SEARCH FOR POINT SOURCES OF HIGH-ENERGY NEUTRINOS WITH AMANDA
    J. Ahrens,
    1
    X. Bai,
    2
    G. Barouch,
    3
    S. W. Barwick,
    4
    R. C. Bay,
    5
    T. Becka,
    1
    K.­H. Becker,
    6
    D. Bertrand,
    7
    F. Binon,
    7
    A. Biron,
    8
    S. Boeser,
    8
    O. Botner,
    9
    A. Bouchta,
    8,10
    O. Bouhali,
    7
    T. Burgess,
    11
    S. Carius,
    12
    T. Castermans,
    13
    D. Chirkin,
    5,6
    J. Conrad,
    9
    J. Cooley,
    3
    D. F. Cowen,
    14
    A. Davour,
    9
    C. De Clercq,
    15
    T. DeYoung,
    3
    P. Desiati,
    3
    J.­P. Dewulf,
    7
    P. Doksus,
    3
    J. Edsjo
    ¨
    ,
    11
    P. Ekstro
    ¨
    m,
    11
    T. Feser,
    1
    T. K. Gaisser,
    2
    M. Gaug,
    8
    L. Gerhardt,
    4
    A. Goldschmidt,
    16
    A. Hallgren,
    9
    F. Halzen,
    3
    K. Hanson,
    3
    R. Hardtke,
    3
    T. Hauschildt,
    8
    M. Hellwig,
    1
    P. Herquet,
    13
    G. C. Hill,
    3
    P. O. Hulth,
    11
    K. Hultqvist,
    11
    S. Hundertmark,
    4
    J. Jacobsen,
    16
    A. Karle,
    3
    J. Kim,
    4
    L. Ko
    ¨
    pke,
    1
    M. Kowalski,
    8
    K. Kuehn,
    4
    J. I. Lamoureux,
    16
    H. Leich,
    8
    M. Leuthold,
    8
    P. Lindahl,
    12
    J. Madsen,
    17
    P. Marciniewski,
    9
    H. Matis,
    16
    C. P. McParland,
    16
    T. C. Miller,
    2,18
    Y. Minaeva,
    11
    P. Miocinovic
    ´
    ,
    5
    P. C. Mock,
    4,19
    R. Morse,
    3
    T. Neunho
    ¨
    ffer,
    1
    P. Niessen,
    15
    D. R. Nygren,
    16
    H. O
    ¨
    gelman,
    3
    P. Olbrechts,
    15
    C. Pe
    ´
    rez de los Heros,
    9
    A. C. Pohl,
    12
    P. B. Price,
    5
    G. T. Przybylski,
    4
    K. Rawlins,
    3
    E. Resconi,
    8
    W. Rhode,
    6
    M. Ribordy,
    8
    S. Richter,
    3
    J. Rodrı
    ´
    guez Martino,
    11
    P. Romenesko,
    3
    D. Ross,
    4
    H.­G. Sander,
    1
    T. Schmidt,
    8
    D. Schneider,
    3
    R. Schwarz,
    3
    A. Silvestri,
    4
    M. Solarz,
    5
    G. M. Spiczak,
    17
    C. Spiering,
    8
    D. Steele,
    3
    P. Steffen,
    8
    R. G. Stokstad,
    16
    K.­H. Sulanke,
    8
    I. Taboada,
    20
    L. Thollander,
    11
    S. Tilav,
    2
    C. Walck,
    11
    C. Weinheimer,
    1
    C. H. Wiebusch,
    8,10
    C. Wiedemann,
    11
    R. Wischnewski,
    8
    H. Wissing,
    8
    K. Woschnagg,
    5
    W. Wu,
    4
    G. Yodh,
    4
    and S. Young
    4
    (The AMANDA Collaboration)
    Received 2002 July 26; accepted 2002 October 2
    ABSTRACT
    This paper describes the search for astronomical sources of high-energy neutrinos using the
    AMANDA-B10 detector, an array of 302 photomultiplier tubes used for the detection of Cerenkov light
    from upward-traveling neutrino-induced muons, buried deep in ice at the South Pole. The absolute pointing
    accuracy and angular resolution were studied by using coincident events between the AMANDA detector
    and two independent telescopes on the surface, the GASP air Cerenkov telescope and the SPASE extensive
    air shower array. Using data collected from 1997 April to October (130.1 days of live time), a general survey
    of the northern hemisphere revealed no statistically significant excess of events from any direction. The
    sensitivity for a flux of muon neutrinos is based on the effective detection area for through-going muons.
    Averaged over the northern sky, the effective detection area exceeds 10,000 m
    2
    for
    E
    l
    ?
    10 TeV. Neutrinos
    generated in the atmosphere by cosmic-ray interactions were used to verify the predicted performance of the
    detector. For a source with a differential energy spectrum proportional to
    E
    ?
    2
    ?
    and declination larger than
    +40
    ?
    , we obtain
    E
    2
    ð
    dN
    ?
    =
    dE
    Þ?
    10
    ?
    6
    GeV cm
    ?
    2
    s
    ?
    1
    for an energy threshold of 10 GeV.
    Subject headings:
    neutrinos — surveys
    On-line material:
    color figures
    1
    Institute of Physics, University of Mainz, Staudinger Weg 7, D-55099 Mainz, Germany.
    2
    Bartol Research Institute, University of Delaware, Newark, DE 19716.
    3
    Department of Physics, University of Wisconsin, Madison, WI 53706.
    4
    Department of Physics and Astronomy, University of California, Irvine, CA 92697.
    5
    Department of Physics, University of California, Berkeley, CA 94720.
    6
    Fachbereich 8 Physik, BUGH Wuppertal, D-42097 Wuppertal, Germany.
    7
    Universite
    ´
    Libre de Bruxelles, Science Faculty CP 230, Boulevard du Triomphe, B-1050 Brussels, Belgium.
    8
    DESY-Zeuthen, D-15735 Zeuthen, Germany.
    9
    Division of High Energy Physics, Uppsala University, S-75121 Uppsala, Sweden.
    10
    Current address: CERN, CH-1211 Geneva 23, Switzerland.
    11
    Fysikum, Stockholm University, S-11385 Stockholm, Sweden.
    12
    Department of Technology, University of Kalmar, S-39182 Kalmar, Sweden.
    13
    University of Mons-Hainaut, Mons, Belgium.
    14
    Department of Physics, Pennsylvania State University, University Park, PA 16802.
    15
    Vrije Universiteit Brussel, Dienst ELEM, B-1050 Brussels, Belgium.
    16
    Lawrence Berkeley National Laboratory, Berkeley, CA 94720.
    17
    Department of Physics, University of Wisconsin, River Falls, WI 54022.
    18
    Current address: Johns Hopkins University, Applied Physics Laboratory, Laurel, MD 20723.
    19
    Current address: Optical Networks Research, JDS Uniphase, 100 Willowbrook Road, Freehold, NJ 07728-2879.
    20
    Departamento Fı
    ´
    sica, University Simo
    ´
    n Bolı
    ´
    var, Caracas, Venezuela.
    The Astrophysical Journal
    , 583:1040–1057, 2003 February 1
    #
    2003. The American Astronomical Society. All rights reserved. Printed in U.S.A.
    E
    1040

    1. INTRODUCTION
    Nature provides precious few information carriers from
    the deep recesses of space, and it is imperative to develop
    techniques to exploit each one. Throughout history, the
    photon messenger has made vital contributions to the
    understanding of the observable universe. In this paper, we
    present results from a new generation of telescopes designed
    to detect a very different kind of information carrier, high-
    energy neutrinos (where
    E
    ?
    >
    1 TeV). The search for
    astronomical sources of high-energy neutrinos is one of the
    central missions of the Antarctic Muon and Neutrino
    Detector Array (AMANDA; Wischnewski et al. 1999). In
    this paper, we describe a general search for continuous emis-
    sion from a spatially localized direction in the northern
    sky,
    21
    restricted to declinations greater than +5
    ?
    . The search
    technique is conceptually simple: a point source is identified
    by a statistically significant enhancement over expected
    background fluctuations from a particular direction.
    Expected background is readily obtained experimentally
    from off-source sky bins within the same band of declina-
    tion. In contrast, unresolved, or diffuse, signals are charac-
    terized by an isotropic distribution, and backgrounds are
    estimated by detector simulation programs. The most favor-
    able flux predictions for point sources are several orders of
    magnitude lower than the most optimistic predictions for
    diffusely distributed sources. However, atmospheric neu-
    trino background is diffusely distributed as well, so the level
    of intrinsic background in the diffuse search is also several
    orders of magnitude higher. While signal-to-noise ratio con-
    siderations favor the search for diffuse emission over point-
    source searches, the interpretation of a diffusely distributed
    signal is more ambiguous. Thus, the search for point sources
    complements the search for diffuse sources. The latter
    search is described in Hill & Leuthold (2001).
    22
    The more
    specific searches for point emission from gamma-ray
    bursters (Hardtke & Barouch 2001)
    23
    and quasi-pointlike
    emission from Galactic dark matter trapped in the core of
    the Earth (Ahrens et al. 2002b) are presented in separate
    papers since these analyses were optimized for different flux
    spectra and different background characteristics.
    2. MOTIVATION
    The origin of cosmic rays is one of the oldest puzzles in
    particle astrophysics. Shocks from Galactic supernovae are
    widely believed to accelerate cosmic rays to
    ?
    10
    15
    eV, while
    the sources of cosmic rays at the most extreme energies are
    not known. Plausible models of particle acceleration exist
    for many classes of Galactic and extragalactic objects, but
    supporting evidence for any model is largely circumstantial.
    The observation of high-energy neutrinos from point
    sources would unequivocally confirm the hadronic nature
    of such accelerators. Unfortunately, the predicted neutrino
    fluxes from Galactic and extragalactic point sources are
    too low to be detected with AMANDA-B10, although
    uncertainties in the model parameters lead to considerable
    variation in the flux predictions.
    Supernova remnants (SNRs) are one of the few classes of
    Galactic sites that have the capability to supply sufficient
    power to accelerate Galactic cosmic rays. The diffusive
    shock mechanism naturally produces a power-law spectrum
    of
    dN
    =
    dE
    /
    E
    ?
    2
    :
    1
    , which is consistent with the deduced
    spectral index of cosmic rays.
    Recent observations of TeV gamma rays from plerions
    such as the Crab Nebula and SNRs provide direct evidence
    for particle acceleration to high energies. However, these
    observations do not provide compelling evidence for
    hadronic
    acceleration because of an unfortunate ambiguity:
    it is possible (and even probable) that electrons are solely
    responsible for the high-energy gamma-ray production. But
    if SNRs are the accelerators of Galactic cosmic rays, they
    must also accelerate hadrons. A class of models exploits this
    idea by suggesting that both protons and electrons are accel-
    erated by the supernova shock. Pions, both neutral and
    charged, are produced in the nuclear collisions between pro-
    tons and ambient material (a cosmic equivalent of a ‘‘ beam
    dump ’’ commonly used by terrestrial accelerators) and then
    decay to high-energy gamma rays and neutrinos.
    While the notion of particle acceleration by supernova
    shocks provides a credible and largely consistent picture,
    not all observations neatly fit this scheme. Alternative sites
    for cosmic-ray acceleration may emerge from a detailed
    study of the neutrino sky. For example, Galactic microqua-
    sars, a subclass of X-ray binary systems that exhibit relativ-
    istic radio jets, have been identified as possible sources of
    high-energy neutrino emission (Levinson & Waxman 2001)
    and potential sources of the highest energy cosmic rays. If
    they accelerate cosmic rays to high energies, then their dense
    environment creates suitable conditions for an efficient
    beam dump.
    Turning to extragalactic sources, active galactic nuclei
    (AGNs) are among the most luminous objects in the uni-
    verse and promising sources of neutrinos. In these models,
    high-energy neutrino fluxes are generated near the central
    engine or in the jets of radio-loud AGNs (e.g., blazars, a
    class of objects where the jet intersects the line of sight of the
    observer). The fact that gamma-ray emission has been
    detected (Cantanese & Weekes 1999) from nearby blazars
    Mrk 421 and Mrk 501 provides strong evidence for particle
    acceleration to high energies.
    24
    The time-averaged energy
    spectrum from Mrk 501 during 1997 is consistent with an
    unbroken power-law energy spectrum up to 10 TeV
    (Weekes 2001; Konopelko et al. 1999). Beam dump models
    of neutrino production predict comparable fluxes of
    gamma-ray photons and neutrinos. However, gamma-ray
    photons at TeV energies may interact with material or pho-
    ton fields in the source or interact with the diffuse infrared
    background photons during their flight. Because of this
    reprocessing, the measured energy spectrum for gamma-ray
    photons may not trace the energy spectrum of the source.
    Consequently, it is possible for the ratio of neutrino flux to
    gamma-ray flux from a given source to exceed unity. Con-
    straints on this ratio are discussed in
    x
    7.
    Recently, it has been argued (Buckley et al. 1998a, 1998b)
    that the rapid time variability of high-energy photon
    21
    Although the flux limits reported in this paper are computed assuming
    continuous emission, upper bounds could be generated for periodic or
    episodic emission as well.
    22
    Available at http://area51.berkeley.edu/manuscripts/
    20010604xx-diffuse-sens.pdf.
    23
    Available at http://area51.berkeley.edu/manuscripts/
    20010606xx-grbdoc.pdf.
    24
    The review by Cantanese & Weekes (1999) presents a current list of
    detected very high energy gamma-ray sources.
    SEARCH FOR HIGH-ENERGY NEUTRINO SOURCES 1041

