Flux limits on ultra high energy neutrinos
    with AMANDA-B10
    M. Ackermann
    a
    , J. Ahrens
    b
    , H. Albrecht
    a
    , D. Atlee
    c
    , X. Bai
    d
    , R. Bay
    e
    ,
    M. Bartelt
    f
    , S.W. Barwick
    g
    , T. Becka
    b
    , K.H. Becker
    f
    , J.K. Becker
    f
    ,
    E. Bernardini
    a
    , D. Bertrand
    h
    , D.J. Boersma
    a
    ,S.Bo
    ¨
    ser
    a
    , O. Botner
    i
    ,
    A.Bouchta
    i
    ,O.Bouhali
    h
    ,J.Braun
    j
    ,C.Burgess
    k
    ,T.Burgess
    k
    ,T.Castermans
    l
    ,
    D. Chirkin
    e
    , T. Coarasa
    c
    , B. Collin
    c
    , J. Conrad
    i
    , J. Cooley
    j
    , D.F. Cowen
    c
    ,
    A.Davour
    i
    ,C.DeClercq
    m
    ,T.DeYoung
    n
    ,P.Desiati
    j
    ,P.Ekstro
    ¨
    m
    k
    ,T.Feser
    b
    ,
    T.K. Gaisser
    d
    , R. Ganugapati
    j
    , H. Geenen
    f
    , L. Gerhardt
    g
    , A. Goldschmidt
    o
    ,
    A. Groß
    f
    , A. Hallgren
    i
    , F. Halzen
    j
    , K. Hanson
    j
    , D. Hardtke
    e
    , R. Hardtke
    j
    ,
    T. Harenberg
    f
    , T. Hauschildt
    a
    , K. Helbing
    o
    , M. Hellwig
    b
    , P. Herquet
    l
    ,
    G.C. Hill
    j
    , J. Hodges
    j
    , D. Hubert
    m
    , B. Hughey
    j
    , P.O. Hulth
    k
    , K. Hultqvist
    k
    ,
    S. Hundertmark
    k,
    *
    , J. Jacobsen
    o
    , K.H. Kampert
    f
    , A. Karle
    j
    , J. Kelley
    j
    ,
    M. Kestel
    c
    ,L.Ko
    ¨
    pke
    b
    , M. Kowalski
    a
    , M. Krasberg
    j
    , K. Kuehn
    g
    , H. Leich
    a
    ,
    M. Leuthold
    a
    , J. Lundberg
    i
    , J. Madsen
    p
    , K. Mandli
    j
    , P. Marciniewski
    i
    ,
    H.S. Matis
    o
    , C.P. McParland
    o
    , T. Messarius
    f
    , Y. Minaeva
    k
    , P. Mioc
    ˇ
    inovic
    ´
    e
    ,
    R. Morse
    j
    , S. Movit
    c
    ,K.Mu
    ¨
    nich
    f
    , R. Nahnhauer
    a
    , J.W. Nam
    g
    ,
    T. Neunho
    ¨
    ffer
    b
    , P. Niessen
    d
    , D.R. Nygren
    o
    ,H.O
    ¨
    gelman
    j
    , Ph. Olbrechts
    m
    ,
    C. Pe
    ´
    rez de los Heros
    i
    , A.C. Pohl
    q
    , R. Porrata
    e
    , P.B. Price
    e
    , G.T. Przybylski
    o
    ,
    K. Rawlins
    j
    , E. Resconi
    a
    , W. Rhode
    f
    , M. Ribordy
    l
    , S. Richter
    j
    ,
    J. Rodrı
    ´
    guez Martino
    k
    , D. Rutledge
    c
    , H.G. Sander
    b
    , K. Schinarakis
    f
    ,
    S. Schlenstedt
    a
    , D. Schneider
    j
    , R. Schwarz
    j
    , A. Silvestri
    g
    , M. Solarz
    e
    ,
    G.M. Spiczak
    p
    , C. Spiering
    a
    , M. Stamatikos
    j
    , D. Steele
    j
    , P. Steffen
    a
    ,
    R.G. Stokstad
    o
    , K.H. Sulanke
    a
    , I. Taboada
    r
    , O. Tarasova
    a
    , L. Thollander
    k
    ,
    S. Tilav
    d
    , L.C. Voicu
    c
    , W. Wagner
    f
    , C. Walck
    k
    , M. Walter
    a
    , Y.R. Wang
    j
    ,
    C.H. Wiebusch
    f
    , R. Wischnewski
    a
    , H. Wissing
    a
    , K. Woschnagg
    e
    , G. Yodh
    g
    0927-6505/$ - see front matter
    ?
    2004 Elsevier B.V. All rights reserved.
    doi:10.1016/j.astropartphys.2004.09.008
    *
    Corresponding author.
    E-mail address:
    stephan.hundertmark@physto.se(S. Hundertmark).
    Astroparticle Physics 22 (2005) 339–353
    www.elsevier.com/locate/astropart

