0% found this document useful (0 votes)
94 views25 pages

IceCube Review

Uploaded by

Byrnes
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
94 views25 pages

IceCube Review

Uploaded by

Byrnes
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 25

NS64CH04-Gaisser ARI 6 September 2014 14:0

ANNUAL
REVIEWS Further IceCube
Click here for quick links to
Annual Reviews content online,
including:
Thomas Gaisser1 and Francis Halzen2
• Other articles in this volume 1
Bartol Research Institute and Department of Physics and Astronomy, University of Delaware,
• Top cited articles Newark, Delaware 19716
Annu. Rev. Nucl. Part. Sci. 2014.64:101-123. Downloaded from www.annualreviews.org
Access provided by University of Texas - Arlington on 09/24/20. For personal use only.

• Top downloaded articles 2


• Our comprehensive search Department of Physics, University of Wisconsin–Madison, Madison, Wisconsin 53703

Annu. Rev. Nucl. Part. Sci. 2014. 64:101–23 Keywords


First published online as a Review in Advance on neutrino astronomy, cosmic rays, dark matter, neutrino properties
June 19, 2014

The Annual Review of Nuclear and Particle Science Abstract


is online at nucl.annualreviews.org
IceCube is the first kilometer-scale neutrino detector. Built primarily for
This article’s doi: neutrino astronomy, it has recently discovered events with energies above
10.1146/annurev-nucl-102313-025321
100 TeV that are likely to be from distant sources beyond the solar system.
Copyright  c 2014 by Annual Reviews. Among the events are three with deposited energies of more than 1 PeV,
All rights reserved
the highest-energy neutrinos ever detected. We review the astrophysical
arguments that motivate such a large detector, and we describe how it works
and how the high-energy events are reconstructed and identified above the
background of atmospheric neutrinos. We also describe the broad range of
neutrino physics and particle astrophysics topics addressed by IceCube, as
well as its potential for the future.

101
NS64CH04-Gaisser ARI 6 September 2014 14:0

Contents
1. INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
2. RATIONALE FOR THE CONSTRUCTION OF KILOMETER-SCALE
NEUTRINO DETECTORS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
3. NEUTRINO ASTRONOMY I . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
4. THE FIRST KILOMETER-SCALE NEUTRINO DETECTOR. . . . . . . . . . . . . . . . 108
5. ATMOSPHERIC MUONS AND NEUTRINOS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
6. DISCOVERY OF COSMIC NEUTRINOS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
7. NEUTRINO ASTRONOMY II . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
8. COSMIC RAYS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
9. SEARCH FOR DARK MATTER . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
Annu. Rev. Nucl. Part. Sci. 2014.64:101-123. Downloaded from www.annualreviews.org
Access provided by University of Texas - Arlington on 09/24/20. For personal use only.

10. IceCube AS A DISCOVERY INSTRUMENT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118


10.1. Neutrino Oscillations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
10.2. Supernovae and Solar Flares . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
10.3. IceCube, The Facility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
11. OUTLOOK . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120

1. INTRODUCTION
Soon after the 1956 observation of the neutrino (1), the idea emerged that it represented the ideal
astronomical messenger (2–4). The concept has since been demonstrated: Neutrino detectors have
“seen” the Sun and detected a supernova in the Large Magellanic Cloud in 1987. Both observations
were of tremendous importance; the former showed that neutrinos have a tiny mass, opening the
first chink in the armor of the Standard Model of particle physics, and the latter confirmed the
basic nuclear physics of the death of stars.
High-energy neutrinos have distinct potential to probe the extreme Universe. Neutrinos reach
us from the edge of the Universe without absorption and with no deflection by magnetic fields.
They can escape unscathed from the inner neighborhood of black holes and from the accelerators in
which cosmic rays are born. Their weak interactions also make neutrinos very difficult to detect.
Immense particle detectors are required to collect cosmic neutrinos in statistically significant
numbers (5). By the 1970s, researchers had already understood (6) that a kilometer-scale detector
was needed to observe the “cosmogenic” neutrinos produced in the interactions of cosmic rays
with background microwave photons (7).
Above a threshold of ∼4 × 1019 eV, cosmic rays interact with the microwave background, in-
troducing an absorption feature in the cosmic-ray flux, the Greisen–Zatsepin–Kuzmin (GZK)
cutoff (8, 9). The mean free path of extragalactic cosmic rays propagating in the microwave back-
ground is limited to <100 Mpc. Therefore, secondary neutrinos produced in these interactions
are the only probe of the still-enigmatic sources at further distances. Realistic calculations (10)
of the neutrino flux associated with the observed flux of extragalactic cosmic rays appeared in the
1970s and predicted on the order of one event per year in a kilometer-scale detector, subject to
astrophysical uncertainties. Today’s estimates of the sensitivity for observing potential cosmic-ray
accelerators such as galactic supernova remnants (SNRs), active galactic nuclei (AGN), and γ -
ray bursts (GRBs) unfortunately point to the same extreme requirement (5). Building a neutrino
telescope has been a daunting technical challenge.
Given a detector’s required size, early efforts concentrated on instrumenting large volumes
of natural water with photomultipliers that detect the Cherenkov light emitted by the secondary

102 Gaisser · Halzen


NS64CH04-Gaisser ARI 6 September 2014 14:0

particles produced when neutrinos interact with nuclei inside or near the detector (11). After a two-
decade-long effort, building the Deep Underwater Muon and Neutrino Detector (DUMAND) in
the sea off the main island of Hawaii unfortunately failed (12). However, DUMAND pioneered
many of the detector technologies in use today and inspired the deployment of a smaller instrument
in Lake Baikal (13), as well as efforts to commission neutrino telescopes in the Mediterranean
(14–16). These have paved the way toward the planned construction of KM3NeT (16).
The first telescope on the scale envisaged by the DUMAND Collaboration was realized instead
by transforming a large volume of deep Antarctic ice into a particle detector, the Antarctic Muon
and Neutrino Detector Array (AMANDA). In operation from 2000 to 2009 (17), AMANDA
represented the proof of concept for the kilometer-scale neutrino observatory IceCube (18, 19),
completed in 2010 (Figure 1). We write this review at the critical time when IceCube data taken
with the completed detector have revealed the first evidence for a flux of high-energy neutrinos
reaching us from beyond the solar system (20).
Annu. Rev. Nucl. Part. Sci. 2014.64:101-123. Downloaded from www.annualreviews.org
Access provided by University of Texas - Arlington on 09/24/20. For personal use only.

IceCube Lab

IceTop
81 stations
50 m 324 optical sensors

IceCube array
86 strings including 8 DeepCore strings
5,160 optical sensors

AMANDA II array
1,450 m (precursor to IceCube)

DeepCore
8 strings, spacing optimized for lower energies
480 optical sensors

Eiffel Tower
324 m

2,450 m

2,820 m

Bedrock

Figure 1
Artist’s drawing of IceCube at the National Science Foundation’s Amundsen–Scott South Pole Station. The former AMANDA
detector is shown in blue and the DeepCore subarray in green.

www.annualreviews.org • IceCube 103


NS64CH04-Gaisser ARI 6 September 2014 14:0

2. RATIONALE FOR THE CONSTRUCTION OF KILOMETER-SCALE


NEUTRINO DETECTORS
The construction of kilometer-scale neutrino detectors is motivated primarily by the prospect of
detecting neutrinos associated with the sources of high-energy cosmic rays. Cosmic-ray accelera-
tors produce particles with energies in excess of 100 EeV; we still do not know where or how (21).
The bulk of the cosmic rays are Galactic in origin. Any association with our Galaxy presumably
disappears at EeV energy, when the gyroradius of a proton in the Galactic magnetic field exceeds
its size. The cosmic-ray spectrum exhibits a rich structure above an energy of ∼0.1 EeV, but where
exactly the transition to extragalactic cosmic rays occurs is a matter of debate.
The detailed blueprint for a cosmic-ray accelerator must meet two challenges: The highest-
energy particles in the beam must reach beyond 103 TeV (108 TeV) for Galactic (extragalactic)
sources, and their luminosities must be able to accommodate the observed flux. Both requirements
Annu. Rev. Nucl. Part. Sci. 2014.64:101-123. Downloaded from www.annualreviews.org
Access provided by University of Texas - Arlington on 09/24/20. For personal use only.

represent severe constraints that have guided theoretical speculations.


