ARTICLE IN PRESS
    Nuclear Instruments and Methods in Physics Research A 524 (2004) 169–194
    Muon trackreconstruction and data selection techniques
    in
    AMANDA
    J. Ahrens
    a
    , X. Bai
    b
    , R. Bay
    c
    , S.W. Barwick
    d
    , T. Becka
    a
    , J.K. Becker
    e
    , K.-
    H. Becker
    e
    , E. Bernardini
    f
    , D. Bertrand
    g
    , A. Biron
    f
    , D.J. Boersma
    f
    ,S.B
    .
    oser
    f
    ,
    O. Botner
    h
    , A. Bouchta
    h
    , O. Bouhali
    g
    , T. Burgess
    i
    , S. Carius
    j
    , T. Castermans
    k
    ,
    D. Chirkin
    c
    , B. Collin
    l
    , J. Conrad
    h
    , J. Cooley
    m
    , D.F. Cowen
    l
    , A. Davour
    h
    ,
    C. De Clercq
    n
    , T. DeYoung
    o
    , P. Desiati
    m
    , J.-P. Dewulf
    g
    , P. Ekstr
    .
    om
    e
    , T. Feser
    a
    ,
    M. Gaug
    f
    , T.K. Gaisser
    b
    , R. Ganugapati
    m
    , H. Geenen
    e
    , L. Gerhardt
    d
    , A. Gro
    X
    e
    ,
    A. Goldschmidt
    p
    , A. Hallgren
    h
    , F. Halzen
    m
    , K. Hanson
    m
    , R. Hardtke
    m
    ,
    T.Harenberg
    e
    ,T.Hauschildt
    f
    ,K.Helbing
    p
    ,M.Hellwig
    a
    ,P.Herquet
    k
    ,G.C.Hill
    m
    ,
    D. Hubert
    n
    , B. Hughey
    m
    , P.O. Hulth
    i
    , K. Hultqvist
    i
    , S. Hundertmark
    i
    ,
    J. Jacobsen
    p
    , A. Karle
    m
    , M. Kestel
    l
    ,L.K
    .
    opke
    a
    , M. Kowalski
    f
    , K. Kuehn
    d
    ,
    J.I. Lamoureux
    p
    , H. Leich
    f
    , M. Leuthold
    f
    , P. Lindahl
    j
    , I. Liubarsky
    q
    , J. Madsen
    r
    ,
    P. Marciniewski
    h
    , H.S. Matis
    p
    , C.P. McParland
    p
    , T. Messarius
    e
    , Y. Minaeva
    i
    ,
    P. Mio
    W
    inovi
    !
    c
    c
    , P.C. Mock
    d
    , R. Morse
    m
    , K.S. M
    .
    unich
    e
    , J. Nam
    d
    , R. Nahnhauer
    f
    ,
    T. Neunh
    .
    offer
    a
    , P. Niessen
    n
    , D.R. Nygren
    p
    ,H.
    .
    Ogelman
    m
    , Ph. Olbrechts
    n
    ,
    C. P
    !
    erez de los Heros
    h
    , A.C. Pohl
    i
    , R. Porrata
    d
    , P.B. Price
    c
    , G.T. Przybylski
    p
    ,
    K. Rawlins
    m
    , E. Resconi
    f
    , W. Rhode
    e
    , M. Ribordy
    k
    , S. Richter
    m
    , J. Rodr
    !
    ıguez
    Martino
    i
    , D. Ross
    d
    , H.-G. Sander
    a
    , K. Schinarakis
    e
    , S. Schlenstedt
    f
    , T. Schmidt
    f
    ,
    D.Schneider
    m
    ,R.Schwarz
    m
    ,A.Silvestri
    d
    ,M.Solarz
    c
    ,G.M.Spiczak
    r
    ,C.Spiering
    f
    ,
    M. Stamatikos
    m
    , D. Steele
    m
    , P. Steffen
    f
    , R.G. Stokstad
    p
    , K.-H. Sulanke
    f
    ,
    O. Streicher
    f
    , I. Taboada
    s
    , L. Thollander
    i
    , S. Tilav
    b
    , W. Wagner
    e
    , C. Walck
    i
    ,
    Y.-R. Wang
    m
    , C.H. Wiebusch
    e,
    *, C. Wiedemann
    i
    , R. Wischnewski
    f
    , H. Wissing
    f
    ,
    K. Woschnagg
    c
    , G. Yodh
    d
    a
    Institute of Physics, University of Mainz, D­55099 Mainz, Germany
    b
    Bartol Research Institute, University of Delaware, Newark, DE 19716, USA
    c
    Department of Physics, University of California, Berkeley, CA 94720, USA
    d
    Department of Physics and Astronomy, University of California, Irvine, CA 92697, USA
    e
    Fachbereich 8 Physik, BU Wuppertal, Gaussstrasse 20, D­42119 Wuppertal, Germany
    f
    DESY­Zeuthen, D­15738 Zeuthen, Germany
    g
    Universit
    !
    e Libre de Bruxelles, Science Faculty, Brussels, Belgium
    h
    Division of High Energy Physics, Uppsala University, S­75121 Uppsala, Sweden
    *Corresponding author. Tel.: +49-0-202-439-3531; fax: +49-0-202-439-2662.
    E-mail address:
    wiebusch@physik.uni-wuppertal.de (C.H. Wiebusch).
    0168-9002/$ - see front matter
    r
    2004 Elsevier B.V. All rights reserved.
    doi:10.1016/j.nima.2004.01.065

    i
    Department of Physics, Stockholm University, SE­10691 Stockholm, Sweden
    j
    Department of Technology, Kalmar University, S­39182 Kalmar, Sweden
    k
    University of Mons­Hainaut, 7000 Mons, Belgium
    l
    Department of Physics, Pennsylvania State University, University Park, PA 16802, USA
    m
    Department of Physics, University of Wisconsin, Madison, WI 53706, USA
    n
    Vrije Universiteit Brussel, Dienst ELEM, B­1050 Brussels, Belgium
    o
    Department of Physics, University of Maryland, College Park, MD 20742, USA
    p
    Lawrence Berkeley National Laboratory, Berkeley, CA 94720, USA
    q
    Blackett Laboratory, Imperial College, London SW7 2BW, UK
    r
    Physics Department, University of Wisconsin, River Falls, WI 54022, USA
    s
    Department de F
    !
    ısica, Universidad Sim
    !
    on Bol
    !
    ıvar, Caracas, 1080, Venezuela
    The AMANDA Collaboration
    Received 24 September 2003; received in revised form 20 January 2004; accepted 27 January 2004
    Abstract
    The Antarctic Muon And Neutrino Detector
    A
    rray (AMANDA) is a high-energy neutrino telescope operating at the
    geographic South Pole. It is a lattice of photo-multiplier tubes buried deep in the polar ice between 1500 and 2000 m
    :
    The primary goal of this detector is to discover astrophysical sources of high-energy neutrinos. A high-energy muon
    neutrino coming through the earth from the Northern Hemisphere can be identified by the secondary muon moving
    upward through the detector.
    The muon tracks are reconstructed with a maximum likelihood method. It models the arrival times and amplitudes of
    Cherenkov photons registered by the photo-multipliers. This paper describes the different methods of reconstruction,
    which have been successfully implemented within
    AMANDA
    . Strategies for optimizing the reconstruction performance and
    rejecting background are presented. For a typical analysis procedure the direction of tracks are reconstructed with
    about 2
    ?
    accuracy.
    r
    2004 Elsevier B.V. All rights reserved.
    PACS:
    95.55.Vj; 95.75.Pq; 29.40.Ka; 29.85.+c
    Keywords:
    AMANDA
    ; Trackreconstruction; Neutrino telescope; Neutrino astrophysics
    1. Introduction
    The Antarctic Muon And Neutrino Detector
    Array
    [1],
    AMANDA
    , is a large volume neutrino
    detector at the geographic South Pole. It is a lattice
    of photo-multiplier tubes (PMTs) buried deep in
    the optically transparent polar ice. The primary
    goal of this detector is to detect high-energy
    neutrinos from astrophysical sources, and deter-
    mine their arrival time, direction and energy.
    When a high-energy neutrino interacts in the
    polar ice via a charged current reaction with a
    nucleon
    N
    :
    n
    c
    þ
    N
    ­
    c
    þ
    X
    ;
    ð
    1
    Þ
    it creates a hadronic cascade, X, and a lepton,
    c
    ¼
    e
    ;
    m
    ;
    t
    :
    These particles generate Cherenkov
    photons, which are detected by the PMTs. Each
    lepton flavor generates a different signal in the
    detector. The two basic detection modes are
    sketched in Fig. 1.
    A high-energy
    n
    m
    charged current interaction
    creates a muon, which is nearly collinear with the
    neutrino direction; having a mean deviation angle
    of
    c
    ¼
    0
    :
    7
    ?
    E
    n
    =
    TeV
    Þ
    ?
    0
    :
    7
    [2], which implies an
    accuracy requirement of
    t
    1
    ?
    for reconstructing
    the muon direction.
    The high-energy muon emits a cone of Cher-
    enkov light at a fixed angle
    y
    c
    :
    It is determined by
    cos
    y
    c
    ¼ð
    n
    b
    Þ
    ?
    1
    ;
    where
    n
    C
    1
    :
    32 is the index of
    ARTICLE IN PRESS
    J. Ahrens et al. / Nuclear Instruments and Methods in Physics Research A 524 (2004) 169–194
    170

    refraction in the ice. For relativistic particles,
    b
    C
    1
    ;
    and
    y
    c
    E
    41
    ?
    :
    The direction of the muon is
    reconstructed from the time and amplitude in-
    formation of the PMTs illuminated by the
    Cherenkov cone.
    Radiative energy loss processes generate sec-
    ondary charged particles along the muon trajec-
    tory, which also produce Cherenkov radiation.
    These additional photons allow an estimate of the
    muon energy. However, the resolution is limited
    by fluctuations of these processes. This estimate is
    a lower bound on the neutrino energy, because it is
    based on the muon energy at the detector. The
    interaction vertex may be far outside the detector.
    The
    n
    e
    and
    n
    t
    channels are different. The
    electron from a
    n
    e
    will generate an electro-
    magnetic cascade, which is confined to a volume
    of a few cubic meters. This cascade coincides with
    the hadronic cascade X of the primary interaction
    vertex. The optical signature is an expanding
    spherical shell of Cherenkov photons with a lar-
    ger intensity in the forward direction. The tau
    from a
    n
    t
    will decay immediately and also gene-
    rate a cascade. However, at energies
    >
    1 PeV this
    cascade and the vertex are separated by several
    tens of meters, connected by a single track. This
    signature of two extremely bright cascades is
    unique for high-energy
    n
    t
    ;
    and it is called a
    double
    bang event
    [3].
    The measurement of cascade-like events is
    restricted to interactions close to the detector,
    thus requiring larger instrumented volumes than
    for
    n
    m
    detection. Also the accuracy of the direction
    measurement is worse for cascades than for long
    muon tracks. However, when the flux is diffuse,
    the
    n
    e
    and
    n
    t
    channels also have clear advantages.
    The backgrounds from atmospheric neutrinos are
    smaller. The energy resolution is significantly
    better since the full energy is deposited in or near
    the detector. The cascade channel is sensitive to all
    neutrino flavors because the neutral current
    interactions also generate cascades. In this paper,
    we focus on the reconstruction of muon tracks;
    details on the reconstruction of cascades are
    described in Ref. [4].
    The most abundant events in
    AMANDA
    are atmo-
    spheric muons, created by cosmic rays interacting
    with the Earth’s atmosphere. At the depth of
    AMANDA
    their rate exceeds the rate of muons from
    atmospheric neutrinos by five orders of magni-
    tude. Since these muons are absorbed by the earth,
    a muon trackfrom the lower hemisphere is a
    unique signature for a neutrino-induced muon.
    1
    The reconstruction procedure must have good
    angular resolution, good efficiency, and allow
    excellent rejection of down-going atmospheric
    muons.
    This paper describes the methods used to
    reconstruct muon tracks recorded in the
    AMANDA
    experiment. The
    AMANDA-II
    detector is introduced
    in Section 2. The reconstruction algorithms and
    their implementation are described in Sections 3–5.
    ARTICLE IN PRESS
    cascade
    muon
    PMTs
    c
     
    spherical Cherenkov front
    Cherenkov cone
    Fig. 1. Detection modes of the
    AMANDA
    detector: Left: muon tracks induced by muon-neutrinos; Right: Cascades from electron- or tau-
    neutrinos.
    1
    Muon neutrinos above 1 PeV are absorbed by the Earth. At
    these ultra-high-energies (UHE), however, the muon back-
    ground from cosmic rays is small and UHE muons coming
    from the horizon and above are most likely created by UHE
    neutrinos. The search for these UHE neutrinos is described in
    Refs. [5,6].
    J. Ahrens et al. / Nuclear Instruments and Methods in Physics Research A 524 (2004) 169–194
    171