    a
    DESY, 15735 Zeuthen, Germany
    b
    Institute of Physics, University of Mainz, D-55099 Mainz, Germany
    c
    Department of Physics, Pennsylvania State University, University Park, PA 16802, USA
    d
    Bartol Research Institute, University of Delaware, Newark, DE 19716, USA
    e
    Department of Physics, University of California, Berkeley, CA 94720, USA
    f
    Department of Physics, Bergische Univerita
    ¨
    t Wuppertal, 42097 Wuppertal, Germany
    g
    Department of Physics and Astronomy, University of California, Irvine, CA 92697, USA
    h
    Universite
    ´
    Libre de Bruxelles, Science Faculty CP230, Boulevard du Triomphe, B-1050 Brussels, Belgium
    i
    Division of High Energy Physics, Uppsala University, S-75121 Uppsala, Sweden
    j
    Department of Physics, University of Wisconsin, Madison, WI 53706, USA
    k
    Department of Physics, Stockholm University, SE-10691 Stockholm, Sweden
    l
    University of Mons-Hainaut, 7000 Mons, Belgium
    m
    Vrije Universiteit Brussel, Dienst ELEM, B-1050 Brussels, Belgium
    n
    Department of Physics, University of Maryland, College Park, MD 20742, USA
    o
    Lawrence Berkeley National Laboratory, Berkeley, CA 94720, USA
    p
    Physics Department, University of Wisconsin, River Falls, WI 54022, USA
    q
    Department of Chemistry and Biomedical Sciences, University of Kalmar, S-39182 Kalmar, Sweden
    r
    Department of Physics, Universidad Simo
    ´
    n Bolı
    ´
    var, Caracas, 1080, Venezuela
    Received 17 August 2004; received in revised form 22 September 2004; accepted 30 September 2004
    Available online 13 November 2004
    Abstract
    Data taken during 1997 with the AMANDA-B10 detector are searched for a diffuse flux of neutrinos of all flavors
    with energies above 10
    16
    eV. At these energies the Earth is opaque to neutrinos, and thus neutrino induced events are
    concentrated at the horizon. The background are large muon bundles from down-going atmospheric air shower events.
    No excess events above the background expectation are observed and a neutrino flux following
    E
    ?
    2
    , with an equal mix
    of all flavors, is limited to
    E
    2
    U
    (10
    15
    eV <
    E
    <3
    ·
    10
    18
    eV)
    6
    0.99
    ·
    10
    ?
    6
    GeVcm
    ?
    2
    s
    ?
    1
    sr
    ?
    1
    at 90% confidence level. This
    is the most restrictive experimental bound placed by any neutrino detector at these energies. Bounds to specific extra-
    terrestrial neutrino flux predictions are also presented.
    ?
    2004 Elsevier B.V. All rights reserved.
    PACS:
    95.55.Vj; 95.85.Ry; 96.40.Tv
    Keywords:
    Neutrino telescopes; Neutrino astronomy; UHE neutrinos; AMANDA
    1. Introduction
    Neutrino telescopes like AMANDA [1] are uni-
    que astronomy tools, originally designed to look
    downwards for Cherenkov light emitted by up-
    ward traveling muons from charged-current
    muon–neutrino interactions. The Earth is used to
    filter the muon flux from cosmic ray air showers.
    As muons cannot pass through the Earth, any up-
    ward traveling muon in the vicinity of the detector
    is neutrino induced. Due to the sparse instrumen-
    tation of the detection medium and the optical
    properties of the medium itself, misidentification
    of the direction of the muon by the reconstruction
    algorithm is possible. To reduce the amount of
    down-going muons, and therefore the probability
    for misidentification, these detectors are located
    under as much overburden as feasible.
    As the energy of the neutrino increases, the neu-
    trino–nucleon cross section and the muon range
    grow. Both effects result in a large detection prob-
    ability for high energy neutrinos. Above 40TeV,
    the Earth diameter exceeds the neutrino charged-
    current interaction length and high energy neutri-
    340
    M. Ackermann et al. / Astroparticle Physics 22 (2005) 339–353

    nos are significantly absorbed
    [2].At UHE
    1
    the
    Earth is essentially opaque to muon- and elec-
    tron–neutrinos[3]and only downward to horizon-
    tally traveling neutrinos can be detected. The
    overburden above the AMANDA detector is lim-
    ited, which limits the interaction probability for
    neutrinos, concentrating the muons from UHE
    neutrinos at the horizon. Klein and Mann[3]esti-
    mate that a few muons per year above 10
    15
    eV
    from UHE neutrino sources might be detected
    with an 0.1km
    2
    detector, an effective area that is
    achieved by this analysis (as shown in[4]). The en-
    ergy range above 10
    16
    eV is interesting, as experi-
    mentally there are currently no strong limits on
    neutrino fluxes from models describing active
    galactic nuclei (AGN) at these energies. At higher
    energies, above 10
    18
    eV, one expects the guaran-
    teed flux of neutrinos from interactions of high
    energy cosmic rays with the cosmic microwave
    background, or the more speculative flux from de-
    cays of topological defects (for an overview see
    [5–7]). Cosmological sources are expected to pro-
    duce neutrinos in pp or p
    c
    collisions, resulting in
    a flavor ratio of
    m
    e
    :
    m
    l
    :
    m
    s
    = 1:2:0 at the source. As
    these neutrinos propagate over cosmological dis-
    tances, oscillations between the different flavors re-
    sult in a ratio at Earth of
    m
    e
    :
    m
    l
    :
    m
    s
    = 1:1:1 [8].
    Therefore all three flavors are initially of the same
    importance for detection, but for muon–neutrinos
    the detection probability is higher, due to the long
    path length of the resulting muons.
    The search for a neutrino signal in the upper
    hemisphere has to be performed in the presence
    of the large flux of down-going atmospheric muon
    bundles. This analysis exploits the fact that these
    events have different characteristics than neutrino
    induced events. The primary cosmic ray spectrum
    (protons and heavier elements) falls rapidly with
    energy, following approximately
    E
    ?
    2.7
    below the
    knee and steepening thereafter to
    E
    ?
    3.1
    . An atmo-
    spheric air shower produced by an UHE primary
    particle results in a muon bundle consisting of
    hundreds to thousands of muons. At 10
    16
    10
    20
    eV approximately 90% of the primary energy
    is consumed by the electromagnetic cascade of
    the air shower and only a fraction of the remaining
    energy is carried by the resulting muons[9]. When
    comparing a single muon and a muon bundle of
    the same energy, the bundle spreads the generated
    Cherenkov light over a larger volume. In the detec-
    tor, both classes of events can result in a large
    number of photomultiplier tubes (PMT) being hit
    by photons, but neutrino induced events have
    more photon hits overall, i.e. more multiple hits
    are found in single PMTs. Often, however, single
    photons cannot be resolved. This can be compen-
    sated for by exploiting the afterpulse behavior of
    the PMTs. Each photoelectron has a small proba-
    bility to generate an afterpulse delayed by several
    microseconds. These afterpulses are used as a
    ‘‘low gain’’ outlet to identify high energy events
    with a large number of photons incident on the
    PMT. Using this and other event properties, it is
    possible to reduce the number of events caused
    by down-going atmospheric muons to a level
    where a search for a signal from UHE neutrinos
    becomes feasible. In[4]and[10]part of the data
    taken in 1997 with the AMANDA-B10 array was
    used to establish this technique and a preliminary
    limit was set. In this analysis an improved neutrino
    simulation is used, selection criteria are refined,
    systematic uncertainties are assessed and the anal-
    ysis is applied to the full 1997 data set.
    2. The AMANDA-B10 detector
    The data used in this analysis were taken with
    the AMANDA-B10 detector [1] in 1997. This
    detector consists of 302 Optical Modules (OMs)
    on 10 vertical strings. The instrumented parts of
    the strings are located between 1500m and
    2000m below the South Pole ice surface. The
    strings are arranged in two concentric circles.
    The outer circle with a diameter of 120m is formed
    by six strings equipped with a total of 216 OMs
    connected to the surface via twisted pair cables.
    The vertical separation of the OMs located on
    these strings is 10m. Three strings lie on a 60m
    diameter inner circle, the remaining string being
    located at the center. These strings use coaxial
    cable to connect 86 OMs, with 20m spacing, to
    the surface. The coaxial or twisted pair cable
    1
    U
    ltra
    H
    igh
    E
    nergy is used here for energies above 10
    16
    eV.
    M. Ackermann et al. / Astroparticle Physics 22 (2005) 339–353
    341