SNRs were proposed as a likely source of cosmic rays as early as 1934 by Baade & Zwicky (22).
Notably, they assumed the sources were extragalactic because the most recent supernova in the
Milky Way dates to 1572. After diffusion in the interstellar medium was understood, supernova
explosions in the Milky Way became the source of choice for the origin of Galactic cosmic rays
(23), although after 50 years the issue is still being debated (24). The idea is widely accepted
because of energetics. Three Galactic supernova explosions per century converting a reasonable
fraction of a solar mass into particle acceleration can accommodate the steady flux of cosmic rays
in the Galaxy.
Energetics also guides speculations on the origin of extragalactic cosmic rays. By integrating
the cosmic-ray spectrum above the ankle at ∼4 EeV, it is possible to estimate the energy density
in extragalactic cosmic rays as ∼3 × 10−19 erg cm−3 (25). This value is rather uncertain because
of our ignorance of the precise energy where the transition from Galactic to extragalactic sources
occurs. The power required for a population of sources to generate this energy density over the
Hubble time of 1010 years is 2 × 1037 erg s−1 Mpc−3 . A GRB fireball converts a fraction of a solar
mass into the acceleration of electrons, observed as synchrotron photons. The observed energy
in extragalactic cosmic rays can be accommodated with the reasonable assumption that shocks
in the expanding GRB fireball convert roughly equal energy into the acceleration of electrons
and cosmic rays (26). It so happens that 2 × 1051 erg per GRB will yield the observed energy
density in cosmic rays after 1010 years, given that their rate is on the order of 300 Gpc−3 year−1 .
Hundreds of bursts per year over a Hubble time produce the observed cosmic-ray density, just as
three supernovae per century accommodate the steady flux in the Galaxy.
Problem solved? Not really: It turns out that the same result can be achieved assuming that
AGN convert, on average, 2 × 1044 erg s−1 each into particle acceleration (27). As is the case for
GRBs, this is an amount that matches the AGN output in electromagnetic radiation. Whether
GRBs or AGN, the observation that cosmic-ray accelerators radiate similar energies in photons
and cosmic rays may not be an accident.
Neutrinos will be produced at some level in association with the cosmic-ray beam. Cosmic rays
accelerated in regions of high magnetic fields near black holes or neutron stars inevitably interact
with radiation surrounding them. Thus, cosmic-ray accelerators are beam dumps. Cosmic rays
accelerated in supernova shocks interact with gas in the Galactic disk, producing equal numbers of
pions of all three charges that decay into pionic photons and neutrinos. A larger source of secon-
daries is likely to be gas near the sources, for example, cosmic rays interacting with high-density
molecular clouds that are ubiquitous in the star-forming regions where supernovae are more likely
to explode. For extragalactic sources, the neutrino-producing target may be electromagnetic, for

104 Gaisser · Halzen


NS64CH04-Gaisser ARI 6 September 2014 14:0

10–1
IceCube νe
Frejus νμ
10–2 Frejus νe
Super-Kamiokande νμ
10–3 Conventional νμ AMANDA νμ
Unfolding
Eν2 Φν (GeV cm–2 s–1 sr–1)

Conventional νe Forward folding


10–4 IceCube νμ
Unfolding
Forward folding
10–5

10–6
Annu. Rev. Nucl. Part. Sci. 2014.64:101-123. Downloaded from www.annualreviews.org
Access provided by University of Texas - Arlington on 09/24/20. For personal use only.

10–7 Prompt νμ, νe

GRB GZK
100 10–8 Galactic
supernovae

10 10–9
–1 0 1 2 3 4 5 6 7 8 9
events
km–2 year–1 log10 [Eν (GeV)]

Figure 2
Anticipated astrophysical neutrino fluxes are compared with measured and calculated fluxes of atmospheric
neutrinos. The shaded band indicates the level of model-dependent expectations for high-energy neutrinos
of astrophysical origin. Measurements of νμ from Super-Kamiokande (28), Frejus (29), AMANDA (30, 31),
and IceCube (32, 33) are shown along with the νe spectrum at high energy ( green open triangles) (34).
Calculations of conventional νe (red line) and νμ (blue line) (35), νe (red dotted line) (36), and charm-induced
neutrinos (magenta band ) (37) are also shown. Abbreviations: GRB, γ -ray burst; GZK, Greisen–Zatsepin–
Kuzmin cutoff.

instance, photons radiated by the accretion disk of an AGN, or synchrotron photons that coexist
with protons in the expanding fireball producing a GRB. Figure 2 compares estimates of astro-
physical neutrino fluxes with measurements of atmospheric neutrinos. The estimates, discussed
briefly in the following paragraphs, set the level at

dN ν
Eν2  10−8 GeV cm−2 s−1 sr−1 1.
dEν
per flavor, or somewhat less.
Estimating the neutrino flux associated with cosmic rays accelerated in SNRs and GRBs is
relatively straightforward because both the beam, identified with the observed cosmic-ray flux,
and the targets, observed by astronomers, are known. In the case of SNRs, the main uncertainty is
the availability of nearby target material. In the case of GRBs, the main uncertainty is the fraction
of the extragalactic cosmic-ray population that comes from this source.
Active galaxies are complex systems with many possible sites for accelerating cosmic rays and for
targets to produce neutrinos. For example, if acceleration occurs mainly at an outer termination
shock in intergalactic space (38), there would be little target material available. One generic
picture in which the neutrino luminosity is directly related to the contribution of the sources
to extragalactic cosmic rays arises if acceleration occurs in the jets of AGN (or GRBs) (39, 40).
High-energy protons interact in the intense radiation fields inside the jets. In the pγ → pπ 0

www.annualreviews.org • IceCube 105


NS64CH04-Gaisser ARI 6 September 2014 14:0

1e+46

TeV Cat
IC40+IC59 limits
1e+45

Luminosity estimate (erg s–1)


1e+44

1ES1959+650
1e+43

1e+42

M87
1e+41
Cen A
Annu. Rev. Nucl. Part. Sci. 2014.64:101-123. Downloaded from www.annualreviews.org
Access provided by University of Texas - Arlington on 09/24/20. For personal use only.

1e+40
0.001 0.010 0.100 1.000
Redshift, z
Figure 3
Limits on the neutrino flux from selected active galaxies derived from IceCube data taken during
construction, when the instrument was operating with 40 and 59 strings of the total 86 instrumented strings
of digital optical modules (43). These are compared with the TeV photon flux for nearby active galactic
nuclei. Note that energy units are in erg, not TeV.

channel, the protons remain in the accelerator. In the pγ → nπ + channel, however, the neutrons
escape and eventually decay to produce cosmic-ray protons, while the pions decay to neutrinos.
The luminosity of neutrinos from photopion production is then directly related by kinematics to
the cosmic-ray protons that come from decay of the escaping neutrons.
TeV γ -rays are measured from many AGN blazars (http://tevcat.uchicago.edu/). Although
the observed γ -rays are likely to be from accelerated electrons, which radiate more efficiently
than protons, the γ -ray luminosity may indicate the overall cosmic-ray luminosity and hence the
possible level of neutrino production (41). In this context, we introduce Figure 3 (42), showing
IceCube upper limits (43) on the neutrino flux from nearby AGN as a function of their distance.
The sources at redshifts between 0.03 and 0.2 are Northern Hemisphere blazars for which distances
and intensities are listed in TeVCat (http://tevcat.uchicago.edu/) and for which IceCube also
has upper limits. In several cases, the νμ limits have reached the level of the TeV photon flux. One
can sum the sources shown in the figure into a diffuse flux. The result, after accounting for the
distances and luminosities, is 3 × 10−9 GeV cm−2 s−1 sr−1 , or ∼10−8 GeV cm−2 s−1 sr−1 for all
neutrino flavors. This value is at the level of the generic astrophysical neutrino flux of Equation 1.
At this intensity, neutrinos from theorized cosmic-ray accelerators will cross the steeply falling
atmospheric neutrino flux above an energy of ∼300 TeV (Figure 2). The level of events observed
in a cubic-kilometer neutrino detector is ∼10–100 νμ -induced events per year. Such estimates
reinforced the logic for building a cubic-kilometer neutrino detector. A more detailed description
of the theoretical estimates can be found in Reference 44.

3. NEUTRINO ASTRONOMY I
The IceCube neutrino detector (Figure 1) consists of 5,160 digital optical modules (DOMs)
viewing a cubic kilometer of clear ice between 1,450 and 2,450 m below the surface. The depth
of the detector and its projected area determine the trigger rate of ∼3 kHz due to penetrating

106 Gaisser · Halzen


NS64CH04-Gaisser ARI 6 September 2014 14:0

a b
Annu. Rev. Nucl. Part. Sci. 2014.64:101-123. Downloaded from www.annualreviews.org

Figure 4
Access provided by University of Texas - Arlington on 09/24/20. For personal use only.

(a) A neutrino-induced muon from below the horizon crossing the detector (59). (b) A cascade event starting
in the detector (58). The colors of the dots indicate arrival time, from early (red ) to late ( purple), following
the rainbow. The sizes of the dots indicate the number of photons detected.

