
    of events per day, a rate small compared to the event rate
    from cosmic ray muons, as shown in Fig. 3. The main task of
    AMANDA data analysis is to separate these neutrino events
    from the background of cosmic­ray muons. Monte Carlo
    ~
    MC
    !
    simulations of the detector response to muons pro­
    duced by neutrinos or by cosmic rays were undertaken to
    develop techniques of background rejection.
    Downgoing muons were generated by atmospheric
    shower simulations of isotropic protons with
    BASIEV
    @
    6
    #
    or
    protons and heavier nuclei with
    CORSIKA using the QGSJET
    generator
    @
    7,8
    #
    , and tracked to the detector with the muon
    propagation code
    MUDEDX
    @
    9,10
    #
    . Two other muon propaga­
    tion codes were used to check for systematic differences:
    PROPMU
    @
    11
    #
    with a 30% lower rate and
    MMC
    @
    12
    #
    with a
    slightly higher rate. A total of 0.9
    3
    10
    8
    events were simu­
    lated. Most characteristics of the events generated with
    BASIEV
    were found to be similar to the more accurate
    CORSIKA­based simulation. For the latter, the primary cosmic
    ray flux as described by Wiebel­Sooth and Biermann
    @
    13
    #
    was used. The curvature of the Earth has been implemented
    in
    CORSIKA to correctly describe the muon flux at large ze­
    nith angles. The event rate based on this Monte Carlo was 75
    Hz and compares reasonably well with the observed rate of
    100 Hz
    ~
    after deadtime correction
    !
    . The detector response to
    muons was modeled by calculating the photon fields pro­
    duced by continuous and stochastic muonic energy losses
    @
    14
    #
    , and simulating the response of the hardware to these
    photons
    @
    15,16
    #
    . Upgoing muons were generated by a propa­
    gation of atmospheric neutrinos, which were tracked through
    the Earth and allowed to interact in the ice in or around the
    detector or in the bedrock below
    @
    17,18
    #
    . Muons that were
    generated in the bedrock were propagated using
    PROPMU
    @
    11
    #
    until they reached the rock­ice boundary at the depth of 2800
    m. The muons were then propagated through the ice in the
    same way as those from cosmic ray showers. The atmo­
    spheric neutrino flux was taken from Lipari
    @
    19
    #
    .
    The Cherenkov photon propagation through the ice was
    modeled to create multidimensional tables of density and
    FIG. 2. Variation of the optical properties with depth. The effec­
    tive scattering coefficient at a wavelength of 532 nm is shown as a
    function of depth. The
    z
    axis is pointing upwards and denotes the
    vertical distance from the origin of the detector coordinate system
    located at a depth of 1730 m. The shaded areas on the side indicate
    layers of constant scattering coefficient as used in the Monte Carlo
    simulation.
    FIG. 3. The zenith angle distribution of simulated AMANDA
    triggers per 130.1 days of lifetime. The solid line represents triggers
    from downgoing cosmic ray muons generated by
    CORSIKA. The
    dashed line shows triggers produced by atmospheric neutrinos.
    OBSERVATION OF HIGH ENERGY ATMOSPHERIC... PHYSICAL REVIEW D
    66
    , 012005
    ~
    2002
    !
    012005­3

    arrival time probability distributions of the photon flux.
    These photon fields were calculated for pure muon tracks
    and for cascades of charged particles. A real muon track was
    modeled as a superposition of the photon fields of a pure
    muon track and the stochastic energy losses based on cas­
    cades. The photon fields were calculated out to 400 m from
    the emission point, taking into account the orientation of the
    OM with respect to the muon or cascade. In the detector
    simulation, the ice was modeled as 16 discrete layers, as
    indicated by the shaded areas in Fig. 2. The spectral proper­
    ties of the photomultiplier sensitivity, the glass, the gel, and,
    most importantly, the ice itself were included in the simula­
    tion of the photon propagation. The probability of photon
    detection depends on the Fresnel reflectance at all interfaces,
    transmittances of various parts, and quantum and collection
    efficiencies of the PMT. The relevant physical parameters
    have been measured in the laboratory, so that the spectral
    sensitivity of the OM could be evaluated. Two types of OMs,
    differing in the type of pressure vessel, were used in the
    construction of AMANDA­B10. The inner four strings
    ~
    AMANDA­B4
    !
    use Billings housings while the outer six
    strings use Benthos housings.
    ~
    Benthos Inc. and Billings In­
    dustries are the manufacturers of the glass pressure vessels.
    Benthos and Billings are registered trademarks of the respec­
    tive companies.
    !
    The two types of housing have different
    optical properties. The Benthos OMs have an effective quan­
    tum efficiency of 21% at a wavelength of 395 nm for plane­
    wave photons incident normal to the PMT photocathode.
    Ninety percent of the detected photons are in the spectral
    range of 345–560 nm.
    An additional sensitivity effect arises from the ice sur­
    rounding the OMs. The deployment of OMs requires melting
    and refreezing of columns of ice, called ‘‘hole ice’’ hereafter.
    This cycle results in the formation of bubbles in the vicinity
    of the modules, which increase scattering and affects the
    sensitivity of the optical modules in ways that are not under­
    stood in detail. Since the total volume of hole ice is small
    compared to bulk ice in the detector
    ~
    columns of 60 cm di­
    ameter, compared to 30 m spacing between strings
    !
    , its effect
    on optical properties can be treated as a correction to the OM
    angular sensitivity. The increased scattering of photons in the
    hole ice has been simulated and compared to data taken with
    laser measurements
    in situ
    to assess the magnitude of this
    effect. This comparison provides an OM sensitivity correc­
    tion that reduces the relative efficiency in the forward direc­
    tion, but enhances it in the sideways and backward direc­
    tions. The sensitivity in the backward hemisphere
    (90° –180°) relative to the sensitivity integrated over all
    angles (0° –180°) of the optical sensor increases from 20%
    to 27%, due to this correction, while the average relative
    sensitivity in the forward direction (0° –90°) drops from
    80% to 73%. In other words, an OM becomes a somewhat
    more isotropic sensor.
    The effective angular sensitivity of the OMs was also as­
    sessed using the flux of downgoing atmospheric muons as a
    test beam illuminating both the 295 downward facing OMs
    and the 7 upward facing OMs. We assumed that the response
    of the upward facing OMs to light from downward muons is
    equivalent to the response of the downward facing OMs to
    light from upward moving muons. Based on this assumption
    we derived a modified angular response function
    ~
    later re­
    ferred to as
    ANGSENS
    !
    , which resulted in a effective reduc­
    tion of the absolute OM sensitivity in forward direction. In
    this model the effective relative sensitivity is 67% in the
    forward hemisphere, and 33% in the backward hemisphere.
    This correction will be used to estimate the effect of system­
    atic uncertainties in the angular response on the final neu­
    trino analysis.
    The simulation of the hardware response included the
    modeling of gains and thresholds and random noise at the
    levels measured for each OM. The transit times of the cables
    and the shapes of the photomultiplier pulses, ranging from
    170 to 360 ns full width at half maximum
    ~
    FWHM
    !
    , were
    included in the trigger simulation. Multi­photon pulses were
    simulated as superimposed single photoelectron waveforms.
    In all, some 8
    3
    10
    5
    seconds of cosmic rays were simulated,
    corresponding to 7% of the events contained in the 1997 data
    set.
    IV. EVENT RECONSTRUCTION
    The reconstruction of muon events in AMANDA is done
    offline, in several stages. First, the data are ‘‘cleaned’’ by
    removing unstable PMTs and spurious PMT signals
    ~
    or
    ‘‘hits’’
    !
    due to electronic or PMT noise. The cleaned events
    are then passed through a fast filtering algorithm, which re­
    duces the background of downgoing muons by one order of
    magnitude. This reduction allows the application of more
    sophisticated reconstruction algorithms to the remaining data
    set.
    Because of the complexity of the task, and in order to
    increase the robustness of the results, two separate analyses
    of the 1997 data set were undertaken. Both proceeded along
    the general lines described above, but differ in the details of
    implementation. The preliminary stages, which are very
    similar in both analyses, are described here. The particulars
    of each analysis will be described in Secs. V and VI. A more
    detailed description of the reconstruction procedure will be
    published elsewhere
    @
    20
    #
    .
    A. Cleaning and filtering
    The first step in reconstructing events is to clean and cali­
    brate the data recorded by the detector. Unstable channels
    ~
    OMs
    !
    are identified and removed on a run­to­run basis. On
    average, 260 of the 302 OMs deployed are used in the analy­
    ses. The recorded times of the hits are corrected for delays in
    the cables leading from the OMs to the surface electronics
    and for the amplitude­dependent time required for a pulse to
    cross the discriminator threshold. Hits are removed from the
    event if they are identified as being due to instrumental
    noise, either by their low amplitudes or short pulse lengths,
    or because they are isolated in space by more than 80 m and
    time by more than 500 ns from the other hits recorded in the
    event. Pulses with short duration, measured as the time over
    threshold
    ~
    TOT
    !
    , are often related to electronic cross­talk in
    the signal cables or the surface electronics. In analysis II,
    TOT cuts are applied to individual channels beyond the stan­
    dard cleaning common to both analyses
    ~
    see Sec. VI
    !
    .
    J. AHRENS
    et al.
    PHYSICAL REVIEW D
    66
    , 012005
    ~
    2002
    !
    012005­4