    supplies the PMT inside the OM-pressure-sphere
    with high voltage power and transmits the signal
    to the amplifier and data acquisition electronics
    (DAQ) located at the surface. The electrical signal
    from the PMT is dispersed in the cable, which
    strongly limits the capability to resolve single pho-
    tons arriving at the PMT separated by less than a
    few hundred nanoseconds. The PMT and/or the
    amplifier saturate when more than some tens of
    photoelectrons are generated in the PMT. Each
    photoelectron produces an afterpulse with an
    approximately 3% probability, well separated by
    approximately 6
    l
    s from the initial pulse. These
    late afterpulses appear when residual gas mole-
    cules, ionized by electrons from the electron multi-
    plier, travel back to the cathode and release
    additional electrons
    [11,12]. The AMANDA-B10
    detector and its response to a simulated horizontal
    2
    ·
    10
    19
    eV muon is shown in Fig. 1.
    3. Signal and background simulations
    3.1. UHE neutrino simulation
    The neutrino generator ANIS
    [13]
    is used to
    generate neutrinos and anti-neutrinos of all flavors
    following an
    E
    ?
    1
    spectrum between 10
    13
    eV and
    10
    20
    eV. All relevant standard model processes,
    like charged- and neutral-current
    m
    N interactions,
    resonant
    m
    e
    e
    ?
    scattering and tau–neutrino regener-
    ation are simulated. Cross-sections are evaluated
    up to 10
    21
    eV following the framework of pQCD
    and with structure functions according to CTEQ5.
    The neutrinos are propagated through the Earth
    with a density profile taken from the Preliminary
    Reference Earth Model
    [14]. In case of a
    charged-current interaction in the vicinity of the
    detector, the resulting lepton and hadronic energy
    deposition is simulated. Throughout this paper a
    neutrino flavor ratio at Earth of
    m
    e
    :
    m
    l
    :
    m
    s
    = 1:1:1
    and a ratio of
    m
    =
    m
    ¼
    1
    2
    is assumed. The events,
    generated by ANIS, are re-weighted to a diffuse
    AGN-like flux
    E
    2
    U
    =10
    ?
    6
    GeVcm
    ?
    2
    s
    ?
    1
    sr
    ?
    1
    or
    other UHE neutrino flux predictions.Fig. 2shows
    the zenith angle distribution for muon–neutrinos
    which underwent a charged-current interaction
    within the muon path length from the detector.
    With increasing neutrino energy, the absorption
    of the neutrino flux becomes stronger and the
    resulting muon flux concentrates near the horizon.
    The event selection is developed using a simpler
    single muon simulation, re-weighted to an
    E
    ?
    2
    neutrino spectrum. In this simulation the Earth is
    assumed to have the density of ice and the hadro-
    nic energy deposition from the charged-current
    interaction is neglected. Nevertheless, this simpli-
    fied simulation leads to a good agreement in the
    Fig. 1. A simulated horizontal 2
    ·
    10
    19
    eV muon passes the
    AMANDA-B10 detector. The different colors of the OMs
    correspond to different hit times from yellow (early) to blue
    (late). Only the hits depicted in yellow, orange and red are
    generated by the muon, while the green and blue colored hits
    are caused by afterpulses. In this event, all working OMs are hit
    and nearly all of them report at least one hit from afterpulsing.
    (For interpretation of the references in color in this figure
    legend, the reader is referred to the web version of this article.)
    2
    The production of high energy
    m
    e
    is suppressed in
    p
    c
    !
    n
    p
    +
    interactions, altering the ratio
    m
    =
    m
    . For a discussion
    of this effect see [35].
    342
    M. Ackermann et al. / Astroparticle Physics 22 (2005) 339–353

    final neutrino energy and angular distributions
    when compared to the results using the ANIS
    generator.
    3.2. Cosmic ray air shower generation
    Cosmic ray air showers are generated using the
    CORSIKA program (version 5.7001, 1988)
    [15]
    with the QGSJET interaction model. To develop
    the selection criteria, two sets of simulations are
    performed. The first uses the primary composition
    and the spectral slopes for the individual elements
    from
    [16], with energies of the primary particles
    ranging from 8
    ·
    10
    11
    eV to 10
    20
    eV. This simula-
    tion corresponds to the standard AMANDA
    background simulation. The second set is tailored
    to more efficiently simulate events generated by
    high energy primaries. Protons and iron primaries
    are sampled from an
    E
    ?
    2
    spectrum from
    8
    ·
    10
    13
    eV to 10
    20
    eV. Events resulting from
    315
    ·
    10
    6
    proton- and 145
    ·
    10
    6
    iron-primaries
    are re-weighted to the model taken from
    [17],
    which is shown in
    Fig. 3. This model describes
    the averaged measured primary cosmic ray flux
    with proton and iron primaries. As the developed
    selection criteria reject events generated by low en-
    ergy primaries, the higher energy threshold and the
    softer spectral slope allows to reduce the required
    cpu time while generating a large number of high
    energy air shower events.
    3.3. Lepton propagation
    Muons and taus produced in charged-current
    neutrino interactions are propagated through
    the ice and rock using MMC
    [18]
    while for
    muons generated in cosmic ray air showers an
    algorithm developed by Lipari and Stanev
    [19]
    is used. The simulation takes into account the re-
    sponse of the detector to additional Cherenkov
    photons generated by energy depositions due to
    ionization and radiative processes. Table 1 shows
    the mean path length of muons in ice as evalu-
    ated with the algorithm presented in
    [19].At
    Fig. 3. The two-component model describing the averaged
    measured primary cosmic ray flux taken from[17]is used to
    simulate the cosmic ray air shower events.
    Fig. 2. Zenith angle distribution for muon–neutrinos which
    produced a muon able to reach the detector (A zenith angle of
    180
    ?
    corresponds to a upward-going particle). The probability
    distribution is shown for a muon–neutrino flux following
    E
    ?
    2
    and energies above 10
    14
    eV and between 10
    15
    –10
    16
    eV, 10
    17
    10
    18
    eV and 10
    19
    –10
    20
    eV.
    Table 1
    The mean path length for muons (
    h
    L
    l
    i
    ) with energies from
    10
    15
    eV to 10
    20
    eV
    E
    l
    (eV) 10
    15
    10
    16
    10
    17
    10
    18
    10
    19
    10
    20
    h
    L
    l
    i
    (kmwe) 17 21 26 30 33 37
    M. Ackermann et al. / Astroparticle Physics 22 (2005) 339–353
    343