    emission from AGN blazars and the correlated variation
    between the X-ray and TeV regimes disfavor hadronic
    acceleration models for this particular class of objects, but
    others have shown that rapid and correlated variability can
    be accommodated by modest extensions to the existing
    hadronic acceleration models (Rachen 2000; Dermer 1999).
    The vigorous debate suggests that high-energy neutrino
    detectors can play a central role in deciphering the accelera-
    tion mechanism.
    Figure 1 provides a survey of model predictions for
    fluxes of high-energy neutrinos. The models were selected
    to highlight the variation in spectral characteristics. The
    line labeled 3C 273 is representative of several recent neu-
    trino flux predictions, e.g., a similar differential flux is
    predicted for microquasars (Distefano et al. 2002) in the
    region of sky visible to AMANDA and the flat-spectrum
    radio quasar 3C 279 (Dermer & Atoyan 2001). The figure
    also shows the differential neutrino flux limit for an
    assumed source spectrum proportional to
    E
    ?
    2
    for
    AMANDA-B10 and the anticipated corresponding sensi-
    tivity of AMANDA-II and IceCube (Karle et al. 2003).
    The AMANDA-B10 result, the subject of this paper, is
    valid for declinations greater than +40
    ?
    . Many theoreti-
    cal models of potential astronomical neutrino sources
    predict a very hard energy spectrum, approximately
    E
    ?
    2
    (Learned & Mannheim 2000), which leads to a most
    probable energy for a detected neutrino well above 1
    TeV (typically 10–30 TeV). This high energy is a conse-
    quence of three facts: the neutrino cross section for weak
    interactions increases with neutrino energy, and the prop-
    agation length of the secondary muon increases and the
    effective detection area increases as light emission along
    the muon track increases with energy.
    Even though the cosmic-ray puzzle provides a powerful
    motivation to explore the sky for neutrino emission, not all
    high-energy neutrino sources need to contribute to the
    cosmic-ray flux. In particular, a powerful accelerator may
    be surrounded by too much material to emit high-energy
    photons or cosmic rays (they would interact and cascade
    down to lower energies), but this accelerator could be dis-
    covered by exploiting the neutrino messenger. Several
    models of such ‘‘ hidden ’’ sources have appeared in the liter-
    ature. For example, the predicted flux of neutrinos from
    pre-AGN objects (Berezinsky & Dokuchaev 2001) leads to
    a muon detection rate of
    ?
    10 yr
    ?
    1
    km
    ?
    2
    .
    3. DESCRIPTION OF THE AMANDA DETECTOR
    The AMANDA telescope is located below the surface of
    the Antarctic ice sheet at the geographic South Pole. The
    neutrino detection technique relies on the detection of Cer-
    enkov light from upward-traveling neutrino-induced
    muons. Figure 2 shows the current configuration of the
    AMANDA detector. The shallow array, AMANDA-A,
    was deployed to depths between 800 and 1000 m in an
    exploratory phase of the project. The deeper array of 10
    strings, referred to as AMANDA-B10, was deployed during
    the austral summers between 1995 and 1997, to depths
    between 1500 and 2000 m. At this depth, the optical proper-
    ties are suitable for track reconstruction (Woschnagg et al.
    10
    ­10
    10
    ­8
    10
    ­6
    10
    ­4
    10
    2
    10
    3
    10
    4
    10
    5
    10
    6
    E
    2
    (dN/dE) [GeVcm
    ­2
    s
    ­1
    ]
    E
    ν
    [GeV]
    AMANDA­II (1 yr l.t.)
    AMANDA­B10 (this work)
    IceCube
    3C273
    Crab
    AGN Core
    Mkn 501 (
    ν
    =
    γ
    )
    Atm.
    ν
     
    Fig.
    1.
    —Representative survey of
    ?
    þ
    ?
    flux predictions from cosmic
    accelerators of high-energy neutrinos. The AMANDA-B10 result is pre-
    sented here. The dashed horizontal lines give preliminary estimates of the
    minimum detectable flux by AMANDA-II after 1 yr of live time (Barwick
    et al. 2001) and IceCube (Spiering 2001). The atmospheric neutrino fluxes
    (Agrawal et al. 1996) are appropriate for a circular patch of 1
    ?
    (
    lower curve
    )
    and 3
    ?
    radius. The curves do not include the normalization uncertainty,
    possibly 30% in magnitude (Gaisser 2002; T. K. Gaisser 2002, private com-
    munication). Models: 3C 273 (Nellen, Mannheim, & Biermann 1993), Crab
    model I (Bednarek & Protheroe 1999), AGN core (Stecker & Salamon
    1996), and Mrk 501 assuming neutrino spectrum is identical to observed
    gamma spectrum during flaring phase (Weekes 2001). [
    See the electronic
    edition of the Journal for a color version of this figure.
    ]
    120 m
    snow layer
    optical module (OM)
    housing
    pressur
    e
    Optical
    Module
    silicon gel
    HV divider
    light diffuser bal
    l
    60 m
    AMANDA as of 2000
    zoomed in on one
    (true scaling)
    200 m
    Eiffel Tower as comparison
    Depth
    surface
    50 m
    1000 m
    2350 m
    2000 m
    1500 m
    810 m
    1150 m
    AMANDA­A (top)
    zoomed in on
    AMANDA­B10 (bottom)
    AMANDA­A
    AMANDA­B10
    main cable
    PMT
    Fig.
    2.
    —Schematic view of the AMANDA neutrino telescope. This
    paper describes an analysis of data taken in 1997 with AMANDA-B10, the
    10 inner strings shown in the expanded view in the center. Each dot
    represents one optical module in the array. [
    See the electronic edition of the
    Journal for a color version of this figure.
    ]
    1042 AHRENS ET AL. Vol. 583

    1999).
    25
    The strings are arranged in a circular pattern when
    viewed from the surface. The instrumented volume of
    AMANDA-B10 forms a vertical cylinder with a diameter of
    120 m. Most electronics are housed on the surface in a
    research facility located within a kilometer of the Amundsen-
    Scott South Pole Station. The detector was commissioned in
    1997 February (Wischnewski et al. 1999; Barwick et al. 2000)
    and expanded by adding nine strings of optical modules
    (OMs) between 1997 December and 2000 January. The com-
    posite array of 19 strings forms the AMANDA-II array,
    which was commissioned in 2000 February.
    AMANDA-B10 consists of 302 OMs arranged on 10 ver-
    tical strings. Each OM contains an 8 inch (21 cm) diameter
    photomultiplier tube (PMT) controlled by passive elec-
    tronics and housed in a glass pressure vessel. The OMs are
    connected to the surface by dedicated electrical cables,
    which supply high voltage and carry the anode signals from
    the PMTs. For each event, the amplitudes and arrival times
    of the pulses from the OMs are digitized by peak analog-to-
    digital converter and time-to-digital converter (TDC) val-
    ues. The TDCs are capable of measuring eight distinct
    pulses per channel. The precision of the arrival time mea-
    surement is 5 ns. Details of deployment, timing resolution,
    and detector operation can be found in Andres et al. (2000,
    2001). Readout of the entire array was triggered by a major-
    ity logic system, which demanded that at least 16 OMs pro-
    duce signals, or ‘‘ hits,’’ within a time window of 2.2
    l
    s. This
    window takes into account the rather large time variation
    introduced by the large geometric size of the detector and
    the cable propagation delays. Random signals from the
    OMs (or ‘‘ noise ’’) were observed at a rate of 300 Hz on the
    inner four strings and 1.5 kHz for OMs on the outer six
    strings, the difference being due to different levels of radio-
    active potassium in the glass pressure vessels. On average,
    random noise contributed one count per event to the major-
    ity trigger.
    Optical absorption and scattering properties of the glacial
    ice that encapsulates the AMANDA detector have been
    studied using light sources buried with the strings and
    Cerenkov light from atmospheric muons. These studies
    (Woschnagg et al. 1999) confirm that the ice is not homoge-
    neous but consists of horizontal strata correlated with cli-
    matological events in the past, such as ice ages.
    26
    Variations
    in the concentration of insoluble impurities between the
    strata produce a strong modulation of the optical proper-
    ties. The absorption length, averaged over depth within the
    AMANDA-B10 array, is 110 m at a wavelength of 400 nm,
    and the average effective scattering length is approximately
    20 m.
    The detection of neutrinos relies on the observation of
    Cerenkov photons generated by muons created in charged-
    current interactions. At the energies of interest, muons typi-
    cally propagate for distances in excess of several kilometers
    (e.g., a muon with
    E
    ¼
    10 TeV will travel 8 km in water).
    Therefore, neutrino interactions outside the instrumented
    volume can be inferred by the presence of a muon, providing
    a method to extend the volume of the detector beyond the
    instrumented boundary of the array. The average angle
    between the muon direction and the parent neutrino direc-
    tion,
    h
    ?
    ?
    l
    i
    , is approximately 0
    =
    65
    =
    ð
    E
    ?
    =
    TeV
    Þ
    0
    :
    48
    for
    E
    ?
    less
    than 100 TeV (Oppelt 2001).
    27
    However, nearly independ-
    ent of the muon energy, the precision of the measured muon
    direction in AMANDA-B10 is approximately 4
    ?
    (see
    x
    6),
    which dominates the angular uncertainty in the neutrino
    direction.
    The AMANDA-B10 data analyzed here were collected
    between 1997 April and October. Once construction was
    completed in 1997 February, calibration and data manage-
    ment activities continued until April. Operations ceased
    between 1997 late October and 1998 February, because of
    the beginning of construction of the AMANDA-II array.
    Furthermore, limitations in the data acquisition and archiv-
    ing system during that first year of operation reduced the
    total live time to approximately 130.1 days.
    4. SIMULATIONS
    Astronomical signals are unlikely to produce more than a
    few tens of upgoing neutrino events per year in AMANDA-
    B10. Data are therefore overwhelmingly dominated by two
    types of background: downgoing atmospheric muons gener-
    ate essentially all of the recorded events, and atmospheric
    neutrinos contribute a few tens of events per day. The point-
    source search relies on a good understanding of both signal
    and background through simulations based on Monte
    Carlo techniques.
    Atmospheric muon events are generated from the mea-
    sured flux of cosmic rays (Wiebel-Sooth & Biermann 1998).
    Two different air shower simulation packages were used to
    assess systematic uncertainty: BASIEV (Bosiev et al. 1989)
    and CORSIKA (Version 5.6; Heck et al. 1998, 1999) using
    the QGSJET hadronic interaction model. CORSIKA was
    modified to include the curved geometry of the Earth and
    atmosphere to provide a more accurate description of the
    flux at zenith angles close to the horizon. Most characteris-
    tics of the events generated with BASIEV were found to be
    similar to those from the more accurate but computation-
    ally more intensive CORSIKA simulation. The density pro-
    file of the atmosphere was modified for polar conditions,
    but no attempt was made to replicate the small seasonal var-
    iations of the trigger rate (Bouchta et al. 1999).
    28
    Muon
    tracking from the surface to the detector was handled by the
    muon propagation program MUDEDX (Lohmann, Kopp,
    & Voss 1985; W. Lohmann, R. Kopp, & R. Voss 1995, pri-
    vate communication [MUDEDX Version 2.02]), and the
    energy-loss characteristics were compared against two addi-
    tional propagation programs that are available for general
    use: PROPMU (Lipari & Stanev 1991) and MMC (Rhode
    & Chirkin 2001). Integral lateral distributions of muons at
    the depth of AMANDA were simulated for proton and iron
    showers (X. Bai et al. 2003, in preparation) and used for ver-
    ification of the detector performance as described below.
    The propagation of upward-traveling muons from neu-
    trino interactions was treated differently from that of down-
    going atmospheric muons because the energies of signal
    neutrinos were expected to extend to much higher energies.
    25
    Available at http://krusty.physics.utah.edu/~icrc1999/root/vol2/
    h4_1_15.pdf.
    26
    See P. B. Price, K. Woschnagg, & D. Chirkin (2000), AMANDA
    public manuscript 20000201, available at http://amanda.berkeley.edu/
    manuscipts.
    27
    Available at http://www-zeuthen.desy.de/~apohl/files/doktor/ps.gz.
    28
    Available at http://krusty.physics.utah.edu/~icrc1999/root/vol2/
    h3_2_11.pdf.
    No. 2, 2003 SEARCH FOR HIGH-ENERGY NEUTRINO SOURCES 1043