muons produced by interactions of cosmic rays in the atmosphere above. The ratio of neutrino-
induced signal to atmospheric background is on the order of one per million. The neutrino rate is
dominated by atmospheric neutrinos produced around the globe. The first challenge is to select a
sufficiently pure sample of neutrinos, and the second is to identify the small fraction of astrophysical
neutrinos.
Neutrino events may be broadly classified into two groups, tracks and cascades, which reflect
the patterns of Cherenkov light emitted by the charged particles produced when the neutrinos in-
teract. Tracks are produced by charged-current interactions of νμ s, whereas cascades are produced
by charged-current interactions of νe s and ντ s and neutral-current interactions of all neutrino fla-
vors. Figure 4 shows high-energy examples of each, discussed in Section 5 below. The νμ -induced
muons typically travel several kilometers, whereas the characteristic length scale for the electro-
magnetic showers that dominate the cascade events is tens of meters. A complementary way to
classify neutrino events is to distinguish events that start inside the detector from those in which
the neutrino interacts outside the detector. The largest neutrino sample consists of νμ -induced
muons entering the detector from zenith angles too large to be atmospheric muons (θ ≥ 85◦ ).
The rate of such events in the full IceCube detector is ∼100 per day or more, depending on how
the threshold for the analysis is set. The mean energy of this sample, dominated by atmospheric
neutrinos, is ∼1 TeV, of which only a fraction is deposited in the detector.
The starting event sample includes both tracklike and cascade events. At 2 PeV, the mean
decay length of a τ lepton is comparable to the 125-m spacing between the strings of IceCube.
Somewhere in the PeV range it should be possible to identify a ντ by its characteristic double-bang
signature (45). Such an event has not yet been identified. For Eν < 1 PeV, the interaction of a ντ
in IceCube will look much like that of a νe .
Reconstruction of events depends on accurate timing (<3 ns) and on the ability to measure the
amount of Cherenkov light generated along the tracks of charged particles produced by the neu-
trino interactions. Basically, the arrival time of photons at the DOMs determines the trajectory,
and the amount of light is a proxy for the deposited energy. The number of photons produced per
unit path length and their distribution in wavelength are well-known quantities (46, section 30.7).
An understanding of the properties of the ice (47) is crucial to relate light generated to light
observed in the DOM as a function of distance and orientation relative to the emission point.
Reconstruction of tracks in ice has been well studied (48). For typical kilometer tracks, the

www.annualreviews.org • IceCube 107


NS64CH04-Gaisser ARI 6 September 2014 14:0

angular resolution is better than 1◦ . Reconstruction of cascade events is a topic of current study
at IceCube (49). Determining the deposited energy from the observed light pool is, in principle,
straightforward once the vertex is located by symmetry (modulo ice properties). Angular resolu-
tion is significantly poorer than for tracks. In the large cascades studied by detailed simulations
on an event-by-event basis, it is possible to determine the directions to within 15◦ on the basis of
the shapes of the waveforms, which reflect the directionality of the cascade electrons.

4. THE FIRST KILOMETER-SCALE NEUTRINO DETECTOR


Before the first string of IceCube was deployed in January 2005, considerable effort went into
designing the hot water drilling system and developing the design and assembly procedure for
the DOMs. The first working season at the South Pole was 2003–2004, when major components
of the drill system were shipped in and test tanks were set up for IceTop, the surface array of
Annu. Rev. Nucl. Part. Sci. 2014.64:101-123. Downloaded from www.annualreviews.org
Access provided by University of Texas - Arlington on 09/24/20. For personal use only.

IceCube.
Optical modules consist of a 10-inch Hamamatsu photomultiplier (50), a main computer board,
and a board with 12 light-emitting diodes for calibration, all enclosed in a watertight glass sphere.
The sphere consists of two hemispherical sections, partially evacuated and sealed, with a sin-
gle penetrator connecting to an external cable. DOMs were assembled at three sites: Madison,
Wisconsin; Stockholm/Uppsala, Sweden; and DESY–Zeuthen, Germany.
The drilling system consisted of a number of components, starting with fuel tanks and gen-
erators followed by sets of hot water heaters, high-pressure pumps, and the drill control center,
arranged in a series of buildings termed the drill camp. The drill camp was excavated from its
storage location at the beginning of each austral summer, starting in late October, and relocated
to the center of the drilling area for the season. Start-up included establishing a working reservoir
(Rodriguez well) for the hot water drilling system. Water was recycled from the high-pressure
pumps through an insulated surface hose with a reach of 300 m into a single 3-km-long, 10-cm-
diameter drilling hose that moved down at a rate of 2 to 3 m min−1 . As drilling proceeded, water
was brought back to the surface system through a return hose from the hole. Hot water continued
to flow through the system as the drill hose was rewound so as to finish with a water-filled cylinder
2.5 km deep by 60 cm in diameter. The standard time for preparing a hole was 36 h.
Optical modules were staged in a tower structure adjacent to the drill tower so that deploy-
ment could begin soon after drilling was complete. The use of two systems allowed drilling and
deployment to proceed in leapfrog fashion (Figure 5). During deployment, the drill tower was
used to carry the main cable. DOMs were connected to breakouts on the cable as it descended into
the hole. Two IceTop tanks, each instrumented with two DOMs, were deployed 25 m from the
top of each hole. After drilling procedures were established, up to 20 strings could be deployed in
a season (Table 1).
Data-acquisition programs running on the DOMs process photomultiplier tube (PMT) output
to provide time-stamped waveforms (51). Timing at the nanosecond level is achieved through
automated synchronization of a global clock distributed through kilometer-length copper wires.
Surface cables bring signals from the DOMs to computers in the IceCube Lab (ICL), an elevated
building in the center of the array. The computers process the DOM signals to form triggers and
preliminary reconstructions of events. Events are filtered and a subset selected for transmission
by satellite to the north.

5. ATMOSPHERIC MUONS AND NEUTRINOS


Muons and neutrinos from decays of mesons produced by cosmic-ray interactions in the atmo-
sphere are the background in the search for neutrinos of extraterrestrial origin. The 3-kHz trigger

108 Gaisser · Halzen


NS64CH04-Gaisser ARI 6 September 2014 14:0
Annu. Rev. Nucl. Part. Sci. 2014.64:101-123. Downloaded from www.annualreviews.org
Access provided by University of Texas - Arlington on 09/24/20. For personal use only.

Figure 5
Photo of the drilling operation at the South Pole. The tower in the foreground is being used for drilling,
while a string of digital optical modules is being deployed at the more distant tower. The South Pole
Telescope (left) and the Martin A. Pomerantz Observatory (right) are visible on the horizon.

rate of IceCube is dominated by atmospheric muons from decays of pions and kaons produced
in the atmosphere above the detector. The distribution peaks near the zenith and decreases with
increasing angle as the muon energy required to reach the deep detector increases. Most atmo-
spheric muons are easily identified as entering tracks from above and are rejected. Because of
the large ratio of muons to neutrinos, however, misreconstructed atmospheric muons remain an
important source of background for all searches.
Measurement of the spectrum of atmospheric neutrinos is an important benchmark for a neu-
trino telescope. The spectrum of atmospheric νμ has been measured by unfolding the measured
rate and energy deposition of neutrino-induced muons entering the detector from below the hori-
zon (Figure 2) (32). More challenging is the measurement of the flux of atmospheric νe s. This has
been done by making use of DeepCore, the more densely instrumented subarray in the deep center
of IceCube, to identify contained cascade events. The known spectrum of νμ s is used to calculate
the contribution of neutral-current interactions to the observed rate of cascades. Subtracting the
neutral-current contribution leads to the measurement of the spectrum of atmospheric νe s from
100 GeV to 10 TeV (Figure 2) (34).
In general, atmospheric neutrinos are indistinguishable from astrophysical neutrinos. An im-
portant exception occurs in the case of νμ s from above when the neutrino energy is sufficiently

Table 1 IceCube deployment by season (cumulative)


Season Strings IceTop stations
2004–2005 1 4
2005–2006 9 16
2006–2007 22 26
2007–2008 40 40
2008–2009 59 59
2009–2010 79 73
2010–2011 86 81

www.annualreviews.org • IceCube 109


NS64CH04-Gaisser ARI 6 September 2014 14:0

high and the zenith angle sufficiently small that the muon produced in the same decay as the neu-
trino is guaranteed to reach the detector (52). In this case, the atmospheric neutrino provides its
own self-veto. By use of a Monte Carlo simulation, the atmospheric neutrino passing rate can be
evaluated more generally by including other high-energy muons produced in the same cosmic-ray
shower as the neutrino. In this way, the method can be extended to νe s. In practice, the passing
rate is significantly reduced for zenith angles θ < 70◦ and Eν > 100 TeV.
The spectrum of atmospheric neutrinos becomes one power steeper than the spectrum of pri-
mary nucleons at high energy as the competition between interaction and decay of pions and kaons
increasingly suppresses their decay. For the dominant kaon channel, the characteristic energy for
the steepening is Eν ∼ 1 TeV/ cos θ. A further steepening occurs above 100 TeV at the knee
in the primary spectrum. Astrophysical neutrinos should reflect the cosmic-ray spectrum in the
source and are therefore expected to have a significantly harder spectrum than that of atmospheric
neutrinos. Establishing an astrophysical signal above the steep atmospheric background requires
Annu. Rev. Nucl. Part. Sci. 2014.64:101-123. Downloaded from www.annualreviews.org
Access provided by University of Texas - Arlington on 09/24/20. For personal use only.

an evaluation of the atmospheric neutrino spectrum around 100 TeV and above.
Although there is some uncertainty associated with the composition through the knee region
(53), the major uncertainty in the spectrum of atmospheric neutrinos at high energy is the level
of charm production. The short-lived charmed hadrons preferentially decay up to a characteristic
energy of 107 GeV, producing prompt muons and neutrinos with the same spectrum as their parent
cosmic rays. This prompt flux of leptons has not yet been measured. Existing limits (54, 55) allow
a factor of two or three around the level predicted by a standard calculation (37), after correction
for steepening at the knee. For reasonable assumptions, the charm contribution is expected to
dominate the conventional spectrum above ∼10 TeV for νe s, above ∼100 TeV for νμ s, and
above ∼1 PeV for muons (56).
The expected hardening in the spectrum of atmospheric neutrinos due to prompt neutrinos is
partially degenerate with a hard astrophysical component. However, the spectrum of astrophysical
neutrinos should reflect the spectrum of cosmic rays at their sources, which is expected to be harder
than the spectrum of cosmic rays arriving at Earth. It should eventually be possible with IceCube
to measure the charm contribution by requiring a consistent interpretation of neutrino flavors
and muons for which there is no astrophysical component. An additional signature of atmospheric
charm is the absence of seasonal variations for this component (57).