    UHE energies, muons can travel up to several
    tens of kilometers, which allows to monitor large
    volumes of ice.
    4. Data selection and analysis
    During the 1997 data taking period a total of
    1.05
    ·
    10
    9
    events satisfied a majority trigger of at
    least 16 hit OMs within about 2
    l
    s. The data tak-
    ing rate was 100Hz. Out of the 302 OMs, 41
    (14%) were either not functioning or showed
    abnormal behavior and were excluded from the
    analysis. Signals with a short time over threshold
    (TOT) are mostly caused by electrical cross talk
    between OMs on the same string and are removed
    from the analysis. In order to enhance high energy
    events, events that deposited a large amount of
    photons in and around the detector are selected
    by requiring hits in more than 95 OMs, with four
    of them located on the inner strings. During the
    detector
    ?
    s operational time of 174 days, 4
    ·
    10
    6
    events fulfilling these criteria were collected. Tak-
    ing into account the detector dead time of 25%
    the data taking period corresponds to 131 days
    of live time. The goal of this analysis is to reject
    atmospheric air shower events and retain a maxi-
    mum number of UHE neutrino induced events.
    Due to the long path length of UHE muons (see
    Table 1), it is expected that events caused by
    muon–neutrinos contribute a larger fraction to
    the final event sample than electron– or tau–neu-
    trino induced events. The development of selection
    criteria is therefore performed with simulated
    muon–neutrino induced events and only for the
    optimization of the final criterion the electron–
    and tau–neutrino induced events are added to
    the analysis.
    4.1. Variables
    To separate UHE neutrino induced signal
    events from air shower induced background events
    eight variables are used. Two neural nets (NN)
    combine subsets of these variables to achieve bet-
    ter separation. The variables are defined as
    follows:
    NCHis the number of OMs that reported one
    or more hits. The time-to-digital converters con-
    nected to each OM can distinguish up to eight hits.
    The variable NHITS is the sum of all hits for all
    OMs.
    Fig. 4(left) shows the NHITS distribution
    of the experiment, air shower and signal simula-
    tions. The atmospheric air shower simulation
    describes the experimental data well, while the sig-
    nal simulation extends to larger values of NHITS.
    For bright events with many photons arriving at
    the OM, dispersion of the electrical signal in the
    cables to the surface effectively merges signals from
    primarily unscattered or only slightly scattered
    photons. F1His the fraction of hit OMs with ex-
    actly one hit, divided by NCH. A faint signal in-
    duces mostly single photoelectrons in each hit
    OM and F1His close to one, while bright events
    induce many photoelectrons per OM and therefore
    more afterpulses, both effects resulting in F1Hclo-
    ser to zero. This variable is therefore effective in
    selecting bright events.
    Fig. 4(right) shows F1H
    for experimental data and simulated signal and
    background. The air shower simulations describe
    the experimental data and the simulated signal
    events populate lower values of F1H. The mean
    amplitude per hit OM, MA, is the sum of ampli-
    tudes in hit OMs divided by NCH. Bright events
    have, on average, a higher mean amplitude. The
    first guess reconstruction (FG) and likelihood
    reconstruction (LR) return directional informa-
    tion. The quality of the likelihood reconstruction
    is reflected in the likelihood
    L
    . The underlying
    model for the likelihood reconstruction is a mini-
    1. NCHNumber of hit OMs
    2. NHITS Number of hits for all OMs
    3. F1HFraction of hit OMs with exactly
    one hit
    4. MA Mean amplitude for hit OMs
    5.
    h
    (FG) Zenith angle obtained from
    first guess
    6.
    h
    (LR) Zenith angle obtained from
    likelihood reconstruction
    7.
    L
    Likelihood for likelihood
    reconstruction
    8.
    S
    Smallest moment of tensor of inertia
    344
    M. Ackermann et al. / Astroparticle Physics 22 (2005) 339–353

    mal ionizing muon that induces a comparatively
    small amount of Cherenkov light in the detector
    medium, while the first guess assumes light travel-
    ing at a certain speed along a line
    3
    . At high ener-
    gies each OM receives hundreds to thousands of
    photons, significantly altering the arrival time
    probability as a function of distance. This leads
    to a poor resolution in the directional reconstruc-
    tion for UHE events. Using the tensor of inertia
    for the events, the different shapes of the air
    shower and UHE neutrino induced events can be
    used. From the available variables the smallest
    eigenvalue of the tensor of inertia
    S
    shows the best
    discrimination power.
    4.2. Level 1
    Selecting events with F1H< 0.65 defines the Le-
    vel 1 in this analysis. The experimental data sam-
    ple is reduced to 263,000 events, or 6.5% of the
    initial sample. These events are reconstructed with
    the first guess and the likelihood reconstruction.
    The reconstructed zenith angles (
    h
    (FG),
    h
    (LR)),
    the likelihood parameter
    L
    and the F1Hvariable
    are used in the neural net NN1 to distinguish sim-
    ulated signal from background. The simulation of
    air shower events gives reasonable agreement in
    the zenith angle distribution, for both first guess
    and likelihood reconstructions (Fig. 5). The distri-
    butions show that the reconstructions are not able
    to resolve the directional information of the UHE
    neutrino induced events, which are predominantly
    horizontal. Nevertheless, the direction of air
    shower muon events is better reconstructed and
    this is exploited by the neural net NN1.
    Fig. 6
    shows the neural net output for experimental data,
    signal Monte Carlo and the two sets of air shower
    simulations. Both samples, the standard set of
    CORSIKA generated data with a limited amount
    of events (the outlier at NN1
    ?
    0.95 corresponds
    to one event) and the re-weighted high energy
    CORSIKA sample using only proton and iron cos-
    mic ray primaries, show good agreement with the
    experimental data.
    4.3. Level 2
    Reducing the Monte-Carlo background to
    approximately 1% of the previous level, Level 2 is
    defined by selecting events with NN1 > 0.37. The
    experimental data set is reduced to 3326 events,
    while the air shower simulation predicts 2976 ± 51
    events. This is 11% lower and within the systematic
    Fig. 4. The NHITS distributions for the experiment, air shower and signal simulation to the left and to the right F1H. At F1H = 0.65
    the position of the selection criteria leading to the next level is shown.
    3
    For more details on the reconstruction methods used in
    AMANDA see [20].
    M. Ackermann et al. / Astroparticle Physics 22 (2005) 339–353
    345

    uncertainty, as discussed later. At this level, the fi-
    nal selection criterion is defined, based on a second
    neural net (NN2) using the variables F1H, NCH,
    NHITS, MA and
    S
    . Other variables and combina-
    tions have been investigated, but the ones used are
    found to be the most effective.
    Fig. 7
    shows the
    NN2 output for the experiment, the air shower
    and the combined electron–, muon– and tau–neu-
    trino simulation. A Kolmogorov test, using the
    CORSIKA generated events and the experimental
    data yields a probability close to one for both cases
    resulting from the same parent distribution.
    Fig. 5. The zenith angle distribution for the first guess (FG) (left) and the likelihood reconstruction (LR) (right).
    Fig. 6. The normalized neural net (NN1) output for the two
    sets of cosmic ray air shower simulation (weighted as dotted
    crosses, unweighted as triangles), the experiment and the UHE
    neutrino induced events.
    E
    Fig. 7. The experimental data and the prediction from the air
    shower simulation for the neural net NN2. The number of
    events expected from an 10
    ?
    6
    E
    ?
    2
    GeVcm
    ?
    2
    s
    ?
    1
    sr
    ?
    1
    neutrino
    source (all flavors equally mixed) is superimposed.
    346
    M. Ackermann et al. / Astroparticle Physics 22 (2005) 339–353