    Following the cleaning and the calibration, a ‘‘line fit’’ is
    calculated for each event. This fit is a simple
    x
    2
    minimiza­
    tion of the apparent photon flux direction, for which an ana­
    lytic solution can be calculated quickly
    @
    21
    #~
    see also
    @
    1
    #!
    .It
    contains no details of Cherenkov radiation or propagation of
    light in the ice. Hits arriving at time
    t
    i
    at PMT
    i
    located at
    r
    i
    W
    are projected onto a line. The minimization of
    x
    2
    5
    (
    i
    (
    r
    W
    i
    2
    r
    W
    0
    2
    v
    W
    lf
     
    t
    i
    )
    2
    gives a solution for
    r
    W
    0
    and a velocity
    v
    W
    lf
    . The
    results of this fit—at the first stage the direction
    v
    W
    lf
    /
    u
    v
    W
    lf
    u
    ,at
    later stages the absolute value of the velocity—are used to
    filter the data set. Approximately 80–90 % of the data, for
    which the line fit solution is steeply downgoing, are rejected
    at this stage.
    B. Maximum likelihood reconstruction
    After the data have been passed through the fast filter,
    tracks are reconstructed using a maximum likelihood
    method. The observed photon arrival times do not follow a
    simple Gaussian distribution attributable to electronic jitter;
    instead, a tail of delayed photons is observed. The photons
    can be delayed predominantly by scattering in the ice that
    causes them to travel on paths longer than the length of the
    straight line inclined at the Cherenkov angle to the track.
    Also, photons emitted by scattered secondary electrons gen­
    erated along the track will have emission angles other than
    the muon Cherenkov angle. These effects generate a distri­
    bution of arrival times with a long tail of delayed photons.
    We construct a probability distribution function describ­
    ing the expected distribution of arrival times, and calculate
    the likelihood
    L
    time
    of a given reconstruction hypothesis as
    the product of the probabilities of the observed arrival times
    in each hit OM:
    L
    time
    5
    )
    i
    5
    1
    N
    hit
    p
    ~
    t
    res
    (
    i
    )
    u
    d
    (
    i
    )
    ,
    u
    ori
    (
    i
    )
    !
    ~
    1
    !
    where
    t
    res
    5
    t
    obs
    2
    t
    Cher
    is the time residual
    ~
    the delay of the
    observed hit time relative to that expected for unscattered
    propagation of Cherenkov photons emitted by the muon
    !
    ,
    and
    d
    and
    u
    ori
    are the distance of the OM from the track and
    the orientation of the module with respect to the track. The
    probability distribution function
    p
    includes the effects of
    scattering and absorption in the bulk ice and in the refrozen
    ice around the modules. The functional form of
    p
    is based on
    a solution to a transport equation of the photon flux from a
    monochromatic point source in a scattering medium
    @
    22,23
    #
    .
    The free parameters of this function are then fit to the ex­
    pected time profiles that are obtained by a simulation of the
    photon propagation from muons in the ice
    @
    14,22
    #
    . Varying
    the track parameters of the reconstruction hypothesis, we
    find the maximum of the likelihood function, corresponding
    to the best track fit for the event. The result of the fit is
    described by five parameters: three (
    x
    ,
    y
    ,
    z
    ) to determine a
    reference point, and two (
    u
    ,
    f
    ) for the zenith and azimuth of
    the track direction. Figure 4 shows an event display of two
    upgoing muon events together with the reconstructed tracks.
    C. Quality parameters
    The set of apparently upgoing tracks provided by the re­
    construction procedure exceeds the expected number of up­
    going tracks from atmospheric neutrino interactions by one
    to three orders of magnitude, depending on the details of the
    reconstruction algorithm
    ~
    see Secs. V and VI
    !
    . In order to
    reject the large number of ‘‘fake events’’—events generated
    by a downgoing muon or cascade, but seemingly having an
    upgoing structure—we impose additional requirements on
    the reconstructed events to obtain a relatively pure neutrino
    sample. These requirements consist of cuts on observables
    derived from the reconstruction and on topological event pa­
    rameters. Below, we describe the most relevant of the param­
    eters used.
    FIG. 4. Event display of an upgoing muon event. The gray scale
    indicates the flow of time, with early hits at the bottom and the
    latest hits at the top of the array. The arrival times match the speed
    of light. The sizes of the circles correspond to the measured ampli­
    tudes.
    OBSERVATION OF HIGH ENERGY ATMOSPHERIC... PHYSICAL REVIEW D
    66
    , 012005
    ~
    2002
    !
    012005­5

    1. Reduced likelihood, L
    In analogy to a reduced
    x
    2
    , we define a reduced likeli­
    hood
    L
    5
    2
    ln
    L
    time
    N
    hit
    2
    5
    ~
    2
    !
    where
    N
    hit
    2
    5, the number of recorded hits in the event less
    the five track fit parameters, is the number of degrees of
    freedom. A smaller
    L
    corresponds to a higher quality of the
    fit.
    2. Number of direct hits, N
    dir
    The number of direct hits is defined as the number of hits
    with time delays
    t
    res
    smaller than a certain value. We use
    time intervals of
    @
    2
    15 ns,
    1
    25 ns
    #
    and
    @
    2
    15 ns,
    1
    75 ns
    #
    , and denote the corresponding parameters as
    N
    dir
    (25)
    and
    N
    dir
    (75)
    , respectively. The negative extent of the window
    allows for jitter in PMT rise times and for small errors in
    geometry and calibration, while the positive side includes
    these effects as well as delays due to scattering of the pho­
    tons. Events with many direct hits
    ~
    i.e. , only slightly delayed
    photons
    !
    are likely to be well reconstructed.
    3. Track length, L
    dir
    The track length is defined by projecting each of the direct
    hits onto the reconstructed track, and measuring the distance
    between the first and the last hit. A cut on this parameter
    rejects events with a small lever arm for the reconstruction.
    Direct hits with time residuals of
    @
    2
    15 ns,
    1
    75 ns
    #
    are
    used for the measurement of the track length. Cuts on the
    absolute length, as well as zenith angle dependent cuts
    ~
    which take into account the cylindrical shape of the detec­
    tor
    !
    have been used. The requirement of a minimum track
    length corresponds to imposing a muon energy threshold.
    For example, a track length of 100 m translates into a muon
    energy threshold of about 25 GeV.
    4. Smoothness, S
    The ‘‘smoothness’’ parameter is a check on the self­
    consistency of the fitted track. It measures the constancy of
    light output along the track. Highly variable apparent emis­
    sion of light usually indicates that the track either has been
    completely misreconstructed or that an underlaying muonic
    Cherenkov light was obscured by a very bright stochastic
    light emission, which usually leads to poor reconstruction.
    The smoothness parameter was inspired by the Kolmogorov­
    Smirnov test of the consistency of two distributions; in our
    case the consistency of the observed hit pattern with the hy­
    pothesis of constant light emission by a muon.
    Figure 5 shows two events to illustrate the characteristics
    of the smoothness parameter. One event is a long uniform
    track, which was well reconstructed. The other event is a
    background event which displays a very poor smoothness.
    The simplest definition of the smoothness is given by
    S
    5
    max
    ~
    u
    S
    j
    u
    !
    , where
    S
    j
    5
    j
    2
    1
    N
    2
    1
    2
    l
    j
    l
    N
    .
    ~
    3
    !
    Figure 6 illustrates the smoothness parameter for the two
    events displayed in Fig. 5. Here
    l
    j
    is the distance along the
    track between the points of closest approach of the track to
    the first and the
    j
    th
    hit modules, with the hits taken in order
    of their projected position on the track.
    N
    is the total number
    FIG. 5. Two muon events: The upgoing muon event shown on
    the left has a smooth distribution of hits along the track. The track­
    like hit topology of this event can be used to distinguish it from
    background events. The event on the right is a background event
    with a poor smoothness value.
    J. AHRENS
    et al.
    PHYSICAL REVIEW D
    66
    , 012005
    ~
    2002
    !
    012005­6