    Neutrinos were tracked through the Earth and allowed to
    interact in the ice within or near the instrumented volume of
    the detector or in the bedrock below (Hill 1997). Muons
    with energies above 10
    5
    :
    5
    GeV at production were propa-
    gated using PROPMU until they reached the rock-ice boun-
    dary and then propagated through the ice in exactly the
    same way as downgoing atmospheric muons. For energies
    below 10
    5
    :
    5
    GeV at the production vertex, muons were
    tracked with MUDEDX. The background fluxes from at-
    mospheric
    ?
    l
    and
    ?
    l
    (Agrawal et al. 1996) were included
    (AMANDA cannot differentiate the charge sign of the
    muon).
    In addition to background from atmospheric muons and
    muon neutrinos, the detection efficiency for atmospheric
    electron neutrinos has been simulated. At the relevant ener-
    gies, the flux of
    ?
    e
    is far smaller than for
    ?
    l
    , so the back-
    ground contribution is small a priori. Furthermore, the
    topology of electron neutrino events, reflecting the electro-
    magnetic shower generated by the secondary electron, is
    spherical rather than linear, and this characteristic has been
    exploited to further increase the rejection. Simulations show
    that the detection rate of atmospheric
    ?
    e
    in the point-source
    analysis is only 0.3% of
    ?
    l
    (Young 2001)
    29
    and therefore
    negligible in this context. We note that a separate analysis
    was devised to search for electron neutrinos and conse-
    quently achieved a much larger sensitivity (Ahrens et al.
    2002c).
    An overview of the simulation of the detector response is
    given in Hundertmark (1999). The depth dependence of the
    optical properties of the bulk ice (Woschnagg et al. 1999)
    30
    is included along with a realistic treatment of trigger condi-
    tions, intrinsic noise rates, and hardware-dependent pulse
    shapes. The linearity and saturation of the photomultiplier
    response is included. The angle-dependent sensitivity of the
    optical module was computed from a convolution of PMT
    quantum and collection efficiencies as estimated by the man-
    ufacturer, detection efficiency, wavelength-dependent trans-
    parency through the pressure housing and buffer gel, and
    obscuration by cables and mechanical support hardware.
    The relative angular dependence of the sensitivity of the
    photomultiplier tubes was measured in the laboratory. The
    local optical properties of the refrozen ice were included in
    the photon tables that describe the probability and the tim-
    ing characteristics of photon propagation (Ahrens et al.
    2002a).
    Figure 3 presents the differential distribution of the multi-
    plicity of optical modules, or channels, participating in an
    event,
    N
    ch
    , for the full detector simulation and for experi-
    mental data after known detector-related artifacts were
    removed. The integrated rates differ by less than 25%, which
    is within the systematic uncertainties associated with the
    flux of the primary cosmic rays (Gaisser 2002; T. K. Gaisser
    2002, private communication) and uncertainties associated
    with the absolute sensitivity of the optical module (see
    x
    8).
    The agreement in the shape of the distribution demonstrates
    a good understanding of the overall sensitivity of the array
    for the most common events that trigger AMANDA-B10.
    As the inset of Figure 3 shows, the largest values of
    N
    ch
    are produced by events with more than one muon. This
    information provides indirect evidence that the response of
    AMANDA to single high-energy muons is correctly mod-
    eled, by the following argument. Multimuon bundles that
    reach AMANDA mainly consist of muons below the critical
    energy of 600 GeV, which implies that energy loss due to
    ionization is near minimum. The Cerenkov light production
    from muons well above the critical energy is dominated by
    electromagnetic showers, and the total light from a muon
    with
    E
    l
    is approximately equal to the light of a bundle of
    N
    muons, where
    N
    ?
    E
    l
    =
    E
    crit
    . Therefore, the light production
    by a multimuon event can be related to the light production
    by a single-muon event. For example, the average energy
    loss per unit length for a muon with energy 10
    13
    eV is
    approximately a factor of 15 larger than for a single muon,
    as long as the energy is below 600 GeV. There are several
    modest limitations to this line of reasoning. One is that the
    lateral distribution of multimuon events generates Ceren-
    kov photons over a much larger cylinder than a single
    muon. Another difference is that multimuon events deposit
    Cerenkov photons more uniformly than the equivalent
    high-energy muon, for which energy loss is dominated by
    occasional pair production and bremsstrahlung. However,
    optical scattering by the ice mitigates the effects of nonuni-
    form photon generation. Simulations show that bundles of
    20 muons generate an
    N
    ch
    distribution similar to that of
    single muons with an energy of 10 TeV. The correlation
    between muon multiplicity and
    N
    ch
    multiplicity is shown in
    Figure 4. Note that bias introduced by the majority logic in
    the trigger prevents the muon multiplicity from converging
    to zero as
    N
    ch
    approaches zero.
    29
    Available at http://area51.berkeley.edu/manuscripts/
    20010702xx-young-main.pdf.
    30
    Available at http://krusty.physics.utah.edu/~icrc1999/
    proceedings.html.
    N
    ch
    Differential Rate [Hz]
    experiment (99 Hz)
    simulation (77 Hz)
    N
    ch
    Relative Frequency
    single­
    μ
    multi­
    μ
    10
    ­3
    10
    ­2
    10
    ­1
    1
    0 20 40 60 80 100 120 140
    10
    ­5
    10
    ­4
    10
    ­3
    10
    ­2
    10
    ­1
    0 50 100
    Fig.
    3.
    —Differential distribution of observed (
    solid curve
    ) and predicted
    (
    dashed curve
    ) trigger rates as a function of the event multiplicity
    N
    ch
    (i.e.,
    the number of optical modules that participate in each event). The inte-
    grated rates are given in parentheses. Note that
    N
    ch
    extends below the
    majority logic threshold of 16 because of removal of data caused by experi-
    mental artifacts.
    Inset
    : Relative contribution to the trigger rates from single
    muons (
    solid curve
    ) and multiple-muon bundles (
    dashed curve
    ) that traverse
    the fiducial volume of the array. [
    See the electronic edition of the Journal for
    a color version of this figure.
    ]
    1044 AHRENS ET AL. Vol. 583

    5. ANALYSIS PROCEDURE
    The analysis procedure exploits two essential characteris-
    tics of the signal to simplify the analysis relative to atmo-
    spheric neutrino measurements. First, the sources are
    assumed to be point sources in the sky, so only events within
    a restricted angular region are considered. Second, we use
    the topological and directional characteristics of the spec-
    trally hard neutrino signal to help reject poorly recon-
    structed atmospheric muons (i.e., downward-traveling
    muons reconstructed as upward traveling) and atmospheric
    neutrinos, both of which have softer spectra. Unlike many
    neutrino detectors, the effective sensitivity of AMANDA
    varies dramatically as a function of the background-
    rejection requirements. By concentrating on harder spectra,
    the effective area of the detector can be increased by relaxing
    the background-rejection criteria. Since the point-source
    analysis tolerates a larger background (
    B
    ) contamination in
    the final data sample, the analysis procedure optimizes on
    signal to noise (
    S
    =
    ffiffiffiffi
    B
    p
    ) rather than signal purity (
    S
    =
    B
    ).
    Prior to track reconstruction and event selection, experi-
    mental data were selected from runs that exhibited no
    abnormal behavior, and various instrumentally induced
    artifacts were removed. Once the data in the runs were certi-
    fied, individual OMs in the array were examined to insure
    proper operation. OM channels with hardware malfunction
    (
    ?
    15% of OMs), such as pickup from unusually large exter-
    nal noise sources or fluctuations in the response of the
    amplifier electronics, were rejected. Approximately 85% of
    OMs remained after deselection. Occasional signals induced
    by cross talk in the electrical cables or surface electronics
    exhibited characteristic behavior and could be removed by
    straightforward restrictions on pulse amplitude and width.
    Noise signals generated internally by the photomultiplier
    tubes were readily removed if their time of arrival occurred
    earlier than 5 ms prior to the formation of the event trigger.
    The reconstruction programs stochastically account for
    PMT noise within the event duration.
    After this initial data cleaning, a number of event recon-
    struction techniques (Andres et al. 2000) are applied to the
    data. The most sophisticated technique relies on a search in
    multiparameter space to find the maximum likelihood for a
    track hypothesis given the recorded hits. After reconstruc-
    tion is completed, events are selected according to a set of
    criteria that retain only the highest quality events that pos-
    sess topological and directional information consistent with
    those expected for upgoing neutrino-induced muons. In a
    first step, the data sample of 1
    :
    05
    ?
    10
    9
    events at trigger
    level is reduced to a more manageable size by two filtering
    stages. Most events in data are readily identified as due to
    downward-traveling muons by computationally fast recon-
    struction routines. Removing these events reduces the data
    approximately by a factor of 10
    3
    .
    An iterative analysis procedure was developed to maxi-
    mize
    S
    =
    ffiffiffiffi
    B
    p
    for a simulated signal with an energy spectrum
    proportional to
    E
    ?
    2
    . It ignored the absolute time of the
    event, which helped to minimize bias from potential sources
    in the data. In this optimization the background was deter-
    mined from experimental data by assuming that the fraction
    of signal events in the data sample is negligible. After the fil-
    tering stages, cuts were applied sequentially on a set of selec-
    tion variables, with several variables included more than
    once. The specific value for each cut after stage 2 was chosen
    to retain
    e
    80% of the signal. At each stage, given this
    constraint on signal efficiency, the same cut was made on
    data for the variable with the largest rejection power,
    R
    ¼
    ?
    sig
    =?
    bgr
    , where
    ?
    ¼
    N
    pass
    =
    N
    0
    , and
    N
    0
    and
    N
    pass
    are the
    numbers of events before and after the application of the
    selection cut, respectively. The signal-to-noise ratio was
    then computed as a function of zenith angle to ensure that
    the acceptance of AMANDA-B10 remained as large as pos-
    sible near the horizon. The effective areas for detection of
    background and signal needed for this computation were
    determined from simulations (as described in
    x
    7). After
    each stage, this procedure was repeated on the remaining
    variables.
    Besides restrictions on the reconstructed zenith angle
    h
    ,
    the most effective selection criteria impose a threshold on
    the number
    N
    dir
    of only slightly scattered, or ‘‘ direct,’’ pho-
    tons (i.e., photons that travel between the reconstructed
    track and the OM in nearly a straight line) and the track
    length
    L
    dir
    over which these photons are detected. Further-
    more, the analysis requires a minimum goodness-of-fit value
    from the maximum likelihood procedure. Other criteria
    evaluate the topological distribution of the photon emission
    using variables that describe the granularity of the light pat-
    tern along the trajectory and a related observable that
    assesses the sphericity of the photon pattern.
    31
    Table 1
    shows the selection variables and cuts used in this analysis,
    including a brief technical description of the two filtering
    stages. The selection variables were introduced previously
    (Ahrens et al. 2002a), and a complete description is also
    available (Young 2001). Also shown in the table are efficien-
    cies and rejection factors at each stage of the analysis for
    experimental data, simulated background, and simulated
    signal, averaged over all angles.
    31
    Muons normally generate a linear distribution of Cerenkov
    photons, whereas electromagnetic cascades initiated by pair production or
    bremsstrahlung produce spherically symmetric distributions.
    N
    ch
    N
    μ
    0
    5
    10
    15
    20
    25
    30
    0 20 40 60 80 100 120
    Fig.
    4.
    —Muon multiplicity
    N
    l
    vs. OM multiplicity
    N
    ch
    from a full
    detector simulation. The area of the boxes is linearly proportional to num-
    ber of events. The average values (
    dots
    ) show the correlation between these
    two quantities, and the vertical error bars show the statistical uncertainty.
    No. 2, 2003 SEARCH FOR HIGH-ENERGY NEUTRINO SOURCES 1045