    4.4. The final selection criterion and sensitivity
    The final selection criterion is based on the out-
    put of the neural net NN2. To find the criterion
    that puts the strongest constraint on a given theo-
    retical model the procedure outlined by Hill and
    Rawlins [21] is used. This method uses only infor-
    mation from Monte-Carlo simulations and is
    therefore free from bias introduced by using exper-
    imental data. The simulated background expecta-
    tion
    n
    b
    as a function of the selection variable is
    used to calculate the average event upper limit
    l
    90
    (90% confidence limit) as defined in [22]. The
    strongest constraint on a theoretical model corre-
    sponds to the selection criterion that minimizes
    the model rejection factor
    l
    90
    =
    n
    s
    where
    n
    s
    is the
    number of events expected from this model that
    satisfy the selection criteria. The average flux
    upper limit, or sensitivity, is defined as
    U
    90
    ¼
    U
    l
    90
    n
    s
    , where
    U
    is the neutrino flux predicted by a
    specific theoretical model. The model rejection fac-
    tor states how strongly, on average, a given flux
    U
    is rejected when performing repeated (hypotheti-
    cal) runs of the experiment. The actual experiment
    will obtain a limit based on the observed number
    of events. Fig. 8 shows the model rejection factor
    for an
    E
    2
    U
    =10
    ?
    6
    GeVcm
    ?
    2
    s
    ?
    1
    sr
    ?
    1
    neutrino flux.
    As the neural net assigns signal-like events values
    close to one, it is expected that the largest rejection
    power is achieved for a selection close to this
    value. Indeed, the inset inFig. 8, using finer bin-
    ning, shows a minimum close to NN2 = 1. Here
    NN2 > 0.9 is chosen, ignoring the loss of approxi-
    mately 10% in rejection power compared to the ac-
    tual minimum. For this selection,
    n
    s
    = 7.44 ± 0.14
    and
    n
    b
    = 4.6 ± 1.2 is expected from simulations
    for signal and background, respectively. This re-
    sults in a sensitivity of 5.1 events (90% confidence
    level), or
    E
    2
    U
    90
    ð
    m
    e
    :
    m
    l
    :
    m
    s
    ¼
    1
    :
    1
    :
    1
    Þ¼
    0
    :
    69
    ?
    10
    ?
    6
    GeVcm
    ?
    2
    s
    ?
    1
    sr
    ?
    1
    . At this level, by calculating
    the number of equivalent events, the biased Monte
    Carlo background simulation corresponds to
    approximately 3 times the experimental live time.
    4.5. Effective detector area
    The effective area for neutrinos,
    A
    eff
    , represents
    the area of an ideal detector capable of detecting
    neutrinos with full efficiency. Using the effective
    area for neutrinos one can calculate neutrino in-
    duced event rates for a given neutrino flux. The
    number of events is calculated by:
    N
    event
    ¼
    T
    Z
    d
    E
    m
    d
    XU
    ð
    E
    m
    ,
    H
    m
    Þ
    A
    eff
    ð
    E
    m
    ,
    H
    m
    Þ
    :
    With
    T
    the live time of the experiment,
    E
    m
    and
    H
    m
    denote the energy and the zenith angle of the
    neutrino.
    Fig. 9(left) shows the effective area for
    neutrinos as a function of energy for the three fla-
    vors averaged over the azimuth and zenith angles.
    As the energy increases the muon–neutrino contri-
    bution dominates over those from electron– and
    tau–neutrinos, because of the increasing path
    length of the muons. At 10
    11
    GeV the detector is
    approximately 4–5 times more sensitive for
    muon–neutrinos than for electron–neutrinos, the
    sensitivity for tau–neutrinos being in-between.
    Due to the angular averaging, this figure implicitly
    averages the effects of absorption and interaction
    probability.
    Fig. 9(right) shows that for nearly
    M
    Fig. 8. The model rejection factor for an
    E
    2
    U
    =
    10
    ?
    6
    GeVcm
    ?
    2
    s
    ?
    1
    sr
    ?
    1
    neutrino flux. The
    x
    -axis denotes the
    selection with NN2 larger than the corresponding value.
    Selecting events with NN2 > 0.9 leads to a model rejection
    factor of 0.69.
    M. Ackermann et al. / Astroparticle Physics 22 (2005) 339–353
    347

    horizontal events, the detection area is much larger
    than the averaged area and exceeds 10
    4
    m
    2
    for hori-
    zontal muon–neutrinos.
    4.6. Event rates
    The overall absolute sensitivity of the simulation
    is verified by comparing the predicted event rate
    from background simulations with the experimen-
    tal rate. Table 2 shows the event rates for the differ-
    ent selection criteria. The number of experimental
    events drops from 4
    ·
    10
    6
    to 5 and the simulated
    background rate describes the experimental rate
    at all levels within the systematic uncertainties.
    5. Systematic uncertainties
    Systematic uncertainties in simulation input
    parameters lead to an uncertainty in the final re-
    sult. Changes in these parameters influence the
    number of expected signal or background events
    for a fixed set of selection criteria. In this analysis
    the main uncertainties are related to the properties
    of the ice as medium for photon propagation, the
    muon propagation in the ice, the absolute detector
    sensitivity, the neutrino cross section at high ener-
    gies and the absolute flux and composition of the
    primary cosmic rays. The effects of variations in
    the simulation input parameters on the number
    of events remaining after the final selection are
    studied. The derived systematic uncertainties are
    included in the limit calculation using the method
    of Conrad et al. [23], which is an extension of the
    approach of Cousins and Highland [24].
    5.1. Absolute detector sensitivity
    There are three main uncertainties affecting the
    absolute sensitivity of the detector: the absolute
    Table 2
    Event rates (Hz) for the different analysis levels for the experiment and the background simulation performed with CORSIKA
    Selection criteria
    NCH> 95 F1H< 0.65 NN1 > 0.37 NN2 > 0.9
    Experiment 0.35 0.023 2.9
    ·
    10
    ?
    4
    4.4
    ·
    10
    ?
    7
    CORSIKA 0.31 0.026 2.6
    ·
    10
    ?
    4
    4.1
    ·
    10
    ?
    7
    Fig. 9. The effective area for neutrinos of all three flavors averaged over all angles (left) and for nearly horizontal events (right).
    348
    M. Ackermann et al. / Astroparticle Physics 22 (2005) 339–353