    of hits. Tracks with hits clustered at the beginning or end of
    the track have
    S
    j
    approaching
    1
    1or
    2
    1, leading to
    S
    5
    1.
    High quality tracks such as the event on the left side of Fig.
    5, with
    S
    close to zero, have hits equally spaced along the
    track.
    5. Sphericity
    Treating the hit modules as point masses, we can form a
    tensor of inertia for each event, describing the spatial distri­
    bution of the hits. Diagonalizing the tensor of inertia yields
    as eigenvalues
    I
    i
    the moments of inertia about the principal
    axes of rotation. For a long, cylindrical distribution of hit
    modules, two moments will be much larger than the third.
    We can reject spherical events, such as those produced by
    muon bremsstrahlung, by requiring that the normalized mag­
    nitude of the smallest moment,
    I
    1
    /
    (
    I
    i
    , be small.
    D. Principal methods of the analyses
    The two analyses of the data diverge after the filtering
    stage, following different approaches to event reconstruction
    and background rejection.
    Analysis I uses an improved likelihood function based on
    a more detailed description of the photon response
    @
    22
    #
    , fol­
    lowed by a set of stepwise tightened cuts. Analysis II uses a
    Bayesian reconstruction
    @
    24
    #
    in which the likelihood is mul­
    tiplied by a zenith angle dependent prior function, resulting
    in a strong rejection of downgoing background.
    Rare backgrounds due to unsimulated instrumental ef­
    fects, such as cross­talk between signal channels and un­
    stable voltage supply, were identified in the course of the
    analyses. These effects either produced spurious triggers, or,
    more often, spurious hits that caused the event to be misre­
    constructed. Different but comparably efficient techniques
    were developed to treat these backgrounds. In analysis I the
    event topology is inspected; if the spatial pattern of hit OMs
    is inconsistent with the reconstructed muon trajectory, the
    event is rejected. Analysis II attempts to remove the anoma­
    lous hits or triggers through identification of characteristic
    correlations in signal amplitudes and times, which consider­
    ably reduces the rate of these misreconstructions.
    At this stage the data set in each analysis is reduced to
    several thousand events out of the original 1.05
    3
    10
    9
    , but the
    data are still background dominated. The prediction for at­
    mospheric neutrinos is about 500 at this point.
    For the final selection of a nearly pure sample of neutrino
    induced events, cuts on characteristic observables are tight­
    ened until the remaining background disappears. The two
    analyses use different techniques to choose their final cuts,
    but obtain comparable efficiencies. Further details of the
    analyses can be found in Refs.
    @
    25–27
    #
    .
    V. ANALYSIS I
    In this analysis the data were processed through three lev­
    els of initial cuts, designed to reduce the number of back­
    ground events to a manageable size for the final cut evalua­
    tion. After a first filtering based on the line fit
    ~
    level 1
    !
    , cuts
    on the zenith angle, the number of direct photons, and the
    likelihood of the fitted track obtained by the maximum like­
    lihood reconstruction were applied
    ~
    level 2
    !
    .
    A. Removal of cascade­like events and detector artifacts
    A third filter level used the results of an iterative likeli­
    hood reconstruction with varying track initializations, a fit
    based on the hit probabilities
    @
    see Eq.
    ~
    4
    !#
    and a reconstruc­
    tion to the hypothesis of a high energy cascade, e.g., due to a
    bright seconday muon bremsstrahlung interaction.
    The first two levels of filtering consisted of relatively
    weak cuts on basic parameters like the zenith angle and like­
    lihood. They reduced the data set to about 4
    3
    10
    5
    events. At
    this stage, residual unsimulated instrumental features become
    apparent, e.g., comparatively high amplitude cross­talk pro­
    duced when a downgoing muon emits a bright shower in the
    center of the detector. Such events are predominantly recon­
    structed as moving vertically upward and can be identified in
    the distribution of the center of gravity
    ~
    COG
    !
    of hits. Its
    vertical component (
    z
    COG
    ) shows unpredicted peaks in the
    middle and the bottom of the detector
    @
    see also Fig. 14
    ~
    top
    !
    ,
    demonstrating the effect for analysis II
    #
    , while the horizontal
    components (
    x
    COG
    and
    y
    COG
    ) show an enhancement of hits
    towards the outer strings. These strings are read out via
    twisted pair cables, as opposed to the coaxial cables used on
    the inner strings. The twisted pair cables were found to be
    more susceptible to cross­talk signals. Note that variations in
    the optical parameters of the ice due to past climatological
    episodes also produce some vertical structure.
    We developed additional COG cuts on the topology of the
    events in order to remove these backgrounds. These cuts,
    which depend on the reconstructed zenith angle, use the
    track lengths
    L
    dir
    and the normalized smallest eigenvalues of
    the tensor of inertia (
    I
    1
    /
    (
    I
    i
    ).
    FIG. 6. Illustration of the smoothness parameter, which com­
    pares the observed distribution of hits to that predicted for a muon
    emitting Cherenkov light. In the simplest formulation, shown here,
    the prediction is given as a straight line. A large deviation from a
    straight line
    ~
    0. 68
    !
    is found for the event on the right in Fig. 5. The
    high quality track­like event on the left in Fig. 5 displays a small
    deviation
    ~
    0.09
    !
    .
    OBSERVATION OF HIGH ENERGY ATMOSPHERIC... PHYSICAL REVIEW D
    66
    , 012005
    ~
    2002
    !
    012005­7

    Figure 7 shows the different components of the center of
    gravity of the hits and the reconstructed zenith angle before
    and after application of the COG cuts, and the Monte Carlo
    prediction for fake upward events stemming from misrecon­
    structed downgoing muons. The cuts remove most of the
    unsimulated background—in particular that far from the
    horizon—and bring experiment and simulation into much
    better agreement.
    In order to verify the signal passing rates, these cuts and
    those from the previous levels were applied to a subsample
    of unfiltered
    ~
    i.e., downgoing
    !
    events but with the zenith
    angle dependence of the cuts reversed, thus using the abun­
    dant cosmic ray muons as stand­ins for upgoing muons.
    In all, these three levels of filtering reduced the data set by
    a factor of approximately 10
    5
    ~
    see Table II
    !
    .
    B. Multi­photoelectron likelihood and hit likelihood
    Before the final cut optimization the last, most elaborate
    reconstruction was applied, combining the likelihoods for the
    arrival time of the first of muliple photons in a PMT with the
    likelihoods for PMTs to have been hit or have not been hit.
    The probability densities
    p
    (
    t
    res
    (
    i
    )
    u
    d
    (
    i
    )
    ,
    u
    ori
    (
    i
    )
    )
    @
    see Eq.
    ~
    1
    !
    ,
    Sec. IV B
    #
    describe only the arrival times of single photons.
    Density functions for the multi­photoelectron case have to
    include the effect of repeatedly sampling the distribution of
    photon arrival times. For several detected photons, the first
    of them is usually less scattered than the average photon
    ~
    which defines the single photoelectron case
    !
    . Therefore the
    leading edge of a PMT pulse composed of multiple photo­
    electrons
    ~
    MPE
    !
    will be systematically shifted to earlier
    times compared to a single photoelectron. The
    MPE likeli­
    hood
    L
    time
    MPE
    @
    22
    #
    uses the recorded amplitude information to
    model this shift.
    In the reconstructions mentioned so far, the timing infor­
    mation from hit PMTs was used. However, a PMT which
    was
    not
    hit also delivers information. The
    hit likelihood
    L
    hit
    does not depend on the arrival times but represents the prob­
    ability that the track produced the observed hit pattern. It is
    constructed from the probability densities
    p
    hit
    (
    d
    (
    i
    )
    ,
    u
    ori
    (
    i
    )
    ) that
    a given PMT
    i
    was hit if it was in fact hit, and the probabili­
    ties
    @
    1
    2
    p
    hit
    (
    d
    (
    j
    )
    ,
    u
    ori
    (
    j
    )
    )
    #
    that a given PMT
    j
    was not hit if it
    was not hit:
    L
    hit
    5
    )
    i
    5
    1
    N
    hit
    p
    hit
    ~
    d
    (
    i
    )
    ,
    u
    ori
    (
    i
    )
    !
    )
    i
    5
    N
    hit
    1
    1
    N
    OM
    ~
    1
    2
    p
    hit
    ~
    d
    (
    i
    )
    ,
    u
    ori
    (
    i
    )
    !!
    ~
    4
    !
    where the first product runs over all hit PMTs and the second
    over all non­hit PMTs.
    The likelihood combining these two probabilities is
    FIG. 7. Characteristic distributions of the cen­
    ter of gravity
    ~
    COG
    !
    of events. The figures on the
    left show the distribution of the depth
    z
    COG
    ver­
    sus the reconstructed zenith angle. The figures on
    the right show the horizontal location of events in
    the
    x
    COG
    ­
    y
    COG
    plane of events with 0 m
    ,
    z
    COG
    ,
    50 m. The positions of the strings are marked
    by stars. Top: Experimental data before applica­
    tion of the COG cuts. Middle: Experimental data
    after application of the COG cuts. Bottom: Ex­
    pectation from the BG simulation after cuts.
    J. AHRENS
    et al.
    PHYSICAL REVIEW D
    66
    , 012005
    ~
    2002
    !
    012005­8

    L
    5
    L
    time
    MPE
     
    L
    hit
    .
    ~
    5
    !
    A cut on the reconstructed zenith angle obtained from
    fitting with
    L
    leaves less than 10
    4
    events in the data set,
    defined as level 4 in Fig. 8.
    C. Final separation of the neutrino sample
    For the final stage of filtering, a method
    ~
    CUTEVAL
    !
    was
    developed to select and optimize the cuts taking into account
    correlations between the cut parameters. A detailed descrip­
    tion of this method can be found in
    @
    27
    #
    . The principle of
    CUTEVAL
    is to numerically optimize the ratio of signal to
    A
    background by variation of the selection of cut parameters,
    as well as the actual cut values. Parameters are used only if
    they improve the efficiency of separation over optimized cuts
    on all other already included parameters. A first optimization
    was based purely on Monte Carlo simulations, with simu­
    lated atmospheric neutrinos for signal and simulated down­
    going muons forming the background. This optimization
    yielded four such independent parameters. Two other optimi­
    zations involved experimental data. In both cases, experi­
    mental data have been defined as the background sample. In
    one case, the signal was represented by atmospheric neutrino
    Monte Carlo simulations, in the other by experimental data
    subjected to zenith angle inverted cuts
    ~
    i.e., to downward
    events passing the quality cuts, but being ‘‘good’’ events
    with respect to the upper hemisphere instead—like neutrino
    candidates—with respect to the lower hemisphere
    !
    . These
    latter optimizations yielded two additional parameters, which
    rejected a small contribution of residual unsimulated back­
    grounds: coincident muons from simultaneous independent
    air showers and events accompanied by instrumental artifacts
    such as cross­talk. After application of these two cuts to
    simulated and experimental data, the distributions of observ­
    ables agree to a satisfactory precision.
    Once the minimal set of parameters is found, the optimal
    cut values can be represented as a function of the number of
    background events
    N
    BG
    passing the cuts. The result is a path
    through the cut parameter space which yields the best signal
    efficiency for any desired purity of the signal, characterized
    by
    N
    BG
    . Using this representation, one can calculate the
    number of events passing the cuts as a function of the fitted
    N
    BG
    for signal and for background Monte Carlo program.
    Figure 8
    ~
    top
    !
    shows this dependence for simulations as well
    as for experimental data, with
    N
    BG
    varying from trigger level
    to a level that leaves only a few events in the data set. One
    observes that the actual background expectation falls roughly
    linearly as the fitted
    N
    BG
    is reduced. Below values of a few
    hundred events the signal is expected to dominate the event
    sample. The experimental curve follows the expectation from
    the sum of background and signal Monte Carlo program. For
    large
    N
    BG
    , the observed event rate follows the background
    expectation. At smaller
    N
    BG
    , the experimental shape turns
    over into the signal expectation and follows it nicely down to
    the sample of events with highest quality
    ~
    the smallest values
    of
    N
    BG
    !
    . For a moderate background contamination of
    N
    BG
    5
    10, one gets a total of 223 neutrino candidates. The param­
    eters and cut values as obtained by the
    CUTEVAL procedure
    are summarized in Table I.
    Figure 8
    ~
    bottom
    !
    translates the background parameter
    N
    BG
    into an event quality parameter
    Q
    , defined as
    Q
    [
    ln(
    N
    0
    /
    N
    BG
    )
    5
    ln(1.05
     