    The simulated background from atmospheric muons and
    neutrinos is compared to data at all stages of the analysis to
    establish confidence in the simulation. Figure 5 shows the
    comparison at stage 4, which is sufficiently early in the anal-
    ysis to retain high statistical precision. All distributions are
    equally normalized because the optimized cut value depends
    on the fraction of events that remain (
    bottom panel
    ).
    Figure 6 shows the final stage (13) of the analysis procedure.
    We note that the signal sensitivity determined from simula-
    tions is quite robust against small deviations between the
    simulated event distribution and the actual response of the
    detector (Young 2001).
    Because of the large experimental data sample, precision
    studies of detector performance are possible from the most
    common events in the sample to extremely rare compo-
    nents. The predicted and experimental sample sizes are com-
    pared through all stages of the analysis to establish the
    absolute calibration of detector sensitivity. This method
    assumes that signal from astronomical sources contributes
    negligibly to the sample. Figure 7 shows that the simulated
    background and data agree to within a factor of 2 at all
    stages of the analysis even though the size of the event sam-
    ples varies by 6 orders of magnitude. The relative rates and
    the agreements in shape of the selection variable distribu-
    tions provide evidence that the background generators and
    detector simulation programs are an adequate description
    of the detector physics, including detector response.
    Because of the optimization on signal to noise, all stages of
    the analysis produce event samples that are dominated by
    poorly reconstructed downgoing atmospheric muon events.
    Diffuse backgrounds from atmospheric neutrinos become
    noticeable only after rejecting most of the poorly recon-
    structed atmospheric muons, but they never dominate the
    event sample.
    Obviously, atmospheric muon data are an imperfect tool
    to investigate the sensitivity of the detector to neutrino-
    induced muons, because of the downgoing nature of the
    events and differences in the energy and multiplicity distri-
    butions. This concern is addressed by the measurement of
    atmospheric neutrinos (Andres et al. 2001; Ahrens et al.
    2002a), which were used to verify the basic operational sen-
    sitivity of AMANDA-B10 to a known neutrino signal. In
    TABLE 1
    Description of Variables Used in Analysis
    Stage Selection Cut
    ?
    data
    ?
    bgr
    ?
    sig
    R
    data
    R
    bgr
    0..................... Trigger 1 1 1
    . . . ...
    1..................... Filter 1: 0.0193 0.0190 0.433 22.4 22.8
    1a...................
    ?
    ð
    1
    Þ
    >
    50
    ?
    ... . . . ... . . . ...
    1b...................
    ?
    ð
    2
    Þ
    >
    80
    ?
    ... . . . ... . . . ...
    1c ...................
    N
    ð
    2
    Þ
    dir
    ð
    b
    Þ
    >
    2
    ... . . . ... . . . ...
    2..................... Filter 2: 0.0232 0.0283 0.538 23.2 19.0
    2a...................
    ?
    ð
    2
    Þ
    >
    90
    ?
    ... . . . ... . . . ...
    2b...................
    ?
    0
    :
    43
    <
    S
    ð
    2
    Þ
    mrl
    <
    0
    :
    3
    ... . . . ... . . . ...
    2c ...................
    L
    ð
    2
    Þ
    dir
    ð
    c
    Þ
    >
    75 m
    ... . . . ... . . . ...
    2d...................
    L
    ð
    2
    Þ
    =
    L
    ð
    1
    Þ
    <
    4
    ?
    10
    ?
    6
    ... . . . ... . . . ...
    2e ...................
    ?
    ð
    5
    Þ
    >
    90
    ?
    ... . . . ... . . . ...
    3..................... cos
    ?
    ð
    5
    Þ
    <
    ?
    0
    :
    1 0.867 0.861 0.925 1.07 1.07
    4.....................
    P
    ð
    5
    Þ
    up
    =
    P
    ð
    5
    Þ
    down
    >
    9
    :
    2 0.212 0.178 0.817 3.85 4.59
    5.....................
    L
    ð
    5
    Þ
    =
    L
    ð
    4
    Þ
    <
    1
    :
    02 0.370 0.334 0.947 2.56 2.84
    6.....................
    ?
    0
    :
    21
    <
    S
    ð
    3
    Þ
    Phit
    <
    0
    :
    33 0.458 0.408 0.927 2.02 2.27
    7.....................
    L
    ð
    5
    Þ
    dir
    ð
    c
    Þ
    >
    100 m 0.638 0.657 0.932 1.46 1.42
    8.....................
    N
    ð
    5
    Þ
    dir
    ð
    c
    Þ
    ?
    N
    ð
    4
    Þ
    dir
    ð
    c
    Þ
    >
    3 0.737 0.710 0.912 1.24 1.28
    9.....................
    ?
    0
    :
    25
    <
    S
    ð
    5
    Þ
    mrl
    <
    0
    :
    26 0.711 0.671 0.912 1.28 1.36
    10...................
    L
    ð
    5
    Þ
    dir
    ð
    b
    Þ
    >
    40 m 0.744 0.698 0.955 1.28 1.37
    11...................
    P
    ð
    5
    Þ
    =
    P
    ð
    4
    Þ
    >
    9
    :
    2 0.581 0.562 0.921 1.59 1.64
    12...................
    L
    ð
    3
    Þ
    <
    4
    :
    9 0.515 0.466 0.906 1.76 1.94
    13...................
    N
    ð
    5
    Þ
    dir
    ð
    c
    Þ
    >
    9 0.659 0.616 0.963 1.46 1.56
    Notes.
    —Description of selection criteria applied to reconstruction variables in the data
    reduction procedure. For additional information on the filters and selection variables, consult
    Young (2001). The numerical identification (superscript in parentheses) refers to the reconstruc-
    tion algorithm of the event: (1) line fit used as first guess for likelihood fit; (2) maximum likelihood
    method for muon track; (3) hit probability reconstruction based on radial distribution of OMs
    that detect photons; (4) maximum likelihood method assuming cascade event; (5) iterative applica-
    tion of maximum likelihood for muon track. Direct hits are photons that arrive within (b)
    [
    ?
    15, +25] ns or (c) [
    ?
    15, +75] ns of the unscattered time of flight between track and optical mod-
    ule. The reduced likelihood parameter
    L
    is
    ?
    log
    ð
    P
    Þ
    divided by the number of degrees of freedom,
    where
    P
    is the maximized probability. Passing efficiencies (
    ?
    ) relative to the prior stage are shown
    at each stage for experimental data, simulated background, and simulated signal. Rejection
    factors (
    R
    ) for experimental data and simulated background are shown for each stage in the two
    columns farthest to the right.
    1046 AHRENS ET AL. Vol. 583

    this analysis, the relative agreement between the measured
    and predicted event rates is 30%, which is consistent with
    uncertainties in the measured flux of cosmic-ray primaries,
    theoretical uncertainty in the interaction models, and sys-
    tematic uncertainties in the modeling of the detector
    response. However, because of the steeply falling energy
    spectrum for atmospheric neutrinos, the mean energy of the
    muons induced by charged-current interactions is close to
    the energy threshold of the detector, which implies that they
    cannot be used to reliably probe the high-energy response of
    AMANDA-B10.
    6. POINTING RESOLUTION AND
    POINT-SOURCE SEARCH
    The final stage of the data analysis procedure yields a
    sample of 815 events (as is evident from Fig. 7, atmospheric
    neutrinos contribute about 25% of the events to the simu-
    lated background). Visual inspection of the distribution in
    the sky of the final event sample, shown in Figure 8, reveals
    no obvious clustering. The increase in event density near the
    horizon (decl
    :
    ¼
    0
    ?
    ) is due to the zenith dependence of the
    atmospheric muon component of the background. In order
    to perform a quantitative search for possible sources of
    high-energy neutrinos in the northern hemisphere, the sky
    was divided into nonoverlapping angular bins of approxi-
    mately equal solid-angle coverage. A point source would
    then be revealed by a statistically significant clustering of
    events within a particular angular bin. The optimal bin size
    and shape depend on the space-angle resolution of the
    detector, which can be expressed in terms of a point-spread
    function. The space-angle deviation
    ?
    between the true
    0
    0.02
    0.04
    0.06
    0.08
    0.1
    0.12
    0.14
    0.16
    5 1015 20253035
    data
    signal
    background
    N
    (5)
    dir(c)
    Normalized Units
    Efficiency
    N
    (5)
    dir(c)
    0
    0.2
    0.4
    0.6
    0.8
    1
    1.2
    5 1015 20253035
    Fig.
    6.
    Top
    : Similar to Fig. 5, but for analysis stage 13, which involves
    the number of direct hits in a time window of
    ?
    15 to +75 ns.
    Bottom
    :
    Passing efficiencies—defined as integrated sums, from given value to
    infinity, of distributions shown in the top panel—as a function of cut value.
    The vertical lines indicate the cut applied in the analysis. [
    See the electronic
    edition of the Journal for a color version of this figure.
    ]
    Analysis Stage
    No. of events
    1997 AMANDA­B10 data
    atmospheric muons
    atmospheric neutrinos
    total background
    10
    2
    10
    3
    10
    4
    10
    5
    10
    6
    10
    7
    10
    8
    10
    9
    01 2 3 45 67 89 10 11 12 13
    Fig.
    7.
    —Number of events remaining in the sample as the selection
    criteria in the 13 analysis stages listed in Table 1 are applied sequentially.
    The 1997 AMANDA-B10 data (
    circles
    ) are compared to simulated back-
    ground from atmospheric muons generated by cosmic-ray interactions
    (
    squares
    ) and from muons induced by atmospheric neutrinos (
    asterisks
    ).
    Also shown is the sum of atmospheric muon and neutrino backgrounds
    (
    triangles
    ). No normalization was applied, but systematic uncertainty at the
    trigger level (analysis stage 0) may be as large as
    ?
    30%. [
    See the electronic
    edition of the Journal for a color version of this figure.
    ]
    0
    0.02
    0.04
    0.06
    0.08
    0.1
    0.12
    0.14
    0.16
    0.18
    0 5 10 15 20 25 30
    data
    signal
    background
    P
    (5)
    up
    /P
    (5)
    down
    Normalized Units
    Efficiency
    P
    (5)
    up
    /P
    (5)
    down
    0
    0.2
    0.4
    0.6
    0.8
    1
    1.2
    0 5 10 15 20 25 30
    Fig.
    5.
    Top
    : Equally normalized distributions for experimental data
    (
    circles
    ) and simulated background (
    asterisks
    ) and signal (
    triangles
    ) for
    stage 4 of the point-source analysis, which compares the best likelihood for
    an upgoing track hypothesis with the likelihood for a track in the opposite
    (i.e., downgoing) direction.
    Bottom
    : Passing efficiencies as a function of cut
    value. The efficiency is obtained from the integrated sums of distributions
    shown in the top panel from given value to infinity. The vertical lines
    indicate the cut applied in the analysis. [
    See the electronic edition of the Jour-
    nal for a color version of this figure.
    ]
    No. 2, 2003 SEARCH FOR HIGH-ENERGY NEUTRINO SOURCES 1047