6. DISCOVERY OF COSMIC NEUTRINOS


The generation of underground neutrino detectors preceding construction of AMANDA searched
for cosmic neutrinos without success and established an upper limit on their flux, assuming an E−2
energy dependence (29):
dN
Eν2 ≤ 5 × 10−6 GeV cm−2 s−1 sr−1 . 2.
dEν
Operating for almost one decade, AMANDA improved this limit by two orders of magnitude.
With data taken during its construction, IceCube’s sensitivity rapidly approached the theoretical
flux estimates for candidate sources of cosmic rays, such as SNRs, GRBs, and with a larger un-
certainty, AGN (Figure 2). With its completion, IceCube also positioned itself to observe the
much-anticipated cosmogenic neutrinos, with some estimates predicting as many as two events
per year (10).
Cosmogenic neutrinos were the target of a dedicated search using IceCube data collected
between May 2010 and May 2012. Two events were found (58). However, their energies, rather
than ∼EeV, as expected for cosmogenic neutrinos, are just above 1 PeV. These events are particle

110 Gaisser · Halzen


NS64CH04-Gaisser ARI 6 September 2014 14:0

showers initiated by neutrinos interacting inside the instrumented detector volume. Their light
pool of roughly 100,000 photoelectrons extends over more than 500 m (Figure 4b). With no
evidence of a muon track, both are most likely initiated by νe s or ντ s.
Before this serendipitous discovery, neutrino searches had almost exclusively focused on νμ s that
interact primarily outside the detector to produce kilometer-long muon tracks passing through the
instrumented volume. This approach maximizes the event rate by enlarging the target volume, but
it is necessary to use the Earth as a filter to remove the huge background flux of muons produced by
cosmic-ray interactions in the atmosphere. This approach limits the neutrino view to a single flavor
and half the sky. Inspired by the observation of the two PeV events, a filter was designed to identify
high-energy neutrinos interacting inside the detector. It divides the instrumented volume of ice
into an outer veto shield and a 420-MT inner fiducial volume. The separation between veto and
signal regions was optimized to reduce the background of atmospheric muons and neutrinos to a
handful of events per year while keeping 98% of the signal. The background of atmospheric muon
Annu. Rev. Nucl. Part. Sci. 2014.64:101-123. Downloaded from www.annualreviews.org
Access provided by University of Texas - Arlington on 09/24/20. For personal use only.

punch-through was determined experimentally by measuring the rate at which muons tagged in
the veto region passed an inner veto region of similar size. The great advantage of specializing to
neutrinos interacting inside the instrumented volume of ice is that the detector functions as a total
absorption calorimeter measuring deposited energy with 10–15% resolution. Also, neutrinos from
all directions in the sky, including both muon tracks produced in νμ charged-current interactions
and secondary showers produced by neutrinos of all flavors, can be identified.
Analyzing the data covering the same time period as the cosmogenic neutrino search,
28 candidate neutrino events were identified with in-detector deposited energies between 30 and
1,140 TeV (20). Most of the events, including the two previously discovered PeV events, come
from the southern sky. Some of these events have sufficiently high energy and small zenith angle
so that, if they were produced in the atmosphere, they would probably have been accompanied by
muons and excluded from the data sample.
Of the 28 events, 21 are showers with no evidence of a muon track and with energies measured
to better than 15%, although their directions are determined to 10–15◦ only. The remaining
7 events are muon tracks, which do allow for subdegree angular reconstruction; however, only a
lower limit on their energy can be established because of the unknown fraction carried away by
the exiting muon. Furthermore, the lower-energy muon-like events from above include four that
start near the detector boundary and are consistent with the expected background of atmospheric
muons. The expected background is 10.6+5.0 −3.6 events, which has comparable contributions from
atmospheric muons and atmospheric neutrinos. The sample of 28 events represents an excess of
more than 4σ above background.
Figure 6 shows the energy and zenith-angle dependence of the 28 events. There is a significant
excess of events above 100 TeV compared with the background expectation. Both the energy and
zenith-angle dependence observed are consistent with what is expected for a flux of neutrinos
produced by cosmic-ray accelerators. The flavor composition of the flux is, after corrections for
the acceptances of the detector to the different flavors, consistent with 1:1:1, as anticipated for a
flux originating in cosmic sources.
The large errors on the background are associated with the possible presence of a neutrino
component originating from the production and prompt leptonic decays of charmed particles in
the atmosphere. Such a flux has not been observed so far. Although its energy and zenith-angle
dependence are known, its normalization is not; see Figure 2 for one attempt at calculating the flux
of charm origin. Neither the energy nor the zenith-angle dependence of the 28 events observed
can be described by a charm flux, and in any case, fewer than 3.4 events are allowed at the 1σ level
by the present upper limit on a charm component of the atmospheric flux set by IceCube itself
(55).

www.annualreviews.org • IceCube 111


NS64CH04-Gaisser ARI 6 September 2014 14:0

Southern Sky (downgoing) Northern Sky (upgoing)

102 a 10 b

Events per 662 days


Events per 662 days

101

100
4
Annu. Rev. Nucl. Part. Sci. 2014.64:101-123. Downloaded from www.annualreviews.org
Access provided by University of Texas - Arlington on 09/24/20. For personal use only.

10–1 2

0
102 103 –1.0 –0.5 0.0 0.5 1.0
Deposited EM-equivalent energy in detector (TeV) sin (declination)

Background atmospheric muon flux


Background atmospheric neutrinos (π/K)
Background statistical and systematic uncertainties
Atmospheric neutrinos (benchmark charm flux)
Atmospheric neutrinos (90%-CL charm limit)
Signal + background best-fit astrophysical E–2 spectrum
Data

Figure 6
Distribution of (a) the deposited energies and (b) declination angles of the observed events compared with model predictions. Energies
plotted are in-detector visible energies, which are lower limits on the neutrino energy. Note that deposited energy spectra are always
harder than the spectra of the neutrinos that produced them because the neutrino cross section increases with energy. The expected
rate of atmospheric neutrinos is based on Northern Hemisphere muon neutrino observations. The estimated distribution of the
background from atmospheric muons is shown in red. Due to lack of statistics from data far above the cut threshold, the shape of the
distributions from muons in this figure has been determined using Monte Carlo simulations, with total rate normalized to the estimate
obtained from the in-data control sample. Combined statistical and systematic uncertainties on the sum of backgrounds are shown as a
hatched area. The gray line shows the best-fit E−2 astrophysical spectrum with all-flavor normalization (1:1:1) of Equation 3 and a
spectral cutoff of 2 PeV. Abbreviation: EM, electromagnetic. Modified from Reference 20 with permission.

Fitting the data to a superposition of extraterrestrial neutrinos on an atmospheric background


yields a cosmic-neutrino flux of
dN
Eν2 = 3 × 10−8 GeV cm−2 s−1 sr−1 3.
dEν
for the sum of the three neutrino flavors. As discussed in Section 2, this is the level of flux anticipated
for neutrinos accompanying the observed cosmic rays.
So, where do the neutrinos come from? A map of their arrival directions is shown in Figure 7.
A test statistic, TS = 2 × log L/L0 , was used, where L is the signal-plus-background maximized
likelihood and L0 is the background-only likelihood obtained by scrambling the data. No signifi-
cant spot on the sky was found when the data were compared with randomized pseudoexperiments.
A repeated analysis only for showers showed a hot spot at a right ascension of 281◦ and a declina-
tion of 23◦ close to the Galactic center. After correcting for trials, the probability corresponding
to its TS is 8%. Searches were also made for clustering of the events in time and for a possible

112 Gaisser · Halzen


NS64CH04-Gaisser ARI 6 September 2014 14:0

13
9
26
17

360° 5 1 0°
11
23 27
25 24
8
18 22 16 6 21
2 14 3 10
Annu. Rev. Nucl. Part. Sci. 2014.64:101-123. Downloaded from www.annualreviews.org

7
Access provided by University of Texas - Arlington on 09/24/20. For personal use only.

15 4
12
19
20 Equatorial
28
coordinates

0 TS = 2log (L/L0) 12.4

Figure 7
Sky map in equatorial coordinates of the test statistic (TS) that measures the probability of clustering among
the 28 events. The most significant cluster consists of five events—all showers and including the second-
highest-energy event in the sample—with a final significance of only 8%. The Galactic plane is shown as a
gray line, and the Galactic center appears as a filled gray square. Best-fit locations of individual events are
indicated with vertical crosses (+) for showers and angled crosses (x) for muon tracks. Modified from
Reference 20.

correlation with the times of observed GRBs. No statistically significant correlation was found.
Fortunately, more data are already available, and the analysis, performed blind, can be optimized
for searches of future data samples.