    10
    9
    /
    N
    BG
    ). The plot shows the ratios
    FIG. 8. The fitted background parameter
    N
    BG
    . Top: Number of
    events versus
    N
    BG
    . Smaller values of
    N
    BG
    correspond to harder
    cuts. Below
    N
    BG
    5
    1500 the
    CUTEVAL
    parametrization was used to
    calculate the cut values corresponding to
    N
    BG
    . For larger values of
    N
    BG
    the data points correspond to the cuts from the filter levels:
    Level 4
    ~
    see Sec. V B
    !
    , level 3, level 2, level 1, and trigger level
    ~
    Table II
    !
    . Bottom: Ratios of events passing in the experimental
    data compared to various Monte Carlo expectations for signal and
    background as a function of event quality. The dashed line indicates
    the final cuts.
    OBSERVATION OF HIGH ENERGY ATMOSPHERIC... PHYSICAL REVIEW D
    66
    , 012005
    ~
    2002
    !
    012005­9

    of events from the upper figure as a function of
    Q
    . At higher
    qualities (
    Q
    .
    17), the ratio of observed events to the atmo­
    spheric neutrino simulation flattens out with a further varia­
    tion of only 30%. The value at
    Q
    5
    17 is approximately 0.6
    for the standard Monte Carlo program
    ~
    chosen in Fig. 8, top
    !
    and approximately unity for the
    ANGSENS Monte Carlo pro­
    gram
    ~
    chosen in Fig. 8, bottom
    !
    .
    Table II lists the cut efficiencies for the atmospheric neu­
    trino simulation
    ~
    with and without the implementation of the
    angular sensitivity fitted model
    ANGSENS
    of the OMs—see
    Secs. III and VII
    !
    , the background simulation of atmospheric
    muons from air showers
    ~
    without
    ANGSENS
    !
    and the experi­
    mental data. Again, the experimental numbers agree well
    with the background simulation up to the first two filter lev­
    els. Later, the Monte Carlo program underestimates the ex­
    perimental passing rates slightly. The last row shows the ex­
    pected numbers of events for the last stage of filtering. If, in
    addition, the effect of neutrino oscillations
    ~
    see Sec. VII
    !
    is
    included, the atmospheric neutrino simulation including the
    ANGSENS
    model predicts 224 events, in closest agreement
    with the experiment. However, the 5% effect due to oscilla­
    tions is smaller than our systematic uncertainty
    ~
    see Sec.
    VII
    !
    .
    D. Characteristics of the neutrino candidates
    1. Time distribution
    Figure 9 shows the cumulative number of neutrino events
    as well as the cumulative number of event triggers plotted
    versus the day number in 1997. One can observe that the
    neutrino events follow the number of triggers, albeit with a
    small deficit during the Antarctic winter. This deficit is con­
    sistent with statistical fluctuations.
    ~
    Actually, seasonal varia­
    tions slightly
    de
    crease the downward muon rate during the
    Antarctic winter
    @
    28
    #
    and should result in a 10% deficit of
    triggers with respect to upward neutrino events.
    !
    TABLE I. Final quality parameters and cuts obtained from the cut evaluation procedure. The ‘‘direct’’
    time interval for variables
    N
    dir
    ,
    L
    dir
    , and
    S
    dir
    is
    @
    2
    15 ns,
    1
    75 ns
    #
    . The first four rows show cut parameters
    obtained by all
    ~
    Monte Carlo and experimental
    !
    searches; the last two rows show two additional
    ~
    weaker
    !
    cuts, which were found to remove unsimulated backgrounds.
    Parameter Cut Explanation
    u
    S
    u
    ,
    0.28 See Sec. IV C 4
    u
    S
    P
    hit
    u
    /(
    u
    mpe
    2
    90°)
    ,
    0.01 Tightens the requirement on the smoothness for
    tracks
    close to the horizon where background is high
    (
    N
    dir
    2
    2)
     
    L
    dir
    .
    750 m Lever arm of the track times the number of
    supporting points
    log(
    L
    up
    /
    L
    down
    )
    ,2
    7.7 Ratio of the likelihoods of the best
    upgoing and best downgoing hypotheses
    C
    (mpe,lf)
    ,
    35° Space angle between the results from the
    multi­photon
    likelihood reconstruction and the line fit. This cut
    effectively removes cross­talk features.
    A
    (
    S
    dir
    )
    2
    1
    (
    S
    dir
    P
    hit
    )
    2
    ,
    0.55 Parameter combining the two smoothness
    definitions
    ~
    here calculated using only direct hits
    !
    .
    This cut effectively removes coincident muon
    events from independent air showers.
    TABLE II. The cut efficiencies for the atmospheric neutrino Monte Carlo
    ~
    MC
    !
    prediction, the atmo­
    spheric muon background Monte Carlo prediction, and the experimental data for 130 days of detector
    lifetime. Efficiencies are given for filter levels L1 to L4. L4 is the final selection. All errors are purely
    statistical. The final background prediction of 7 events has been normalized at trigger level.
    Filter level Atm.
    n
    Atm.
    n
    MC Atm.
    m
    MC Experimental
    MC
    ANGSENS
    ~
    Background
    !
    data
    Events at trigger level 8978 5759 9.03
    3
    10
    8
    1.05
    3
    10
    9
    Efficiency at level 1 0.34 0.37 0.4
    3
    10
    2
    1
    0.5
    3
    10
    2
    1
    Efficiency at level 2 0.15 0.15 0.4
    3
    10
    2
    3
    0.4
    3
    10
    2
    3
    Efficiency at level 3 0.7
    3
    10
    2
    1
    0.7
    3
    10
    2
    1
    0.7
    3
    10
    2
    5
    0.1
    3
    10
    2
    4
    Efficiency after final cuts 0.4
    3
    10
    2
    1
    0.4
    3
    10
    2
    1
    0.6
    3
    10
    2
    8
    0.2
    3
    10
    2
    6
    No. of events 362
    6
    4 237
    6
    67
    6
    5 223
    passing final cuts normalized
    J. AHRENS
    et al.
    PHYSICAL REVIEW D
    66
    , 012005
    ~
    2002
    !
    012005­10

    2. Zenith angle distribution
    Figure 10 shows the zenith angle distribution of the 223
    neutrino candidates compared to the Monte Carlo prediction
    for atmospheric neutrinos
    @
    17
    #
    and the few remaining events
    predicted by background simulations. Note that the Monte
    Carlo prediction is normalized to experiment.
    ~
    The total
    number of events is 362 for the atmospheric neutrino simu­
    lation and 223 for experiment, i.e., there is a deficit of 39
    percent in the absolute number of events.
    !
    There is good
    agreement between the prediction and the experiment in the
    shape of the angular distribution.
    3. Characteristic distributions and visual inspection
    Four methods were used to evaluate the effectiveness of
    the analysis and the level of residual backgrounds:
    ~
    a
    !
    N
    2
    1
    cuts
    ,
    ~
    b
    !
    unbiased variables
    ,
    ~
    c
    !
    low level distributions
    ,
    and
    ~
    d
    !
    visual inspection
    .
    ~
    a
    !
    The
    N
    2
    1
    test
    evaluates the
    N
    final cuts one by one
    and yields an estimate of the background contamination in
    the final sample. One applies all but one of the final cuts
    ~
    the
    one in the selected variable
    !
    , and plots the data in this vari­
    able. In the signal region of this variable
    ~
    defined by the later
    applied cut
    !
    shapes of experiment and signal Monte Carlo
    program should agree. In the background region, the experi­
    mental data should approach the expected background shape.
    Figure 11 shows four of these distributions. The applied cut
    is shown by a dotted line. All four cuts satisfy the test: the
    shape of the distributions agree reasonably well on both sides
    of the applied cuts. Two
    N
    2
    1 distribution from analysis II
    are shown in Fig. 19.
    ~
    b
    !
    An obvious test is the investigation of distributions of
    unbiased variables
    ~
    i.e., variables to which no cuts have
    been applied
    !
    in the final neutrino sample. Here, the experi­
    mental distributions follow the Monte Carlo signal expecta­
    tions nicely. Some deviations are observed, especially in the
    number of OMs hit and the velocity
    v
    lf
    obtained from the
    line fit
    ~
    see Sec. IV A
    !
    . However, as can be seen from Fig.
    20, part of these disagreements disappear if the standard at­
    mospheric neutrino MC program is replaced by the
    ANGSENS
    MC version.
    ~
    c
    !
    In order to account for possible pathological
    low level
    features in the data sample
    ~
    especially cross­talk
    !
    ,we
    ~
    i
    !
    in­
    vestigated basic pulse amplitude and pulse width
    ~
    TOT
    !
    dis­
    tributions and
    ~
    ii
    !
    re­fitted all events after the cross­talk hit
    cleaning procedure applied in analysis II
    ~
    which is tighter
    than the standard cross­talk cleaning introduced in Sec.
    IV A
    !
    . Both these distributions and that for the recalculated
    zenith angles show no significant deviation from the previ­
    ous ones. No cross­talk features are found in the resulting
    neutrino sample.
    ~
    d
    !
    Finally, a
    visual inspection
    of the full neutrino sample
    was performed, by visually displaying each event like in Fig.
    4. The visual inspection gives consistent results with the
    other methods of background estimation and yields an upper
    limit on the background contamination of muons from ran­
    dom coincident air showers
    ~
    see below
    !
    .
    E. Background estimation
    The results of four independent methods of background
    estimation are summarized in Table III.
    First, the background Monte Carlo program itself gives an
    estimate. It yields 7 events if rates are normalized to the
    trigger level
    ~
    see Table II
    !
    . Because the passing rates differ
    slightly between the experiment
    ~
    higher
    !
    and the background
    Monte Carlo program
    ~
    lower
    !
    , we made the conservative
    choice to renormalize the background Monte Carlo program
    to the level 3 experimental passing rate. This gives an esti­
    mate of about 16 background events in our final sample.
    From the
    N
    2
    1 distributions we obtained an alternative
    approximation of the residual background. We re­normalized
    both signal and background MC events in the background
    region to fit the number of experimental events in the back­
    ground region. The number of re­normalized background
    FIG. 9. The integrated exposure of the AMANDA detector in
    1997. The figure shows the cumulative number of triggers
    ~
    upper
    curve
    !
    and the number of observed neutrino events
    ~
    lower curve
    !
    versus the day number. The intervals with zero gradient correspond
    to periods where the detector was not operating stably; data from
    these periods were excluded from the analysis.
    FIG. 10. Zenith angle distribution of the experimental data com­
    pared to simulated atmospheric neutrinos and a simulated back­
    ground of downgoing muons produced by cosmic rays. In this fig­
    ure the Monte Carlo prediction is normalized to the experimental
    data. The error bars report only statistical errors.
    OBSERVATION OF HIGH ENERGY ATMOSPHERIC... PHYSICAL REVIEW D
    66
    , 012005
    ~
    2002
    !
    012005­11