    Section 6 summarizes event classes for which the
    reconstruction may fail and strategies to identify
    and eliminate such events. The performance
    of the reconstruction procedure is shown in
    Section 7. We discuss possible improvements
    in Section 8.
    2. The AMANDA detector
    The
    AMANDA-II
    detector (see
    Fig. 2) has been
    operating since January 2000 with 677 optical
    modules (OM) attached to 19 strings. Most of the
    OMs are located between 1500 and 2000 m below
    the surface. Each OM is a glass pressure vessel,
    which contains an 8-in. hemispherical PMT and its
    electronics.
    AMANDA-B10
    ,
    2
    the inner core of 302
    OMs on 10 strings, has been operating since 1997.
    One unique feature of
    AMANDA
    is that it
    continuously measures atmospheric muons in
    coincidence with the
    South Pole Air Shower
    Experiment
    surface arrays
    SPASE-1
    and
    SPASE-2
    [7]. These muons are used to survey the detector
    and calibrate the angular resolution (see Section 7
    and
    Refs. [8,9]), while providing
    SPASE
    with
    additional information for cosmic ray composition
    studies [10].
    The PMT signals are processed in a counting
    room at the surface of the ice. The analog signals
    are amplified and sent to a majority logic trigger
    [11]. There the pulses are discriminated and a
    trigger is formed if a minimum number of hit
    PMTs are observed within a time window of
    typically 2
    m
    s
    :
    Typical trigger thresholds were 16
    hit PMT for
    AMANDA-B10
    and 24 for
    AMANDA-II
    .
    For each trigger the detector records the peak
    amplitude and up to 16 leading and trailing edge
    times for each discriminated signal. The time
    resolution achieved after calibration is
    s
    t
    C
    5ns
    for the PMTs from the first 10 strings, which are
    read out via coaxial or twisted pair cables. For the
    remaining PMTs, which are read out with optical
    fibers the resolution is
    s
    t
    C
    3
    :
    5ns
    :
    In the cold
    environment of the deep ice the PMTs have low
    noise rates of typically 1 kHz
    :
    The timing and amplitude calibration, the array
    geometry, and the optical properties of the ice are
    determined by illuminating the array with known
    optical pulses from in situ sources [11]. Time
    offsets are also determined from the response to
    through-going atmospheric muons[12].
    ARTICLE IN PRESS
    light diffuser ball
    HV divider
    silicon gel
    Module
    Optical
    pressure
    housing
    Depth
    120 m
    AMANDA­II
    AMANDA­B10
    Inner 10 strings:
    zoomed in on one
    optical module (OM)
    main cable
    PMT
    200 m
    1000 m
    2350 m
    2000 m
    1500 m
    1150 m
    Fig. 2. The
    AMANDA-II
    detector. The scale is illustrated by the Eiffel tower at the left.
    2
    Occasionally in the paper we will refer to this earlier
    detector instead of the full
    AMANDA-II
    detector.
    J. Ahrens et al. / Nuclear Instruments and Methods in Physics Research A 524 (2004) 169–194
    172

    The optical absorption length in the ice is
    typically 110 m at 400 nm with a strong wave-
    length dependence. The effective scattering length
    at 400 nm is on average
    C
    20 m
    :
    It is defined as
    l
    s
    =
    ð
    1
    ?
    /
    cos
    y
    s
    S
    Þ
    ;
    where
    l
    s
    is the scattering length
    and
    y
    s
    is the scattering angle. The ice parameters
    vary strongly with depth due to horizontal ice
    layers, i.e., variations in the concentration of
    impurities which reflect past geological events
    and climate changes [13–19].
    3. Reconstruction algorithms
    The muon trackreconstruction algorithm is a
    maximum likelihood procedure. Prior to recon-
    struction simple pattern recognition algorithms,
    discussed in Section 4, generate the initial esti-
    mates required by the maximum likelihood recon-
    structions.
    3.1. Likelihood description
    The reconstruction of an event can be general-
    ized to the problem of estimating a set of unknown
    parameters
    f
    a
    g
    ;
    e.g. trackparameters, given a set
    of experimentally measured values
    f
    x
    g
    :
    The
    parameters,
    f
    a
    g
    ;
    are determined by maximizing
    the likelihood
    L
    ð
    x
    j
    a
    Þ
    which for independent
    components
    x
    i
    of
    x
    reduces to
    L
    ð
    x
    j
    a
    Þ¼
    Y
    i
    p
    ð
    x
    i
    j
    a
    Þð
    2
    Þ
    where
    p
    ð
    x
    i
    j
    a
    Þ
    is the probability density function
    (p.d.f.) of observing the measured value
    x
    i
    for
    given values of the parameters
    f
    a
    g
    [20].
    To simplify the discussion we assume that the
    Cherenkov radiation is generated by a single
    infinitely long muon track(with
    b
    ¼
    1) and forms
    a cone. It is described by the following parameters:
    a
    ¼ð
    r
    0
    ;
    t
    0
    ;
    #
    p
    ;
    E
    0
    Þð
    3
    Þ
    and illustrated inFig. 3. Here,
    r
    0
    is an arbitrary
    point on the track. At time
    t
    0
    ;
    the muon passes
    r
    0
    with energy
    E
    0
    along a direction
    #
    p
    :
    The geome-
    trical coordinates contain five degrees of freedom.
    Along this track, Cherenkov photons are emitted
    at a fixed angle
    y
    c
    relative to
    #
    p
    :
    Within the
    reconstruction algorithm it is possible to use a
    different coordinate system, e.g.
    a
    ¼ð
    d
    ;
    Z
    ;
    y
    Þ
    :
    The
    reconstruction is performed by minimizing
    ?
    log
    ð
    L
    Þ
    with respect to
    a
    :
    The values
    f
    x
    g
    presently recorded by
    AMANDA
    are
    the time
    t
    i
    and duration
    TOT
    i
    (
    Time Over
    Threshold
    ) of each PMT signal, as well as the
    peakamplitude
    A
    i
    of the largest pulse in each
    PMT. PMTs with no signal above threshold are
    also accounted for in the likelihood function. The
    hit times give the most relevant information.
    Therefore we will first concentrate on
    p
    ð
    t
    j
    a
    Þ
    :
    3.1.1. Time likelihood
    According to the geometry in Fig. 3, photons
    are expected to arrive at OM
    i
    (at
    r
    i
    ) at time
    t
    geo
    ¼
    t
    0
    þ
    #
    p
    r
    i
    ?
    r
    0
    Þþ
    d
    tan
    y
    c
    c
    vac
    ð
    4
    Þ
    with
    c
    vac
    the vacuum speed of light.
    3
    It is
    convenient to define a relative arrival time, or
    time residual
    t
    res
    ?
    t
    hit
    ?
    t
    geo
    ð
    5
    Þ
    which is the difference between the observed hit
    time and the hit time expected for a ‘‘direct
    photon’’, a Cherenkov photon that travels un-
    delayed directly from the muon to an OM without
    scattering.
    ARTICLE IN PRESS
    c
    θ
    c
    θ
    OM
    µ
    d
    x
    Cherenkov light
    t , , E
    0 0 0
    PMT­axis
    η
    p
    r
    r
    i
    Fig. 3. Cherenkov light front: definition of variables.
    3
    We note that Eq. (4) neglects the effect that Cherenkov light
    propagates with group velocity as pointed out in Ref. [21].It
    was shown inRef. [14]that for AMANDA this approximation
    is justified.
    J. Ahrens et al. / Nuclear Instruments and Methods in Physics Research A 524 (2004) 169–194
    173

    In the ideal case, the distribution,
    p
    ð
    t
    res
    j
    a
    Þ
    ;
    would be a delta function. However, in a realistic
    experimental situation this distribution is broa-
    dened and distorted by several effects, which are
    illustrated in
    Fig. 4. The PMT jitter limits the
    timing resolution
    s
    t
    :
    Noise, e.g. darknoise of the
    PMT, leads to additional hits which are random in
    time. These effects can generate negative
    t
    res
    values, which would mimic unphysical causality
    violations. Secondary radiative energy losses along
    the muon trajectory create photons that arrive
    after the ideal Cherenkov cone. These processes
    are stochastic, and their relative photon yield
    fluctuates.
    In
    AMANDA
    , the dominant effect on photon
    arrival times is scattering in the ice.
    4
    The effect
    of scattering depends strongly on the distance,
    d
    ;
    of the OM from the trackas illustrated in
    Fig. 4.
    Since the PMTs have a non-uniform angular
    response,
    p
    ð
    t
    res
    Þ
    also depends on the orientation,
    Z
    ;
    of the OM relative to the muon track(see
    Fig.
    3). OMs facing away from the trackcan only see
    light that scatters backtowards the PMT face. On
    average this effect shifts
    t
    res
    to later times and
    modifies the probability of a hit.
    The simplest time likelihood function is based
    on a likelihood constructed from
    p
    1
    ;
    the p.d.f. for
    arrival times of single photons
    i
    at the locations of
    the hit OMs
    L
    time
    ¼
    Y
    N
    hits
    i
    ¼
    1
    p
    1
    ð
    t
    res
    ;
    i
    j
    a
    ¼
    d
    i
    ;
    Z
    i
    ;
    y
    Þ
    :
    ð
    6
    Þ
    Note that one OM may contribute to this product
    with several hits. The function
    p
    1
    ð
    t
    res
    ;
    i
    j
    a
    Þ
    is
    obtained from the simulation of photon propaga-
    tion through ice (see Section 3.2). However, this
    description is limited, because the electrical and
    optical signal channels can only resolve multiple
    photons separated by a few 100 ns and
    C
    10 ns
    ;
    respectively. Within this time window, only the
    arrival time of the first pulse is recorded.
    This first photon is usually less scattered than
    the average single photon, which modifies the
    probability distribution of the detected hit time.
    The arrival time distribution of the first of
    N
    photons is given by
    p
    1
    N
    ð
    t
    res
    Þ¼
    Np
    1
    ð
    t
    res
    Þ
    Z
    N
    t
    res
    p
    1
    ð
    t
    Þ
    d
    t
    ??
    ð
    N
    ?
    1
    Þ
    ¼
    Np
    1
    ð
    t
    tres
    Þð
    1
    ?
    P
    1
    ð
    t
    res
    ÞÞ
    ð
    N
    ?
    1
    Þ
    ð
    7
    Þ
    P
    1
    is the cumulative distribution of the single
    photon p.d.f. The function
    p
    1
    N
    ð
    t
    res
    Þ
    is called the
    multi-photo-electron (MPE) p.d.f. and corre-
    spondingly defines
    L
    MPE
    :
    This concept can be extended to the more
    general case of
    p
    k
    N
    ð
    t
    res
    Þ
    ;
    the p.d.f. for the
    k
    th
    photon out of a total of
    N
    to arrive at
    t
    res
    ;
    given by
    p
    k
    N
    ð
    t
    res
    Þ¼
    N
    N
    ?
    1
    k
    ?
    1
    !
    p
    1
    ð
    t
    res
    Þð
    1
    ?
    P
    1
    ð
    t
    res
    ÞÞ
    ð
    N
    ?
    k
    Þ
    P
    1
    ð
    t
    res
    ÞÞ
    ð
    k
    ?
    1
    Þ
    ð
    8
    Þ
    p
    k
    N
    ð
    t
    res
    Þ
    specifies the likelihood of arrival times of
    individual photoelectrons for averaged time series
    of
    N
    photoelectrons. With waveform recording the
    arrival times and amplitudes of individual pulses
    can be resolved.
    When the number of photoelectrons,
    N
    ;
    is not
    measured precisely enough, multi-photon informa-
    tion can be included via another method. Instead
    ARTICLE IN PRESS
    t
    res
    t
    res
    t
    res
    t
    res
    00
    high
    low
    00
    + showers + scattering
    close track
    far track
    + noise
    t
    jitter
    jitter jitter
    jitter
    Fig. 4. Schematic distributions of arrival times
    t
    tes
    for different
    cases: Top left: PMT jitter. Top right: the effect of jitter and
    random noise. Bottom left: The effect of jitter and secondary
    cascades along the muon track. Bottom right: The effect of
    jitter and scattering.
    4
    In water detectors this effect is neglected[22]or treated as a
    small correction[23].
    J. Ahrens et al. / Nuclear Instruments and Methods in Physics Research A 524 (2004) 169–194
    174

    of measuring
    N
    ;
    the p.d.f. of the first photoelec-
    tron can be calculated by convolving the MPE
    p.d.f.
    p
    1
    N
    ð
    t
    ;
    d
    Þ
    with the Poisson probability
    P
    Poisson
    N
    ð
    m
    Þ
    ;
    where
    m
    is the mean expected number
    of photoelectrons as a function of the distance,
    d
    :
    p
    1
    m
    ð
    t
    res
    Þ¼
    1
    N
    X
    N
    i
    ¼
    1
    m
    i
    e
    ?
    m
    i
    !
    p
    1
    i
    ð
    t
    res
    Þ
    ¼
    m
    1
    ?
    e
    ?
    m
    p
    1
    ð
    t
    res
    Þ
    e
    ?
    m
    P
    1
    ð
    t
    res
    Þ
    ð
    9
    Þ
    This result is called the Poisson Saturated Ampli-
    tude (PSA) p.d.f.
    [24,25]
    and correspondingly
    defines
    L
    PSA
    :
    The constant
    N
    ¼
    1
    ?
    e
    ?
    m
    renorma-
    lizes the p.d.f. to unity.
    The probability of (uncorrelated) noise hits is
    small. They are further suppressed by a
    hit cleaning
    procedure (Section 5.3), which is applied before
    reconstruction. They are included in the likelihood
    function by simply adding a constant p.d.f.
    p
    0
    :
    3.1.2. Hit and no­hit likelihood
    The likelihood in the previous section relies only
    on the measured arrival times of photons. How-
    ever, the topology of the hits is also important.
    PMTs with no hits near a hypothetical trackor
    PMTs with hits far from the trackare unlikely.
    A likelihood utilizing this information can be
    constructed as
    L
    hit
    ¼
    Y
    N
    ch
    i
    ¼
    1
    P
    hit
    ;
    i
    Y
    N
    OM
    i
    ¼
    N
    ch
    þ
    1
    P
    no
    ?
    hit
    ;
    i
    ð
    10
    Þ
    where
    N
    ch
    is the number of hit OMs and
    N
    OM
    the
    number of operational OMs. The probabilities
    P
    hit
    and
    P
    no
    ?
    hit
    of observing or not observing a hit
    depend on the trackparameters
    a
    :
    Additional hits
    due to random noise are easily incorporated:
    P
    no
    ?
    hit
    ­
    *
    P
    no
    ?
    hit
    ?
    P
    no
    ?
    hit
    P
    no
    ?
    noise
    and
    P
    hit
    ­
    *
    P
    hit
    ¼
    1
    ?
    *
    P
    no
    ?
    hit
    :
    Assuming that the probability
    P
    hit
    1
    is known for
    a single photon, the hit and no-hit probabilities of
    OMs for
    n
    photons can be calculated:
    P
    no
    ?
    hit
    n
    ¼ð
    1
    ?
    P
    hit
    1
    Þ
    n
    and
    P
    hit
    n
    ¼
    1
    ?
    P
    no
    ?
    hit
    n
    ¼
    1
    1
    ?
    P
    hit
    1
    Þ
    n
    :
    ð
    11
    Þ
    The number of photons,
    n
    ;
    depends on
    E
    m
    ;
    the
    energy of the muon:
    n
    ¼
    n
    ð
    E
    m
    Þ
    :
    For a fixed track
    geometry, the likelihood (Eq. (10)) can be used to
    reconstruct the muon energy.
    3.1.3. Amplitude likelihood
    The peakamplitudes recorded by
    AMANDA
    can be
    fully incorporated in the likelihood [26], which is
    particularly useful for energy reconstruction. The
    likelihood can be written as
    L
    ¼
    W
    N
    OM
    Y
    N
    OM
    i
    ¼
    1
    w
    i
    P
    i
    ð
    A
    i
    Þð
    12
    Þ
    where
    P
    i
    ð
    A
    i
    Þ
    is the probability that OM
    i
    observes
    an amplitude
    A
    i
    ;
    with
    A
    i
    ¼
    0 for unhit OMs.
    W
    and
    w
    i
    are weight factors, which describe devia-
    tions of the individual OM and the total number of
    hit OMs from the expectation.
    P
    i
    depends on the
    mean number
    m
    of expected photoelectrons:
    P
    i
    ð
    A
    i
    Þ¼
    P
    hit
    ð
    1
    ?
    P
    th
    i
    Þ
    P
    ð
    A
    i
    ;
    m
    Þ
    P
    ð
    /
    A
    i
    S
    ;
    m
    Þ
    :
    ð
    13
    Þ
    The probability
    P
    i
    ð
    A
    i
    Þ
    is normalized to the
    probability of observing the most likely amplitude
    /
    A
    i
    S
    :
    P
    th
    i
    ð
    m
    Þ
    is the probability that a signal of
    m
    does not produce a pulse amplitude above the
    discriminator threshold. As before,
    P
    hit
    ¼
    1
    ?
    P
    no
    ?
    hit
    ;
    where the no-hit probability is given
    by Poisson statistics:
    P
    no
    ?
    hit
    ¼
    exp
    ð?
    m
    Þð
    1
    ?
    P
    noise
    Þ
    :
    The probability of
    A
    i
    ¼
    0 is a special case:
    P
    i
    ð
    0
    Þ¼
    P
    no
    ?
    hit
    þ
    P
    hit
    P
    th
    i
    :
    Energy reconstructions based on
    this formulation of the likelihood will be referred
    to as
    Full E
    reco
    :
    An alternative energy reconstruction technique
    (see Section 3.2.4) uses a neural net which is fed
    with energy sensitive parameters.
    3.1.4. Zenith weighted (Bayesian) likelihood
    Another extension of the likelihood
    [27–29]
    incorporates external information about the muon
    flux via Bayes’ Theorem. This theorem states that
    for two hypotheses
    A
    and
    B
    ;
    P
    ð
    A
    j
    B
    Þ¼
    P
    ð
    B
    j
    A
    Þ
    P
    ð
    A
    Þ
    P
    ð
    B
    Þ
    :
    ð
    14
    Þ
    Identifying
    A
    with the trackparameters
    a
    and
    B
    with the observations
    x
    ;
    Eq. (14) gives the prob-
    ability that the inferred muon track
    a
    was in fact
    ARTICLE IN PRESS
    J. Ahrens et al. / Nuclear Instruments and Methods in Physics Research A 524 (2004) 169–194
    175