    angular coordinates of a muon (
    ?
    true
    ;
    ?
    true
    ) and the recon-
    structed coordinates (
    ?
    rec
    ;
    ?
    rec
    ) is given by
    cos
    ?
    ¼
    cos
    ?
    rec
    cos
    ?
    true
    þ
    sin
    ?
    rec
    sin
    ?
    true
    cos
    ð
    ?
    rec
    ?
    ?
    true
    Þ
    :
    ð
    1
    Þ
    Figure 9 shows the distribution of
    ?
    and the corresponding
    point-spread function, computed from the
    ?
    -distribution
    by dividing by the appropriate solid angle, for the simulated
    sample of upward-traveling muons generated by neutrinos
    with an
    E
    ?
    2
    energy spectrum. A median value of
    ?
    ¼
    3
    =
    9,
    averaged over positive declinations, is achieved. A function
    involving the sum of two Gaussian distributions was fitted
    to the point-spread function. It yields an amplitude ratio of
    A
    2
    =
    A
    1
    ¼
    0
    :
    25, indicating the importance of the second com-
    ponent related to the degrading angular resolution at large
    muon energies (Young 2001). Given this point-spread func-
    tion and the relatively small number of background events
    shown in Figure 8, the optimal slice in zenith angle is 11
    =
    25
    (Young 2001). For azimuth angle, a weak optimum occurs
    for a width of 12
    ?
    for the declination band closest to the
    horizon. These angular dimensions of the bins were chosen
    to maximize the signal to noise.
    Two studies were performed to check the predicted space-
    angle resolution and absolute pointing accuracy. The first
    uses AMANDA events that were also tagged by the GASP
    air Cerenkov telescope (Barbagli et al. 1993). GASP deter-
    mines the direction of the air shower, and AMANDA meas-
    ures the direction of the penetrating muon component. At
    AMANDA depths, these events are almost entirely single
    muons. Unfortunately, the duty cycle of operation is low, so
    the sample size is relatively small. To improve the statistical
    accuracy of the angular resolution studies, a second method
    based on extensive air showers was developed. This method
    utilized events that triggered both the SPASE array and
    AMANDA (X. Bai et al. 2003, in preparation). SPASE
    responds to the electron and photon content of the shower
    front that reaches the surface. Since the direction of muons
    within the air shower event is nearly perpendicular to the
    shower front, the difference between the direction of the air
    shower and the reconstructed muon direction can be used to
    deduce the angular resolution of AMANDA. SPASE meas-
    ures the direction of an air shower with a pointing resolution
    of approximately 1
    ?
    –2
    ?
    (Dickinson et al. 2000; depending
    on shower size), which is small enough to calibrate the
    AMANDA pointing resolution.
    In this study, AMANDA data were analyzed using the
    procedure outlined in Table 1, with the exception that
    angle-dependent cuts were inverted to account for the
    downgoing direction of travel of SPASE/AMANDA coin-
    cidence events (e.g., the cut at stage 3 was changed to
    cos
    ?
    ð
    5
    Þ
    >
    0
    :
    1). The absolute pointing accuracy is character-
    ized by the average of
    D
    ?
    , the difference between the true
    and the reconstructed zenith angle. Because of the excellent
    zenith-angle resolution of SPASE, SPASE/AMANDA
    coincidence data were used to deduce
    D
    ?
    using the
    reconstructed zenith angles of both detectors,
    D
    ?
    ¼
    ?
    AMANDA
    ?
    ?
    SPASE
    . Figure 10 shows the measured zenith-
    angle resolution using SPASE/AMANDA coincidence
    events, together with the resolution obtained for a SPASE/
    AMANDA simulation of air showers initiated by protons
    and iron nuclei. Iron primaries produce a larger fraction of
    coincidence events with more than one muon penetrating to
    AMANDA depths, which accounts for the small difference
    between protons and iron nuclei. The coincidence data sup-
    port the predicted angular resolution and show that the
    angular offset is small compared to the angular dimensions
    of the sky bins. These conclusions are nearly independent of
    the choice of cosmic-ray primary.
    Also shown in Figure 10 is the expected angular resolu-
    tion as function of declination for single muons with ener-
    gies of 0.1 and 4 TeV within the detector volume. These
    muon energies were chosen to be representative of the aver-
    age muon energy that were initiated by atmospheric neutri-
    nos and by a source with a differential energy spectrum
    proportional to
    E
    ?
    2
    . The predicted value for the absolute
    offset is less than 1
    =
    5, which is consistent with results
    obtained by additional study of the SPASE/AMANDA
    8
    4
    9
    1
    3
    2
    6
    5
    7
    10
    1
    ­ Mkn 501
    2
    ­ Mkn 421
    3
    ­ NGC4151
    4
    ­ 1ES2344
    5
    ­ 3C66A
    6
    ­ 1ES1959+650
    7
    ­ Crab Nebula
    8
    ­ Cassiopeia A
    9
    ­ Cygnus X­3
    10
    ­ Geminga
    Fig.
    8.
    —Sky plot of 815 events obtained from the point-source analysis.
    Horizontal coordinates are right ascension, and vertical coordinates are
    declination. Also shown are the sky coordinates for 10 potential high-
    energy neutrino sources. [
    See the electronic edition of the Journal for a color
    version of this figure.
    ]
    0
    0.05
    0.1
    0.15
    0.2
    0.25
    0 2468 10 12 14 16 18 20
    mean
    RMS
    median
    = 5.00
    °
    = 3.82
    °
    = 3.94
    °
    Ψ
    [degrees]
    Arbitrary Units
    Ψ
    [degrees]
    Point spread function
    A
    1
    = 0.33
    σ
    1
    = 1.90
    °
    A
    2
    = 0.08
    σ
    2
    = 4.45
    °
    0
    0.1
    0.2
    0.3
    0.4
    0.5
    0.6
    0.7
    0 2468 10 12 14 16 18 20
    Fig.
    9.
    Top
    : Distribution of space angle
    ?
    between true and recon-
    structed muon track direction for simulated signal with neutrino energy
    spectrum proportional to
    E
    ?
    2
    , averaged over direction.
    Bottom
    : Point-
    spread function for signal in AMANDA-B10, deduced from the space-
    angle distribution in the top panel. A function composed of the sum of two
    Gaussians was used to characterize this distribution.
    1048 AHRENS ET AL. Vol. 583

    and GASP/AMANDA coincidence events (Rawlins 2001;
    X. Bai et al. 2003, in preparation).
    32
    The offset is due to bias
    in the AMANDA event reconstruction, which tends to pro-
    duce more vertical events. This effect is most evident from
    the high-energy muon simulation, which shows that the
    angular offset changes sign for positive and negative decli-
    nations. Since the absolute offset is substantially smaller
    than the angular resolution and small compared to the size
    of the search bin, the offset has minimal impact on the signal
    efficiency. For a zenith (declination) offset of 1
    =
    5, 6% of the
    signal events are shifted to the neighboring bin. The bottom
    panel of the figure also shows that the angular resolution for
    upward-traveling events is slightly better than for events
    traveling in the downward direction, presumably because of
    the asymmetry in the response of the photomultiplier tubes,
    which are oriented toward the center of the Earth.
    To obtain approximately equal solid-angle coverage for
    all bins, the northern sky is divided into 154 nonoverlapping
    bins, using the calculated optimal declination slice (11
    =
    25)
    and a varying number of bins in azimuth for the resulting
    eight declination bands—from 30 near the horizon to three
    near +90
    ?
    declination. Each angular bin is then tested for
    an excess of events by computing the significance,
    ?
    ¼?
    log
    10
    ð
    P
    Þ
    ;
    ð
    2
    Þ
    where
    P
    ¼
    X
    1
    n
    ¼
    N
    0
    e
    ?
    l
    l
    n
    n
    !
    ð
    3
    Þ
    is the probability of the bin containing at least the
    observed number of events
    N
    0
    , assuming that fluctuations
    are described by a Poisson distribution. The expected
    mean number of events
    l
    is obtained by taking the aver-
    age of the number of events in all other bins in the same
    declination slice. The polar location of AMANDA
    assures equal sky coverage for all declinations, independ-
    ent of time gaps in the collection of data. Figure 11
    shows the distribution of significance for the experimental
    data and for random fluctuation of the background
    events. This noise estimate is obtained by randomizing
    the right ascension coordinate of the data events, then
    recalculating
    ?
    for each bin. A point-source candidate is
    identified by a large observed value of significance with a
    large ratio to the significance expected from random fluc-
    tuations of background. To avoid the statistical problem
    of a potential source near a bin boundary distributing
    signal between two adjacent bins, the procedure was
    repeated with the grid shifted by one-half of a bin in
    both declination and azimuth. The largest value of signif-
    icance,
    ?
    ¼
    1
    :
    85, appears in the bottom panel. Taking
    into account the 154 bins in the sky and the two versions
    of the sky grid, there is a 40% probability that the most
    significant sky bin is produced by random fluctuation of
    background. Therefore, the distribution of significance
    shows no evidence of a source.
    Another approach was also investigated, using the angu-
    lar correlation function between event pairs to avoid the
    problem of a source near a bin boundary, but this alternate
    approach did not reveal sources either (Young 2001).
    32
    Available at http://area51.berkeley.edu/manuscripts/
    20011002xx-kaths_thesis.pdf.
    ­3
    ­2
    ­1
    0
    1
    2
    3
    4
    ­80 ­60 ­40 ­20 0 20 40 60 80
    signal (
    E
    μ
    = 4 TeV)
    signal (
    E
    μ
    = 100 GeV)
    SPASE/AMANDA coincidence data
    proton simulation
    iron simulation
    Declination [degrees]
    Δθ
    [degrees]
    Declination [degrees]
    Ψ
    median
    [degrees]
    0
    1
    2
    3
    4
    5
    6
    7
    8
    ­80 ­60 ­40 ­20 0 20 40 60 80
    Fig.
    10.
    —Offset in reconstructed zenith angle (
    top
    ) and median space
    angle (
    bottom
    ) as a function of declination. Positive declination
    corresponds to upward-traveling events in the AMANDA array. SPASE/
    AMANDA coincidence data (
    squares
    ) are compared to expectation assum-
    ing that the cosmic-ray elemental composition is entirely protons (
    circles
    )
    or iron nuclei (
    triangles
    ). Also shown is the expectation for signal (i.e.,
    neutrino-induced muons) with two different energies within the detector.
    [
    See the electronic edition of the Journal for a color version of this figure.
    ]
    10
    ­2
    10
    ­1
    1
    10
    10
    2
    0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
    Significance (
    ξ
    )
    No. of bins
    data
    background
    nominal grid
    shifted grid
    Significance (
    ξ
    )
    No. of bins
    10
    ­2
    10
    ­1
    1
    10
    10
    2
    0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
    Fig.
    11.
    —Distribution of significance for the 154 sky bins. Data
    (
    squares
    ) are compared to expectation from randomized background
    (
    circles
    ). The statistical uncertainty of the randomized background is negli-
    gible. The top panel shows the results for the nominal sky grid. The sky grid
    used in the bottom panel has been shifted from the nominal sky grid by
    one-half of a bin in both right ascension and declination. [
    See the electronic
    edition of the Journal for a color version of this figure.
    ]
    No. 2, 2003 SEARCH FOR HIGH-ENERGY NEUTRINO SOURCES 1049

    7. FLUX LIMITS
    The absence of a detected source translates into an upper
    limit on the high-energy neutrino flux. Neutrino flux and
    neutrino-induced muon flux limits depend on the effective
    area of the detector
    A
    eff
    for a muon with energy
    E
    l
    . The
    effective area, obtained by dividing the signal rate by the
    incident flux, is determined from simulations by
    A
    eff
    ð
    E
    l
    Þ¼
    f
    ev
    ð
    E
    l
    Þ
    A
    GEN
    ;
    ð
    4
    Þ
    where
    A
    GEN
    is the cross-sectional area of the cylinder in the
    simulation that contains all neutrino interaction vertices
    and
    f
    ev
    is the fraction of generated muon events that survive
    the 13-stage data analysis procedure. As an example, results
    are shown in Figure 12 for muon vertices located near the
    detector. The effective area is computed as a function of
    declination for muons with energies of 1, 10, and 100 TeV.
    These effective areas can be contrasted to the geometric
    cross section of the instrumented volume of deep ice, which
    spans from 1
    :
    1
    ?
    10
    4
    m
    2
    in the vertical direction to
    6
    :
    1
    ?
    10
    4
    m
    2
    in the horizontal direction.
    The
    average
    effective area, folding in the source spectra,
    detector response, and the probability of a neutrino-nucleon
    interaction, is given by
    A
    i
    eff
    ¼
    N
    A
    ?
    ice
    A
    GEN
    "
    Z
    E
    max
    ?
    E
    min
    ?
    ?
    ?
    l
    ð
    E
    ?
    Þ
    S
    ð
    E
    ?
    Þ
    ?h
    R
    ð
    E
    ?
    ;
    E
    min
    l
    Þi
    f
    ev
    d
    ?
    ?
    dE
    ?
    dE
    ?
    #
    ?
    Z
    E
    max
    ?
    E
    min
    ?
    d
    ?
    i
    dE
    ?
    dE
    ?
    !
    ?
    1
    ;
    ð
    5
    Þ
    where the index
    i
    denotes either
    l
    or
    ?
    ,and
    E
    min
    ?
    ¼
    10 GeV
    and
    E
    max
    ?
    ¼
    10
    7
    GeV define the energy range in the simula-
    tion, which safely brackets the interval of interest for most
    theoretical models. The other variables are
    N
    A
    , Avogadro’s
    number,
    ?
    ice
    , the molar density of nucleons in ice, and
    ?
    ?
    l
    ,
    the charged-current cross section for
    ?
    l
    (or
    ?
    l
    ) interactions.
    The term
    S
    ð
    E
    ?
    Þ
    accounts for neutrino absorption in the
    Earth, and
    h
    R
    ð
    E
    ?
    ;
    E
    min
    l
    Þi
    is the average propagation length
    for a muon created by a neutrino with energy
    E
    ?
    , corrected
    for the energy threshold of the detector. In this calculation,
    the neutrino interaction vertices are located randomly
    within a volume that is large compared to the instrumented
    volume of the detector. The propagation programs properly
    account for muon energy losses.
    The muon differential flux is related to the neutrino differ-
    ential flux by
    d
    ?
    l
    dE
    ?
    ¼
    ?
    ?
    l
    ð
    E
    ?
    Þ
    S
    ð
    E
    ?
    Þh
    R
    ð
    E
    ?
    ;
    E
    min
    l
    Þi
    d
    ?
    ?
    dE
    ?
    :
    ð
    6
    Þ
    The energy-averaged muon effective area is presented in
    Figure 13 for different spectral indexes. The energy response
    of AMANDA is described by the distribution of
    E
    l
    at the
    detector, which depends on the spectral index of the neu-
    trino spectrum. Figure 14 shows the distribution for differ-
    ential spectra proportional to
    E
    ?
    2
    and
    E
    ?
    3
    . For
    E
    ?
    2
    , the
    most probable muon energy (mode) is 5 TeV, and the cen-
    tral 90% of the muon events are within the energy interval
    80 GeV to 200 TeV. For a spectral index of 3, appropriate
    for atmospheric neutrinos, the most probable detected
    energy of the muon is much lower.
    If neutrino emission from a point source is described by a
    differential energy spectrum, the energy-averaged flux limit
    is calculated from
    ?
    limit
    i
    ¼
    l
    s
    N
    0
    ;
    N
    bgr
    ??
    T
    live
    ?
    bin
    A
    i
    eff
    :
    ð
    7
    Þ
    Declination [degrees]
    A
    μ
    eff
    [m
    2
    ]
    E
    μ
    = 100 TeV
    E
    μ
    = 10 TeV
    E
    μ
    = 1 TeV
    0
    10000
    20000
    30000
    40000
    50000
    0 102030405060708090
    Fig.
    12.
    —AMANDA-B10 effective area for muon detection as a func-
    tion of declination (90
    ?
    is vertically up) for three different muon energies at
    the detector. The vertical error bars are statistical. [
    See the electronic edition
    of the Journal for a color version of this figure.
    ]
    E
    ν
    ­2
    E
    ν
    ­2.3
    E
    ν
    ­2.5
    E
    ν
    ­2.7
    E
    ν
    ­3
    Input spectrum:
    Declination [degrees]
    A
    μ
    eff
    [m
    2
    ]
    0
    2000
    4000
    6000
    8000
    10000
    12000
    0 102030 40506070 8090
    Fig.
    13.
    —AMANDA-B10 average effective area for muon detection as
    a function of declination (90
    ?
    is vertically up) for assumed differential
    neutrino spectral indexes between 2 and 3. The vertical error bars indicate
    the uncertainty obtained from studies described in
    x
    8. [
    See the electronic
    edition of the Journal for a color version of this figure.
    ]
    1050 AHRENS ET AL. Vol. 583