    MC events in the signal region is then a background esti­
    mate. This estimate was performed
    N
    times
    ~
    once for each
    N
    2
    1 distribution
    !
    . The average over all
    N
    estimations yields
    14 background events. Note that this averaging procedure is
    reasonable only for the case of independent cuts. With the
    method by which we have chosen the cut parameters, this
    condition is satisfied to first approximation.
    We have found that cross­talk hits are related to the char­
    acteristic triple­peak structure in the distribution of the ver­
    tical component of the center of gravity of hits (
    z
    COG
    ) which
    has been discussed in Sec. V A—see Fig. 7 and also Fig. 14
    ~
    top
    !
    . Since there are remaining cross talk hits which have
    survived the standard cleaning
    ~
    see Sec. IV A
    !
    , this distribu­
    tion was studied in detail. As shown in Fig. 12, the final
    experimental sample of neutrino candidates shows no statis­
    tically significant excess with respect to the atmospheric neu­
    trino Monte Carlo prediction in the regions of the character­
    istic peaks. Therefore, an upper limit on this special class of
    background was derived and yields
    ,
    35 events.
    The visual inspection of the neutrino sample yields 13
    events. Seven of them show the signature of coincident
    muons from independent air showers; i.e., two well separated
    spatial concentrations of hits, each with a downward time
    flow but with the lower group appearing earlier than the up­
    per one. Taking into account the scanning efficiencies which
    were determined by scanning signal and background Monte
    Carlo events, an upper limit of 23 events is obtained from
    visual inspection.
    Combining the results from the above methods, the ex­
    pected background is estimated to amount to 4 to 10% of the
    223 experimental events.
    FIG. 11. Two distributions of variables used as cut parameters in
    the last filter level
    ~
    see Table I for an explanation of the variables
    !
    .
    In both cases, all final cuts with the exception of the variable plotted
    have been applied. The cuts on the displayed parameters are indi­
    cated by the dashed vertical lines. Arrows indicate the accepted
    parameter space.
    FIG. 12. Distributions of
    z
    COG
    for the experiment and atmo­
    spheric neutrino signal Monte Carlo program
    MC standard
    and
    MC
    bulk ice
    denote two different ice models. The first includes vertical
    ice layers in accordance with Fig. 2; the second uses homogeneous
    ice.
    TABLE III. Various estimates of the background remaining in
    the experimental data sample of 223 neutrino candidates.
    BG estimation method Estimation
    BG MC 16
    6
    8
    N
    2
    1 cuts 14
    6
    4
    z
    COG
    distributions
    ,
    35
    Visual inspection
    ,
    23
    J. AHRENS
    et al.
    PHYSICAL REVIEW D
    66
    , 012005
    ~
    2002
    !
    012005­12

    VI. ANALYSIS II
    The second analysis follows a different approach; instead
    of optimizing cuts to reject misreconstructed cosmic ray
    muons, this analysis concentrates on improving the recon­
    struction algorithm with respect to background rejection. The
    large downgoing muon flux implies that even a small frac­
    tion of downgoing muons misreconstructed as upgoing will
    produce a very large background rate. Equivalently, for each
    apparently upgoing event, there were many more downgoing
    muons passing the detector than there were upgoing muons;
    even though any single downgoing muon had only a small
    probability of faking an upgoing event, the total probability
    that the event was a fake is quite high.
    A. Bayesian reconstruction
    This analysis of the problem motivates a Bayesian ap­
    proach
    @
    24
    #
    to event reconstruction. Bayes’ theorem in prob­
    ability theory states that for two assertions
    A
    and
    B
    ,
    P
    ~
    A
    u
    B
    !
    P
    ~
    B
    !
    5
    P
    ~
    B
    u
    A
    !
    P
    ~
    A
    !
    ,
    where
    P
    (
    A
    u
    B
    ) is the probability of assertion
    A
    given that
    B
    is true. Identifying
    A
    with a particular muon track hypothesis
    m
    and
    B
    with the data recorded for an event in the detector,
    we have
    P
    ~
    m
    u
    data
    !
    5
    L
    time
    ~
    data
    u
    m
    !
    P
    ~
    m
    !
    ,
    where we have dropped a normalization factor
    P
    (data)
    which is a constant for the observed event. The function
    L
    time
    is the regular likelihood function of Eq.
    ~
    1
    !
    , and
    P
    (
    m
    )
    is the so­called prior function, the probability of a muon
    m
    5
    m
    (
    x
    ,
    y
    ,
    z
    ,
    u
    ,
    f
    ) passing through the detector.
    For this analysis, we have used a simple one­dimensional
    prior function, containing the zenith angle information at
    trigger level in Fig. 3. By accounting in the reconstruction
    for the fact that the flux of downgoing muons from cosmic
    rays is many orders of magnitude larger than that of upgoing
    neutrino­induced muons, the number of downgoing muons
    that are misreconstructed as upgoing is greatly reduced. It
    should be noted that the objections that are often raised with
    respect to the use of Bayesian statistics in physics are not
    relevant to this problem: the prior function is well defined
    and normalized and independently known to relatively good
    precision, consisting only of the fluxes of cosmic ray muons
    and atmospheric muon neutrinos.
    B. Removal of instrumental artifacts
    The Bayesian reconstruction algorithm is highly efficient
    at rejecting downgoing muon events. Of 2.6
    3
    10
    8
    events
    passing the fast filter, only 5.8
    3
    10
    4
    are reconstructed as up­
    going. By contrast, the standard maximum likelihood recon­
    struction produces about 2.4
    3
    10
    7
    false upgoing reconstruc­
    tions. However, less than a thousand neutrino events are
    predicted by Monte Carlo program, so it is clear that a sig­
    nificant number of misreconstructions remain.
    Detailed inspection of the 5.8
    3
    10
    4
    events reveals that the
    vast majority is produced by cross­talk overlaid on triggers
    from downgoing muons emitting bright stochastic light near
    the detector. This cross­talk confuses the reconstruction al­
    gorithm, producing apparently upgoing tracks. Because
    cross­talk is not included in the detector simulation, the char­
    acteristics of the fakes are not predicted well by the simula­
    tion, and the rate of misreconstruction is much higher than
    predicted.
    The cross­talk is removed by additional hit cleaning rou­
    tines developed by examination of this cross­talk enriched
    data set. For example, cross­talk in many channels can be
    identified in scatter plots of pulse width vs amplitude, as
    shown in Fig. 13. The pulse width is measured as time over
    threshold
    ~
    TOT
    !
    . Real hits form the distribution shown on
    the left. High amplitude pulses should have large pulse
    width. This is not the case for cross­talk induced pulses. In
    channels with high levels of cross­talk, an additional vertical
    FIG. 13. Pulse amplitude vs duration for modules on the outer
    strings. Normal hits lie in the distribution shown in the upper figure.
    High amplitude pulses of more than a few photoelectrons are valid
    only if the pulse width is also large. Cross­talk induced pulses of
    high amplitude are characterized by small time over threshold
    ~
    TOT
    !
    . The cutoff seen at high amplitude is due to saturation of the
    amplitude readout electronics.
    OBSERVATION OF HIGH ENERGY ATMOSPHERIC... PHYSICAL REVIEW D
    66
    , 012005
    ~
    2002
    !
    012005­13