    responsible for the observed event
    x
    :
    P
    ð
    x
    j
    a
    Þ
    is the
    probability that
    a
    ;
    assumed to be true, would
    generate the event
    x
    —in other words, the like-
    lihood described in the previous sections.
    P
    ð
    a
    Þ
    is
    the prior probability of observing the track
    a
    ;
    i.e.,
    the relative frequencies of different muon tracks as
    a function of their parameters.
    P
    ð
    x
    Þ
    ;
    which is
    independent of the trackparameters
    a
    ;
    is a
    normalization constant which ensures that
    Eq. (14) defines a proper probability. Because the
    likelihood is only defined up to an arbitrary
    constant factor, this normalization may be ignored
    in the present context.
    In order to obtain
    P
    ð
    a
    j
    x
    Þ
    ;
    one thus has to
    determine the prior probability distribution,
    P
    ð
    a
    Þ
    ;
    of how likely the various possible track directions
    are a priori. The reconstruction maximizes the
    product of the p.d.f.
    and
    the prior.
    The flux of muons deep underground is reason-
    ably well known from previous experiments. Any
    point source of muons would be at most a small
    perturbation on the flux of penetrating atmo-
    spheric muons and muons created by atmospheric
    neutrinos. The most striking feature of the back-
    ground flux from atmospheric muons is the strong
    dependence on zenith angle. For vertically down-
    going tracks it exceeds the flux from neutrino
    induced muons by about 5 orders of magnitude
    but becomes negligible for up-going tracks. This
    dependence, which is modeled by a Monte Carlo
    calculation [30], acts as a zenith dependent weight
    to the different muon hypotheses,
    a
    :
    With this
    particular choice, some tracks, which would
    otherwise reconstruct as up-going, reconstruct as
    down-going tracks. This greatly reduces the rate at
    which penetrating atmospheric muons are mis-
    reconstructed as up-going neutrino events [31].In
    principle, a more accurate prior could be used. It
    would need to include the depth and energy
    dependence of the atmospheric muons as well as
    the angular dependence of atmospheric neutrino-
    induced muons.
    Upon completion of this work, we learned that
    this technique was developed independently by the
    NEVOD neutrino detector collaboration[32]who
    were able to extract an atmospheric neutrino from
    a background of 10
    10
    atmospheric muons in a
    small
    ð
    6
    ?
    6
    ?
    7
    :
    5m
    3
    Þ
    surface detector.
    3.1.5. Combined likelihoods
    The likelihood function
    L
    time
    of the hit times is
    the most important for trackreconstruction.
    However, it is useful to include other information
    like the hit probabilities. The combined p.d.f. from
    Eqs. (7)–(10) is
    L
    MPE
    "
    P
    hit
    P
    no
    ?
    hit
    ¼
    L
    MPE
    ð
    L
    hit
    Þ
    w
    ð
    15
    Þ
    which is particularly effective. Here
    w
    is an
    optional weight factor which allows the adjust-
    ment of the relative weight of the two likelihoods.
    This likelihood is sensitive not only to the track
    geometry but also to the energy of the muon.
    As discussed in Section 3.1.4, the zenith angle-
    dependent prior function,
    P
    ð
    y
    Þ
    ;
    can be included as
    a multiplicative factor. This combination
    L
    Bayes
    ¼
    P
    ð
    y
    Þ
    L
    time
    ð
    16
    Þ
    has been used in the analysis of atmospheric
    neutrinos
    [30]. However, all of these improved
    likelihoods are limited by the underlying model
    assumption of a single muon track.
    3.2. Likelihood implementation
    The actual implementation of the likelihoods
    requires detailed knowledge of the photon propa-
    gation in the ice. On the other hand, efficiency
    considerations and numeric problems favor a
    simple and robust method.
    The photon hit probabilities and arrival time
    distributions are simulated as functions of all
    relevant parameters with a dedicated Monte Carlo
    simulation and archived in large look-up tables.
    This simulation is described in Refs. [26,30,33]
    The
    AMANDA
    Collaboration has followed differ-
    ent strategies for incorporating this data into the
    reconstruction. In principle the probability density
    functions are taken directly from these archives.
    However, one has to face several technical
    difficulties due to the memory requirements of
    the archived tables, as well as numeric problems
    related to the normalization of interpolated bins
    and the calculation of multi-photon likelihoods.
    Alternatively, one can simplify the model and
    parametrize these archives with analytical func-
    tions, which depend only on a reduced set of
    parameters. Comparisons of two independent
    ARTICLE IN PRESS
    J. Ahrens et al. / Nuclear Instruments and Methods in Physics Research A 524 (2004) 169–194
    176

    parametrizations [24,34] show that the direct and
    parametrized approaches yield similar results in
    terms of efficiency. This indicates that the para-
    metrization itself is not limiting the reconstruction
    quality; rather, as mentioned earlier, the recon-
    struction is limited by the assumptions of the
    model being fit. Therefore, we will concentrate on
    only one parametrization.
    3.2.1. Analytical parametrization
    A simple parametrization of the arrival time
    distributions can be achieved with the following
    function, which we call
    Pandel function
    .Itisa
    gamma distribution and its usage is motivated by
    an analysis of laser light signals in the BAIKAL
    experiment[35]. There, it was found that for the
    case of an isotropic, monochromatic and point-
    like light source,
    p
    1
    ð
    t
    res
    Þ
    can be expressed in the
    form
    p
    ð
    t
    res
    Þ?
    1
    N
    ð
    d
    Þ
    t
    d
    =
    l
    Þ
    t
    ð
    d
    =
    l
    ?
    1
    Þ
    res
    G
    ð
    d
    =
    l
    Þ
    ?
    e
    ?
    t
    res
    1
    t
    þ
    c
    medium
    l
    a
    ??
    þ
    d
    l
    a
    ??
    ð
    17
    Þ
    N
    ð
    d
    Þ¼
    e
    ?
    d
    =
    l
    a
    1
    þ
    t
    c
    medium
    l
    a
    ??
    ?
    d
    =
    l
    ð
    18
    Þ
    without special assumptions on the actual optical
    parameters. Here,
    c
    medium
    ¼
    c
    vac
    =
    n
    is the speed of
    light in ice,
    l
    a
    the absorption length,
    G
    ð
    d
    =
    l
    Þ
    the
    Gamma function and
    N
    ð
    d
    Þ
    a normalization factor,
    which is given by Eq. (18). This formulation has
    free parameters
    l
    and
    t
    ;
    which are unspecified
    functions of the distance
    d
    and the other geome-
    trical parameters. They are empirically determined
    by a Monte Carlo model.
    The Pandel function has some convenient
    mathematical properties: it is normalized, it is
    easy to compute, and it can be integrated
    analytically over the time,
    t
    res
    ;
    which simplifies
    the construction of the multi-photon (MPE) time
    p.d.f.. For small distances the function has a pole
    at
    t
    ¼
    0 corresponding to a high probability of an
    unscattered photon. Going to larger values of
    d
    ;
    longer delay times become more likely. For
    distances larger than the critical value
    d
    ¼
    l
    ;
    the
    power index to
    t
    res
    changes sign, reflecting that the
    probability of undelayed photons vanishes: essen-
    tially all photons are delayed due to scattering.
    The large freedom in the choice of the two
    parameter functions
    t
    (units of time) and
    l
    (units
    of length) and the overall reasonable behavior is
    the motivation to use this function to parametrize
    not only the time p.d.f. for point-like sources, but
    also for muon tracks [34]. The Pandel function is
    fit to the distributions of delay times for fixed
    distances
    d
    and angles
    Z
    (between the PMT axis
    and the Cherenkov cone). These distributions are
    previously obtained from a detailed photon
    propagation Monte Carlo for the Cherenkov light
    from muons. The free fit parameters are
    t
    ;
    l
    ;
    l
    a
    and the effective distance
    d
    eff
    ;
    which will be
    introduced next.
    When investigating the fit results as a function
    of
    d
    and angle
    Z
    (see
    Fig. 3), we observe that
    already for a simple ansatz of constant
    t
    ;
    l
    and
    l
    a
    the optical properties in
    AMANDA
    are described
    sufficiently well within typical distances. The
    dependence on
    Z
    is described by an effective
    distance
    d
    eff
    which replaces
    d
    in Eq. (17). This
    means that the time delay distributions for back-
    ward illumination of the PMT is found to be
    similar to a head-on illumination at a larger
    distance. The following parameters are obtained
    for a specific ice model, and are currently used in
    the reconstruction:
    t
    ¼
    557 ns
    ;
    d
    eff
    ¼
    a
    0
    þ
    a
    1
    d
    l
    ¼
    33
    :
    3m
    ;
    a
    1
    ¼
    0
    :
    84
    l
    a
    ¼
    98 m
    ;
    a
    0
    ¼
    3
    :
    1m
    ?
    3
    :
    9 m cos
    ð
    Z
    Þþ
    4
    :
    6 m cos
    2
    ð
    Z
    Þ
    :
    ð
    19
    Þ
    A comparison of the results from this parame-
    trization with the full simulation is shown inFig. 5
    for two extreme distances. The simple approxima-
    tion describes the behavior of the full simulation
    reasonably well. However, this simple overall
    description has a limited accuracy, especially for
    d
    E
    l
    (not shown). Reconstructions, based on the
    Pandel function with different ice models, and a
    generic reconstruction, that uses the full simula-
    tion results, yield similar results. These compar-
    isons indicate that the results of the reconstruction
    do not critically depend on the fine tuning of the
    ARTICLE IN PRESS
    J. Ahrens et al. / Nuclear Instruments and Methods in Physics Research A 524 (2004) 169–194
    177