7. NEUTRINO ASTRONOMY II
During construction, IceCube collected more than 100,000 νμ s from the Northern Hemisphere,
detected by their secondary, upgoing muon tracks originating in neutrino interactions inside or
near the detector. Above neutrino energies of tens of TeV, where potential sources may dominate
the steeply falling atmospheric neutrino background, these tracks are reconstructed to better
than 0.5◦ ; at these energies neutrino and muon directions are aligned. The muons lose energy
predominantly by catastrophic interactions in the ice, and the measured dE/dx of the muon inside
the detector can be used as a proxy for the neutrino energy. A straightforward way to search
for sources of neutrinos is to look for clustering of events in arrival direction; the energy of the
events in a cluster can subsequently be used to further separate very high energy cosmic neutrinos
from atmospheric background. As discussed in the previous section, at energies above 300 TeV,
atmospheric neutrino events are very rare and energy is sufficient to establish an extraterrestrial
source, even in the Southern Hemisphere.
The status of the point source search is summarized in the 3-year sky map (60) obtained when
IceCube took data between April 2008 and May 2011 with partial configurations of 40, 59, and
79 out of 86 strings when completed (Figure 8). As in Figure 7, the color scheme shows at each

www.annualreviews.org • IceCube 113


NS64CH04-Gaisser ARI 6 September 2014 14:0

Point sources: significance sky map (IC79+59+40)

+85°

Atmospheric neutrinos
TeV–PeV
Northern Hemisphere
+45° 108,317 events

24 h 0h
Atmospheric muons
Annu. Rev. Nucl. Part. Sci. 2014.64:101-123. Downloaded from www.annualreviews.org

–45°
Access provided by University of Texas - Arlington on 09/24/20. For personal use only.

146,018 events
Southern Hemisphere
PeV–EeV –85°

0.0 0.6 1.2 1.8 2.4 3.0 3.6 4.2 4.8 5.4 6.0
–log10 p

Figure 8
Pretrial significance sky map in equatorial coordinates ( J2000) of the all-sky point source scan for the
combined IC79+IC59+IC40 data sample (60). The dashed line indicates the Galactic plane. Modified from
Reference 60.

location in the sky the probability of a source based on TS = 2 × log L/L0 , where, in this case, L
is the likelihood that a source with spectral index γ produces a signal with the significance shown.
The likelihood of background L0 is evaluated using the data, which are dominated by background.
For the detailed definition of the probability distribution functions, we refer the reader to
Reference 60. In brief, for a source location xs , each event is assigned a source probability density
corresponding to the probability of the event belonging to a source at xs . The source probability
density is the product of a spatial density function describing the potential of an event reconstructed
with direction xi to have true direction xs and the probability of observing reconstructed muon
energy Ei given the source spectral index. The source is assumed to emit neutrinos according to
an E −γ power law energy spectrum.
To evaluate the posttrial probability, also known as the look-elsewhere effect, one performs
pseudoexperiments in which the data are randomized to determine the significance of a possible
excess. This unbinned likelihood analysis results in sensitivity (upper limit in the absence of
an excess) and discovery potential as a function of zenith angle (Figure 9). Neutrino limits on
northern sources still exceed expectations by a factor of a few, but IceCube point source analysis is
approaching the sensitivity to reveal sources associated with the cosmic-neutrino flux in Figure 7.
Promising targets include the hard-spectrum sources observed by the MILAGRO γ -ray tele-
scope at tens of TeV with a differential slope γ  2. Such sources produce a neutrino flux on the
order of Eν2 (dN/dEν )  10−12 TeV cm−2 s−1 , assuming that the observed γ -rays are of pionic
origin (62). This is definitely the case for γ -ray production by cosmic rays in molecular clouds.
At this level, only about one event per year is expected in IceCube. However, such a low rate may
be sufficient to produce a statistically significant signal after several years because the atmospheric

114 Gaisser · Halzen


NS64CH04-Gaisser ARI 6 September 2014 14:0

IC79+IC59+IC40 sensitivity (90% CL)


IC79+IC59+IC40 discovery potential (5σ)
–9
10 IC79+IC59+IC40 upper limits (90% CL)
ANTARES sensitivity (90% CL)
ANTARES upper limits (90% CL)
E2dN/dE (TeV cm–2 s–1)

10–10

10–11
Annu. Rev. Nucl. Part. Sci. 2014.64:101-123. Downloaded from www.annualreviews.org
Access provided by University of Texas - Arlington on 09/24/20. For personal use only.

10–12
–1.0 –0.5 0.0 0.5 1.0
sin (δ)
Figure 9
Muon neutrino and antineutrino flux 90%-CL upper limits and sensitivities for an E−2 spectrum (60) and
the published limits of ANTARES (61). The different likelihood function and method to derive upper limits
used by ANTARES may account for differences in the limits from the two experiments at the level of 20%.
The points show the actual upper limits for a list of selected point sources. Modified from Reference 60.

flux within a ∼0.5◦ bin defined by the resolution is smaller yet, especially at the higher energies.
Therefore, not only the slope of the energy spectrum but also how high a source extends in energy
is critical. Sources of the still-unidentified “PeVatrons” that accelerate the highest-energy Galactic
cosmic rays to the knee would be expected to produce secondary pionic γ -rays and neutrinos of
several hundred TeV, provided that their environment is suitable for pion production.
GRBs constitute a special class of point sources because they cluster in both space and time. A
powerful model-independent analysis method has been developed to search for neutrinos coinci-
dent in time (up to 1,000 s) and direction with GRB flares observed by satellites, most prominently
Fermi and Swift. Data taken with the 40- and 59-string configurations led to an upper limit on the
GRB flux that is a factor of four below the predictions (63). This result implies either that GRBs
alone cannot account for the flux of extragalactic cosmic rays or that the efficiency of neutrino
production is much lower than has been predicted. Scanning a wider time window, an event with
109 TeV deposited energy and angular resolution within 0.2◦ of a Swift GRB was found to have
the most significant correlation, even though it was 14 h before the GRB. Subsequent analysis
showed that the event had hits in IceTop, marking it as a cosmic-ray background event and at
the same time illustrating the sparseness of the background for the GRB search with νμ -induced
muons. With GRBs on probation, the stock rises for the alternative speculation that associates
supermassive black holes at the centers of galaxies with the unaccounted-for cosmic-ray accelera-
tors. Suggestions to reduce the GRB neutrino flux prediction (64) will be confronted with rapidly
improving limits from a completed detector.

8. COSMIC RAYS
The high-energy cosmic-ray spectrum is related to neutrino astrophysics with IceCube in two
ways. At Earth, it produces the atmospheric muons and neutrinos that are the background for

www.annualreviews.org • IceCube 115


NS64CH04-Gaisser ARI 6 September 2014 14:0

a b

360° 0° 107 108 109

(GeV1.7 m–2 sr–1 s–1)


–1.5 –1 –0.5 0 0.5 1 1.5
Relative intensity (×10–3)

360° 0°
104

dE dA dΩ dt
–1.5 –1 –0.5 0 0.5 1 1.5
dN IceTop 73, SIBYLL 2.1, H4a composition assumption
Annu. Rev. Nucl. Part. Sci. 2014.64:101-123. Downloaded from www.annualreviews.org
Access provided by University of Texas - Arlington on 09/24/20. For personal use only.

Relative intensity (×10–3) KASCADE-Grande, SIBYLL 2.1


KASCADE, SIBYLL 2.1
GAMMA 2008
E 2.7 ×

Tunka-133
Tibet III, SIBYLL 2.1
360° 0°
6.5 7.0 7.5 8.0 8.5 9.0
–3 –2 –1 0 1 2 3 log10 E (GeV)
Relative intensity (×10–3)

Figure 10
(a) Maps of cosmic-ray intensity measured with IceCube with (top to bottom) mean energies of 20 (65), 400 (66), and 2,000 TeV (67).
(b) Primary cosmic-ray spectrum measured with IceCube. The black points represent the energy spectrum so far measured with
IceTop. The shaded gray band represents the largest contribution to the systematic uncertainty in the spectrum from IceTop alone.
Panel b modified from Reference 72.

astrophysical neutrinos. In the cosmos, in or near their sources and during propagation, cosmic
rays interact to produce secondary particles, including γ -rays and neutrinos.
Two aspects of the cosmic-ray spectrum at Earth are being measured by IceCube. One is the
anisotropy in the arrival directions of the cosmic rays in the energy range from 20 TeV to 2 PeV
(Figure 10a). The lower-energy measurements [∼20 (65) and ∼400 TeV (66)] are made using
penetrating muons reconstructed in the deep array of IceCube. The high-energy plot (∼2 PeV)
is made with the IceTop air shower array on the surface (67). These are the first measurements
of the cosmic-ray anisotropy in the southern sky and complete the view of cosmic-ray arrival
directions in this energy range observed by detectors in the north (68, 69). The data here are
smoothed on a 20◦ angular scale. Anisotropies have also been measured on a smaller scale (70). A
noteworthy feature of the large-scale anisotropy in Figure 10 is that its phase changes between
20 and 400 TeV and then persists as energy increases further. The amplitude increases as well,
going to 2 PeV.
With its surface component, IceTop (71), IceCube is a three-dimensional array for cosmic-
ray physics. The main goal is to measure the spectrum and composition of the cosmic rays from
300 TeV to above 1 EeV. Figure 10b shows the energy spectrum measured so far with IceTop
(72) in comparison with several other measurements made in the past decade (73–77). The fine
resolution obtained with IceTop is in part due to the high altitude at the South Pole, which allows
the showers to be measured near their maximum, reducing shower-to-shower fluctuations. As a
consequence, the measurement is able to reveal deviations from smooth power law behavior with
a high degree of resolution. In addition to the steepening of the spectrum in the well-known knee

116 Gaisser · Halzen


NS64CH04-Gaisser ARI 6 September 2014 14:0

region above 3 PeV, there is a structure around 100 PeV where the spectrum first hardens, starting
at 20 PeV, and then steepens above 140 PeV.
The largest contribution to the systematic uncertainty in the spectrum from IceTop alone
(Figure 10) is from uncertainty in the assumption about primary composition that is made to
obtain the energy spectrum. Information about the composition can be obtained using the full
three-dimensional capability of IceCube to measure the energy deposition in the deep detector,
as well as the shower at the surface for the subset of showers with trajectories that pass through
both parts of the detector. The ratio of shower size at the surface to energy deposited by the muon
bundle in the deep ice is sensitive to primary composition because heavy nuclei produce more
muons at a given energy than do light nuclei. A preliminary analysis (78) uses a neutral network
to unfold energy spectrum and composition. The energy spectrum obtained agrees well with that
measured by IceTop alone, whereas the proportion of heavy nuclei increases from 1 to 300 PeV.
Annu. Rev. Nucl. Part. Sci. 2014.64:101-123. Downloaded from www.annualreviews.org
Access provided by University of Texas - Arlington on 09/24/20. For personal use only.