    band is found at high amplitudes but short pulse widths, as
    seen in the lower figure.
    Other hit cleaning algorithms use the time correlation and
    amplitude relationship between real and cross­talk pulses and
    a map of channels susceptible to cross­talk and the channels
    to which they are coupled. An additional instrumental effect,
    believed to be caused by fluctuating high voltage levels, pro­
    duces triggers with signals from most OMs on the outer
    strings but none on the inner four strings; some 500 of these
    bogus triggers were also removed from the data set. The
    5.8
    3
    10
    4
    upgoing events were again reconstructed after the
    additional hit cleaning was applied. Only 4.9
    3
    10
    3
    ~
    8.4%
    !
    of
    the events remained upgoing, compared to an expectation
    from Monte Carlo program of 1855 atmospheric muon
    events
    ~
    37.8% of the total before the additional cleaning
    !
    ,
    and 555 atmospheric neutrino events. Figure 14
    ~
    top
    !
    shows
    that while there has been a significant reduction in the instru­
    mental backgrounds, an unsimulated structure still remains
    in the center­of­gravity
    ~
    COG
    !
    distribution for these remain­
    ing data events. The application of additional quality criteria
    brings this distribution in agreement, as shown in Fig. 14
    ~
    bottom
    !
    .
    C. Quality cuts
    The improvements in the reliability of the reconstruction
    algorithm described above obviated the need for large num­
    bers of cut parameters or for careful optimization of the cuts.
    Because the signal­to­noise ratio of the upward­reconstructed
    data is quite high to begin with, we have the possibility of
    comparing the behavior of real and simulated data over a
    wide range of cut strengths to verify that the data agree with
    the predictions for upgoing neutrino­induced muons, not
    only in number but also in their characteristics. Using the cut
    parameters described in Sec. IV C
    ~
    with the likelihood re­
    placed by the Bayesian posterior probability
    !
    and a require­
    ment that events fitted as relatively horizontal by the line fit
    filtering algorithm not be reconstructed as steeply upgoing
    by the full reconstruction
    ~
    a requirement that suppresses re­
    sidual cross­talk misreconstructions
    !
    , an index of event qual­
    ity was formed.
    To do so, we rescale the six quality parameters described
    above by the cumulative distributions of the simulated atmo­
    spheric neutrino signal, and consider the six­dimensional cut
    space formed by the rescaled parameters. A point in this
    space corresponds to fixed values of the quality parameters,
    and events can be assigned to locations based on their track
    length, sphericity, and so forth.
    It is difficult to compare the distributions of data and
    simulated up­ and downgoing muons directly because of the
    high dimensionality of the space. We therefore project the
    space down to a single— ‘‘quality’’ — dimension by divid­
    ing it into concentric rectangular shells, as illustrated in Fig.
    15. The vertex of each shell lies on a line from the origin
    through a reference set of cuts which are believed to isolate a
    fairly pure set of neutrino events. Events in the full cut space
    are assigned an overall quality value, based on the shell in
    which they lie.
    FIG. 14. Top: Event center of gravity distribution after recon­
    struction with special cross­talk cleaning algorithms applied to the
    events. Unsimulated background remains. Bottom: The data agree
    with the neutrino signal after application of additional quality cuts.
    FIG. 15. Definition of event quality. Events are plotted in
    N
    ­dimensional cut space
    ~
    two dimensions are shown here for clar­
    ity
    !
    . A line is drawn from the origin
    ~
    no cuts
    !
    through a selected set
    of cuts, and the space is divided into rectangular shells of equal
    width. Events are assigned a quality
    q
    according to the shell in
    which they are found.
    J. AHRENS
    et al.
    PHYSICAL REVIEW D
    66
    , 012005
    ~
    2002
    !
    012005­14

    With this formulation we can compare the characteristics
    of the data to simulated neutrino and cosmic­ray muon
    events. Figure 16 compares the number of events passing
    various levels of cuts; i.e., the integral number of events
    above a given quality. At low qualities,
    q
    <
    3, the data set is
    dominated by misreconstructed downgoing muons, data as
    well as the simulated background exceed the predicted neu­
    trino signal. At higher qualities, the passing rates of data
    closely track the simulated neutrino events, and the predicted
    background contamination is very low.
    We can investigate the agreement between data and
    Monte Carlo simulations more systematically by comparing
    the differential number of events
    within
    individual shells,
    rather than the total number of events passing various levels
    of cuts. This is done in Fig. 17, where the ratios of the
    number of events observed to those predicted from the com­
    bined signal and background simulations are shown. One can
    see that at low quality levels there is an excess in the number
    of misreconstructed events observed. This is mainly due to
    remaining cross­talk. There is also an excess, though statis­
    tically less significant, at very high quality levels, which is
    believed to be caused by slight inaccuracies in the descrip­
    tion of the optical parameters of the ice. Nevertheless, over
    the bulk of the range there is close agreement between the
    data and the simulation, apart from an overall normalization
    factor of 0.58. The absolute agreement is consistent with the
    systematic uncertainties. It should be emphasized that the
    quality parameter is a convolution of all six quality param­
    eters, and so the flat line in Fig. 17 demonstrates agreement
    in the correlations between cut parameters.
    D. Background estimation and signal description
    If we reduce the 4917 upward­reconstructed events by
    requiring a quality of at least 7 on the scale of Fig. 16, we
    obtain a set of 204 neutrino candidates. The background con­
    tamination, which is due to misreconstructed downgoing
    muons, was estimated in three ways. The first way is to
    simulate the downgoing muon flux, bearing in mind that we
    are looking at a very low tail (10
    2
    8
    ) of the total muon dis­
    tribution. The second way is to renormalize the signal simu­
    lation by the factor of 0.58 obtained from Fig. 17 and sub­
    tract the predicted events from the observed data set
    ~
    accepting the excess at extremely high qualities, however,
    as signal
    !
    . The third way, a cross check on the first two
    methods, is to examine the data looking for fakes due to
    unsimulated effects such as cross­talk, independent coinci­
    dent downgoing muons, and so forth. All three methods yield
    estimates of 5–10 % contamination.
    The zenith angle distribution for the 204 events is shown
    in Fig. 18, and compared to that for the simulation of atmo­
    spheric neutrinos. In the figure the Monte Carlo events are
    normalized to the number of observed events to facilitate
    comparison of the shapes of the distributions. The agreement
    in absolute number is consistent with the systematic uncer­
    tainties in the absolute sensitivity and the flux of high energy
    atmospheric neutrinos. The shape of the distribution of data
    is statistically consistent with the prediction from atmo­
    spheric neutrinos. Figure 14
    ~
    bottom
    !
    shows the distribution
    of the
    z
    COG
    parameter for the 204 events. The level 7 quality
    cuts have removed the remainder of the instrumental events
    left after the Bayesian reconstruction with the improved
    cross­talk cleaning algorithm, bringing the data events in line
    with the atmospheric neutrino expectations. The efficiencies
    corresponding to the three steps of the data analysis:
    ~
    1
    !
    events reconstructed upward,
    ~
    2
    !
    events reconstructed up­
    ward with cross­talk cleaning, and
    ~
    3
    !
    with additional level 7
    quality cuts are summarized in Table IV.
    Figure 19
    ~
    top
    !
    shows the smoothness distribution for
    events that have passed the quality level 7 cuts for the five
    observables except smoothness. The vertical dashed line at
    smoothness
    ;
    0.29 shows the value of the level 7 smooth­
    FIG. 16. Numbers of events above a certain quality level, for
    downgoing muon Monte Carlo simulations, atmospheric neutrino
    Monte Carlo simulations, and experimental data.
    FIG. 17. Ratio of data to Monte Carlo simulations
    ~
    cosmic ray
    muons plus atmospheric neutrinos
    !
    . Unlike Fig. 16, the plot is
    differential—the ratio at a particular quality does not include events
    at higher or lower qualities.
    OBSERVATION OF HIGH ENERGY ATMOSPHERIC... PHYSICAL REVIEW D
    66
    , 012005
    ~
    2002
    !
    012005­15

    ness cut. This cut removes the tail of fake events leaving a
    good agreement between remaining data and the Monte
    Carlo expectation. Figure 19
    ~
    bottom
    !
    shows the same plot
    for the direct length variable. Again, a clear tail of fake
    events is removed by requiring a direct length of greater than
    70 m.
    VII. SYSTEMATIC UNCERTAINTIES
    As a novel instrument, AMANDA poses a unique chal­
    lenge of calibration. There are no known natural sources of
    high energy neutrinos, apart from atmospheric neutrinos,
    whose observation could be used to measure the detector’s
    response. Understanding the behavior of the detector is thus
    a difficult task, dependent partly on laboratory measurements
    of the individual components, partly on observations of arti­
    ficial light sources embedded in the ice, and partly on obser­
    vations of downgoing muons. Even with these measure­
    ments, uncertainties in various properties that systematically
    affect the response of the detector persist, which prevent us
    at this time from making a precise measurement of the atmo­
    spheric neutrino flux. The primary sources of systematic un­
    certainties, and their approximate effects on the number of
    upgoing atmospheric neutrinos in the final data sample, as
    TABLE IV. Event numbers for experimental data and Monte Carlo simulations for four major stages in
    the analysis. The errors quoted are statistical only.
    Monte Carlo Monte Carlo Data
    downgoing
    m
    atmospheric
    n
    Events triggered 8.8
    3
    10
    8
    8978 1.05
    3
    10
    9
    Efficiency: Reconstructed upgoing 0.55
    3
    10
    2
    5
    0.55
    3
    10
    2
    4
    Efficiency: Reconstructed upgoing (2.1
    6
    0.08)
    3
    10
    2
    6
    (6.2
    6
    0.06)
    3
    10
    2
    2
    4.7
    3
    10
    2
    6
    ~
    with cross­talk cleaning
    !
    Efficiency: Final cuts (
    q
    >
    7) (1.9
    6
    0.6)
    3
    10
    2
    8
    (3.1
    6
    0.03)
    3
    10
    2
    2
    1.9
    3
    10
    2
    7
    No. of events: Quality
    >
    717
    6
    5 279
    6
    3 204
    FIG. 18. The zenith angle distribution of upward reconstructed
    events. The size of the hatched boxes indicates the statistical preci­
    sion of the atmospheric neutrino simulation. The Monte Carlo pre­
    diction is normalized to the data.
    FIG. 19. Smoothness and direct length variables where quality
    level 7 cuts have been applied in all but the displayed variable (
    N
    2
    1 cuts; see also Sec. V D 3, Fig. 11
    !
    . The vertical dashed lines
    with the arrow indicate the region of acceptance in the displayed
    variable. In each case, a clear tail of fake events is removed by
    application of the cut, leaving good agreement in shape between the
    remaining events and the Monte Carlo expectation.
    J. AHRENS
    et al.
    PHYSICAL REVIEW D
    66
    , 012005
    ~
    2002
    !
    012005­16