    underlying ice models, and it justifies the use of the
    above simple model.
    3.2.2. Extension to realistic PMT signals
    Although the Pandel function is the basis of a
    simple normalized likelihood, it has several defi-
    ciencies. It is not defined for negative
    t
    res
    ;
    it ignores
    PMT jitter, and it has a pole at
    t
    res
    ¼
    0
    ;
    which
    causes numerical difficulties. These problems can
    be resolved by convolving the Pandel function,
    Eq. (17), with a Gaussian, which accounts for the
    PMT jitter. Unfortunately such convolution re-
    quires significant computing time.
    Instead the Pandel function is modified by
    extending it to negative times,
    t
    res
    o
    0
    ;
    with a (half)
    Gaussian of width
    s
    g
    :
    The effects of PMT jitter are
    only relevant for small values of
    t
    res
    :
    For times
    t
    res
    X
    t
    1
    the original function is used, and the two
    parts are connected by a spline interpolation (3rd
    order polynomial). The result,
    #
    P
    ð
    t
    res
    Þ
    ;
    is called
    upandel function
    .
    Using
    t
    1
    ¼
    ffiffiffiffiffiffi
    2
    p
    p
    s
    g
    and requiring further a
    smooth interpolation and the normalization to
    be unchanged, the polynomial coefficients
    a
    j
    and
    normalization of the Gaussian
    N
    g
    can be calcu-
    lated analytically
    [34]. The free parameter
    s
    g
    includes all timing uncertainties, not just the
    PMT jitter. Good reconstruction results are
    achieved for a large range 10
    p
    s
    g
    p
    20 ns
    :
    3.2.3. P
    hit
    P
    no
    ?
    hit
    Parametrization
    The normalization
    N
    ð
    d
    Þ
    in Eq. (18) is used to
    construct a hit probability function,
    P
    hit
    :
    The
    function
    P
    hit
    n
    with
    P
    hit
    1
    ?
    N
    ð
    d
    Þ
    ;
    is fit to the hit
    probability determined by the full
    AMANDA
    detector
    simulation, as a function of distance, orientation
    and muon energy. The free parameters are the
    Pandel parameters
    t
    and
    l
    ;
    l
    a
    ;
    #
    d
    and
    ˜
    n
    :
    The
    effective distance
    #
    d
    ;
    is similar to the effective
    distance in the Pandel parametrization. We define
    ˜
    n as the power index of Eq. (11), which corre-
    sponds to an effective number of photons. It is
    important to understand that
    N
    ð
    d
    Þ
    is not a hit
    probability and
    ˜
    n is not just a number of photons.
    They are constructs, that are calibrated with a
    Monte Carlo simulation. Technically, the power
    index,
    ˜
    n
    ?
    N
    in Eq. (11), factorizes into
    ˜
    n
    ð
    Z
    ;
    E
    m
    Þ¼
    e
    ð
    Z
    ;
    E
    m
    Þ
    n
    ð
    E
    m
    Þ
    :
    The variable
    n
    ;
    where
    n
    ¼
    n
    ð
    E
    Þ
    ;
    is
    related to the number of photons incident on the
    PMT and its absolute efficiency. The factor
    e
    ð
    Z
    ;
    E
    m
    Þ
    is related to the orientation dependent
    PMT sensitivity but also accounts for the energy-
    dependent angular emission profile of photons
    with respect to the bare muon.
    3.2.4. Energy reconstruction
    The reconstruction of the trackgeometry is a
    search for five parameters. If the muon energy is
    added as a fit parameter, the minimization is
    significantly slower. Therefore, the energy recon-
    struction is performed in two steps. First, the track
    geometry is reconstructed without the energy
    parameter. Then these geometric parameters are
    used in an energy reconstruction, that only
    determines the energy. However, if the time
    likelihood utilizes amplitude information, e.g. in
    the combined likelihood, Eq. (7), or the PSA
    likelihood, Eq. (9), the track parameters also
    ARTICLE IN PRESS
    time delay / ns
    d = 8m
    Delay prob / ns
    time delay / ns
    d = 71m
    Delay prob / ns
    10
    4
    10
    3
    10
    2
    10
    1
    0 200 400
    10
    7
    10
    6
    10
    5
    10
    4
    10
    3
    0 500 1000 1500
    Fig. 5. Comparison of the parametrized Pandel function (dashed curves) with the detailed simulation (blackhistograms) at two
    distances
    d
    from the muon track.
    J. Ahrens et al. / Nuclear Instruments and Methods in Physics Research A 524 (2004) 169–194
    178

    depend on the energy. In this case the energy and
    geometric parameters must be reconstructed to-
    gether. Currently, three different approaches are
    used to reconstruct the muon energy. They are
    compared in Section 7.3.
    (1) The simplest method utilizes the
    P
    hit
    P
    no
    ?
    hit
    reconstruction (see Sections 3.1.2 and 3.2.3).
    (2) The
    Full E
    reco
    method (see Section 3.1.3)
    models the measured amplitudes in a like-
    lihood for reconstructing the energy. This
    algorithm performs better, but it is more
    dependent on the quality of the amplitude
    calibration of the OMs.
    (3) An alternative way to measure the energy is
    based on a neural network
    [36]. The neural
    networkuses 6-6-3-1 and 6-3-5-1 feed-forward
    architecture for
    AMANDA-B10
    and
    AMANDA-II
    ,
    respectively. The energy correlated variables
    which are used as input are the mean of the
    measured amplitudes (ADC), the mean and
    RMS of the arrival times (LE) or pulse
    durations (TOT), the total number of signals,
    the number of OMs hit and the number of
    OMs with exactly one hit.
    Less challenging than a full reconstruction, a
    lower energy threshold is determined by requiring
    a minimum number of hit OMs. The number of hit
    OMs is correlated with the energy of the muon.
    Since celestial neutrinos are believed to have a
    substantially harder spectrum than atmospheric
    neutrinos, an excess of high multiplicity events
    would indicate that a hard celestial source exists.
    Values for this parameter determined from
    AMANDA
    data already set a tight upper limit on the diffuse
    flux of high-energy celestial neutrinos [37].
    3.2.5. Cascade reconstruction
    The reconstruction of
    cascade like events
    is
    described in detail elsewhere
    [4]. The basic
    approach is similar to the trackreconstruction. It
    assumes events form a point light source with
    photons propagating spherically outside with a
    higher intensity in the forward direction. The
    cascade reconstruction also uses the Pandel func-
    tion (see Eq. (17)) with parameters that are specific
    for cascades.
    In several muon analyses, a cascade fit is used as
    a competing model. In cases where the cascade fit
    achieves a better likelihood than the track
    reconstruction, the trackhypothesis is rejected.
    In particular this is used as a selection criterion to
    reject background events which are mis-recon-
    structed due to bright secondary cascades.
    4. First guess pattern recognition
    The likelihood reconstructions need an initial
    trackhypothesis to start the minimization. The
    initial trackis derived from
    first guess methods
    ,
    which are fast analytic algorithms that do not
    require an initial track.
    4.1. Direct walk
    A very efficient
    first guess method
    is the
    direct
    walk
    algorithm. It is a pattern recognition algo-
    rithm based on carefully selected hits, which were
    most likely caused by direct photons.
    The four step procedure starts by selecting
    track
    elements
    , the straight line between any two hit
    OMs at distance
    d
    ;
    which are hit with a time
    difference
    j
    D
    t
    j
    o
    d
    c
    vac
    þ
    30 ns with
    d
    >
    50 m
    :
    ð
    20
    Þ
    The known positions of the OMs define the track
    element direction
    ð
    y
    ;
    f
    Þ
    :
    The vertex position
    ð
    x
    ;
    y
    ;
    z
    Þ
    is taken at the center between the two
    OMs. The time at the vertex
    t
    0
    is defined as the
    average of the two hit times.
    In a next step, the number of
    associated hits
    (AH) are calculated for each trackelement.
    Associated hits are those with
    ?
    30
    o
    t
    res
    o
    300 ns
    and
    d
    o
    25 m
    ð
    t
    res
    þ
    30
    Þ
    1
    =
    4
    (
    t
    in ns), where
    d
    is the
    distance between hit OM and trackelement
    and
    t
    res
    is the
    time residual
    , which is defined in
    Eq. (5). After selecting these associated hits, track
    elements of poor quality are rejected by requiring:
    N
    AH
    X
    10 and
    s
    L
    ?
    ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
    ðð
    1
    =
    N
    AH
    Þ
    X
    i
    ð
    L
    i
    ?
    /
    L
    S
    Þ
    2
    Þ
    q
    X
    20 m
    :
    ARTICLE IN PRESS
    J. Ahrens et al. / Nuclear Instruments and Methods in Physics Research A 524 (2004) 169–194
    179

    Here, the ‘‘lever arm’’
    L
    i
    is the distance between
    the vertex of the trackelement and the point on
    the trackelement which is closest to OM
    i
    and
    /
    L
    S
    is the average of all
    L
    i
    -values. Track
    elements that fulfill these criteria qualify as
    track
    candidates
    (TC).
    Frequently, more than one trackcandidate is
    found. In this case, a cluster search is performed for
    all trackcandidates that fulfill the quality criterion
    Q
    TC
    X
    0
    :
    7
    Q
    max
    where
    Q
    max
    ¼
    max
    ð
    Q
    TC
    Þ
    and
    Q
    TC
    ¼
    min
    ð
    N
    AH
    ;
    0
    :
    3m
    ?
    1
    ?
    s
    L
    þ
    7
    Þ
    :
    ð
    21
    Þ
    In the cluster search, the ‘‘neighbors’’ of each track
    candidate are counted, where neighbors are track
    candidates with space angle differences of less than
    15
    ?
    :
    The cluster with the largest number of track
    candidates is selected.
    In the final step, the average direction of all
    trackcandidates inside the cluster defines the
    initial trackdirection. The trackvertex and time
    are taken from the central track candidate in the
    cluster. Well separated clusters can be used to
    identify independent muon tracks in events which
    contain multiple muons (see Section 6.1).
    4.2. Line­fit
    The
    line­fit
    [38] algorithm produces an initial
    trackon the basis of the hit times with an optional
    amplitude weight. It ignores the geometry of the
    Cherenkov cone and the optical properties of the
    medium and assumes light traveling with a velocity
    v
    along a 1-dimensional path through the detector.
    The locations of each PMT,
    r
    i
    ;
    which are hit at a
    time
    t
    i
    can be connected by a line
    r
    i
    E
    r
    þ
    v
    ?
    t
    i
    :
    ð
    22
    Þ
    A
    w
    2
    to be minimized is defined as
    w
    2
    ?
    X
    N
    hit
    i
    ¼
    1
    ð
    r
    i
    ?
    r
    ?
    v
    ?
    t
    i
    Þ
    2
    ð
    23
    Þ
    where
    N
    hit
    is the number of hits. The
    w
    2
    is
    minimized by differentiation with respect to the
    free fit parameters
    r
    and
    v
    :
    This can be solved
    analytically
    r
    ¼
    /
    r
    i
    S
    ?
    v
    ?
    /
    t
    i
    S
    and
    v
    ¼
    /
    r
    i
    ?
    t
    i
    S
    ?
    /
    r
    i
    S
    ?
    /
    t
    i
    S
    /
    t
    2
    i
    S
    ?
    /
    t
    i
    S
    2
    ð
    24
    Þ
    where
    /
    x
    i
    S
    1
    =
    N
    hit
    Þ
    P
    N
    hit
    i
    x
    i
    denotes the mean
    of parameter
    x
    with respect to all hits.
    The line-fit thus yields a vertex point
    r
    ;
    and a
    direction
    e
    ¼
    v
    LF
    =
    j
    v
    LF
    j
    :
    The zenith angle is given
    by
    y
    LF
    ??
    arccos
    ð
    v
    z
    =
    j
    v
    LF
    :
    The time residuals (Eq. (5)) for this initial track
    generally do not follow the distribution expected
    for a Cherenkov model. If the
    t
    0
    parameter of the
    initial trackis shifted to better agree with a
    Cherenkov model, subsequent reconstructions
    converge better (see Section 5.2.3).
    The absolute speed
    v
    LF
    ?j
    v
    j
    ;
    of the line-fit is the
    mean speed of the light propagating through the
    one-dimensional detector projection. Spherical
    events (cascades) and high energy muons have
    low
    v
    LF
    values, and thin, long events (minimally
    ionizing muon tracks) have large values.
    4.3. Dipole algorithm
    The
    dipole algorithm
    considers the
    unit
    vector
    from one hit OM to the subsequently hit OM as an
    individual dipole moment. Averaging over all
    individual dipole moments yields the global mo-
    ment
    M
    :
    It is calculated in two steps. First, all hits
    are sorted according to their hit times. Then a
    dipole­moment
    M
    is calculated
    M
    ?
    1
    N
    ch
    ?
    1
    X
    N
    ch
    i
    ¼
    2
    r
    i
    ?
    r
    i
    ?
    1
    j
    r
    i
    ?
    r
    i
    ?
    1
    j
    :
    ð
    25
    Þ
    It can be expressed via an absolute value
    M
    DA
    ?
    j
    M
    j
    and two angles
    y
    DA
    and
    f
    DA
    :
    These angles
    define the initial track.
    The dipole algorithm does not generate as good
    an initial trackas the direct walkor the line-fit, but
    it is less vulnerable to a specific class of back-
    ground events: almost coincident atmospheric
    muons from independent air showers in which
    the first muon hits the bottom and the second
    muon hits the top of the detector.
    ARTICLE IN PRESS
    J. Ahrens et al. / Nuclear Instruments and Methods in Physics Research A 524 (2004) 169–194
    180

    4.4. Inertia tensor algorithm
    The
    inertia tensor algorithm
    is based on a
    mechanical picture. The pulse amplitude from a
    PMT at
    r
    i
    corresponds to a virtual mass
    a
    i
    at
    r
    i
    :
    One can then define the tensor of inertia
    I
    of that
    virtual mass distribution. The origin is the center
    of gravity
    ð
    COG
    Þ
    of the mass distribution. The
    COG
    -coordinates and the tensor of inertia com-
    ponents are given by
    COG
    ?
    X
    N
    ch
    i
    ¼
    1
    ð
    a
    i
    Þ
    w
    ?
    r
    i
    and
    I
    k
    ;
    l
    ?
    X
    N
    ch
    i
    ¼
    1
    ð
    a
    i
    Þ
    w
    d
    kl
    r
    i
    Þ
    2
    ?
    r
    k
    i
    ?
    r
    l
    i
    ?
    :
    ð
    26
    Þ
    The amplitude weight
    w
    X
    0 can be chosen
    arbitrarily. The most common settings are
    w
    ¼
    0
    (ignoring the amplitudes) and
    w
    ¼
    1 (setting the
    virtual masses equal to the amplitudes). The tensor
    of inertia has three eigenvalues
    I
    j
    ;
    j
    e
    f
    1
    ;
    2
    ;
    3
    g
    ;
    corresponding to its three main axes
    e
    j
    :
    The
    smallest eigenvalue
    I
    1
    corresponds to the longest
    axis
    e
    1
    :
    In the case of a long track-like event
    I
    1
    5
    f
    I
    2
    ;
    I
    3
    g
    and
    e
    1
    approximates the direction of
    the track. The ambiguity in the direction along the
    e
    1
    axis is resolved by choosing the direction where
    the average OM hit time is latest. In the case of a
    cascade-like event,
    I
    1
    E
    I
    2
    E
    I
    3
    :
    The ratios between
    the
    I
    j
    can be used to determine the sphericity of the
    event.
    5. Aspects of the technical implementation
    5.1. Reconstruction framework
    The basic reconstruction procedure, sketched in
    Fig. 6, is sequential. A fast reconstruction program
    calculates the initial trackhypothesis for the
    likelihood reconstruction. All reconstruction pro-
    grams may use a reduced set of hits in order to
    suppress noise hits and other detector artifacts.
    Event selection criteria can be applied after each
    step to reduce the event sample, and allow more
    time consuming calculations at later reconstruc-
    tion stages. This procedure may iterate with more
    sophisticated but slower algorithms analyzing
    previous results. The final step is usually the
    production of
    Data Summary Tape
    (DST) like
    information, usually in form of PAW
    N
    -tuples
    [39]. A detailed description of this procedure can
    be found inRef. [40].
    The reconstruction frameworkis implemented
    with the
    recoos
    program[41], which is based on the
    rdmc
    library [42] and the SiEGMuND software
    package [43]. The
    recoos
    program is highly
    modular, which allows flexibility in the choice
    and combination of algorithms.
    5.2. Likelihood maximization
    The aim of the reconstruction is to find the
    trackhypothesis which corresponds to the max-
    imum likelihood. This is done by minimizing
    ?
    log
    ð
    L
    Þ
    with respect to the trackparameters.
    We have implemented several minimization
    procedures.
    The likelihood space for
    AMANDA
    events is often
    characterized by several minima. Local likelihood
    minima can arise due to symmetries in the
    detector, especially in the azimuth angle, or due
    to unexpected hit times caused by scattering. In the
    example, shown in Fig. 7, the reconstruction
    converged on a local minimum, because of non-
    optimal starting values. Several techniques, which
    are used to find the global minimum, are here
    presented. In particular, the iterative reconstruc-
    tion, Section 5.2.2, solves the problem and
    converges to the global minimum. One generally
    assumes that the global minimum corresponds to
    the true solution, but this is not always correct due
    to stochastic nature of light emission and detec-
    tion. Such events cannot be reconstructed properly
    and have to be rejected using quality parameters
    (see Section 6.2).
    ARTICLE IN PRESS
    Hit­cleaning Hit­cleaning Selections
    ↓↓
     