    The quantity
    l
    s
    ð
    N
    0
    ;
    N
    bgr
    Þ
    is the upper limit on the number
    of signal events at the 90% confidence level, calculated fol-
    lowing the unified procedure of Feldman & Cousins (1998),
    where
    N
    0
    is the observed number of events in a potential
    source bin and
    N
    bgr
    is the expected number of background
    events. For the binning technique employed in our point-
    source search,
    N
    bgr
    is determined by averaging the number
    of observed events over the declination band, excluding the
    bin being considered. The efficiency factor
    ?
    bin
    accounts for
    the finite angular resolution and the possible noncentral
    location within the bin of a potential source. The factor
    T
    live
    is the operational live time of the detector. The resulting
    muon and neutrino flux limits are shown in Figures 15 and
    16, respectively, for various assumed spectral indexes
    between 2 and 3.
    Figure 17 shows a comparison of the AMANDA flux lim-
    its with a representative sample of neutrino detectors
    located in the Northern Hemisphere. The AMANDA-B10
    detector approaches its maximum sensitivity for declina-
    tions greater than +30
    ?
    , which complements the sky regions
    covered by neutrino detectors such as MACRO (Montaruli
    et al. 1999; Ambrosio et al. 2001; Perrone et al. 2001) and
    Super-Kamiokande (Matsuno et al. 2001).
    33
    With only
    130.1 days of detector live time, the muon flux limits for pos-
    itive declinations approach those achieved for the southern
    sky. The AMANDA flux limits were calculated for
    E
    ?
    >
    10
    GeV, while both Super-Kamiokande and MACRO present
    fluxes for
    E
    l
    >
    1–2 GeV, but for relatively hard differential
    neutrino spectra, such as
    E
    ?
    2
    ?
    , the impact of energy thresh-
    old on muon flux limits is modest (Biron 2002).
    34
    In addition to the general search for a point source, a
    number of potential sources of particular interest were
    investigated by performing the significance test while cen-
    tering the search bin on their sky coordinates. The result-
    ing flux limits are presented in Table 2. As one
    interesting example, we compare the AMANDA limit on
    log
    10
    (E
    μ
    /GeV)
    Normalized Units
    E
    ν
    ­2
    input spectrum
    E
    ν
    ­3
    input spectrum
    mode
    E
    μ
    = 5 TeV
    mode
    E
    μ
    = 80 GeV
    0
    0.02
    0.04
    0.06
    0.08
    0.1
    0.12
    12 345 67
    Fig.
    14.
    —Neutrino-induced muon energy distribution at the detector.
    The differential neutrino spectra used as input for the simulation are pro-
    portional to
    E
    ?
    2
    (
    solid curve
    ) and
    E
    ?
    3
    (
    dashed curve
    ), respectively. The
    most probable energy for each distribution is shown.
    E
    ν
    ­2
    input spectrum
    E
    ν
    ­2.3
    input spectrum
    E
    ν
    ­2.5
    input spectrum
    E
    ν
    ­2.7
    input spectrum
    E
    ν
    ­3
    input spectrum
    Declination [degrees]
    Φ
    ν
    limit
    [cm
    ­2
    s
    ­1
    ]
    10
    ­7
    10
    ­6
    10
    ­5
    10
    ­4
    10
    ­3
    0 102030 405060 708090
    Fig.
    15.
    —Neutrino flux limit (90% CL) for various spectral indexes. The
    results are shown as a function of declination, averaged over right
    ascension. Power-law exponent refers to the differential neutrino energy
    spectrum. The vertical error bars indicate the uncertainty obtained from
    systematic studies [
    See the electronic edition of the Journal for a color version
    of this figure.
    ]
    E
    ν
    ­2
    input spectrum
    E
    ν
    ­2.3
    input spectrum
    E
    ν
    ­2.5
    input spectrum
    E
    ν
    ­2.7
    input spectrum
    E
    ν
    ­3
    input spectrum
    Declination [degrees]
    Φ
    μ
    limit
    [cm
    ­2
    s
    ­1
    ]
    10
    ­14
    10
    ­13
    10
    ­12
    0 102030 405060 708090
    Fig.
    16.
    —Neutrino-induced muon flux limit (90% CL) for various
    spectral indexes. The results are shown as a function of declination, aver-
    aged over right ascension. Note that the power-law exponent refers to the
    differential neutrino energy spectrum. [
    See the electronic edition of the Jour-
    nal for a color version of this figure.
    ]
    33
    Available at http://www.copernicus.org/icrc/papers/ici7384_p.pdf.
    34
    Available at http://area51.berkeley.edu/manuscripts.
    No. 2, 2003 SEARCH FOR HIGH-ENERGY NEUTRINO SOURCES 1051

    the neutrino flux from Mrk 501 to the observed gamma-
    ray flux and this flux corrected for intergalactic absorp-
    tion by infrared photons in Figure 18. Assuming that the
    neutrino energy spectrum is proportional to the inferred
    gamma-ray spectrum at the source, the AMANDA limit
    constrains the proportionality factor. For example, the
    ratio of the flux of neutrinos to the flux of gamma rays
    must be less than 10 if the source spectrum of Konopelko
    et al. (1999) is assumed. Recent work (de Jager & Stecker
    2002) suggests that the ratio may be smaller.
    8. IMPACT OF SYSTEMATIC UNCERTAINTIES
    ON FLUX LIMITS
    In the absence of a well-understood source of high-energy
    neutrinos, the sensitivity of the AMANDA-B10 detector, as
    expressed in terms of the effective area and angular resolu-
    tion, had to be estimated from detector simulations. The
    required input relies on knowledge of detector performance
    extracted from, e.g., laboratory measurements of the indi-
    vidual components, in situ measurements of the optical
     
     
    Fig.
    17.
    —Upper limit on the muon flux (90% CL) as a function of decli-
    nation. The solid curve is the AMANDA-B10 limit, averaged over right
    ascension, and the region delineated by the long-dashed curves provides a
    guide to the statistical fluctuation within the declination interval (see Table
    4). The band defined by the short-dashed curves indicates the range of limits
    presented by MACRO (Montaruli et al. 1999; Ambrosio et al. 2001;
    Perrone et al. 2001) and Super-Kamiokande (Matsuno et al. 2001). [
    See the
    electronic edition of the Journal for a color version of this figure.
    ]
    TABLE 2
    Muon and Neutrino Flux Limits on Selected Point Sources
    Source Model
    N
    0
    N
    bgr
    D
    E
    (TeV)
    ?
    limit
    ?
    (10
    ?
    8
    cm
    ?
    2
    s
    ?
    1
    )
    ?
    limit
    l
    (10
    ?
    15
    cm
    ?
    2
    s
    ?
    1
    )
    Mrk 501 ....................... 1 7 3.5 0.3–20 86.0 38.9
    Mrk 501 ....................... 2 7 3.5 1–1000 9.5 14.6
    Mrk 421 ....................... 3 4 3.7 1–1000 11.2 9.7
    NGC 4151.................... 3 5 3.6 1–1000 12.9 10.9
    NGC 4151.................... 4 5 3.6 60–2500 0.0042 5.6
    1ES 2344...................... 5 5 2.9 1–400 12.5 10.3
    3C 66A......................... 5 3 3.5 0.8–250 7.2 6.6
    1ES 1959+650.............. 5 4 1.7 0.8–250 13.2 9.7
    Crab Nebula ................ 5 2 5.6 1–1000 4.2 5.0
    Cas A........................... 5 3 2.2 1.8–1000 9.8 7.6
    Cyg X-3 ....................... 5 2 3.4 1–1000 4.9 4.6
    Geminga ...................... 5 4 7.1 1.8–1000 6.8 9.1
    Notes.
    —Muon and neutrino flux limits on selected sources for
    E
    ?
    >
    10 GeV. The term
    N
    0
    is the number
    of observed events in the search bin, and
    N
    bgr
    is the expected background. The energy interval
    D
    E
    contains
    90% of the neutrino events, and the flux limits are corrected for systematic uncertainty (see
    x
    8). Representa-
    tive survey of models (second column): (1) neutrino spectrum identical to measured photon spectrum
    (Aharonian et al. 1999); (2)
    d
    ?
    ?
    =
    dE
    /
    E
    ?
    1
    :
    92
    ?
    ; (3) Szabo & Protheroe 1992; (4) Stecker et al. 1991; (5)
    d
    ?
    ?
    =
    dE
    /
    E
    ?
    2
    ?
    .
    log
    10
    (
    E
    γ
    / TeV)
    log
    10
    (
    E
    2
    d
    N
    /d
    E
    / TeV m
    ­2
    s
    ­1
    )
    Markarian 501,
    z
    = 0.0336
    AMANDA flux limit
    observed gamma flux (HEGRA)
    corrected gamma flux (Konopelko, et al.)
    ­10
    ­8
    ­6
    ­4
    ­2
    0
    ­1 ­0.5 0 0.5 1 1.5 2
    Fig.
    18.
    —Time-averaged spectrum of gamma rays from Mrk 501
    observed in 1997 (Aharonian et al. 1999; Krennrich et al. 1999) and cor-
    rected for intergalactic absorption by the diffuse infrared background
    (Konopelko et al. 1999). Gamma-ray flux is compared to the AMANDA
    neutrino limit assuming an energy dependence on the neutrino flux
    proportional to
    E
    ?
    2
    .[
    See the electronic edition of the Journal for a color
    version of this figure.
    ]
    1052 AHRENS ET AL. Vol. 583