    sensitivity of the OMs, the shadowing of the
    OMs by the string-cable, and the effect of the ex-
    act optical properties of the re-frozen ice enclos-
    ing the OMs in the 60cm wide drill hole. The
    determination of the uncertainty in the absolute
    sensitivity of the OM cannot easily be decoupled
    from the uncertainty in the ice properties. Fol-
    lowing the discussion in
    [25]
    the uncertainty in
    the absolute sensitivity of the OM is estimated
    to be 15%.
    5.2. Optical ice parameters
    High energy muons generate a large number of
    photons in the ice. These photons are scattered
    and absorbed. Due to the large number of pho-
    tons, some photons can be detected over long dis-
    tances in the ice. This analysis uses an average
    effective scattering and absorption length (
    k
    eff
    =
    24m,
    k
    abs
    = 130m) to describe the propagation of
    photons. The error on both parameters is esti-
    mated to be about 10%. The parameters are not
    independent, i.e. a larger scattering length implies
    a larger absorption length. To estimate the system-
    atic effect, we simulate the muon–neutrino signal
    with (
    k
    eff
    =19m,
    k
    abs
    = 90m), (
    k
    eff
    =22m,
    k
    abs
    =
    110m) and (
    k
    eff
    =26m,
    k
    abs
    = 145m) in addition
    to the standard combination. The variation of
    the parameters span a wider range than allowed
    by the estimated uncertainty, to account for an
    overall shift seen in these values when comparing
    earlier[1]and newer analyses[26]of experimental
    data. The variation of the ice parameters changes
    the absolute number of events predicted by the
    simulation. If the number of events predicted by
    the simulation were to differ from the number of
    observed events by significantly more than allowed
    by the overall error one would introduce a normal-
    ization procedure. This is taken into account when
    determining the systematic uncertainty due to var-
    iation of the optical parameters, by normalizing
    the number of events in the leftmost (most back-
    ground like) bin in the NN2 distribution arbi-
    trarily to 100. UsingTable 3, the relative error of
    the average number of events, 34%, is taken as sys-
    tematic uncertainty introduced by the uncertainty
    in the description of the optical properties of the
    ice.
    5.3. Muon propagation
    The path length and energy loss distribution for
    muons during their passage through ice is subject
    to uncertainties. To investigate this effect, muons
    caused by UHE neutrinos are propagated with
    two different muon propagation codes [19,18]
    and the number of signal events passing the final
    selection NN2 > 0.9 are counted. From this, the
    uncertainty related to muon propagation simula-
    tion is estimated to be 6%.
    5.4. Neutrino cross section
    In[27]neutrino cross sections are calculated up
    to 10
    21
    eV. Below 10
    16
    eV all standard sets of par-
    ton distributions yield very similar cross sections.
    Above this energy, the cross sections are sensitive
    to assumptions made about the behavior for
    x
    !
    0. The authors of [27] conclude that at
    10
    20
    eV the uncertainty reaches a factor of 2. Here
    the charged-current cross section is multiplied/di-
    vided by a factor that increases linearly to 2 with
    the energy increasing from 10
    16
    eV to 10
    20
    eV.
    Again, the number of signal events passing the
    final selection NN2 > 0.9 is counted and an un-
    certainty of 8% is derived.
    5.5. Primary cosmic ray flux
    The uncertainty in the absolute primary cosmic
    ray flux enters as a scaling factor for the number
    of events expected from atmospheric air shower
    Table 3
    Number of events passing the final selection criterion
    NN2 > 0.9 for different combinations of effective scattering
    and absorption length
    Ice model Events passing NN2 > 0.9
    k
    eff
    =19m,
    k
    abs
    = 90m 22.2
    k
    eff
    =22m,
    k
    abs
    = 110m 26.7
    k
    eff
    =24m,
    k
    abs
    = 130m 22.7
    k
    eff
    =26m,
    k
    abs
    = 145m 42.7
    Average 28.6 ± 9.6
    Relative error 34%
    The relative difference in event numbers is an estimate of the
    systematic error introduced by uncertainties in the description
    of the optical ice properties.
    M. Ackermann et al. / Astroparticle Physics 22 (2005) 339–353
    349

    simulations. Ref.
    [28]
    summarizes the integrated
    flux averaged for different experiments and gives
    the spread between the data from different experi-
    ments as error. In the energy range of interest the
    error on the absolute flux does not exceed 20%.
    Different elements of the same energy as primary
    cosmic ray particles lead to a different shower
    development and can affect the number of events
    expected from these air showers. The model used
    for the primary cosmic ray composition taken
    from
    [17]
    is shown in
    Fig. 3. This model fits a
    ‘‘heavy’’ composition for high energy primaries.
    Using data from the combined SPASE-AMAN-
    DA experiment, a similar trend to heavier prima-
    ries at higher energies is seen [29]. To investigate
    the sensitivity of the analysis to the composition
    of the primary cosmic rays, the proton and iron
    content was varied between 0% and 100%. The
    two extremes, protons only or iron only, together
    with the model from Fig. 3 are shown at Level 2
    in
    Fig. 10. The shape of the distribution is not
    influenced by the primary composition. After nor-
    malizing to the number of events in experimental
    data (3326 events), the variation in the number
    of events passing the NN2 > 0.9 selection while
    varying the proton/iron content is used to deter-
    mine the uncertainty. Allowing the two extremes
    (proton or iron only) this uncertainty is about
    25%. This is an overestimation, as the proton or
    iron only compositions are contradicted by other
    experiments. Allowing a minimum of 20% of pro-
    tons or iron as extremes decreases the uncertainty
    to about 15%, the value used in this analysis.
    5.6. Summary of systematic uncertainties
    This analysis accounts for the major sources of
    systematic uncertainties, including uncertainties in
    the background simulation. The variation of the
    optical properties leads to the largest error of
    34%, the absolute sensitivity of the OM contrib-
    utes 15%, the neutrino cross section 8% and the
    muon propagation 6%. Treating these errors as
    independent and adding them quadratically, the
    signal uncertainty becomes 39%. For down-going
    atmospheric air showers (i.e. background), we
    add in quadrature the statistical uncertainty of
    26% (4.6 ± 1.2 expected events), uncertainties from
    composition of 15% and normalization of the
    absolute flux of 20% yielding a total uncertainty
    of 36%.
    6. Neutrino flux limits
    Fig. 7
    shows the distribution of NN2 for
    experimental data, the simulated air showers,
    and a neutrino flux simulation with
    E
    2
    U
    =
    10
    ?
    6
    GeVcm
    ?
    2
    s
    ?
    1
    sr
    ?
    1
    . Applying the selection cri-
    terion NN2 > 0.9, the simulation gives a back-
    ground expectation of 4.6 ± 1.2 events while the
    experiment yields 5 events for a live time of 131
    days. Using the tables by Feldman and Cousins
    [22] and disregarding systematic uncertainties, this
    results in an upper limit of 4.7 events at 90% con-
    fidence level. With this, an all flavor neutrino
    source with a spectrum proportional to
    E
    ?
    2
    is
    limited by
    E
    2
    U
    = 0.63
    ·
    10
    ?
    6
    GeVcm
    ?
    2
    s
    ?
    1
    sr
    ?
    1
    .
    Including the systematic uncertainties as evaluated
    above and following the method described in [23],
    the event upper limit increases to 7.35 (90% confi-
    Fig. 10. NN2 for cosmic ray air shower simulations and
    different primary compositions at Level 2. The composition
    from [17] is also shown in Fig. 3, the other two represent the
    extreme cases of only proton or iron primaries. The total
    number of events is normalized to the experimental number of
    events.
    350
    M. Ackermann et al. / Astroparticle Physics 22 (2005) 339–353