    Data
    first guess
    Likelihood
    Analysis
    Selections Selections
    Fig. 6. Schematic principle of the reconstruction chain.
    J. Ahrens et al. / Nuclear Instruments and Methods in Physics Research A 524 (2004) 169–194
    181

    5.2.1. Minimization algorithms
    The reconstruction frameworkallows us to use
    and compare these numerical minimization algo-
    rithms:
    Simplex
    [44],
    Powell’s
    [44],
    Minuit
    [45]
    (using the
    minimize
    method), and
    Simulated
    annealing
    [44]. The Simplex algorithm is the fastest
    algorithm. Powell’s method and Minuit are
    B
    5
    times slower than the Simplex algorithm. The
    reconstruction results from Minuit and the Sim-
    plex algorithm are nearly identical and almost as
    good as the Powell results. Exceptions occur in less
    than 1% of the cases, when these methods fail and
    stop at the extreme zenith angles
    y
    ¼
    0
    ?
    and 80
    ?
    :
    The Simulated annealing algorithm is less sensitive
    to local minima than the other algorithms, but it is
    much slower and requires fine-tuning.
    5.2.2. Iterative reconstruction
    The
    iterative reconstruction
    algorithm success-
    fully copes with the problem of local minima and
    extreme zenith angles by performing multiple
    reconstructions of the same event. Each recon-
    struction starts with a different initial track.
    Therefore, the fast Simplex algorithm is sufficient.
    The ability to find the global minimum depends
    strongly on the quality of the initial track. A
    systematic scan of the full parameter space for
    initial seeds is not feasible. Instead the iterative
    algorithm concentrates on the direction angles,
    zenith and azimuth, and uses reasonable values for
    the spatial coordinates. The following procedure
    yields good results.
    The result of a first minimization is saved as a
    reference. Then both direction angles are ran-
    domly selected. The trackpoint,
    r
    0
    ;
    is transformed
    to the point on the new track, which is closest to
    the center of gravity of hits. The time,
    t
    0
    ;
    of this
    point is shifted to match the Cherenkov expecta-
    tion (see Section 5.2.3). Then a new minimization
    is started. If the minimum is less than the reference
    minimum, it is saved as the new reference. This
    procedure is iterated
    n
    times, and the best minima
    found for zenith angles
    above
    and
    below
    the
    horizon, are saved, and used to generate an
    important selection parameter (see Section 6.2).
    This algorithm substantially reduces the number
    of false minima found, after a few iterations. For
    n
    ¼
    6 roughly 95% of the results are in the vicinity
    of the asymptotic optimum for
    n
    ­
    N
    :
    For
    n
    C
    20
    more than 99% of the results are the global
    minimum. Despite the fast convergence, the
    iterative reconstruction
    requires significant CPU
    time, which limits its use to reduced data sets.
    5.2.3. Lateral shift and time residual
    The efficiency of finding the global minimum of
    the likelihood function can be improved by
    translating the arbitrary vertex and/or time origin
    of the output trackfrom the first guess algorithm
    before application of the full maximum likelihood
    method.
    Transformation of
    r
    0
    :
    In general this ‘‘vertex’’
    point is arbitrary in the infinite trackapproxima-
    tion used, and first guess methods may produce
    positions distant from the detector. During the
    likelihood minimization, numerical errors can be
    avoided and the convergence improved by shifting
    this point along the direction of the tracktowards
    the point closest to the
    center of gravity of hits
    (see
    Eq. (26)). The vertex time
    t
    0
    is transformed
    accordingly:
    D
    ð
    t
    0
    Þ¼
    D
    ð
    r
    0
    Þ
    =
    c
    :
    Transformation of t
    0
    :
    The time
    t
    0
    obtained from
    first guess algorithms is not calculated from a full
    Cherenkov model. The efficiency of the likelihood
    reconstruction can be improved by shifting the
    t
    0
    such that the time residuals, Eq. (5), fit better to a
    ARTICLE IN PRESS
    180
    160
    140
    120
    100
    80
    60
    40
    20
    0
    170
    175
    180
    185
    190
    Parabola fit
    experimental data
    log (likelihood)
    theta [
    °
    ]
    Fig. 7. An example of the likelihood space (one-dimensional
    projection) for a specific
    AMANDA
    event. Shown is
    ?
    log
    ð
    L
    Þ
    as
    function of the zenith angle. Each point represents a fit, for
    which the zenith angle was fixed and the other trackparameters
    were allowed to vary in order to find the best minimum. A local
    minimum which was found by a gradient likelihood minimiza-
    tion is indicated by a fitted parabola. Improved methods that
    avoid this are described in the text.
    J. Ahrens et al. / Nuclear Instruments and Methods in Physics Research A 524 (2004) 169–194
    182

    Cherenkov model. In particular it is useful to
    avoid negative
    t
    res
    ;
    which would correspond to
    causality violations. This can be achieved by
    transforming
    t
    0
    ­
    t
    0
    ?
    t
    ?
    res
    ;
    where
    t
    ?
    res
    is the most
    negative time residual.
    5.2.4. Coordinates and restricted parameters
    The trackcoordinates
    a
    ;
    which are used by the
    likelihood, are independent of the coordinates
    actually chosen for the minimization. Therefore,
    the coordinate system can be chosen arbitrarily.
    Any of the parameters in this coordinate system
    can be kept fixed. During the minimization
    parameterization
    functions translate the coordi-
    nates as necessary. The most commonly used
    coordinates are
    r
    0
    and the zenith and azimuth
    angles
    y
    ;
    f
    :
    The freedom in the choice of coordinates can be
    used to improve the numerical minimization, for
    systematic studies, or to fix certain parameters
    according to external knowledge. An example is
    the reconstruction of coincident events with the
    SPASE
    surface arrays [8]. Here, we fix the location
    of the trajectory to coincide with the core location
    at the surface as measured by
    SPASE
    . Then, the
    direction is determined with the
    AMANDA
    recon-
    struction subject to this constraint.
    Under certain circumstances the allowed range
    of the reconstruction parameters is restricted. The
    most important example here is to restrict the
    reconstructed zenith angle to above or below the
    horizon, to find the most likely up- or down-going
    tracks, respectively. Comparing the quality of the
    two solutions can be used for background rejec-
    tion. Technically the constrained fit is accom-
    plished by multiplying the likelihood by a prior,
    which is zero outside the allowed parameter range.
    5.3. Preprocessing and hit cleaning
    The data must be filtered and calibrated before
    reconstruction. Defective OMs are removed, and
    the amplitudes and hit times are calibrated. A
    hit
    cleaning
    procedure identifies and flags hits which
    appear to be noise or electronic effects, such as
    cross talkor after-pulsing. These hits are not used
    in the reconstruction, but they are retained for
    post-reconstruction analysis.
    The hit cleaning procedure can be based on
    simple and robust algorithms, because the PMTs
    have low noise rates. Noise and after-pulse hits are
    strongly suppressed by rejecting hits that are
    isolated in time and space from other signals in
    the detector. Typically a hit is considered to be
    noise if there is no hit within a distance of 60–
    100 m and a time of
    7
    300 to
    7
    600 ns
    :
    Cross talk
    hits are identified by examining the amplitudes and
    pulse widths of the individual pulses and by
    analyzing the correlations of uncalibrated hit times
    with hits of large amplitude in channel combina-
    tions which are known to cross talk to each other.
    The required cross talkcorrelation map was
    determined independently in a dedicated calibra-
    tion campaign.
    5.4. Processing speeds
    The first guess algorithms are sufficiently fast
    that the execution time is dominated by file input/
    output and the software framework. Typical fit
    times are
    E
    20 ms per event on a 850 MHz
    Pentium-III Linux PC. The processing speed of
    the likelihood reconstructions can vary signifi-
    cantly depending on the number of free para-
    meters, the number of iterations, the minimization
    algorithm, and the experimental parameters like
    the number of hit OMs. These effects dominate the
    differences in processing speeds due to different
    reconstruction algorithms. The typical execution
    time for a 16-fold iterative likelihood reconstruc-
    tions using the simplex minimizer to reconstruct
    the five free trackparameters is
    C
    250 ms per
    event.
    6. Background rejection
    The performance of the reconstruction depends
    strongly on the quality and background selection
    criteria. The major classes of background events in
    AMANDA
    (see Section 6.1) are suppressed by the
    quality parameters presented in Section 6.2.
    Optimization strategies for the selection criteria
    are summarized in Section 6.3. Finally, we
    evaluate the reconstruction performance in
    Section 7.
    ARTICLE IN PRESS
    J. Ahrens et al. / Nuclear Instruments and Methods in Physics Research A 524 (2004) 169–194
    183

    6.1. Background classes
    Most background events from atmospheric
    muons are well reconstructed and can be
    rejected by selecting up-going reconstructed
    events. However, there is a small fraction of mis-
    reconstructed events, amounting to about 10
    ?
    2
    for
    the unbiased and about 10
    ?
    4
    for the zenith-
    weighted reconstruction. These events are rejected
    by additional selection criteria described in Section
    6.2. These background events are classified as
    follows.
    Nearly horizontal muons
    : These events have true
    incident angles close to the horizon. A small error
    in the reconstruction causes them to appear as up-
    going. These events are not severely mis-recon-
    structed, but occur due to the finite angular
    resolution.
    Muon bundles
    : The spatial separation between
    multiple muons from a single air shower, a
    muon
    bundle
    , is usually small enough that the event can
    be described by a single bright muon track. If the
    separation is too large, the reconstruction fails.
    Cascades
    : Bright stochastic energy losses (e.g.
    bremsstrahlung) produce additional light, which
    distorts the Cherenkov cone from the muon.
    Cascades emit most of their light with the same
    Cherenkov angle as the muon, but some light is
    emitted at other angles. These secondary events
    can cause the reconstruction to fail, especially
    when the cascade(s) produce more light than the
    muon itself. A special class of these events are
    muons which pass outside the detector and release
    a bright cascade, which can mimic an up-going hit
    pattern.
    Stopping muons
    : Over the depth of the detector
    the muon flux changes by a factor of
    B
    2
    ;
    since
    muons lose their energy and stop. These muons
    can create an up-going hit pattern, especially when
    the muon stops just before entering the detector
    from the side.
    Scattering layers
    : The scattering of light in the
    polar ice cap varies with depth. Light from bright
    events, can mimic an up-going hit pattern, in
    particular when it traverses layers of higher
    scattering.
    Corner clippers
    : These are events where the
    muon passes diagonally below the detector. The
    light travel upwards through the detector mimick-
    ing an up-going muon.
    Uncorrelated coincident muons
    : Due to the large
    size of the
    AMANDA
    detector, the probability of
    muons from two independent air showers forming
    a single event is small on the trigger level but not
    negligible. If an initial muon traverses the bottom
    of the detector and a later muon traverses the top,
    the combination can be reconstructed as an up-
    going muon.
    Electronic artifacts
    : Noise, cross talkand other
    transient electronic malfunctions are generally
    small effects, but they can occasionally produce
    hits, which distort the time pattern. Such effects
    become important after a selection process of
    several orders of magnitude.
    6.2. Quality parameters
    Background events, which pass a zenith angle
    selection, need to be rejected by applying selection
    criteria on quality parameters. These parameters
    usually evaluate information, which is not opti-
    mally exploited in the reconstruction. The detailed
    choice of quality parameters is specific to each
    analysis. Here, we summarize the most important
    categories.
    The
    number of direct hits
    ,
    N
    dir
    ð
    t
    1
    :
    t
    2
    Þ
    ;
    is the
    number of hits with small time residuals:
    t
    1
    o
    t
    res
    o
    t
    2
    (see Eq. (5)). Un-scattered photons
    provide the best information for the reconstruc-
    tion, and a large number of
    N
    dir
    indicates high
    quality information in the event. Empirically
    reasonable values are
    t
    1
    C
    ?
    15 ns and
    t
    2
    between
    ¼þ
    25 and
    þ
    150 ns
    ;
    depending on the specific
    analysis.
    The
    length of the event L
    is obtained by
    projecting each hit OM onto the reconstructed
    trackand taking the distance between the two
    outermost of these points.
    L
    can be considered as
    the ‘‘lever arm’’ of the reconstruction. Larger
    values corresponding to a more robust and precise
    reconstruction of the track’s direction. This para-
    meter is particularly powerful when calculated for
    direct hits only, and is then referred to as
    L
    dir
    ð
    t
    1
    :
    t
    2
    Þ
    :
    Length requirements are efficient
    against corner clippers, stopping muons and
    cascades.
    ARTICLE IN PRESS
    J. Ahrens et al. / Nuclear Instruments and Methods in Physics Research A 524 (2004) 169–194
    184