    properties of the ice, and calibration studies. Consequently,
    the predicted sensitivity is affected by uncertainty in this
    information. Table 3 lists the dominant contributions to
    systematic uncertainties. The uncertainty in the right
    column is defined as the variation
    A
    j
    eff
    ?
    A
    nom
    eff
    Þ
    =
    ð
    A
    j
    eff
    þ
    A
    nom
    eff
    Þj
    of the effective area from its value determined
    with the nominal set of input parameters,
    A
    nom
    eff
    , given by the
    area,
    A
    j
    eff
    , obtained by varying the specified parameter
    (index
    j
    ) by its estimated uncertainty.
    The most significant component is generated from the
    uncertainty in the angular dependence of the OM sensitiv-
    ity. It arises mainly from a lack of detailed understanding of
    the physics governing the refreezing process in the water col-
    umn required to be melted for the deployment of OMs. A
    local increase in scattering from air bubbles trapped in the
    vicinity of the OM translates into a modulation of its angle-
    dependent acceptance. This effect is difficult to disentangle
    from the intrinsic angular dependence of the OM sensitivity,
    which was measured in the laboratory. An event sample
    highly enriched in atmospheric muons was used to investi-
    gate the in situ angle dependence of the OM sensitivity
    (Ahrens et al. 2002a). The modification to the angular sensi-
    tivity leads to a 25% uncertainty in the effective area.
    Since the angle-integrated sensitivity of the OM is a
    poorly constrained parameter in this analysis, we also inves-
    tigated the impact of varying the absolute sensitivity of the
    OM. It was parameterized by a wavelength-dependent func-
    tion that included the PMT quantum efficiency, the OM col-
    lection efficiency, obscuration by nearby cables, and
    absorption properties of the glass pressure vessel and cou-
    pling gel. We obtain a fractional uncertainty of 0.15 in the
    effective area after reducing the absolute OM sensitivity by
    15%, a value consistent with the atmospheric neutrino
    results. Further reduction is inconsistent with observed
    experimental trigger rates.
    As mentioned in
    x
    4, two muon propagation routines
    were employed to show that systematic variations in the
    effective area for signal (i.e., upward-traveling) muons
    were between 5% and 10%. This is much less than
    observed for studies of atmospheric muons, presumably
    because of the much weaker angular dependence of the
    average path length. Possible uncertainties in timing and
    position calibration of individual OMs are included by
    varying these parameters to the largest extent allowed by
    the imprecision of the calibration procedures. The effec-
    tive area changes by 10%. A conservative estimate of the
    variation in sensitivity introduced by uncertainties in the
    depth-dependent optical properties and their approximate
    treatment in the detector simulation is obtained by sub-
    stituting the nominal bulk ice model, containing a param-
    eterization of the measured dust strata, with a
    homogeneous ice model. The impact on effective area is
    less than 15%.
    The impact of the most dominant systematic uncertain-
    ties on the average effective area is shown in Figure 19,
    where the systematic variations have been applied one at a
    time, with the exception of the muon propagation curve,
    which also includes the variation of the angular OM
    sensitivity.
    The variation of the detector sensitivity due to systematic
    uncertainties was studied by adjusting the physical parame-
    ters in the detector simulation. The parameters were
    adjusted according to the known or estimated uncertainties
    listed in Table 3, which are assumed to bound the true val-
    ues of the parameters. Systematic uncertainty was included
    according to the prescription of Conrad et al. (2003), which
    is an extension of the method of Cousins & Highland
    (1992). The calculations assumed that the distribution of
    systematic uncertainty was flat and bounded by the maxi-
    mum and minimum values for a given declination bin,
    found in Figures 20 and 21.
    The solid curves in Figures 20 and 21 indicate the flux lim-
    its after adjusting for systematic uncertainty. They are valid
    for declination greater than +5
    ?
    . The limits including sys-
    tematic uncertainties are about 25% worse than those
    obtained from the simulation with nominal input parame-
    ters. Finally, the flux limits change by less than 6% because
    of the effects of zenith offset and the variation in
    ?
    median
    due
    to declination (Young 2001). These small effects were not
    taken into consideration in the limit calculations.
    9. DISCUSSION
    The previous sections have shown that AMANDA-B10
    has unprecedented sensitivity to high-energy neutrinos and
    possesses the necessary angular response and background
    TABLE 3
    Systematic Uncertainty in AMANDA­B10 Effective Area
    Source of Systematic Uncertainty
    Error in
    A
    l
    eff
    (%)
    Angular dependence of OM sensitivity............................
    ?
    25
    Absolute OM sensitivity..................................................
    ?
    15
    Muon propagation..........................................................
    ?
    10
    Calibration (timing and geometry)..................................
    ?
    10
    Hardware simplifications in detector simulation .............
    ?
    <
    10
    Optical properties of bulk ice ..........................................
    ?
    15
    Declination [degrees]
    A
    μ
    eff
    [m
    2
    ]
    nominal (layered ice)
    homogeneous ice
    absolute OM sensitivity
    angular OM sensitivity
    muon propagation
    0
    2000
    4000
    6000
    8000
    10000
    12000
    0 102030405060708090
    Fig.
    19.
    —Comparison of average muon effective area for differential
    neutrino signal proportional to
    E
    ?
    2
    . See text for explanation of legend.
    Statistical uncertainty is indicated by vertical lines, unless it is obscured by
    the symbol. [
    See the electronic edition of the Journal for a color version of this
    figure.
    ]
    No. 2, 2003 SEARCH FOR HIGH-ENERGY NEUTRINO SOURCES 1053

    rejection to search for point emission of these particles from
    astronomical objects; i.e., it is a novel telescope that detects
    the neutrino messenger. The sensitivity and angular
    response were determined by simulation. The reliability of
    these programs was established by utilizing the known sig-
    nals generated by (downgoing) atmospheric muons and
    (upgoing) atmospheric neutrinos. The angular response was
    confirmed by the study of air shower events that triggered
    both AMANDA-B10 and SPASE. Systematic uncertainty
    in the analysis procedure was also addressed.
    The search for point sources of high-energy neutrinos
    revealed no candidates. A set of event selection criteria was
    determined by optimizing the signal-to-noise ratio for a sig-
    nal with a hard energy spectrum, yet this analysis retains
    reasonable sensitivity for softer spectra. The upper limits on
    muon flux for all search bins in the northern hemisphere are
    presented in Table 4.
    The neutrino flux limits in Figure 15 are inferred from the
    assumption of a power-law energy spectrum. This proce-
    dure is reliable if the mean energy of the neutrino-induced
    muon is compatible with the energy response of the detec-
    tor. For example, Figure 14 shows that
    E
    l
    at the detector
    brackets the interval between 0.1 and 10
    3
    TeV for source
    spectra proportional to
    E
    ?
    2
    . Two lines of evidence show
    that the simulated energy response of the detector is valid
    over this interval. First, the agreement between the detected
    and expected rates of atmospheric neutrinos shows that the
    response of AMANDA is being correctly modeled in the
    sub-TeV region. Second, the tails of the
    N
    ch
    -distribution are
    sensitive to brighter events within AMANDA, which
    are roughly equivalent to single muons with energy above 1
    TeV. We know of no reason to doubt the predicted energy
    response for
    E
    l
    <
    10
    3
    TeV. Evaluation and calibration of
    the energy response beyond 10
    3
    TeV remain an ongoing
    activity.
    Not all model predictions are well characterized by
    power-law energy spectra. Therefore, Table 2 shows the
    results for a selection of models in the literature. The
    inferred limits on neutrino flux apply to point sources with
    continuous emission (or episodic emission averaged over
    the time interval of data collection) and power-law energy
    spectra with a fixed spectral index. The limits presented here
    for sources at large positive declination complement exist-
    ing data, so that comparable limits now exist for the entire
    sky.
    During 1997, the TeV gamma-ray emission of two
    nearby AGN blazars (Mrk 421 and 501) were observed
    to exhibit episodic flaring. If neutrino emission follows
    the same time variability, then it may be possible to
    improve the signal-to-noise ratio by eliminating the peri-
    ods of relatively low output. Multiple detections of Mrk
    501 from several air Cerenkov instruments allowed nearly
    continuous monitoring, including periods when the Moon
    was shining. However, monitoring by multiple instru-
    ments extended only from March to late August. Because
    of uncertainties in the details of the time dependence of
    the gamma-ray emission,
    neutrino
    flux limits are not
    greatly improved by restricting the analysis to high-flux
    periods of gamma-ray emission.
    While this paper describes an analysis dedicated to the
    search for point sources, another strategy was developed
    based on the event selection of the atmospheric neutrino
    analysis (Biron 2002). The results of this complementary
    analysis are consistent with the results presented here. The
    absolute efficiency was extracted by comparing to the known
    flux from atmospheric neutrinos. Moreover, the second
    analysis was subject to different systematic uncertainties.
    The method based on the atmospheric neutrino analysis
    retained a smaller event sample of 369 events, of which
    ?
    270 are expected from atmospheric neutrinos. The cut
    selections produce an implicit optimization on more vertical
    Declination [degrees]
    Φ
    μ
    limit
    [cm
    ­2
    s
    ­1
    ]
    nominal (layered ice)
    homogeneous ice
    absolute OM sensitivity
    angular OM sensitivity
    muon propagation
    corrected
    10
    ­14
    0 10 2030 405060 7080 90
    Fig.
    20.
    —Comparison of muon flux calculations for differential neutrino
    signal proportional to
    E
    ?
    2
    . See text for explanation of legend. Statistical
    uncertainty is small and not shown. The solid curve (corrected) includes
    systematic uncertainty and indicates the final result of this work. [
    See the
    electronic edition of the Journal for a color version of this figure.
    ]
    Declination [degrees]
    Φ
    ν
    limit
    [cm
    ­2
    s
    ­1
    ]
    nominal (layered ice)
    homogeneous ice
    absolute OM sensitivity
    angular OM sensitivity
    muon propagation
    corrected
    10
    ­7
    0 10 2030 405060 7080 90
    Fig.
    21.
    —Comparison of neutrino flux calculations for differential
    neutrino signal proportional to
    E
    ?
    2
    . See text for explanation of legend.
    Statistical uncertainty is small and not shown. The solid curve (corrected)
    includes systematic uncertainty and indicates the final result of this work.
    [
    See the electronic edition of the Journal for a color version of this figure.
    ]
    1054 AHRENS ET AL. Vol. 583

    events and/or softer energy spectra. Figure 22 compares the
    average effective area of the two analyses for an assumed
    differential spectra proportional to
    E
    ?
    2
    . The best flux limits
    for soft spectra are obtained by atmospheric neutrino analy-
    sis, but the neutrino and muon flux limits for either analysis
    are much larger than obtained for an assumed power law of
    an index of
    ?
    2.0.
    While the flux limits for any particular source or direction
    in the northern hemisphere can be extracted from this analy-
    sis (see Table 4), flux limits, both integral and pseudodiffer-
    ential, for a preselected list of 62 sources have been reported
    (Biron 2002). These include all known TeV gamma-ray
    blazars, nearby QSOs, and Galactic TeV gamma-ray
    sources in the northern hemisphere. The list also includes
    TABLE 4
    Muon Flux Limits
    Decl.
    (deg)
    R.A.
    (hours)
    ?
    limit
    l
    (10
    ?
    15
    cm
    ?
    2
    s
    ?
    1
    )
    Decl.
    (deg)
    R.A.
    (hours)
    ?
    limit
    l
    (10
    ?
    15
    cm
    ?
    2
    s
    ?
    1
    )
    Decl.
    (deg)
    R.A.
    (hours)
    ?
    limit
    l
    (10
    ?
    15
    cm
    ?
    2
    s
    ?
    1
    )
    85........... 4.0 2.5 39........... 7.8 6.1 17........... 7.9 9.0
    12.0 6.5 8.9 2,5 8.7 9.0
    20.0 10.5 9.9 15.5 9.5 19.9
    73........... 1.3 6.9 11.0 13.5 10.3 14.9
    4.0 4.2 12.0 13.5 11.2 42.6
    6.7 2.1 13.0 8.0 12.0 11.9
    9.3 6.9 14.1 8.0 12.8 34.1
    12.0 13.7 15.1 8.0 13.7 24.6
    14.7 6.9 16.2 10.6 14.5 42.6
    17.3 9.6 17.2 15.5 15.3 8.9
    20.0 4.2 18.3 3.9 16.1 11.9
    22.7 4.2 19.3 6.1 17.0 14.9
    62........... 0.9 7.9 20.3 10.6 17.8 8.9
    2.6 5.5 21.4 3.9 18.6 42.6
    4.3 12.1 22.4 3.9 19.4 19.9
    6.0 3.4 23.5 6.1 20.3 19.9
    7.7 7.9 28........... 0.4 20.5 21.1 48.0
    9.4 5.5 1.3 20.5 21.9 14.9
    11.1 7.9 2.2 7.1 22.8 34.1
    12.9 1.9 3.1 13.9 23.6 29.1
    14.6 5.5 4.0 24.0 6............. 0.4 33.6
    16.3 3.5 4.9 13.9 1.2 64.9
    18.0 3.5 5.8 7.1 2.0 127.2
    19.7 7.9 6.7 20.5 2.8 64.9
    21.4 7.9 7.6 13.9 3.6 47.8
    23.1 9.9 8.4 20.5 4.4 33.6
    51........... 0.6 7.7 9.3 7.1 5.2 96.1
    1.9 10.0 10.2 24.0 6.0 64.9
    3.2 2.2 11.1 20.5 6.8 78.0
    4.4 7.7 12.0 24.0 7.6 78.0
    5.7 5.9 12.9 5.4 8.4 47.8
    6.9 10.0 13.8 13.9 9.2 64.9
    8.2 2.2 14.7 5.4 10.0 64.9
    9.5 10.0 15.6 20.5 10.8 127.2
    10.7 5.9 16.4 16.7 11.6 47.9
    12.0 1.6 17.3 5.4 12.4 33.6
    13.3 10.0 18.2 16.7 13.2 143.5
    14.5 5.9 19.1 7.1 14.0 24.9
    15.8 10.0 20.0 10.2 14.8 47.8
    17.1 7.7 20.9 7.1 15.6 33.6
    18.3 14.1 21.8 13.9 16.4 64.9
    19.6 3.9 22.7 10.2 17.2 47.8
    20.8 5.9 23.6 7.1 18.0 47.8
    22.1 12.4 17........... 0.4 48.0 18.8 112.4
    23.4 7.7 1.2 34.1 19.6 96.1
    39........... 0.5 10.6 2.1 14.9 20.4 33.6
    1.6 15.5 2.9 51.9 21.2 96.1
    2.6 8.0 3.7 19.9 22.0 24.9
    3.7 13.5 4.6 24.6 22.8 64.9
    4.7 3.9 5.4 14.9 23.6 78.0
    5.7 15.5 6.2 4.5
    6.8 6.1 7.0 9.0
    Notes.
    —Neutrino-induced muon flux upper limits for source spectra proportional to
    E
    ?
    2
    . The impact of systematic uncertainty is included. Angu-
    lar coordinates refer to the center of the search bin.
    No. 2, 2003 SEARCH FOR HIGH-ENERGY NEUTRINO SOURCES 1055