    determined by variation of the simulations, are listed below.
    As discusssed in Secs. II and III, AMANDA is embedded
    in a natural medium, which is the result of millennia of cli­
    matological history, that has left its mark in the form of
    layers of particulate matter affecting the optical properties of
    the ice. Furthermore, the deployment of optical modules re­
    quires the melting and refreezing of columns of the ice. This
    cycle results in the formation of bubbles in the vicinity of the
    modules, which increase scattering and affect the sensitivity
    of the optical modules in ways that are not yet fully under­
    stood. The effects of this local hole ice are difficult to sepa­
    rate from the intrinsic sensitivity of the OMs. The uncertain­
    ties in the neutrino rate are approximately 15% from the bulk
    ice layer modeling in the Monte Carlo simulation, and as
    much as 50% from the combined effects of the properties of
    the refrozen hole ice close to the OMs, and the intrinsic OM
    sensitivity, and angular response.
    Figure 20 shows two variables that are sensitive to the
    absolute OM sensitivity: the number of OMs hit and the
    velocity of the line fit. The systematic effects of varying OM
    sensitivity on the hit multiplicity for analysis I are shown on
    the top. The peak of the multiplicity distribution for the stan­
    dard Monte Carlo events
    ~
    nominal efficiency 100%—dashed
    line
    !
    lies at a higher value than for the data. Reducing the
    simulated OM sensitivity by 50% results in a peak at lower
    values than the data. The other variable strongly affected by
    the OM sensitivity—the velocity of the line fit, introduced in
    Sec. IV A
    !
    —is the apparent velocity of the observed light
    front traveling through the ice; see Fig. 20
    ~
    bottom
    !
    .
    As a next step, we investigated the effect of the
    ANGSENS
    OM model
    ~
    first introduced in Sec. III
    !
    on the atmospheric
    neutrino Monte Carlo simulation. The results of this simula­
    tion gave a more consistent description of the experiment for
    several variables—e.g., the hit multiplicity
    ~
    the dotted line in
    Fig. 20
    !
    —and they produced the absolute neutrino event pre­
    diction closer to what was found in Analysis I
    ~
    236.9 events
    predicted, 223 observed
    !
    . Similar effects are seen when this
    Monte Carlo simulation is used with analysis II, however the
    number of predicted events is 25% smaller than observed.
    Thus the
    ANGSENS model, while encouraging, does not com­
    pletely predict the properties of observed events in both
    analyses.
    Another uncertainty lies in the Monte Carlo routines used
    to propagate muons through the ice and rock surrounding the
    detector. A comparison of codes based on
    @
    9
    #
    and
    @
    11
    #
    indi­
    cates that different propagators may change the event rates
    by some 25%.
    Other factors include the simulation of the data acquisi­
    tion electronics and possible errors in the time calibrations of
    individual modules. These effects have been studied by sys­
    tematically varying relevant parameters in the Monte Carlo
    simulations. For realistic levels of variation, these effects are
    well below the 10% level.
    Figure 21 demonstrates how the zenith angle distribution
    depends on different atmospheric neutrino event generators
    ~
    our standard generator
    NUSIM
    @
    17
    #
    and another generator
    NU2MU
    @
    29
    #!
    , and also on the chosen angular sensitivity of
    the optical module. Neutrino flavor oscillations lead to a fur­
    FIG. 20. Distributions of two variables that are affected by the
    OM sensitivity, comparing different signal Monte Carlos events to
    the observed data. Top: the number of OMs hit (
    N
    ch
    ); bottom: the
    event velocity for a simplified fit
    ~
    line fit,
    v
    linefit
    ).
    FIG. 21. Distribution of simulated zenith angles for different
    neutrino generators and also for a modified angular sensitivity of
    the OM.
    OBSERVATION OF HIGH ENERGY ATMOSPHERIC... PHYSICAL REVIEW D
    66
    , 012005
    ~
    2002
    !
    012005­17

    ther reduction of the
    ANGSENS
    prediction by 5.4%
    ~
    in par­
    ticular, close to the vertical direction
    !
    , assuming sin
    2
    2
    u
    5
    1
    and
    D
    m
    2
    5
    2.5
    3
    10
    2
    3
    eV
    2
    @
    30,38
    #
    . The prediction is re­
    duced by 11% if the largest allowed
    D
    m
    2
    is used.
    The combined effect of all these systematic uncertainties
    is sufficiently large that simulations of a given atmospheric
    neutrino flux can produce predictions for the event rate vary­
    ing by a factor of two. By contrast, the estimated theoretical
    uncertainty in the atmospheric neutrino flux, at the energies
    probed by these analyses, is 30%
    @
    31
    #
    . The effect of neutrino
    oscillations with the Super­K preferred parameters would be
    less than 10% at these energies.
    VIII. SYNTHESIS AND GENERAL OVERVIEW
    Both analyses I and II are able to separate more than 200
    neutrino event candidates from the 130.1 days of AMANDA­
    B10 detector lifetime in 1997. Based on atmospheric neu­
    trino simulations we find that about 4% of the total number
    of events triggered by upward moving neutrinos passed the
    final selection. A total deficit in the event rate of about 35%
    with respect to the standard neutrino Monte Carlo prediction
    is found for both analyses. An event overlap of 102 experi­
    mental events is observed, consistent with a predicted over­
    lap of 119
    6
    13 from the atmospheric neutrino Monte Carlo
    prediction. Thus, the combined sample of data provides
    about 300 neutrino candidates. Both analyses estimate their
    residual background to be about 10% of the number of neu­
    trino event candidates.
    Figure 22 shows the energy distribution of the simulated
    neutrinos and the corresponding muon events. Ninety per­
    cent of all Monte Carlo signal events have muon
    ~
    neutrino
    !
    energies between 48
    ~
    66
    !
    GeV and 1.8
    ~
    3.4
    !
    TeV. The domi­
    nant part of the signal events in this analysis comes from
    neutrino energies below 1 TeV. Figure 23 shows the effective
    area as a function of the zenith angle for two ranges of the
    muon energy at the point of closest
    ~
    POC
    !
    approach to the
    detector. The effective area for muons with energies at POC
    between 100 and 1000 GeV is 3.9
    3
    10
    4
    m
    2
    at trigger level
    and 2800 m
    2
    after application of the neutrino selection cuts.
    It should be noted that much higher effective areas are pos­
    sible when searching for neutrinos from astrophysical point
    sources
    @
    32
    #
    or from gamma ray bursts
    @
    33
    #
    .
    Figure 24 shows the point spread function of the recon­
    structed muon trajectory with respect to the true muon direc­
    tion. Based on Monte Carlo simulations, we find a median
    angular resolution of muons from atmospheric neutrinos of
    3.2° for the final sample. A more detailed study of the angu­
    lar resolution can be found in
    @
    25,34,35
    #
    . Figure 25 shows
    the skyplot
    ~
    equatorial coordinates
    !
    of all the candidate neu­
    trino events found across both analyses. The distribution of
    the events on the skyplot is consistent with a random distri­
    bution.
    IX. CONCLUSIONS
    The AMANDA­B10 data from 130.1 days of livetime
    during the austral winter of 1997 have been analyzed in an
    FIG. 22. Energy distributions for simulated atmospheric neu­
    trino events which pass the final neutrino cuts. The effect of neu­
    trino oscillations has not been taken into account. The figure shows
    the neutrino and muon energies at the interaction vertex and the
    energy of the muons at the point of closest approach to the detector
    center.
    FIG. 23. Effective area for muons versus zenith angle. The en­
    ergy of the muons is given at the point of closest approach to the
    detector.
    FIG. 24. Monte Carlo simulation of the angular resolution for
    muons that pass the final selection criteria. The median error is 3.2°.
    J. AHRENS
    et al.
    PHYSICAL REVIEW D
    66
    , 012005
    ~
    2002
    !
    012005­18