    The absolute value of the likelihood at the
    maximum is a good parameter to evaluate the
    quality of a reconstruction. Here, a useful ob-
    servable is the
    likelihood parameter
    L
    which is
    defined as
    L
    ??
    log
    ð
    L
    Þ
    N
    free
    ð
    27
    Þ
    where
    N
    free
    is the degrees of freedom (e.g.
    N
    free
    ¼
    N
    hits
    ?
    5 for a trackreconstruction). For Gaussian
    probability distributions this expression corre-
    sponds to the reduced chi-square.
    L
    can be used
    as a selection parameter, smaller values corre-
    sponding to higher quality. A selection of events
    with good
    L
    P
    hit
    P
    no
    ?
    hit
    values is efficient against
    stopping muons.
    Comparing
    L
    from different reconstructions is a
    powerful technique. Cascade-like events will have
    a better likelihood from a cascade reconstruction
    than one from a trackreconstruction.
    Another efficient rejection method is to compare
    L
    for the best up-going versus the best down-going
    reconstruction of a single event. If the up-going
    reconstruction is not significantly better than the
    down-going reconstruction, the event is rejected.
    These values can be obtained from the iterative
    reconstruction method (Section 5.2.2) or by
    restricting the parameter space. This method is
    particularly powerful when the down-going recon-
    struction uses a zenith weighted likelihood (Sec-
    tion 3.1.4).
    The reconstruction methods consider the
    p.d.f. for each hit separately but ignore correla-
    tions. Therefore, the reconstructions assign the
    same likelihood to tracks where all hits cluster at
    one end of the reconstructed trackand tracks
    where the same number of hits are smoothly
    distributed along the track. The latter hit
    pattern indicates a successful trackreconstruction,
    while the former hit pattern may be caused
    by a background event. The
    smoothness
    parameter
    S
    was inspired by the Kolmogorov–Smirnov
    test of the consistency of two distributions.
    S
    is a measure of the consistency of the observed
    hit pattern with the hypothesis of constant light
    emission by a muon. The simplest definition
    of the smoothness
    S
    is
    S
    ¼
    S
    max
    j
    ;
    where
    S
    max
    j
    is
    that
    S
    j
    ;
    which has the largest absolute value, and
    S
    j
    is defined as
    S
    j
    ?
    j
    ?
    1
    N
    ?
    1
    ?
    l
    j
    l
    N
    :
    ð
    28
    Þ
    l
    j
    is the distance along the trackbetween the points
    of closest approach of the trackto the first and the
    j
    th hit module, with the hits taken in order of their
    projected position on the track.
    N
    is the total
    number of hits. Tracks with hits clustered at the
    beginning or end of the trackhave
    S
    approaching
    þ
    1or
    ?
    1
    ;
    respectively. High-quality tracks with
    S
    close to zero, have hits equally spaced along the
    track. A graphical representation of the smooth-
    ness construction can be found inRef. [30].
    Extensions of this smoothness parameter in-
    clude the restriction of the calculation to direct hits
    only or using the distribution of hit times
    t
    i
    instead
    of the distances
    l
    i
    :
    A particularly important extension is
    S
    P
    hit
    :
    In
    order to account for the granularity and asym-
    metric geometry of the detector one can replace
    the above formulation with one that models the hit
    smoothness expectation for the actual geometry of
    the assumed muon track. This can be accom-
    plished by using the hit probabilities of all
    N
    OM
    ;
    the number of operational OMs, (ordered along
    the track) as weights:
    S
    P
    hit
    ¼
    max
    ð
    S
    P
    hit
    j
    Þ
    with
    S
    P
    hit
    j
    ?
    P
    j
    i
    ¼
    1
    L
    i
    P
    N
    OM
    i
    ¼
    1
    L
    i
    ?
    P
    j
    i
    ¼
    1
    P
    hit
    ;
    i
    P
    N
    OM
    i
    ¼
    1
    P
    hit
    ;
    i
    ð
    29
    Þ
    L
    i
    ¼
    1
    ;
    if the OM
    i
    was hit and 0 otherwise and
    P
    hit
    ;
    i
    is the probability for OM
    i
    to be hit given the
    reconstructed track. The hit probabilities are
    calculated according to the algorithm in Section
    3.2.3. Smoothness selections are very efficient
    against secondary cascades, stopping muons and
    coincident muons from independent air showers.
    Interesting
    AMANDA
    events are analyzed with
    multiple reconstruction algorithms. An event is
    most likely to have been reconstructed correctly, if
    the different algorithms produce consistent results.
    For two reconstructions with directions
    e
    1
    and
    e
    2
    ;
    the space angle between them is given by
    C
    ¼
    arccos
    ð
    e
    1
    ?
    e
    2
    Þ
    ;
    which should be reasonably small
    for successful reconstructions. This concept can be
    extended to multiple reconstructions and their
    angular deviations from the average direction. For
    n
    different reconstructed directions,
    e
    i
    ;
    the average
    ARTICLE IN PRESS
    J. Ahrens et al. / Nuclear Instruments and Methods in Physics Research A 524 (2004) 169–194
    185

    reconstructed direction,
    E
    ;
    is given by
    E
    ¼
    P
    n
    i
    e
    i
    =
    j
    P
    n
    i
    e
    i
    j
    :
    We can define the parameter
    C
    w
    ¼
    X
    i
    ½
    arccos
    ð
    e
    i
    ?
    E
    Þ?
    w
    !
    1
    =
    w
    :
    ð
    30
    Þ
    C
    1
    describes the average space angle between the
    individual reconstructions and
    E
    :
    C
    2
    is a different
    parameter, which treats the deviations between
    E
    and the
    e
    i
    as ‘‘errors’’ and adds them quadrati-
    cally. Small values of
    C
    1
    or
    C
    2
    indicate consistent
    reconstruction results.
    The
    C
    parameters are a mathematically correct
    consistency checkonly when comparing the results
    of uncorrelated reconstructions of the same
    intrinsic resolutions. This is not the case when
    comparing different
    AMANDA
    reconstructions. Irre-
    spective of the validity of such an interpretation,
    C
    1
    or
    C
    2
    are very efficient selection criteria,
    especially against almost horizontal muons and
    wide muon bundles.
    A few additional selection parameters are
    closely related to first guess methods. The ratio
    of the eigenvalues of the
    tensor of inertia
    (see
    Section 4.4) are a measure of the sphericity of the
    event topology, which is an efficient selection
    parameter against cascade backgrounds. Tracks
    reconstructed as down-going by the
    dipole fit
    (see
    Section 4.3) that have a non-negligible
    dipole
    moment
    ,
    M
    DA
    ?j
    ~
    MM
    j
    ;
    indicate coincident muons
    from independent air showers. Larger values of the
    line-fit speed
    v
    LF
    (see Section 4.2) are an indication
    for longer muon-like, smaller values for more
    spherical cascade-like events.
    Finally, two approaches evaluate the ‘‘intrinsic
    resolution’’ or ‘‘stability’’ of the reconstruction of
    each event. One approach quantifies the sharpness
    of the minimum found by the minimizers in
    ?
    log
    ð
    L
    Þ
    by fitting a paraboloid to it. The fitted
    parameters can then be used to classify the
    sharpness of the minimum. The other approach
    splits an event into sub-events (for example,
    containing odd- vs. even-numbered hits) and
    reconstructs the sub-events. If the reconstructed
    directions of the sub-events are different, then the
    reconstruction of the full event has a larger
    uncertainty.
    6.3. Analysis strategies
    Analyses that search for neutrino induced
    muons must cope with a large background of
    atmospheric muons. The optimal choice of recon-
    struction and selection criteria varies strongly with
    different expectations for the energy and angular
    distribution of the signal events. The goal is to
    optimize the signal efficiency over the background
    or noise (square root of the background) based on
    sets of signal and background data.
    ?
    The selection criteria for background sensitive
    variables may be adjusted individually such
    that a specified fraction of signal events pass.
    After these
    first level criteria
    are set, the
    adjustment is repeated until the desired back-
    ground rejection is reached. Each iteration
    defines a ‘‘cut-level’’, which corresponds to
    data sets of increasing purity. This simple
    method is used to derive a defined set of
    selection parameters for the performance Sec-
    tion 7. However, less efficient criteria are mixed
    with more efficient criteria, and correlations of
    the variables are not taken into account.
    Therefore, this method does not achieve the
    optimum signal efficiency.
    ?
    An improvement to this method has been
    demonstrated in an
    AMANDA
    point source
    analysis
    [46,47]. Here, a selection criterion is
    only applied to the most sensitive variable, and
    the most sensitive variable is determined at each
    cut level. An interesting aspect of this point
    source search is that the experimental data
    themselves can be used as a background
    sample, which reduces systematic uncertainties
    from the background simulation. The selection
    criteria are not optimized with respect to signal
    purity but with respect to an optimal signifi-
    cance of a possible signal.
    ?
    Another approach is to combine the selection
    parameters into a single selection parameter,
    called
    event quality
    . This can be done by
    rescaling and normalizing each of the selection
    parameters according to the cumulative dis-
    tribution of the signal expectation. The
    AMANDA
    analyses of atmospheric neutrinos [30,48] used
    this technique.
    ARTICLE IN PRESS
    J. Ahrens et al. / Nuclear Instruments and Methods in Physics Research A 524 (2004) 169–194
    186

    ?
    Additional approaches use discriminant analy-
    sis [49] or neural nets [6,50,51] to optimize the
    efficiency while taking into account the correla-
    tions between selection criteria and their
    individual selectivity. However, both methods
    depend critically on a good agreement between
    experiment and simulation. These methods
    quantify the efficiency of each parameter by
    including and excluding it from the optimiza-
    tion procedure.
    ?
    The ‘‘C
    ut
    E
    val
    ’’ method finds the optimum
    combination of selection parameters and cut
    values by numerically maximizing a significance
    function,
    Q
    :
    An example is
    Q
    ¼
    S
    =
    ffiffiffiffi
    B
    p
    ;
    where
    S
    is the number of signal events, and
    B
    is the
    number of background events after selection.
    The implementation proceeds in several
    steps. First, the most efficient selection para-
    meter,
    C
    1
    ;
    is the parameter that individually
    maximizes
    Q
    :
    The next parameter,
    C
    2
    ;
    is the
    parameter that maximizes
    Q
    in conjunction
    with
    C
    1
    :
    More parameters are successively
    determined until the addition of a new para-
    meter fails to improve
    Q
    :
    This procedure takes
    correlations between the selection criteria into
    account. The final number of selection para-
    meters is reduced to a minimum, while max-
    imizing the efficiency. Next, the optimal
    selection for this combination parameters is
    computed as a function of a boundary condi-
    tion (e.g. the maximum number of accepted
    background events). This boundary condition
    is also used to define a single quality parameter.
    Such a formalized procedure has to be
    carefully monitored, e.g. to handle potentially
    un-simulated experimental effects. The C
    ut
    E-
    val
    procedure is monitored by defining differ-
    ent, complementary optimization functions
    Q
    ;
    which allow real and simulated data to be
    compared
    [30,52–54]
    7. Performance
    This section describes the performance of the
    reconstruction methods. It is based on illustrative
    data selections, and the actual performance of a
    dedicated analysis can be different. Unless noted
    otherwise, the data shown is from Monte
    Carlo simulations of atmospheric neutrinos for
    AMANDA-II
    .
    7.1. First guess algorithms
    Since the first guess algorithms are used as a
    starting point for the full reconstruction, they
    should provide a reasonable estimate of the track
    coordinates. Also, these algorithms are used as the
    basis of early level filtering, and therefore need to
    be sufficiently accurate for that purpose, i.e. they
    should at least reconstruct the events in the correct
    hemisphere.
    As an example,
    Table 1
    gives the passing
    efficiencies with respect to the
    AMANDA-II
    trigger
    for atmospheric neutrinos (signal) and atmo-
    spheric muons (background) for the first guess
    methods (see Section 4), after the selection of
    events with calculated zenith angles larger 80
    ?
    :
    The
    direct walk
    algorithm gives the best background
    suppression and the highest atmospheric neutrino
    passing rate. Correspondingly, it also gives the
    best initial tracks to the likelihood reconstructions.
    7.2. Pointing accuracy of the track reconstruction
    The angular accuracy of the reconstruction can
    be expressed in terms of a point spread function,
    which is given by the space angle deviation
    C
    between the true and the reconstructed direction of
    a muon corrected for solid angle. The space angle
    deviation is a combined result of two effects: a
    systematic shift in the direction and a random
    spread around this shift. In a point source analysis,
    for example, it is possible to correct for systematic
    ARTICLE IN PRESS
    Table 1
    The atmospheric muon and atmospheric neutrino detection
    efficiencies for a selection at
    y
    X
    80
    ?
    for the first guess
    algorithms
    Reconstruction atm.
    m
    (%) atm.
    n
    (%)
    Direct walk1.5% 93%
    Line-fit 4.8% 85%
    Dipole algorithm 16.8% 78%
    J. Ahrens et al. / Nuclear Instruments and Methods in Physics Research A 524 (2004) 169–194
    187

    shifts and be limited by the point spread function
    alone [47].
    The zenith and space angular deviations are
    shown in Figs. 8 and 9. They are obtained by the
    reconstruction algorithms as used in
    AMANDA-B10
    .
    The same event selection is used for all. As a
    general observation, the distributions of deviations
    for different reconstruction algorithms is surpris-
    ingly similar after a particular selection. Larger
    differences are usually seen in the selection
    efficiencies. A similar behavior is observed for
    AMANDA-II
    .
    The dependence of the space angle deviation for
    the full
    AMANDA-II
    detector on the cut level
    5
    for
    the LH reconstruction is shown in
    Fig. 10. The
    tighter the selection criteria, the better the angular
    resolution. The same general trend is true for the
    other reconstructions. Tight criteria select events
    with unambiguous hit topologies, which are
    reconstructed better. The results for cut level 6
    are shown in Figs. 11–13 as function of the energy
    and the zenith angle.
    The angular resolution (see Fig. 11) has a weak
    energy dependence. The energy of the muon is
    taken at the point of its closest approach to the
    detector center. Best results are achieved for
    energies of 100 GeV–10 TeV
    :
    At energies
    o
    100 GeV
    ;
    the muons have paths shorter than
    the full detector, which limits the angular
    resolution. At energies
    >
    10 TeV
    ;
    more light is
    emitted due to individual stochastic energy loss
    processes along the muon track. Here, the hit
    pattern is not correctly described by the underlying
    reconstruction assumption of a bare muon track
    (see Section 6.1).
    The space angular resolution depends on the
    incident muon zenith angle (see
    Fig. 12). Again
    ARTICLE IN PRESS
    Fig. 8. The zenith angle deviations for various reconstructions
    of
    AMANDA-B10
    . The result of an atmospheric neutrino
    simulation after the selection criteria of (Ahrens et al.)
    [30]
    is
    shown. The fits are a
    line­fit
    (LF), an iterated
    upandel
    fit (LH),
    an iterated zenith-weighted
    upandel
    fit and a MPE fit.
    Fig. 9. The distribution of space angle deviations for various
    reconstructions of
    AMANDA-B10
    . The result of an atmospheric
    neutrino simulation after the selection criteria of (Ahrens et al.)
    [30]
    is shown. The fits are a
    line­fit
    (LF), an iterated
    upandel
    fit
    (LH), an iterated zenith-weighted
    upandel
    fit and a MPE fit.
    5
    The cut levels defined here are typical and intended as
    demonstrating example. We use typical selection parameters
    from Section 6.2: the reconstructed zenith angle,
    y
    DW
    >
    80
    ?
    ;
    y
    LH
    >
    80
    ?
    ;
    N
    ch
    ;
    N
    LH
    dir
    ð?
    15
    :
    25
    Þ
    ;
    L
    LH
    dir
    ð?
    15
    :
    75
    Þ
    ;
    L
    LH
    ;
    S
    LH
    and
    C
    1
    ð
    DW
    ;
    LH
    ;
    MPE
    Þ
    :
    Our goal here is to illustrate the analysis,
    and we do not optimize with respect to efficiency and angular
    resolution. Instead each individual criterion is enforced in such
    a way that 95% of the events from the previous level would
    pass, and correlations between the parameters are ignored.
    Specific physics analyses will use selection criteria of higher
    efficiency and will achieve better angular resolutions than the
    C
    2
    ?
    ;
    shown here.
    J. Ahrens et al. / Nuclear Instruments and Methods in Physics Research A 524 (2004) 169–194
    188

