Results from the ANTARES neutrino telescope with six years of data

Similar documents
Search for diffuse cosmic neutrino fluxes with the ANTARES detector

Lawrence Berkeley National Laboratory Lawrence Berkeley National Laboratory

Results from the ANTARES neutrino telescope

Search for high-energy neutrinos from GRB130427A with the ANTARES neutrino telescope

Recent results of the ANTARES neutrino telescope

Search for a diffuse cosmic neutrino flux with ANTARES using track and cascade events

A Summary of recent Updates in the Search for Cosmic Ray Sources using the IceCube Detector

Recent Results from the ANTARES experiment

Dept. of Physics and Astronomy, Michigan State University, 567 Wilson Rd., East Lansing, MI 48824, USA

PoS(NOW2016)041. IceCube and High Energy Neutrinos. J. Kiryluk (for the IceCube Collaboration)

High energy neutrino astronomy with the ANTARES Cherenkov telescope

Detection of transient sources with the ANTARES telescope. Manuela Vecchi CPPM

High Energy Neutrino Astrophysics Latest results and future prospects

Indirect detection of Dark Matter with the ANTARES Neutrino Telescope

Search for high energy neutrino astrophysical sources with the ANTARES Cherenkov telescope

Multi-Messenger Programs in ANTARES: Status and Prospects. Véronique Van Elewyck Laboratoire AstroParticule et Cosmologie (Paris)

NEUTRINO ASTRONOMY AT THE SOUTH POLE

Astroparticle Physics with IceCube

Study of the high energy Cosmic Rays large scale anisotropies with the ANTARES neutrino telescope

IceCube Results & PINGU Perspectives

SELECTED RESULTS OF THE ANTARES TELESCOPE AND PERSPECTIVES FOR KM3NET. D. Dornic (CPPM) on behalf the ANTARES Coll.

On the scientific motivation for a wide field-of-view TeV gamma-ray observatory in the Southern Hemisphere

Follow-up of high energy neutrinos detected by the ANTARES telescope

Neutrino Astronomy. Ph 135 Scott Wilbur

High-energy neutrino detection with the ANTARES underwater erenkov telescope. Manuela Vecchi Supervisor: Prof. Antonio Capone

Brazilian Journal of Physics ISSN: Sociedade Brasileira de Física Brasil

PHY326/426 Dark Matter and the Universe. Dr. Vitaly Kudryavtsev F9b, Tel.:

in multimessenger astrophysics.

Status and Perspectives for KM3NeT/ORCA

IceCube: Dawn of Multi-Messenger Astronomy

PoS(EPS-HEP2017)008. Status of the KM3NeT/ARCA telescope

Gamma-rays, neutrinos and AGILE. Fabrizio Lucarelli (ASI-SSDC & INAF-OAR)

SEARCH FOR NUCLEARITES WITH THE ANTARES DETECTOR *

Search for Point-like. Neutrino Telescope

Secluded Dark Matter search in the Sun with the ANTARES neutrino telescope

Search for GeV neutrinos associated with solar flares with IceCube

THE KM3NET NEUTRINO TELESCOPE IN THE MEDITERRANEAN SEA

PoS(ICRC2017)945. In-ice self-veto techniques for IceCube-Gen2. The IceCube-Gen2 Collaboration

Searches for astrophysical sources of neutrinos using cascade events in IceCube

IceCube. francis halzen. why would you want to build a a kilometer scale neutrino detector? IceCube: a cubic kilometer detector

Galactic diffuse neutrino component in the astrophysical excess measured by the IceCube experiment

neutrino astronomy francis halzen university of wisconsin

Multi-messenger studies of point sources using AMANDA/IceCube data and strategies

Search for GeV neutrinos associated with solar flares with IceCube

The ANTARES neutrino telescope:

A Search for Point Sources of High Energy Neutrinos with AMANDA-B10

A M A N DA Antarctic Muon And Neutrino Detector Array Status and Results

Gustav Wikström. for the IceCube collaboration

MACRO Atmospheric Neutrinos

First results of the ANTARES neutrino telescope

Understanding High Energy Neutrinos

Search for Dark Matter from the Galactic Halo with the IceCube Neutrino Observatory Paper Review