    dence level) and the upper flux limit to
    E
    2
    U
    =
    0.99
    ·
    10
    ?
    6
    GeVcm
    ?
    2
    s
    ?
    1
    sr
    ?
    1
    . This is 57% above
    the limit without systematic uncertainties. Fig. 11
    shows the energy distribution for a diffuse neutrino
    flux corresponding to the above limit. Ninety per-
    cent of the events are contained in the energy re-
    gion from 10
    15
    eV to 3
    ·
    10
    18
    eV, with peaks just
    above 10
    16
    eV for the muon–neutrino contribution
    and at the Glashow resonance (6.3PeV) for the
    electron–neutrino contribution.
    The experimental data can be used to set limits
    on neutrino flux predictions other than a generic
    E
    ?
    2
    spectrum. Several AGN models in the litera-
    ture predict fluxes that might be detectable with
    the sensitivity derived above. The pioneering
    AGN core model by Stecker et al. (S91) [30] has
    been updated by the more recent model (S96)
    [31]. Others are the AGN jet model by Protheroe
    (P97) [32] and a model by Mannheim (M95) [33].
    With systematic uncertainties included, this analy-
    sis excludes at 90% confidence level the neutrino
    fluxes predicted by the models S91, S96 and P97
    with model rejection factors between 0.24 and
    0.97, while M95 is not quite excluded (see
    Table
    4). Other models generally predict lower fluxes,
    while neutrinos from cosmic ray interactions with
    E
    Fig. 11. Energy distribution for the sum and the individual con-
    tributions of the three neutrino flavors for a diffuse
    E
    ?
    2
    neutrino flux with a strength of
    E
    2
    U
    = 0.99
    ·
    10
    ?
    6
    GeVcm
    ?
    2
    s
    ?
    1
    sr
    ?
    1
    after the final selection criterion. The num-
    ber of events from
    m
    e
    ,
    m
    l
    and
    m
    s
    are 2.2, 3.6 and 1.5.
    Table 4
    Number of events expected (
    n
    s
    ) for 131 days after applying the
    selection NN2 > 0.9 for a generic
    E
    ?
    2
    source and four different
    predicted AGN models (references given in the text)
    Model
    n
    s
    NN2 > 0.9
    mrf mrf excl.
    sys.
    mrf incl.
    sys.
    10
    ?
    6
    E
    ?
    2
    7.44 ± 0.14 0.69 0.63 0.99
    S91 7.56 ± 0.25 0.67 0.62 0.97
    S96 30.5 ± 1.0 0.17 0.15 0.24
    P97 13.45 ± 0.21 0.38 0.35 0.55
    M95 6.17 ± 0.04 0.83 0.76 1.19
    UL – 5.1 4.7 7.35
    The model rejection factor is the factor by which the source
    strength needs to be multiplied to equal the event upper limit
    (UL). The middle column
    ð
    mrf
    Þ
    gives the sensitivity, while the
    last two columns give the experimental results for the model
    rejection factor excluding and including systematic uncertain-
    ties. A rejection factor <1 indicates an exclusion of the neutrino
    flux caused by the model at 90% confidence level.
    Fig. 12. Three models of AGN neutrino emission, S91 [30], S96
    [31] and P97 [32] which are excluded by this analysis and the not
    quite excluded model M95
    [33]
    are shown as well as the
    experimental 90% confidence level upper limit to an diffuse
    E
    ?
    2
    all flavor neutrino flux as a bold solid line. The dotted line
    shows the sensitivity of this analysis. The thin line to the left
    shows the AMANDA-II all flavor limit [35] for lower energies.
    The range of the presented limit corresponds to the region that
    contains 90% of the expected signal.
    M. Ackermann et al. / Astroparticle Physics 22 (2005) 339–353
    351

    the microwave background or from decays of
    topological defects are too high in energy and
    too low in flux for the achieved sensitivity.
    7. Summary and discussion
    Data recorded with the AMANDA-B10 detec-
    tor in 1997 are searched for leptons caused by a
    diffuse flux of UHE neutrinos. In contrast to the
    analyses [26] and [34] which used the AMANDA
    detector to search for an upward traveling flux of
    neutrinos, or the analysis
    [35]
    which optimized
    for cascades at lower energies, this analysis
    searches for a UHE signal from horizontal and
    down-going events. Restricting the analysis to very
    bright events, the flux of down-going muons
    caused by atmospheric air showers can be suffi-
    ciently suppressed while retaining a large sensitiv-
    ity to neutrinos. At all levels, the experimental
    data are described well by the air shower simula-
    tion. The sensitivity (excluding systematic errors)
    of this analysis to an equally mixed all flavor neu-
    trino flux is
    E
    2
    U
    ¼
    0
    :
    69
    ?
    10
    ?
    6
    GeVcm
    ?
    2
    s
    ?
    1
    sr
    ?
    1
    .
    Including a combined error (statistical and system-
    atic) of 36% for the background and 39% for the
    signal simulation, this analysis sets a flux upper
    limit for 131 days of
    E
    2
    U
    (
    m
    e
    :
    m
    l
    :
    m
    s
    = 1:1:1) =
    0.99
    ·
    10
    ?
    6
    GeVcm
    ?
    2
    s
    ?
    1
    sr
    ?
    1
    at 90% confidence
    level, shown as bold line in
    Fig. 12. The energy
    range 10
    15
    –3
    ·
    10
    18
    eV contains 90% of the neu-
    trino induced events, with the remainder being
    equally divided above and below this range. This
    energy range is well suited to exploring neutrino
    emission models from AGN blazars, and three
    specific model predictions, two by Stecker et al.
    [30,31]
    and one by Protheroe
    [32], are excluded.
    The recent analysis [35] has excluded the two mod-
    els by Stecker et al., but not the model by Prothe-
    roe. An extension to even higher energies would be
    desirable in order to explore neutrino emission
    from the decay of topological defects or to search
    for the guaranteed UHE neutrinos from interac-
    tions of the highest energy cosmic rays with the
    Fig. 13. The same UHE muon–neutrino event in the AMANDA-B10 detector (left) and the IceCube array (right), illustrating the
    amount of additional information gained by the larger size of the detector and the larger number of optical modules in IceCube.
    352
    M. Ackermann et al. / Astroparticle Physics 22 (2005) 339–353