    this is shown only for the LH reconstruction, the
    other reconstructions are similar. Up-going muons
    with cos
    y
    m
    C
    ?
    0
    :
    7 are best reconstructed, and
    horizontal muons are the worst, because of the
    geometry of the
    AMANDA-II
    detector. Nearly
    vertical events with cos
    y
    m
    C
    ?
    1 have a poorer
    angular resolution, because they illuminate
    fewer strings, which can cause ambiguities in the
    azimuth.
    Systematic shifts also degrade the angular resolu-
    tion.
    AMANDA
    observes a small zenith-dependent shift
    ARTICLE IN PRESS
    log
    10
    (E
    µ
    /GeV)
    space angle deviation
    [°]
    Mean
    Median
    1.5
    2
    2.5
    3
    3.5
    4
    4.5
    5
    0
    2
    4
    6
    Fig. 11. The dependence of the space angle deviation of the LH
    fit on the muon energy for
    AMANDA-II
    . Shown are mean (stars)
    and median (circles) for simulated atmospheric neutrinos.
    Cut level
    space angle deviation
    [°]
     
    Mean
    Median
    2
    3
    4
    5
    6
    7
    8
    9
    10
    0246
    Fig. 10. The dependence of the space angle deviation of the LH
    reconstruction in
    AMANDA-II
    on the event selection (cut levels).
    cos (
    Θ
    µ
    )
    space angle deviation
    [°]
     
    Mean
    Median
    1
    1.5
    2
    2.5
    3
    3.5
    4
    4.5
    5
    ­1 ­0.75 ­0.5 ­0.25 0
    Fig. 12. The space angle deviations of the LH fit as a function
    of the cosine of the incident zenith angle (for
    AMANDA-II
    ).
    Shown are the mean (stars) and median (circles) for simulated
    atmospheric neutrinos.
    cos (
    Θ
    µ
    )
    Mean
    Median
    ­2
    ­1.5
    ­1
    ­0.5
    0
    0.5
    1
    1.5
    2
    ­1 ­0.5 0
    Θ
    µ
    ­
    Θ
    LH
    Fig. 13. The zenith angle shift of the reconstruction versus the
    cosine of the incident angle. Shown are mean (stars) and
    median (circles) for simulated atmospheric neutrinos.
    J. Ahrens et al. / Nuclear Instruments and Methods in Physics Research A 524 (2004) 169–194
    189

    of the reconstructed zenith angle and no systema-
    tic shift in azimuth. This is shown in Fig. 13 for
    simulated atmospheric neutrinos in
    AMANDA-II
    .
    The size of this shift depends on the zenith angle
    itself, and it is determined by the geometry of
    AMANDA
    , which has a larger size in vertical than in
    horizontal directions. From a comparison with
    AMANDA-B10
    data
    [46,55], we observe that these
    shifts become smaller with a larger horizontal
    detector size. These shifts are confirmed by
    analyzing
    AMANDA
    events coincident with
    SPASE
    (see
    below).
    These angular deviations have been obtained
    from Monte Carlo simulations. They can be
    experimentally verified by analyzing coincident
    events between
    AMANDA
    and
    SPASE
    . An analysis of
    data from the 10 string
    AMANDA-B10
    detector,
    shown in
    Fig. 14, confirms the estimate of
    C
    3
    ?
    obtained from Monte Carlo studies for
    AMANDA-
    B10
    . Unfolding the estimated
    SPASE
    resolution of
    C
    1
    ?
    confirms the estimated
    AMANDA-B10
    resolu-
    tion of
    C
    3
    ?
    near the
    SPASE-AMANDA
    coincidence
    direction [8–10].
    A simulation-independent estimate can be ob-
    tained by splitting the hits of individual events in
    two parts and reconstructing each sub-event
    separately. The difference in the two results gives
    an estimate of the total angular resolution. Such
    analyses are being performed at present and results
    will be published separately.
    7.3. Energy reconstruction
    The energy resolution of the three methods,
    described in Section 3.2.4, is shown in Fig. 15 as
    function of the muon energy at its closest point
    to the
    AMANDA-B10
    center. The resolution for
    AMANDA-B10
    in
    D
    log
    10
    E
    is
    C
    0
    :
    4
    ;
    for the interest-
    ing energy range of a few TeV to 1 PeV
    :
    Below
    C
    600 GeV the energy resolution is limited,
    because the amount of light emitted by a muon
    is only weakly dependent on its energy. Above
    1 TeV the resolution improves because radiative
    energy losses become dominant. Above 100 TeV
    the resolution degrades, because energy loss
    fluctuations dominate.
    Although these methods are quite different,
    their performances are similar. The full
    E
    reco
    and
    P
    hit
    methods achieve similar resolutions up to
    1 PeV
    :
    The
    P
    hit
    method becomes worse above this
    energy, because in
    AMANDA-B10
    almost all of the
    OMs are hit, and the method saturates. In
    contrast, the Neural Net method shows a slightly
    poorer resolution up to 1 PeV but is better above.
    Its resolution is relatively constant over several
    decades of energy. This is an advantage when
    reconstructing an original energy spectrum with an
    unfolding procedure as in Ref. [36].
    The
    AMANDA-II
    detector contains more than
    twice as many OMs as
    AMANDA-B10
    , and the
    ARTICLE IN PRESS
    0
    Space angle (degrees)
    arbitrary units
    20
    40
    60
    80
    100
    120
    140
    5
    0 101520253
    0
    Fig. 14. Distribution of the space angle deviations between air
    shower directions assigned by
    SPASE-2
    and muon directions
    assigned by
    AMANDA-B10
    for coincident events measured in
    1997. The figure is not corrected for the systematic shift.
    log
    10
    (E
    gen
    /GeV)
    σ
    (log
    10
    (E
    rec
    / E
    gen
    ))
    P
    hit
    method
    Full E
    reco
    method
    Neural Net method
    0
    0.2
    0.4
    0.6
    0.8
    1
    1.2
    1.4
    2345678
    Fig. 15. Comparison of the resolution of three different energy
    reconstruction approaches for
    AMANDA-B10
    .
    E
    gen
    is the gener-
    ated energy (MC) and
    E
    rec
    the reconstructed energy.
    J. Ahrens et al. / Nuclear Instruments and Methods in Physics Research A 524 (2004) 169–194
    190

    energy resolution is better, especially at larger
    energies,
    s
    ð
    D
    log
    10
    E
    Þ
    C
    0
    :
    3
    :
    The neural net recon-
    struction results for
    AMANDA-II
    are shown in Fig.
    16. Finally, the recently installed transient wave-
    form recorders (TWR) allow better amplitude
    measurements, which should significantly improve
    the results of the energy reconstructions, in
    particular, the
    full E
    reco
    method [56].
    As discussed in Section 1, the cascade channel
    can achieve substantially better resolutions, be-
    cause the full energy is deposited inside or close of
    the detector. Energy resolutions in
    D
    log
    10
    E
    of
    p
    0
    :
    2 and
    p
    0
    :
    15 can be achieved by
    AMANDA-B10
    and
    AMANDA-II
    . respectively [16].
    7.4. Systematic uncertainties
    Several parameters of the detector are calibrated
    and therefore only known with limited accuracy.
    These parameters include the time offsets, the OM
    positions and the absolute OM sensitivities. We
    have estimated the effects of these uncertainties on
    the resolution of
    AMANDA
    reconstructions
    [55].As
    an example, Fig. 17 shows the effect of an
    additional contribution to the time calibration
    uncertainty for the 10 string
    AMANDA-B10
    detector.
    The zenith angular resolutions for simulated
    atmospheric neutrino events only degrade when
    the additional timing uncertainties exceed 10 ns
    :
    Additional tests with similar results were done
    with non-random systematic shifts such as a
    depth-dependent shift or a string-dependent shift.
    Therefore, the angular resolution is insensitive to
    the uncertainties in the time calibration. The
    geometry of the detector is known to better than
    30 cm horizontally and to better than 1 m
    vertically, which corresponds to timing uncertain-
    ties of
    t
    1or 3
    :
    5ns
    ;
    respectively. Therefore, the
    geometry calibration is also sufficiently accurate.
    Similarly, the effect of uncertainties on other
    parameters, like the absolute PMT efficiency, has
    been investigated. No indication was found that
    the remaining calibration uncertainties seriously
    affect the angular resolution or the systematic
    zenith angle offset. The combined calibration
    uncertainties are expected to affect the accuracy
    of the reconstruction by less than 5
    %
    in the zenith
    angle resolution and to less than 0
    :
    5
    ?
    in the
    absolute pointing offset.
    ARTICLE IN PRESS
    arbitrary units
    log10 (Energy/GeV)
    1 TeV
    10 TeV
    100 TeV
    1 PeV
    700
    600
    500
    400
    300
    200
    100
    0
    12 345 67
    Fig. 16. Energy reconstruction for simulated muons of
    different fixed energy in
    AMANDA-II
    , using the neural net
    method.
    Random timing error
    [
    ns
    ]
     
    RMS (
    Θ
    µ
    ­
    Θ
    LH
    )
    2
    2.5
    3
    3.5
    4
    4.5
    5
    1
    10
    Fig. 17. The zenith angle deviations (RMS) as function of an
    additional uncertainty in the
    t
    0
    time calibration. Data is shown
    for simulated atmospheric neutrino events in
    AMANDA-B10
    with
    the selection of (Ahrens et al.)
    [30]. The transit time of the
    PMTs has been shifted without correcting for in the reconstruc-
    tion. The shift is a fixed value for each PMT, obtained from a
    random Gaussian distribution.
    J. Ahrens et al. / Nuclear Instruments and Methods in Physics Research A 524 (2004) 169–194
    191

    8. Discussion and Outlook
    We have developed methods to reconstruct and
    identify muons induced by neutrinos
    [30], inspite
    of the challenges of the natural environment and
    large backgrounds. These methods allow us to
    establish
    AMANDA
    as a working neutrino telescope.
    The reconstruction techniques described in this
    paper are still subject to improvement in several
    aspects.
    The likelihood description
    : The likelihood func-
    tions for trackreconstruction are based on the
    assumption of exactly one infinitely long muon
    trackper event. Extensions of this model to
    encompass
    starting muon tracks
    (including the
    description of the hadronic vertex),
    stopping
    muons
    ,
    muon bundles
    of non-negligible width, and
    multiple independent muons
    will be important,
    particularly in the context of larger detectors such
    as Ice Cube. Initial efforts fitting multiple muons
    with the direct walkalgorithm have been useful in
    rejecting coincident down-going muons, and work
    toward reconstructing muon bundles has begun in
    the context of events coincident with
    SPASE
    air
    showers.
    The p.d.f. calculation
    : The likelihood function is
    based on parametrizations of probability density
    functions (p.d.f.). The p.d.f. is obtained from
    Monte Carlo simulations, and its accuracy is
    limited by the accuracy of the simulation. Better
    simulations lead directly to a better p.d.f. and
    hence better reconstructions.
    The p.d.f. parametrizations
    : The p.d.f. is para-
    metrized by functions (e.g. the
    Pandel functions
    )
    which only approximate the full p.d.f. More
    accurate parametrization functions will result in
    better reconstructions. For example, the scattering
    coefficient shows a significant depth dependence
    (see Section 2). The current reconstruction is based
    on an average p.d.f. assuming depth-independent
    ice properties. While the trackreconstruction is
    relatively insensitive to the accuracy of the
    parametrization, we expect a depth-dependent
    p.d.f. to have better energy reconstruction.
    Complementary information
    : The current recon-
    struction algorithms do not include all available
    information in an event. In particular, correlations
    between detected PMT signals are ignored. For
    this reason dedicated selection parameters have
    been designed to exploit this information. They are
    used to discriminate between well reconstructed
    and poorly reconstructed events and improve the
    quality of the data sample. Future workwill try to
    improve these parameters and expand the present
    likelihood description.
    Transient waveform recorders
    : At the beginning
    of the year 2003, the detector readout has been
    upgraded with transient waveform recorders [56].
    We expect a substantial improvement of the
    multiple-photon detection and the dynamic range
    in particular for high muon energies.
    The construction of a much larger detector, the
    IceCube
    detector, will start in the year 2004. It will
    consist of 4800 PMT deployed on 80 vertical
    strings and will surround the
    AMANDA
    detector [57].
    The performance of
    IceCube
    has been studied
    with realistic Monte Carlo simulations and similar
    analysis techniques as described in this paper [58].
    The result is a substantially improved performance
    in terms of sensitivity and reconstruction accuracy.
    A direction accuracy of about 0
    :
    7
    ?
    (median) for
    energies above 1 TeV is achieved. Similar to
    AMANDA
    , we expect a further improvement by
    exploiting the full information, avaliable from
    the recorded wave-forms, in the reconstruction.
    Acknowledgements
    This research was supported by the following
    agencies: Deutsche Forschungsgemeinschaft
    (DFG); German Ministry for Education and
    Research; Knut and Alice Wallenberg Founda-
    tion, Sweden; Swedish Research Council; Swedish
    Natural Science Research Council; Fund for
    Scientific Research (FNRS-FWO), Flanders In-
    stitute to encourage scientific and technological
    research in industry (IWT), and Belgian Federal
    Office for Scientific, Technical and Cultural affairs
    (OSTC), Belgium. UC-Irvine AENEAS Super-
    computer Facility; University of Wisconsin
    Alumni Research Foundation; U.S. National
    Science Foundation, Office of Polar Programs;
    U.S. National Science Foundation, Physics
    Division; U.S. Department of Energy; D.F.
    Cowen acknowledges the support of the NSF
    ARTICLE IN PRESS
    J. Ahrens et al. / Nuclear Instruments and Methods in Physics Research A 524 (2004) 169–194
    192