    effort to detect high energy atmospheric neutrino events, and
    to compare their properties to expectations. Two working
    groups in the collaboration, using differing reconstruction,
    cut optimization and instrumental event rejection techniques,
    produced sets of 223 and 204 neutrino candidates, respec­
    tively. Several methods of background estimation put the re­
    sidual event contamination from downgoing atmospheric
    muons and instrumental artifacts at about 10%. Taking into
    account systematic uncertainties, the observed event num­
    bers are consistent with systematically varied atmospheric
    neutrino Monte Carlo predictions, which are from 150–400
    events. The range of these predictions is dominated by un­
    certainties in the neutrino flux, in the understanding of pho­
    ton propagation through the bulk ice and the refrozen hole
    ice, and in muon propagation and energy loss. The Monte
    Carlo prediction suggests that 90% of the selected events are
    produced by neutrinos in the energy range of
    ;
    66 GeV to
    3.4 TeV. The observation of atmospheric neutrinos in line
    with expectations establishes AMANDA­B10 as a working
    neutrino telescope. We finally note that many of the proce­
    dures for signal separation simplify considerably in larger
    detectors. In particular, first results from AMANDA­II
    @
    36,37
    #
    demonstrate that the neutrino signal is separated with
    much higher efficiency and with fewer cuts than for
    AMANDA­B10.
    ACKNOWLEDGMENTS
    This research was supported by the following agencies:
    U.S. National Science Foundation, Office of Polar Programs;
    U.S. National Science Foundation, Physics Division; Univer­
    sity of Wisconsin Alumni Research Foundation; U.S. Depart­
    ment of Energy; Swedish Natural Science Research Council;
    Swedish Research Council; Swedish Polar Research Secre­
    tariat; Knut and Alice Wallenberg Foundation, Sweden; Ger­
    man Ministry for Education and Research; U.S. National En­
    ergy Research Scientific Computing Center
    ~
    supported by
    the Office of Energy Research of the U.S. Department of
    Energy
    !
    ; UC­Irvine AENEAS Supercomputer Facility; Deut­
    sche Forschungsgemeinschaft
    ~
    DFG
    !
    . D. F. Cowen acknowl­
    edges the support of the NSF CAREER program and C.
    Pe
    ´
    rez de los Heros acknowledges support from the EU 4th
    framework of Training and Mobility of Researchers.
    @
    1
    #
    E. Andre
    ´
    s
    et al.
    , Astropart. Phys.
    13
    ,1
    ~
    2000
    !
    .
    @
    2
    #
    E. Andre
    ´
    s
    et al.
    , Nature
    ~
    London
    !
    410
    , 441
    ~
    2001
    !
    .
    @
    3
    #
    AMANDA Collaboration, K. Woschnagg, in
    Proceedings of
    the
    26
    th
    International Cosmic Ray Conference
    ~
    IUPAP, Salt
    Lake City, UT, 1999
    !
    , Vol. 2, pp. 200–203.
    @
    4
    #
    P. B. Price, K. Woschnagg, and D. Chrikin, Geophys. Res.
    Lett.
    27
    , 2129
    ~
    2000
    !
    .
    @
    5
    #
    J. Ahrens
    et al.
    , Astropart. Phys.
    16
    , 345
    ~
    2002
    !
    .
    @
    6
    #
    S. N. Boziev
    et al.
    , INR Report P­0630
    ~
    unpublished
    !
    .
    @
    7
    #
    D. Heck
    et al.
    , Report FZKA 6019, Forschungszentrum
    Karlsruhe, Karlsruhe, Germany
    ~
    unpublished
    !
    .
    @
    8
    #
    D. Heck, in
    Simulation and Analysis Methods for Large Neu­
    trino Telescopes
    ~
    DESY, Zeuthen, Germany, 1998
    !
    , pp. 228–
    234,
    DESY­PROC­1999­01.
    @
    9
    #
    W. Lohmann, R. Kopp, and R. Voss, Report CERN­85­03,
    Geneva, Switzerland
    ~
    unpublished
    !
    .
    @
    10
    #
    R. Kopp, W. Lohmann, and R. Voss, MUDEDX: Parametriza­
    tion and Monte Carlo generation of energy loss of high energy
    muons
    ~
    1–10000 GeV
    !
    ; for astrophysicists up to 100000 GeV,
    Version 2.02, 1995
    ~
    private communication
    !
    .
    @
    11
    #
    P. Lipari and T. Stanev, Phys. Rev. D
    44
    , 3543
    ~
    1991
    !
    .
    @
    12
    #
    W. Rhode and D. Chirkin, in
    Proceedings of the
    27
    th
    Interna­
    tional Cosmic Ray Conference
    ~
    Copernicus Gesellschaft, Ham­
    burg, Germany, 2001
    !
    , Vol. 3, pp. 1017–1020.
    @
    13
    #
    B. Wiebel­Sooth and P. L. Biermann,
    Numerical Data and
    Functional Relationships in Science and Technology
    , Landolt­
    Bo
    ¨
    rnstein, New Series, Vol. VI 3C
    ~
    Springer­Verlag, Berlin,
    Heidelberg, 1998
    !
    , Chap. 7.6, pp. 37–90.
    @
    14
    #
    A. Karle, in
    Simulation and Analysis Methods for Large Neu­
    trino Telescopes
    ~
    Ref.
    @
    8
    #!
    , pp. 174–185,
    DESY­PROC­1999­01.
    @
    15
    #
    S. Hundertmark, Ph.D. thesis, Humboldt Universita
    ¨
    t zu Berlin,
    Berlin, Germany, 1999, available at URL
    http://amanda.berkeley.edu/manuscripts/
    @
    16
    #
    S. Hundertmark, in
    Simulation and Analysis Methods for Large
    Neutrino Telescopes
    ~
    Ref.
    @
    8
    #!
    , pp. 276–286,
    DESY­PROC­1999­
    01
    .
    @
    17
    #
    G. C. Hill, Astropart. Phys.
    6
    , 215
    ~
    1997
    !
    .
    @
    18
    #
    G. C. Hill, Ph.D. thesis, University of Adelaide, Adelaide, Aus­
    tralia, 1996.
    @
    19
    #
    P. Lipari, Astropart. Phys.
    1
    , 195
    ~
    1993
    !
    .
    @
    20
    #
    J. Ahrens
    et al.
    ~
    unpublished
    !
    .
    @
    21
    #
    V. Stenger, DUMAND Internal Report HDC­1­90.
    @
    22
    #
    C. Wiebusch, in
    Simulation and Analysis Methods for Large
    Neutrino Telescopes
    ~
    Ref.
    @
    8
    #!
    , pp. 302–316,
    DESY­PROC­1999­
    01
    .
    @
    23
    #
    D. Pandel, Diploma thesis, Humboldt­University, Berlin, 1996.
    @
    24
    #
    G. C. Hill, in
    Proceedings of the
    27
    th
    International Cosmic
    Ray Conference
    ~
    Ref.
    @
    12
    #!
    , Vol. 3, pp. 1279–1282.
    @
    25
    #
    A. Biron, Ph.D. thesis, Humboldt Universtita
    ¨
    t zu Berlin, Ber­
    lin, Germany, 2002, available at URL
    FIG. 25. Neutrino skyplot of upgoing events as seen with
    AMANDA­B10 in 1997 in equatorial coordinates. In this figure,
    neutrino events from both analyses are combined. The background
    of non­neutrino events is estimated to be less than 10%.
    OBSERVATION OF HIGH ENERGY ATMOSPHERIC... PHYSICAL REVIEW D
    66
    , 012005
    ~
    2002
    !
    012005­19

    http://amanda.berkeley.edu/manuscripts/
    @
    26
    #
    T. R. DeYoung, Ph.D. thesis, University of Wisconsin, Madi­
    son, 2001, available at URL
    http://amanda.berkeley.edu/manuscripts/
    @
    27
    #
    M. Gaug, Diploma thesis, Humboldt Universita
    ¨
    t zu Berlin,
    Berlin, Germany, 2001,
    DESY­THESIS­2001­022, available at URL
    http://amanda.berkeley.edu/manuscripts/.
    @
    28
    #
    AMANDA Collaboration, A. Bouchta, in
    Proceedings of the
    26th International Cosmic Ray Conference
    ,
    ~
    Ref.
    @
    3
    #!
    ,HE
    3.2.11.
    @
    29
    #
    E. Dalberg, Ph.D. thesis, Stockholm University, Stockholm,
    Sweden, 1999, available at URL
    http://amanda.berkeley.edu/manuscripts/
    @
    30
    #
    Super­Kamiokande Collaboration, J. Kameda, in
    Proceedings
    of the
    27
    th
    International Cosmic Ray Conference
    ~
    Ref.
    @
    12
    #!
    ,
    Vol. 3, pp. 1057–1060.
    @
    31
    #
    T. K. Gaisser, ‘‘Uncertainty in flux of Atmospheric Neutrinos:
    Implications for Upward Muons in Amanda B10,’’ Internal
    Report 20001201, 2000.
    @
    32
    #
    J. Ahrens
    et al.
    , ‘‘Search for point sources of high energy neu­
    trinos with the AMANDA detector.’’
    @
    33
    #
    AMANDA Collaboration, R. Hardtke and G. Barouch, in
    Pro­
    ceedings of the
    27
    th
    International Cosmic Ray Conference
    ~
    Ref.
    @
    12
    #!
    , Vol. 3, pp. 1121–1124.
    @
    34
    #
    AMANDA Collaboration, ‘‘Calibration and Survey of
    AMANDA with the SPASE detectors’’
    ~
    to be published
    !
    .
    @
    35
    #
    AMANDA Collaboration, G. Spiczak, in
    Proceedings of the
    27
    th
    International Cosmic Ray Conference
    ~
    Ref.
    @
    12
    #!
    , Vol. 3,
    pp. 977–980.
    @
    36
    #
    AMANDA Collaboration, R. Wischnewski, in
    Proceedings of
    the
    27
    th
    International Cosmic Ray Conference
    ~
    Ref.
    @
    12
    #!
    , Vol.
    3, pp. 1105–1108.
    @
    37
    #
    AMANDA Collaboration, S. W. Barwick, in
    Proceedings of
    the
    27
    th
    International Cosmic Ray Conference
    ~
    Ref.
    @
    12
    #!
    , Vol.
    3, pp. 1105–1108.
    @
    38
    #
    S. Fukuda
    et al.
    , Phys. Rev. Lett.
    85
    , 3999
    ~
    2000
    !
    .
    J. AHRENS
    et al.
    PHYSICAL REVIEW D
    66
    , 012005
    ~
    2002
    !
    012005­20

    Back to top