    microwave background radiation. This is pre-
    vented by the limited size and relatively small
    number of OMs of the AMANDA-B10 detector,
    where saturation effects become visible. However
    the larger AMANDA-II detector
    [36]
    equipped
    with optical fibers for dispersion free signal trans-
    mission and its advanced technique of waveform
    capture will improve the analysis of UHE neutri-
    nos, as will the IceCube detector [37] on a longer
    time scale.
    Fig. 13
    shows the simulated response
    to a muon caused by an UHE neutrino interaction
    passing the AMANDA-B10 and IceCube array,
    respectively. In particular the IceCube array, with
    construction beginning during 2004/2005, will dra-
    matically increase the amount of information
    exploitable in the UHE regime.
    Acknowledgements
    We acknowledge the support of the following
    agencies: National Science Foundation––Office of
    Polar Programs, National Science Foundation––
    Physics Division, University of Wisconsin Alumni
    Research Foundation, Department of Energy and
    National Energy Research Scientific Computing
    Center (supported by the Office of Energy Re-
    search of the Department of Energy), UC-Irvine
    ANEAS Supercomputer Facility, USA; Swedish
    Research Council, Swedish Polar Research Secre-
    tariat and Knut and Alice Wallenberg Founda-
    tion, Sweden; German Ministry for Education
    and Research, Deutsche Forschungsgemeinschaft
    (DFG), Germany; Fund for Scientific Research
    (FNRS-FWO), Flanderns Institute to encourage
    Scientific and Technological Research in Industry
    (IWT) and Belgian Federal Office for Scientific,
    Technical and Cultural affairs (OSTC), Belgium;
    Fundacio
    ´
    n Venezolana de Promocio
    ´
    n al Investig-
    ador (FVPI), Venezuela; D.F.C acknowledges the
    support of the NSF Career program; E.R.
    acknowledges the support of the Marie-Currie fel-
    lowship program of the European Union.
    References
    [1] E. Andres et al. [AMANDA Collaboration], Nature 410
    (2001) 441.
    [2] R. Gandhi et al., Phys. Rev. D 58 (1998) 093009.
    [3] J.R. Klein, A.K. Mann, Astropart. Phys. 10 (1999) 321.
    [4] S. Hundertmark et al. [AMANDA Collaboration], Proc.
    27th Int. Cosmic Ray Conf., Hamburg, Germany (2001)
    1129.
    [5] E. Waxman, Nucl. Phys. Proc. Suppl. 100 (2001) 314.
    [6] J.G. Learned, K. Mannheim, Ann. Rev. Nucl. Part. Sci. 50
    (2000) 679.
    [7] R.J. Protheroe, arXiv:astro-ph/9612213.
    [8] H. Athar et al., Phys. Rev. D 62 (2000) 103007.
    [9] L. Anchordoqui et al., arXiv:hep-ph/0407020.
    [10] S. Hundertmark et al. [AMANDA Collaboration], DESY-
    PROC-2002-01, 2002, 69.
    [11] Photonis, Photomultiplier tubes,
    http://www.photonis.
    com.
    [12] Hamamatsu, Photomultiplier Tubes, Basics and Applica-
    tions, second ed., 1999.
    [13] M. Kowalski and A. Gazizov, arXiv:astro-ph/0312202.
    [14] A.M. Dziewonski, D.L. Anderson, Phys. Earth Planet.
    Interiors 25 (1981) 297.
    [15] D. Heck, DESY-PROC-1999-01 (1999) 227.
    [16] B. Wiebel-Sooth, P. Biermann, Landolt-Bornstein, vol. VI/
    3c, Springer Verlag, 1999, pp. 37–90.
    [17] R. Glasstetter et al. [KASCADE Collaboration], FZKA-
    6345E (1999).
    [18] D. Chirkin, W. Rhode, arXiv:hep-ph/0407075.
    [19] P. Lipari, T. Stanev, Phys. Rev. D 44 (1991) 3543.
    [20] J. Ahrens et al. [AMANDA Collaboration], Nucl.
    Instrum. Methods A 524 (2004) 169.
    [21] G.C. Hill, K. Rawlins, Astropart. Phys. 19 (2003) 393.
    [22] G.J. Feldman, R.D. Cousins, Phys. Rev. D 57 (1998)
    3873.
    [23] J. Conrad et al., Phys. Rev. D 67 (2003) 012002.
    [24] R.D. Cousins, V.L. Highland, Nucl. Instrum. Methods A
    320 (1992) 331.
    [25] J. Ahrens et al. [AMANDA Collaboration], Astrophys. J.
    583 (2003) 1040.
    [26] J. Ahrens et al. [AMANDA Collaboration], Phys. Rev.
    Lett. 90 (2003) 251101.
    [27] R. Gandhi et al., Astropart. Phys. 5 (1996) 81.
    [28] J.R. Hoerandel, Astropart. Phys. 19 (2003) 193.
    [29] T. Miller et al. [SPASE and AMANDA Collaborations],
    Astropart. Phys. 21 (2004) 565.
    [30] F.W. Stecker et al., Phys. Rev. Lett. 66 (1991) 2697;
    F.W. Stecker et al., Phys. Rev. Lett. 69 (1992) 2738
    (Erratum).
    [31] F.W. Stecker, M.H. Salamon, Space Sci. Rev. 75 (1996)
    341.
    [32] R.J. Protheroe, arXiv:astro-ph/9607165.
    [33] K. Mannheim, Astropart. Phys. 3 (1995) 295.
    [34] J. Ahrens et al. [AMANDA Collaboration], Phys. Rev. D
    67 (2003) 012003.
    [35] M. Ackermann et al. [AMANDA Collaboration], Astro-
    part. Phys. 22 (2004) 127.
    [36] S.W. Barwick et al. [AMANDA Collaboration], arXiv:
    astro-ph/0211269.
    [37] IceCube Project web page, http://icecube.wisc.edu.
    M. Ackermann et al. / Astroparticle Physics 22 (2005) 339–353
    353

    Back to top