9. SEARCH FOR DARK MATTER


IceCube searches indirectly for dark matter by looking for neutrinos from concentrations of weakly
interacting massive particles (WIMPs) in the Sun (79), in the Milky Way (80, 81), and in nearby
external galaxies (82). The neutrinos are secondary products of annihilation of pairs of WIMPs
into Standard Model particles, which include decays to neutrinos. In the case of annihilation in
the Sun or the Earth, only prompt decays allow production of neutrinos.
IceCube is most sensitive to WIMPs with significant spin-dependent cross sections with pro-
tons. These would lead to strong concentrations of WIMPs in the Sun, a nearby and readily
identifiable source. The signal would simply be an excess of neutrinos from the direction of the
Sun over the atmospheric neutrino background in the same angular window. There would be no
alternative astrophysical explanation of such a signal. In most WIMP scenarios, the cross sections
for WIMP capture (σχ , p ) and for WIMP annihilation (σχ ,χ ) are sufficiently large for an equilib-
rium between capture and annihilation to be achieved within the age of the solar system (83). In
this case, limits on neutrinos from the Sun can be expressed in terms of the capture cross section,
σχ , p . If equilibrium is not reached, weaker limits can still be calculated.
Figure 11 shows the current IceCube limits (79). Also shown is a sampling of the WIMP
parameter space that is not ruled out by other experiments. IceCube has produced the most
stringent limits on the cross section for spin-dependent interactions of dark matter particles with
ordinary matter.
A WIMP may interact with ordinary nuclei by spin-independent (e.g., Higgs boson exchange)
and spin-dependent (e.g., Z boson exchange) interactions. The first mechanism favors direct
detection experiments (84) because the WIMP interacts coherently, resulting in an increase in
sensitivity proportional to the square of the atomic number of the detector material. The latter
favors indirect experiments in which the rates are dominated by spin-dependent interactions with
protons. In the case of the Sun, the WIMPs have accumulated over solar timescales, sampling the
dark matter throughout the galaxy and averaging out any structure in the halo that theory may
not have accounted for. Within the context of supersymmetry, direct and indirect experiments
are complementary.
Quantitative interpretation of the IceCube limits is complicated, depending on detailed analysis
of capture rates and annihilation channels. The analysis uses DarkSUSY (83), which builds on the
classic calculations by Gould (85 and references therein) for capture and annihilation rates. The
neutrino spectrum from annihilation of pairs of WIMPs depends on the dominant channel for
coupling to Standard Model particles. The neutrino spectrum at production is also modified
by subsequent interactions and oscillations of neutrinos as they propagate out from the solar

www.annualreviews.org • IceCube 117


NS64CH04-Gaisser ARI 6 September 2014 14:0

MSSM including XENON (2012), ATLAS + CMS (2012)


–35 DAMA no channeling (2008)
COUPP (2012)
Simple (2011)
PICASSO (2012)
–36 Super-K (2011) (bb)
Super-K (2011) (W+W–)
log10(σSD, p cm–2)

–37

–38
Annu. Rev. Nucl. Part. Sci. 2014.64:101-123. Downloaded from www.annualreviews.org
Access provided by University of Texas - Arlington on 09/24/20. For personal use only.

–39

IceCube 2012 (bb)


IceCube 2012 (W+W–)
–40
(τ+τ – for mχ<mW = 80.4 GeV/c2)

1 2 3 4
log10 [mχ (GeV/c2)]

Figure 11
Upper limits at 90% CL on the spin-dependent neutralino–proton cross section, assuming that the neutrinos are produced by b b̄ and
WW annihilation (79). Limits from the Super-Kamiokande (Super-K) and direct detection experiments are shown for comparison. The
shaded region represents supersymmetric models that have not been ruled out by direct experiments. Abbreviation: MSSM, minimal
supersymmetric Standard Model. Modified from Reference 79.

core. Three flavor oscillations with matter effects are included (e.g., Reference 86). The range of
possibilities is bracketed by calculating two extremes,

χ χ → b b̄ (soft) and χ χ → W + W − (hard). 4.


± ±
In the hard channel, W are replaced by τ for mχ < mW .
The IceCube limits in Figure 11 are based on observations of νμ -induced muons with the nearly
complete detector with 79 strings of DOMs, including for the first time the DeepCore subarray.
Data were collected from May 2010 to May 2011, including the austral summer (October–March),
when the Sun is above the horizon. By including events that start inside DeepCore (summer and
winter), the mass range for the WIMP search could be extended down to 20 GeV, which overlaps
some of the allowed region from the DAMA experiment (87). By including muons from below the
horizon entering IceCube from outside (winter only because of the high background of cosmic-ray
muons from above), the mass range is extended to 5 TeV.

10. IceCube AS A DISCOVERY INSTRUMENT

10.1. Neutrino Oscillations


The first IceCube oscillation analysis (88) also uses data from the 79-string detector from May 2010
to May 2011. The analysis is based entirely on νμ -induced muons from below the horizon. Taking

118 Gaisser · Halzen


NS64CH04-Gaisser ARI 6 September 2014 14:0

advantage of the DeepCore subarray of IceCube allowed neutrino oscillations to be observed over
an energy range including the oscillation minimum around 25 GeV for propagation through the
diameter of the Earth. Data were divided into two samples: muon tracks reconstructed using the
entire IceCube detector (Eν > 100 GeV) and events starting in DeepCore (20 < Eν < 100 GeV).
The low-energy sample consists of 719 events, whereas the high-energy sample includes 39,638
events. The high-energy sample, in which standard oscillations do not affect the rates, was used
for calibration. A deficit is observed in the low-energy sample, in which ∼25% more events would
have been detected in the absence of oscillations. Taking account of systematic uncertainties,
including ± 0.05 in the spectral index of the atmospheric neutrino flux at production, the no-
oscillation hypothesis was rejected at more than 5σ. The fitted values of the oscillation parameters
in a two-flavor fit are | m232 | = 2.3+0.5−0.6 × 10
−3
eV2 and sin2 2θ23 > 0.93. For comparison, a
recent global three-flavor analysis (89) gave 2.4 × 10−3 eV2 and 0.95, respectively, with a range
of ±5% at 1σ and a slight dependence on the mass hierarchy. Although the measured oscillation
Annu. Rev. Nucl. Part. Sci. 2014.64:101-123. Downloaded from www.annualreviews.org
Access provided by University of Texas - Arlington on 09/24/20. For personal use only.

parameters agree with previous experiments, it is important to realize that they have been measured
at a characteristic energy that is higher. The measurement is therefore also sensitive to any new
neutrino physics, an important consideration when the precision of the IceCube measurements
will be significantly improved.

10.2. Supernovae and Solar Flares


In addition to the normal acquisition of events that reconstruct as tracks or cascades in the deep
array of IceCube and as air showers in IceTop, the rates at which the PMT voltages cross the
thresholds of discriminators in the DOMs are continuously monitored. A typical rate for DOMs
in the deep ice is 500 Hz (including correlated afterpulses), most of which is noise. Typical rates
in the high-gain DOMs of IceTop are 2–5 kHz, most of which is induced by low-energy photons,
electrons, and muons entering the tanks.
A sudden increase in the total summed counting rate of the deep DOMs would signify a
potential supernova explosion in the Galaxy. Supernova neutrinos of ∼10 MeV interacting within
a few meters of a DOM would generate enough hits to cause a sharp increase in the summed
counting rate followed by a characteristic decline (90). IceCube records a dc current that tracks
the time evolution of the supernova in microsecond time bins. However, the detector records
the time of every photoelectron with nanosecond precision, and the binning can therefore be
improved off-line, which will improve the capability to identify the deleptonization burst. The
additional measurement of the rate of two-photon correlations is sensitive to the energy of the
supernova neutrinos.
In IceTop, sudden changes in rates occur in response to solar events. Forbush decreases,
in which the plasma from a solar flare abruptly reduces the rate of cosmic rays entering the
atmosphere, are frequently detected and can be analyzed. More rare are sudden increases caused
by solar energetic particles that enter the atmosphere with sufficient energy to generate secondary
cosmic rays that reach the IceTop tanks. The event of December 13, 2006, was measured with the
16 tanks (8 stations) then in operation (91). The flare of May 17, 2012, is currently being analyzed.