    CAREER program. I. Taboada acknowledges the
    support of FVPI.
    References
    [1] E. Andr
    !
    es, et al., Nature 410 6827 (2001) 441.
    [2] J.G. Learned, K. Mannheim, Annu. Rev. Nucl. Part. Sci.
    50 (2000) 679.
    [3] J.G. Learned, S. Pakvasa, Astropart. Phys. 3 (1995) 267.
    [4] J. Ahrens, et al., Phys. Rev. D 67 (2003) 012003
    arXive:astro-ph/0206487.
    [5] S. Hundertmark, The
    AMANDA
    Collaboration,
    AMANDA-B10
    limit on UHE muon-neutrinos, in: Proceedings of the 28th
    International Cosmic Ray Conference, Tsukuba, Japan,
    2003.
    [6] S. Hundertmark, et al., A method to detect UHE neutrinos
    with
    AMANDA
    , in: Proceedings of the 27th International
    Cosmic Ray Conference, Hamburg, Germany, Vol. 3,
    2001, pp. 1129–1132, HE.236.
    [7] J.E. Dickinson, et al., Nucl. Instr. and Meth. A 440 (1)
    (2000) 95.
    [8] J. Ahrens, et al., Calibration and Survey of
    AMANDA
    with
    the
    SPASE
    Detectors, Astropart. Phys. accepted for
    publication, URL
    http://www.amanda.uci.edu/docu-
    ments.html
    .
    [9] X. Bai, et al., Calibration and survey of
    AMANDA
    with
    SPASE
    , in: Proceedings of the 27th International Cosmic
    Ray Conference, Hamburg, Germany, Vol. 3, 2001, pp.
    977–980, HE.208.
    [10] K. Rawlins, Measuring the composition of cosmic
    rays with the
    SPASE
    and
    AMANDA
    detectors, Ph.D.
    Thesis, University of Wisconsin, Madison, WI, USA,
    October 2001, URL
    http://amanda.berkeley.edu/
    manuscripts/
    .
    [11] E. Andr
    !
    es, et al., Astropart. Phys. 13 (1) (2000) 1
    arXive:astro-ph/9906203.
    [12] D.F. Cowen, K. Hanson, The
    AMANDA
    Collaboration, Time
    calibration of the
    AMANDA
    neutrino telescope with cosmic
    ray muons, in: Proceedings of the 27th International
    Cosmic Ray Conference, Hamburg, Germany, Vol. 3,
    2001, pp. 1133–1136, HE.237.
    [13] K. Woschnagg, et al., Optical properties of South Pole ice
    at depths from 140 to 2300 meters, in: D. Kieda, M.
    Salamon, B. Dingus (Eds.), Proceedings of the 26th
    International Cosmic Ray Conference, Salt Lake City,
    USA, Vol. 2, 1999, pp. 200–203, HE.4.1.15.
    [14] P.B. Price, K. Woschnagg, Astropart. Phys. 15 (1) (2001)
    97 arXive:hep-ex/0008001..
    [15] K. Woschnagg, P.B. Price, Appl. Opt. 40 (15) (2001) 2496.
    [16] P.B. Price, K. Woschnagg, D. Chirkin, Geophys. Res.
    Lett. 27 (13).
    [17] Y.D. He, P.B. Price, J. Geophys. Res. D 103 (14) (1998)
    17041.
    [18] P.B. Price, L. Bergst
    .
    om, Appl. Opt. 36 (18) (1997) 4181.
    [19] P. Askebjer, et al., Appl. Opt. 36 (18) (1997) 4168.
    [20] Particle Data Group, Review of particle properties, Eur.
    Phys. J. C 15 (1–4).
    [21] L.A. Kuzmichev, On the velocity of light signals in the
    deep underwater neutrino experiments, e-preprint, Mos-
    cow State University, arXive:hep-ex/0005036, March 2000.
    [22] C. Spiering, et al., Trackreconstruction and background
    rejection in the Baikal neutrino telescope, in: Proceedings
    of the Third Nestor International Workshop, Pylos,
    Greece, 1993, p. 234, DESY 94-050.
    [23] E. Carmona, Reconstruction methods for the Antares
    neutrino telescope, in: Proceedings of the Second Work-
    shop on Methodical Aspects of Underwater/Ice Neutrino
    Telescopes, Hamburg, Germany, 2001, p. 111.
    [24] A. Bouchta, Muon analysis with the
    AMANDA-B
    four-string
    detector, Ph.D. Thesis, Stockholms Universitet, Stock-
    holm, Sweden, uSIP Report 1998-07, 1998, URL
    http://
    amanda.berkeley.edu/manuscripts/
    .
    [25] R.A. Porrata, The energy spectrum of pointlike events in
    AMANDA-A
    , Ph.D. Thesis, University of California, Irvine,
    CA, USA, 1997.
    [26] P. Mio
    W
    inovi
    !
    c, Muon energy reconstruction in the
    Antarctic Muon and Neutrino Detector Array (
    AMANDA)
    ,
    Ph.D. Thesis, University of California, Berkeley, CA,
    USA, December, 2001, URL
    http://amanda.berke-
    ley.edu/manuscripts/
    .
    [27] G.C. Hill, Bayesian event reconstruction and background
    rejection in neutrino detectors, in: Proceedings of the 27th
    International Cosmic Ray Conference, Hamburg, Ger-
    many, Vol. 3, 2001, pp. 1279–1282, HE.267.
    [28] Ty R. DeYoung, Gary C. Hill, et al., Application of Bayes’
    theorem to muon trackreconstruction in
    AMANDA
    , in:
    Proceedings Advanced Statistical Techniques in Particle
    Physics, Durham, UK, 2002.
    [29] R.D. Cousins, Conference summary talk, in: Proceedings
    Advanced Statistical Techniques in Particle Physics,
    Durham, UK, 2002.
    [30] J. Ahrens, et al., Phys. Rev. D 66 (1) (2002) 012005
    arXive:astro-ph/0205109.
    [31] T.R. DeYoung, Observation of atmospheric muon neu-
    trinos with
    AMANDA
    , Ph.D. Thesis, University of Wisconsin
    at Madison, Madison, WI, USA, May 2001, URL
    http://amanda.berkeley.edu/manuscripts/
    [32] V.M. Aynutdinov, et al., Phys. Part. Nucl. Lett. 109 (2001)
    43.
    [33] A. Karle, Monte Carlo simulation of photon transport and
    detection in deep ice: muons and cascades, in: Proceedings
    of Workshop on the Simulation and Analysis Methods for
    Large Neutrino Telescopes, DESY-PROC-1999-01, DESY
    Zeuthen, Germany, 1999, pp. 174–185.
    [34] C.H. Wiebusch, Muon reconstruction with
    AMANDA
    , in:
    Proceedings of Workshop on the Simulation and Analysis
    Methods for Large Neutrino Telescopes, DESY-PROC-
    1999-01, DESY Zeuthen, Germany, 1999, pp. 302–316.
    [35] D. Pandel, Bestimmung von Wasser- und Detektorpar-
    ametern und Rekonstruktion von Myonen bis 100 TeV mit
    dem Baikal-Neutrinoteleskop NT-72, Diploma Thesis,
    Humboldt-Universit
    .
    at zu Berlin, Berlin, Germany,
    ARTICLE IN PRESS
    J. Ahrens et al. / Nuclear Instruments and Methods in Physics Research A 524 (2004) 169–194
    193

    February 1996, URL
    http://www-zeuthen.desy.de/
    nuastro/publications/diploma/
    .
    [36] H. Geenen, Energy reconstruction and spectral unfolding
    of atmospheric leptons with the AMANDA detector,
    Diploma Thesis, University of Wuppertal, Wuppertal,
    Germany, November 2002, URL
    http://amanda.
    berkeley.edu/manuscripts
    .
    [37] J. Ahrens, et al., Limits on diffuse fluxes of high energy
    extraterrestrial neutrinos with the
    AMANDA-B10
    detector,
    Phys. Rev. Lett. 90 (2003) 251101 arXive:astro-ph/
    0303218.
    [38] V.J. Stenger, Trackfitting for DUMAND-II Octagon
    Array, External Report HDC-1-90, University of Hawai’i
    at Manoa, Manoa, HI, USA, 1990.
    [39] CERN Information Technology Division, PAW—Physics
    Analysis Workstation; User’s Guide, CERN Program
    Library Long Writeup Q121, Geneva, Switzerland, 1999.
    [40] J. Jacobsen, C. Wiebusch, An overview of offline software
    for
    AMANDA
    , in: Proceedings of Workshop on the Simula-
    tion and Analysis Methods for Large Neutrino Telescopes,
    DESY-PROC-1999-01, DESY Zeuthen, Germany, 1999,
    pp. 194–204.
    [41] O. Streicher, C. Wiebusch, ‘‘recoos’’, muon and neutrino
    reconstruction for underwater/ice Cherenkov telescopes,
    2001, URL
    http://www.ifh.de/nuastro/software/
    siegmund/
    .
    [42] O. Streicher, C. Wiebusch, ‘‘rdmc’’, a library for proces-
    sing neutrino telescope data and MC files, 2001, URL
    http://www.ifh.de/nuastro/software/siegmund/
    .
    [43] G. C. Hill, S. Hundertmark, M. Kowalski, P. Miocinovic,
    T. Neunh
    .
    ofer, P. Niessen, P. Steffen, O. Streicher, C.
    Wiebusch, The SiEGMuND software package, 2001, URL
    http://www.ifh.de/nuastro/software/siegmund/
    .
    [44] W.H. Press, S.A. Teukolsky, W.V. Vetterling, B.P.
    Flannery, Numerical Recipies in C—The Art of Scientific
    Computing, 2nd Edition, Cambridge University Press,
    Cambridge, UK, 1997 URL
    http://www.nr.com/
    .
    [45] F. James, The CERN Computing and Networks Division,
    MINUIT—Function Minimization and Error Analysis;
    Reference Manual, 94th Edition, CERN Program Library
    Long Writeup D506, Geneva, Switzerland, 1994.
    [46] S. Young, A search for point sources of high energy
    neutrinos with the
    AMANDA-B10
    neutrino telescope, Ph.D.
    Thesis, University of California at Irvine, Irvine, CA,
    USA, July 2001, URL
    http://amanda.berkeley.edu/
    manuscripts/
    .
    [47] J. Ahrens, et al., Astrophys. J. 583 (2003) 1040 arXive:as-
    tro-ph/0208006.
    [48] S. B
    .
    oser, Separation of atmospheric neutrinos with the
    AMANDA-II
    detector, Diploma Thesis, Technische
    Universit
    .
    at M
    .
    unchen, Munich, Germany, April 2002.
    [49] J. Ahrens, et al., Phys. Rev. D 66 (3) (2002) 032006
    arXive:astro-ph/0202370.
    [50] P. Niessen, C. Spiering, et al., Search for relativistic
    monopoles with the
    AMANDA
    detector, in: Proceedings of the
    27th International Cosmic Ray Conference, Hamburg,
    Germany, Vol. 4, 2001, pp. 1496–1498, HE.315.
    [51] A. Biron, On the rejection of atmospheric muons in the
    AMANDA
    detector, Diploma Thesis, Humboldt-Universit
    .
    at
    zu Berlin, Berlin, Germany, DESY-THESIS-1998-014,
    March 1998, URL
    http://amanda.berkeley.edu/
    manuscripts/
    .
    [52] M. Gaug, AMANDA event reconstruction and cut
    evaluation methods, in: Proceedings of the Second Work-
    shop on Methodical Aspects of Underwater/Ice Neutrino
    Telescopes, DESY-PROC-2002-01, Hamburg, Germany,
    2001, p. 123.
    [53] M. Gaug, Detection of atmospheric muon neutrinos with
    the
    AMANDA
    neutrino telescope, Diploma Thesis,
    Humboldt-Universit
    .
    at zu Berlin, Berlin, Germany,
    Octobter 2000, URL
    http://amanda.berkeley.edu/
    manuscripts/
    .
    [54] M.J. Leuthold, Search for cosmic high energy neutrinos
    with the
    AMANDA-BIO
    detector, Ph.D. Thesis, Humboldt-
    Universit
    .
    at zu Berlin, Berlin, Germany, September 2001.
    [55] A. Biron, Search for atmospheric muon-neutrinos and
    extraterrestric neutrino point sources in the 1997
    AMANDA-
    B10
    data, Ph.D. Thesis, Humboldt-Universit
    .
    at zu Berlin,
    Berlin, Germany, January 2002, URL
    http://amanda.
    bcerkeley.edu/manuscripts/
    .
    [56] W. Wagner, The
    AMANDA
    Collaboration, New capabilities
    of the AMANDA-II high energy neutrino telescope, in:
    Proceedings of the 28th International Cosmic Ray Con-
    ference, Tsukuba, Japan, 2003.
    [57] S. Yoshida, The
    IceCube
    Collaboration, The IceCube high
    energy neutrino telescope, in: Proceedings of the 28th
    International Cosmic Ray Conference, Tsukuba, Japan,
    2003.
    [58] J. Ahrens, et al., Sensitivity of the IceCube detector to
    astrophysical sources of high energy muon neutrinos,
    Astropart. Phys. accepted for publication, arXive:astro-ph/
    0305196, May 2003.
    ARTICLE IN PRESS
    J. Ahrens et al. / Nuclear Instruments and Methods in Physics Research A 524 (2004) 169–194
    194

    Back to top