    microquasars, the five most luminous AGNs in wavelength
    bands that span across MeV, X-ray, infrared, and radio
    bands. Of particular interest are radio galaxies with strong
    emission at GHz frequencies. We have also investigated BL
    Lac objects that are close to the arrival directions of the very
    highest energy cosmic rays (Tinyakov & Tkachev 2001) and
    the 10 reported cosmic-ray doublets at extreme energies
    (Uchihori et al. 2000).
    10. FUTURE
    The technique employed in this paper optimized the selec-
    tion criteria on signal to noise. Because of the relatively
    large number of sky bins and the relatively low number of
    events in any individual bin, the analysis procedure pro-
    duced event samples that are dominated by poorly recon-
    structed atmospheric muons rather than upward-traveling
    atmospheric neutrino background. However, as the back-
    ground rejection of downgoing events improves with
    larger detectors, such as AMANDA-II, this trend may not
    continue.
    AMANDA-II, completed in 2000 January, surrounds the
    B10 core with nine additional strings, more than doubling
    the number of optical modules. For this broader configura-
    tion, the effective area for neutrino-induced muons remains
    relatively constant over the entire hemisphere (Barwick et
    al. 2001).
    35
    Consequently, AMANDA-II is expected to
    achieve a factor of 5 improvement in sensitivity for nearly
    horizontal events compared to AMANDA-B10 (Wischnew-
    ski et al. 2001).
    36
    The greater statistical sample of atmo-
    spheric neutrinos will allow better tests of the detector
    simulation programs, especially near the horizon. With the
    data already collected on tape, AMANDA-II can observe
    (or exclude) neutrino fluxes that are approximately 1 order
    of magnitude below the limits presented here, as shown in
    Figure 1.
    This research was supported by the following agencies:
    the US National Science Foundation Office of Polar Pro-
    grams and Physics Division; the University of Wisconsin
    Alumni Research Foundation; the US Department of
    Energy; the Swedish Natural Science Research Council; the
    Swedish Polar Research Secretariat; the Knut and Alice
    Wallenberg Foundation, Sweden; the German Ministry for
    Education and Research; the US National Energy Research
    Scientific Computing Center (supported by the Office of
    Energy Research of the US Department of Energy); the
    University of California, Irvine, AENEAS Supercomputer
    Facility; and the Deutsche Forschungsgemeinschaft
    (DFG). C. Pe
    ´
    rez de los Heros received support from the EU
    fourth framework of Training and Mobility of Researchers,
    and D. F. Cowen acknowledges the support of the NSF
    CAREER program. P. Desiati was supported by the
    Koerber Foundation.
    REFERENCES
    Agrawal, V., Gaisser, T., Lipari, P., & Stanev, T. 1996, Phys. Rev. D, 53,
    1314
    Aharonian, F. A., et al. 1999, A&A, 349, 11
    Ahrens, J., et al. (AMANDA Collaboration). 2002a, Phys. Rev. D, 66,
    012005
    ———. 2002b, Phys. Rev. D, 66, 032006
    ———. 2002c, Phys. Rev. D, submitted (astro-ph/0206487)
    Ambrosio, M., et al. (MACRO Collaboration). 2001, ApJ, 546, 1038
    Andres, E., et al. 2000, Astropart. Phys., 13, 1
    ———. 2001, Nature, 410, 441
    Barbagli, G., et al. 1993, Nucl. Phys. B (Proc. Suppl.), 32, 156
    Barwick, S. W., et al. (AMANDA Collaboration). 2000, Nucl. Phys. B
    (Proc. Suppl.), 87, 402
    ———. 2001, Proc. 27th Int. Cosmic-Ray Conf. (Hamburg), 2, 1101
    Bednarek, W., & Protheroe, R. J. 1999, MNRAS, 302, 373
    Berezinsky, V. S., & Dokuchaev, V. I. 2001, Astropart. Phys., 15, 87
    Biron, A. 2002, Ph.D. thesis, Humboldt Univ. Berlin
    Bouchta, A., et al. (AMANDA Collaboration). 1999, Proc. 26th Int.
    Cosmic-Ray Conf. (Salt Lake City), 2, 108
    Boziev, S. N. et al. 1989, preprint (INR preprint P-0630)
    Buckley, J. H., et al. 1998a, Adv. Space Res., 21, 101
    Buckley, J. H., et al. 1998b, Science, 279, 676
    Catanese, M., & Weekes, T. C. 1999, PASP, 111, 1193
    Conrad, J., Botner, O., Hallgren, A., & Pe
    ´
    rez de los Heros, C. 2003, Phys.
    Rev. D, in press (hep-ex/0202013)
    Cousins, R. D., & Highland, V. L. 1992, Nucl. Instrum. Methods Phys.
    Res. A, 320, 331
    de Jager, O. C., & Stecker, F. W. 2002, ApJ, 566, 738
    Dermer, C. D. 1999, Astropart. Phys., 11, 1
    Dermer, C. D., & Atoyan, A. 2001, Phys. Rev. Lett., 87, 221102
    Dickinson, J. E., et al. 2000, Nucl. Instrum. Methods Phys. Res. A, 440,
    95
    Distefano, C., Guetta, D., Waxman, E., & Levinson, A. 2002, ApJ, 575, 378
    Feldman, G. J., & Cousins, R. D. 1998, Phys. Rev. D, 57, 3873
    Gaisser, T. K. 2002, Astropart. Phys., 16, 285
    Hardtke, R., & Barouch, G. (for the AMANDA Collaboration). 2001,
    Proc. 27th Int. Cosmic-Ray Conf. (Hamburg), 2, 1121
    Heck, D., et al. 1998, Rep. FZKA 6019 (Forschungszentrum Karlsruhe)
    ———. 1999, in Simulations and Analysis Methods for Large Neutrino
    Telescopes, ed. C. Spiering (DESY-PROC-1999-01; Zeuthen: DESY),
    228
    Hill, G. 1997, Astropart. Phys., 6, 215
    Hill, G., & Leuthold, M. (for the AMANDA Collaboration). 2001, Proc.
    27th Int. Cosmic-Ray Conf. (Hamburg), 2, 1113
    Hundertmark, S., et al. (AMANDA Collaboration). 1999, Proc. 26th Int.
    Cosmic-Ray Conf. (Salt Lake City), 2, 12
    Karle, A., et al. 2003, Proc. XXth Int. Conf. Neutrino Physics and Astro-
    physics (Amsterdam: North-Holland), in press (astro-ph
    =
    0209556)
    Konopelko, A. K., Kirk, J. G., Stecker, F. W., & Mastichiadis, A. 1999,
    ApJ, 518, L13
    Fig.
    22.
    —Comparison of average muon effective area as a function of
    declinations for differential neutrino signal proportional to
    E
    ?
    2
    . The
    dashed curve is from Biron (2002). Neither curve includes the effect of
    systematic uncertainties. [
    See the electronic edition of the Journal for a color
    version of this figure.
    ]
    35
    Available at http://area51.berkeley.edu/manuscripts/
    20010601xx-ICRC_AmIIv3.pdf.
    36
    Available at http://area51.berkeley.edu/manuscripts/
    20010602xx-icrc2001_aii.pdf.
    1056 AHRENS ET AL. Vol. 583

    Krennrich, F., et al. 1999, ApJ, 511, 149
    Learned, J. G., & Mannheim, K. 2000, Annu. Rev. Nucl. Part. Sci., 50, 679
    Levinson, A., & Waxman, E. 2001, Phys. Rev. Lett., 87, 171101
    Lipari, P., & Stanev, T. 1991, Phys. Rev. D, 44, 3543
    Lohmann, W., Kopp, R., & Voss, R. 1985, preprint (CERN preprint
    CERN-85-03)
    Matsuno, S., et al. (Super-Kamiokande Collaboration). 2001, Proc. 27th
    Int. Cosmic-Ray Conf. (Hamburg), 2, 1065
    Montaruli, T., et al. (MACRO Collaboration). 1999, Proc. 26th Int.
    Cosmic-Ray Conf. (Salt Lake City), 2, 213
    Nellen, L., Mannheim, K., & Biermann, P. L. 1993, Phys. Rev. D, 47, 5270
    Oppelt, A. 2001, Ph.D. thesis, Univ. Me
    ´
    diterrane
    ´
    e Aix-Marseille II
    Perrone, L., et al. (MACRO Collaboration). 2001, Proc. 27th Int. Cosmic-
    Ray Conf. (Hamburg), 2, 1073
    Rachen, J. P. 2000, preprint (astro-ph/0003282)
    Rawlins, K. 2001, Ph.D. thesis, Univ. Wisconsin, Madison
    Rhode, W., & Chirkin, D. 2001, Proc. 27th Int. Cosmic-Ray Conf.
    (Hamburg), 3, 1017
    Spiering, C. 2001, Proc. 27th Int. Cosmic-Ray Conf. (Hamburg), 2, 1242
    Stecker, F. W., Done, C., Salamon, M. H., & Sommers, P. 1991, Phys. Rev.
    Lett., 66, 2697 (erratum 69, 2738 [1992])
    Stecker, F. W., & Salamon, M. H. 1996, Space Sci. Rev., 75, 341
    Szabo, A. P., & Protheroe, R. J. 1992, in Proc. Workshop on High Energy
    Neutrino Astrophysics, ed. V. J. Stenger, J. G. Learned, S. Pakvasa, &
    X. Tata (Singapore: World Scientific), 24
    Tinyakov, P. G., & Tkachev, I. I. 2001, J. Exp. Theor. Phys. Lett., 74, 445
    Uchihori, Y., et al. 2000, Astropart. Phys., 13, 151
    Weekes, T. 2001, AIP Conf. Proc. 558, High Energy Gamma-Ray Astron-
    omy, ed. F. A. Aharonian & H. Vo
    ¨
    lk (Melville: AIP), 15
    Wiebel-Sooth, B., & Biermann, P. L. 1998, in Landolt-Bo
    ¨
    rnstein New
    Series, Group VI, Vol. 3C (Berlin: Springer), 37
    Wischnewski, R., et al. (AMANDA Collaboration). 1999a, Nucl. Phys. B
    (Proc. Suppl.), 75 (1–2), 412
    ———. 2001, Proc. 27th Int. Cosmic-Ray Conf. (Hamburg), 2, 1105
    Woschnagg, K., et al. (AMANDA Collaboration). 1999, Proc. 26th Int.
    Cosmic-Ray Conf. (Salt Lake City), 2, 200
    Young, S. 2001, Ph.D. thesis, Univ. California, Irvine
    No. 2, 2003 SEARCH FOR HIGH-ENERGY NEUTRINO SOURCES 1057

    Back to top