10.3. IceCube, The Facility


During its construction phase, IceCube demonstrated significant potential for facilitating a range
of other research projects. For example, a dust logger provided measurements with millimeter
precision that are valuable for event reconstruction in IceCube but that also provide a record of
surface winds for more than 100,000 years (92).

www.annualreviews.org • IceCube 119


NS64CH04-Gaisser ARI 6 September 2014 14:0

During the construction of AMANDA, antennas forming the RICE detector were already
deployed in some holes to expand the target volume in the search for cosmogenic neutrinos (93).
An acoustic test setup of receivers in the upper portion of four IceCube holes was deployed in
2007 to explore the acoustic technique for detecting ultrahigh-energy neutrinos. A retrievable
transmitter (pinger) was submerged briefly in several newly prepared holes at various depths and
distances from the receivers to measure the attenuation of sound in ice. The attenuation length
of 300 m is significantly shorter than had been expected (94). The Askaryan Radio Array (ARA)
(95) plans to take advantage of the kilometer-scale attenuation for radio signals in ice to construct
a detector with a 200-km2 effective area. That effective area should be sufficient to determine the
level of production of cosmogenic neutrinos, which at present is highly uncertain (10). The initially
deployed ARA detectors send data to computers housed in the ICL for staging and transmission
to the north.
The DM-Ice experiment (96) proposes to repeat the DAMA experiment in the quiet environ-
Annu. Rev. Nucl. Part. Sci. 2014.64:101-123. Downloaded from www.annualreviews.org
Access provided by University of Texas - Arlington on 09/24/20. For personal use only.

ment of the Antarctic ice. An interesting feature of the observation is that the seasonal modulation
of the muon rate has the opposite phase relative to the motion of the Earth through the gas of
dark matter as compared with a detector in the Northern Hemisphere. A test detector to explore
the noise environment for DM-Ice was deployed at the bottom of an IceCube string in December
2010. Its computers and data transmission are also hosted in the ICL.
The enhancement of the low-energy capabilities of IceCube provided by the DeepCore subar-
ray led to the PINGU proposal (97) to deploy an additional 40 strings within the existing DeepCore
detector, which would lower the threshold to below 5 GeV (<25-m muon track length in ice). In
this energy range, matter effects in the Earth lead to resonant oscillations of νμ ↔ νe (ν̄μ ↔ ν̄e )
for normal (inverted) hierarchy (98) that depend on zenith angle. The fact that the neutrino cross
section is larger than that for antineutrinos, coupled with the excess of νμ compared with ν̄μ , allow
the possibility of a measurement sensitive to the neutrino mass hierarchy on a relatively short
timescale. PINGU would also have sensitivity to νμ disappearance, ντ appearance, and maximal
mixing. The lower-energy threshold would also enhance the indirect searches for dark matter
with IceCube, as well as the sensitivity to neutrinos from supernova explosions. In addition, there
is the potential for neutrino tomography of the Earth with PINGU.
The observation of the extraterrestrial neutrinos up to the PeV range has stimulated ideas to
enhance the discovery potential also in that area. One possibility is to improve the atmospheric
neutrino veto by expanding the surface array (99). Expansion of the existing array with strings at
larger separation is also possible in the long term.

11. OUTLOOK
Having found the first few high-energy neutrinos of extraterrestrial origin, the obvious next task
for IceCube is to identify their sources. How long this will take depends on the nature of the
sources. In the case of extragalactic sources, the answer depends on the number and distribution
of sources (100). Because the Universe is transparent to neutrinos, if there is a large number of
sources of comparable strength, then many neutrinos from random directions may be counted
before there are enough to identify a single nearby source. However, if Galactic sources contribute,
then identification of the sources should be possible on a well-defined timescale, as in the example
discussed in Section 7.
More than 30 papers have been written suggesting possible sources for the neutrinos detected
by IceCube since the two PeV events were discovered. Of these, approximately two-thirds favor
an extragalactic origin, and at least one paper proposes a combination of Galactic and extragalactic
sources. IceCube has a design lifetime of more than 15 years, and the detector continues to work

120 Gaisser · Halzen


NS64CH04-Gaisser ARI 6 September 2014 14:0

well. By the time this article is published, an additional 2 years of data will have been accumulated,
compared with the publication in Science (20). An additional cascade event with ∼2 PeV of energy
has already been identified in the later data, although an unbroken E−2 power law appears unlikely.
The question of the spectrum will be clarified as more data are accumulated. Rapid progress in
sorting out the possible models of the sources is expected.

DISCLOSURE STATEMENT
The authors are not aware of any affiliations, memberships, funding, or financial holdings that
might be perceived as affecting the objectivity of this review.

ACKNOWLEDGMENTS
Annu. Rev. Nucl. Part. Sci. 2014.64:101-123. Downloaded from www.annualreviews.org
Access provided by University of Texas - Arlington on 09/24/20. For personal use only.

We are grateful to our colleagues in the IceCube Collaboration who make the science a reality
and to the many individuals and agencies involved in the construction and operation of IceCube.
The operation of IceCube is supported primarily by the US National Science Foundation with
substantial additional support from several national and international agencies. A list of IceCube
institutions is available at http://icecube.wisc.edu/collaboration. A list of agencies providing
support for research with IceCube is posted at http://icecube.wisc.edu/collaboration/funding.

LITERATURE CITED
1. Reines F, Cowan CL Jr. Nature 17:446 (1956)
2. Greisen K. Annu. Rev. Nucl. Part. Sci. 10:63 (1960)
3. Reines F. Annu. Rev. Nucl. Part. Sci. 10:1 (1960)
4. Markov MA. Proc. 1960 Int. Conf. High Energy Phys. 1:578 (1960)
5. Gaisser TK, Halzen F, Stanev T. Phys. Rep. 258:173 (1995); Gaisser TK, Halzen F, Stanev T. Erratum.
Phys. Rep. 271:355 (1995); Learned JG, Mannheim K. Annu. Rev. Nucl. Part. Sci. 50:679 (2000); Halzen
F, Hooper D. Rep. Prog. Phys. 65:1025 (2002); Becker J. Phys. Rep. 458:173 (2008); Katz UF, Spiering C.
Prog. Part. Nucl. Phys. 67:651 (2012); Halzen F. Nuovo Cim. 36:N3 (2012)
6. Roberts A. Rev. Mod. Phys. 64:259 (1992)
7. Berezinsky VS, Zatsepin GT. Phys. Lett. B 28:423 (1969)
8. Greisen K. Phys. Rev. Lett. 16:748 (1966)
9. Zatsepin GT, Kuz’min VA. J. Exp. Theor. Phys. Lett. 4: 78 (1966)
10. Wdowczyk J, Tkaczyk W, Wolfendale AW. J. Phys. A 5:1419 (1972); Stecker FW. Astrophys. Space Sci.
20:47 (1973), Berezinsky VS, Smirnov AY. Astrophys. Space Sci. 32:461 (1975); Ahlers M, et al. Astropart.
Phys. 34:106 (2010)
11. Zheleznykh I. Int. J. Mod. Phys. 21S1:1 (2006); Markov M, Zheleznykh I. Nucl. Phys. 27:385 (1961)
12. Babson J, et al. Phys. Rev. D 42:3613 (1990)
13. Balkanov VA, et al. (BAIKAL Collab.) Nucl. Phys. B Proc. Suppl. 118:363 (2003)
14. Aggouras G, et al. (NESTOR Collab.) Astropart. Phys. 23:377 (2005)
15. Aguilar JA, et al. (ANTARES Collab.) Astropart. Phys. 26:314 (2006)
16. Migneco E. J. Phys. Conf. Ser. 136:022048 (2008)
17. Karle A. Proc. Int. Astron. Union Symp. 288 8:84 (2012)
18. IceCube Collab. Prelimiary Design Document, v1.24. http://www.icecube.wisc.edu/science/
publications/pdd/pdd.pdf (2001)
19. Ahrens J, et al. (IceCube Collab.) Astropart. Phys. 20:507 (2004)
20. Aartsen MG, et al. (IceCube Collab.) Science 342:1242856 (2013)
21. Sommers P. Astropart. Phys. 39/40:88 (2012)
22. Baade W, Zwicky F. Proc. Natl. Acad. Sci. USA 20:259 (1934)

www.annualreviews.org • IceCube 121


NS64CH04-Gaisser ARI 6 September 2014 14:0

23. Ginzburg VL, Syrovatskii SI. The Origin of Cosmic Rays. Oxford, UK: Pergamon (1964)
24. Butt Y. Nature 460:701 (2009)
25. Gaisser TK. AIP Conf. Proc. 558:27 (2001)
26. Waxman E. Phys. Rev. Lett. 75:386 (1995)
27. Becker JK. Phys. Rep. 458:173 (2008)
28. Gonzalez-Garcia MC, Maltoni M, Rojo J. J. High Energy Phys. 0610:075 (2006)
29. Daum K, et al. (Frejus Collab.) Z. Phys. C 66:417 (1995)
30. Abbasi R, et al. (IceCube Collab.) Phys. Rev. D 79:102005 (2009)
31. Abbasi R, et al. (IceCube Collab.) Astropart. Phys. 34:48 (2010)
32. Abbasi R, et al. (IceCube Collab.) Phys. Rev. D 83:012001 (2011)
33. Abbasi R, et al. (IceCube Collab.) Phys. Rev. D 84:082001 (2011)
34. Aartsen MG, et al. (IceCube Collab.) Phys. Rev. Lett. 110:151105 (2013)
35. Honda M, et al. Phys. Rev. D 75:043006 (2007)
36. Barr GD, et al. Phys. Rev. D 70:023006 (2004)
Annu. Rev. Nucl. Part. Sci. 2014.64:101-123. Downloaded from www.annualreviews.org
Access provided by University of Texas - Arlington on 09/24/20. For personal use only.

37. Enberg R, Reno MH, Sarcevic I. Phys. Rev. D 78:043005 (2008)


38. Berezhko EG. Astrophys. J. 684:L69 (2008)
39. Gaisser TK. arXiv:astro-ph/9707283 (1997)
40. Ahlers M, et al. Phys. Rev. D 72:023001 (2005)
41. Alvarez-Muñiz J, Halzen F. Astrophys. J. 576:L33 (2002)
42. Gaisser TK. EPJ Web Conf. 53:01012 (2013)
43. Abbasi R, et al. (IceCube Collab.) Astrophys. J. 732:18 (2011)
44. Halzen F. Nuovo Cim. 036:81 (2013)
45. Learned JG, Pakvasa S. Astropart. Phys. 3:267 (1995)
46. Beringer J, et al. (Part. Data Group) Phys. Rev. D 86:010001 (2012)
47. Aartsen MG, et al. (IceCube Collab.) Nucl. Instrum. Methods A 711:73 (2013)
48. Ahrens J, et al. (AMANDA Collab.) Nucl. Instrum. Methods A 524:169 (2004)
49. Aartsen MG, et al. (IceCube Collab.) J. Instrum. 9:P03009 (2014)
50. Abbasi R, et al. (IceCube Collab.) Nucl. Instrum. Methods A 618:139 (2010)
51. Abbasi R, et al. (IceCube Collab.) Nucl. Instrum. Methods A 601:294 (2009)
52. Schönert S, Gaisser TK, Resconi E, Schulz O. Phys. Rev. D 79:043009 (2009)
53. Gaisser TK, Stanev T, Tilav S. Front. Phys. 8:748 (2013)
54. Aglietta M, et al. (LVD Collab.) Phys. Rev. D 60:112001 (1999)
55. Schukraft A. (IceCube Collab.) Nucl. Phys. Proc. Suppl. 237/238:266 (2013)
56. Gaisser TK. EPJ Web Conf. 52:09004 (2013)
57. Desiati P, Gaisser TK. Phys. Rev. Lett. 105:121102 (2010)
58. Aartsen MG, et al. (IceCube Collab.) Phys. Rev. Lett. 111:021103 (2013)
59. Aartsen MG, et al. (IceCube Collab.) Phys. Rev. D 89:062007 (2014)
60. Aartsen MG, et al. (IceCube Collab.) Astrophys. J. 779:132 (2013)
61. Adriá-Martı́nez S, et al. (ANTARES Collab.) Astrophys. J. Lett. 743:14 (2011)
62. Gonzalez-Garcia MC, Halzen F, Mohapatra S. Astropart. Phys. 31:437 (2009)
63. Abbasi R, et al. (IceCube Collab.) Nature 484:351 (2012)
64. Hummer S, Baerwald P, Winter W. Phys. Rev. Lett. 108:231101 (2012)
65. Abbasi R, et al. (IceCube Collab.) Astrophys. J. 718:L194 (2010)
66. Abbasi R, et al. (IceCube Collab.) Astrophys. J. 746:33 (2012)
67. Aartsen MG, et al. (IceCube Collab.) Astrophys. J. 765:55 (2013)
68. Abdo AA, et al. Astrophys. J. 698:2121 (2009)
69. Amenomori M, et al. (Tibet AS Gamma Collab.) Astrophys. J. 626:L29 (2005)
70. Abbasi R, et al. (IceCube Collab.) Astrophys. J. 740:16 (2011)
71. Abbasi R, et al. (IceCube Collab.) Nucl. Instrum. Methods A 700:188 (2013)
72. Aartsen MG, et al. (IceCube Collab.) Phys. Rev. D 88:042004 (2013)
73. Apel WD, et al. (KASCADE-Grande Collab.) Phys. Rev. Lett. 107:171104 (2011)
74. Antoni T, et al. (KASCADE Collab.) Astropart. Phys. 24:1 (2005)
75. Garyaka AP, et al. J. Phys. G 35:115201 (2008)

122 Gaisser · Halzen


NS64CH04-Gaisser ARI 6 September 2014 14:0

76. Berezhnev SF, et al. Nucl. Instrum. Methods A 692:98 (2012)


77. Amenomori M, et al. (TIBET III Collab.) Astrophys. J. 678:1165 (2008)
78. Aartsen MG, et al. (IceCube Collab.) Proc. 33rd Int. Cosmic Ray Conf. Pap. 0861
79. Aartsen MG, et al. (IceCube Collab.) Phys. Rev. Lett. 110:131302 (2013)
80. Abbasi R, et al. (IceCube Collab.) Phys. Rev. D 84:022004 (2011)
81. Abbasi R, et al. (IceCube Collab.) arXiv:1210.3557 [hep-ex] (2012)
82. Aartsen MG, et al. Phys. Rev. D 88:122001 (2013)
83. Gondolo P, et al. J. Cosmol. Astropart. Phys. 0407:008 (2004)
84. Sadoulet B. Science 315:61 (2007)
85. Gould A. Astrophys. J. 368:610 (1991)
86. Blennow M, Edsjö J, Ohlsson T. J. Cosmol. Astropart. Phys. 01:021 (2008)
87. Bernabei R, et al. Int. J. Mod. Phys. D 22:1360001 (2013)
88. Aartsen MG, et al. (IceCube Collab.) Phys. Rev. Lett. 111:081801 (2013)
89. Fogli GL, et al. Phys. Rev. D 86:013012 (2012)
Annu. Rev. Nucl. Part. Sci. 2014.64:101-123. Downloaded from www.annualreviews.org
Access provided by University of Texas - Arlington on 09/24/20. For personal use only.

90. Abbasi R, et al. (IceCube Collab.) Astron. Astrophys. 535:A109 (2011)


91. Abbasi R, et al. (IceCube Collab.) Astrophys. J. Lett. 689:65 (2008)
92. Bay RC, Rohde RA, Price PB, Bramall NE. J. Geophys. Res. 115:D14126 (2010)
93. Kravchenko I, et al. Phys. Rev. D 85:062004 (2012)
94. Abbasi R, et al. (IceCube Collab.) Astropart. Phys. 34:382 (2011)
95. Allison P, et al. Astropart. Phys. 35:457 (2012)
96. Cherwinka J, et al. Astropart. Phys. 35:749 (2012)
97. IceCube–PINGU Collab. arXiv:1401.2046 [physics] (2014)
98. Akhmedov EK, Razzaque S, Smirnov AY. J. High Energy Phys. 1302:082 (2013); Akhmedov EK, Razzaque
S, Smirnov AY. Erratum. J. High Energy Phys. 1307:026 (2013)
99. Aartsen MG, et al. (IceCube Collab.) arXiv:1309.7010 [astro-ph] (2013)
100. Lipari P. Phys. Rev. D 78:083011 (2008)

www.annualreviews.org • IceCube 123


NS64-FrontMatter ARI 6 September 2014 11:38

Annual Review of
Nuclear and
Particle Science

Contents Volume 64, 2014

A Life in High-Energy Physics: Success Beyond Expectations


Annu. Rev. Nucl. Part. Sci. 2014.64:101-123. Downloaded from www.annualreviews.org
Access provided by University of Texas - Arlington on 09/24/20. For personal use only.

James W. Cronin p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 1
Hadron Polarizabilities
Barry R. Holstein and Stefan Scherer p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p51
Effective Field Theory Beyond the Standard Model
Scott Willenbrock and Cen Zhang p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p83
IceCube
Thomas Gaisser and Francis Halzen p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 101
Fluid Dynamics and Viscosity in Strongly Correlated Fluids
Thomas Schäfer p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 125
Mesonic Low-Energy Constants
Johan Bijnens and Gerhard Ecker p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 149
Superconducting Radio-Frequency Cavities
Hasan S. Padamsee p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 175
TeV-Scale Strings
David Berenstein p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 197
J/ψ and ϒ Polarization in Hadronic Production Processes
Eric Braaten and James Russ p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 221
The First Direct Observation of Double-Beta Decay
Michael Moe p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 247
Weak Polarized Electron Scattering
Jens Erler, Charles J. Horowitz, Sonny Mantry, and Paul A. Souder p p p p p p p p p p p p p p p p p p p 269
Cooling of High-Energy Hadron Beams
Michael Blaskiewicz p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 299
Status and Implications of Beyond-the-Standard-Model
Searches at the LHC
Eva Halkiadakis, George Redlinger, and David Shih p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 319

v
NS64-FrontMatter ARI 6 September 2014 11:38

The Measurement of Neutrino Properties with Atmospheric Neutrinos


Takaaki Kajita p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 343
Properties of the Top Quark
Frédéric Déliot, Nicholas Hadley, Stephen Parke, and Tom Schwarz p p p p p p p p p p p p p p p p p p p p 363
Hard-Scattering Results in Heavy-Ion Collisions at the LHC
Edwin Norbeck, Karel Šafařı́k, and Peter A. Steinberg p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 383

Index

Cumulative Index of Contributing Authors, Volumes 55–64 p p p p p p p p p p p p p p p p p p p p p p p p p p p 413


Annu. Rev. Nucl. Part. Sci. 2014.64:101-123. Downloaded from www.annualreviews.org
Access provided by University of Texas - Arlington on 09/24/20. For personal use only.

Errata

An online log of corrections to Annual Review of Nuclear and Particle Science articles may
be found at http://www.annualreviews.org/errata/nucl

vi Contents

You might also like