High Energy Neutrino Astronomy

Measurement of High Energy Neutrino Nucleon Cross Section and Astrophysical Neutrino Flux Anisotropy Study of Cascade Channel with IceCube

Muon Reconstruction in IceCube

Search for Astrophysical Neutrino Point Sources at Super-Kamiokande

Constraints on atmospheric charmed-meson production from IceCube

arxiv: v1 [astro-ph.he] 28 Jan 2013

High Energy Astrophysics with underwater neutrino detectors. Marco Anghinolfi INFN, Genova, Italia

PoS(NEUTEL2017)079. Blazar origin of some IceCube events

Combined Search for Neutrinos from Dark Matter Annihilation in the Galactic Center using IceCube and ANTARES

The KM3NeT/ARCA detection capability to a diffuse flux of cosmic neutrinos

Particle Physics with Neutrino Telescope Aart Heijboer, Nikhef

Multimessenger test of Hadronic model for Fermi Bubbles Soebur Razzaque! University of Johannesburg

IceCube: Ultra-high Energy Neutrinos

Indirect Dark Matter Detection

Antonio Capone Univ. La Sapienza INFN - Roma

High Energy Neutrino Astrophysics with IceCube

LATEST RESULTS OF THE ANTARES HIGH ENERGY NEUTRINO TELESCOPE

Neutrino induced muons

Investigation of Obscured Flat Spectrum Radio AGN with the IceCube Neutrino Observatory

First Light with the HAWC Gamma-Ray Observatory

Particle Physics Beyond Laboratory Energies

Searches for annihilating dark matter in the Milky Way halo with IceCube

Neutrino Astronomy with IceCube at the Earth's South Pole

A New View of the High-Energy γ-ray Sky with the Fermi Telescope

Detectors for astroparticle physics

Measuring the neutrino mass hierarchy with atmospheric neutrinos in IceCube(-Gen2)

UHE Cosmic Rays and Neutrinos with the Pierre Auger Observatory

High energy events in IceCube: hints of decaying leptophilic Dark Matter?

The positron and antiproton fluxes in Cosmic Rays

Neutrino Astronomy at the South Pole AMANDA and IceCube

Gamma-ray bursts as the sources of the ultra-high energy cosmic rays?

High Energy Emission. Brenda Dingus, LANL HAWC

Recent Results from ANTARES and prospects for KM3NeT. Aart Heijboer. Nikhef, Amsterdam On behalf of the ANTARES and KM3NeT collaborations

PoS(ICRC2017)972. Searches for neutrino fluxes in the EeV regime with the Pierre Auger Observatory. Enrique Zas a for the Pierre Auger Collaboration b

Kurt Woschnagg UC Berkeley

Neutrino Astronomy with AMANDA

Recent Developments of the Real-Time Capabilities of IceCube

A Multimessenger Neutrino Point Source Search with IceCube

Results of the search for magnetic

KM3NeT. P. Piattelli, INFN SciNeGHE 2010, Trieste, september

Cosmic Neutrinos in IceCube. Naoko Kurahashi Neilson University of Wisconsin, Madison IceCube Collaboration

Preliminary results from gamma-ray observations with the CALorimeteric Electron Telescope (CALET)

Searches for Dark Matter Annihilations in the Sun and Earth with IceCube and DeepCore. Matthias Danninger for the IceCube collaboration

Search for neutralino dark matter with the AMANDA neutrino telescope

Multi-wavelength follow-up observations of ANTARES neutrino alerts

From DeepCore to PINGU

Progress and latest results from Baikal, Nestor, NEMO and KM3NeT

Search for top squark pair production and decay in four bodies, with two leptons in the final state, at the ATLAS Experiment with LHC Run2 data

Transcription:

Available online at www.sciencedirect.com ScienceDirect Physics Procedia 61 (2015 ) 450 458 Results from the ANTARES neutrino telescope with six years of data M. Spurio, on the behalf of the ANTARES Collaboration Department of Physics and Astronomy, University of Bologna and INFN, Sezione di Bologna, 40100, Italy Abstract The ANTARES neutrino telescope, completed in 2008, is the largest neutrino telescope in the Northern hemisphere. Located at a depth of 2.5 km in the Mediterranean Sea, 40 km off the Toulon shore, its main goal is the search for high energy neutrinos of astrophysical origin. In this paper we review the main physics results, ranging from searches for steady point sources and diffuse fluxes of neutrinos, to multi-messenger analyses and particle physics. 2015 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/3.0/). c 2011 Published Elsevier Ltd. Selection and peer review is the responsibility of the Conference lead organizers, Frank Avignone, University of South Carolina, and Wick Haxton, University of California, Berkeley, and Lawrence Berkeley Laboratory Keywords: ANTARES neutrino telescope; neutrino astrophysics; cosmic rays; neutrino physics 1. Introduction Neutrino astronomy [1, 2, 3] shares with the γ-ray astronomy the objective to understand the sources and mechanisms of cosmic ray acceleration. Due to their much larger interaction cross-section, γ-rays are easier to detect than neutrinos, and γ-ray astronomy has having a fundamental importance on several topical areas of modern astrophysics and cosmology. Gamma-rays can be produced both by hadronic (through π 0 decay) or leptonic processes (inverse Compton, bremsstrahlung), neutrinos can only be produced by hadronic processes (charged meson decays). In most cases, it is generally possible to fit the γ-ray data with either leptonic or hadronic productions by varying the assumptions of the models (for instance, the intensity of magnetic fields or the environmental matter number density). No single source, either galactic or extragalactic, has been conclusively proven to accelerate cosmic rays at PeV energies. Neutrino astronomy is expected to be decisive for the quest of cosmic ray sources. The small neutrino interaction cross-section allows neutrinos to be not deflected nor absorbed, but it represents also a drawback, as their detection requires a large instrumented target mass (a so-called neutrino telescope ). Neutrino telescopes, at the contrary of any other instrument devoted to astronomy, are looking downward. Up-going muons can only be produced by charged current interactions of (up-going) ν μ. From the bottom hemisphere, the neutrino signal is almost background-free. A major challenge for neutrino telescopes is thus to separate the astrophysical signals from the large background of atmospheric neutrinos produced by cosmic ray interactions with atmospheric nuclei. Email address: spurio@bo.infn.it (on the behalf of the ANTARES Collaboration) 1875-3892 2015 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/3.0/). Selection and peer review is the responsibility of the Conference lead organizers, Frank Avignone, University of South Carolina, and Wick Haxton, University of California, Berkeley, and Lawrence Berkeley Laboratory doi:10.1016/j.phpro.2014.12.107

M. Spurio / Physics Procedia 61 (2015 ) 450 458 451 The detection principle of the ANTARES neutrino telescope, 2, is based on the observation of the Cherenkov light induced by secondary charged particles produced in charged current interactions of neutrinos inside or near the detector volume. The discrimination of cosmic neutrinos from the background rely on three methods: 1) the search for an excess of events in a very small solid angle region over the expected background, 3. 2) the search for an excess of events above a given energy, using the fact that the expected cosmic signal is harder than the atmospheric neutrino spectrum (searches for a diffuse flux), 4. 3) the searches for neutrino candidates correlated in space and/or time with external triggers from other experiments, 5. The first method uses the very good tracking capabilities of the detector, which for the ν μ channel can act as a conventional telescope. The second method uses the calorimetric capabilities: the energy of neutrinoinduced events (both track and shower events) should be estimated with some acceptable resolution. Finally, the third multi-messenger method relies on the time resolution and the information networking with other experiments. Beyond its astrophysics goals, the science reach of a neutrino telescope such as ANTARES includes an extended particle physics program, ranging from the study of neutrino oscillations, the indirect search for dark matter (in the GeV TeV energy range) the search for exotic particles 6. The ANTARES experiment offers also a deep-sea cabled observatory for sea and earth sciences, whose results are not covered in this review. 2. The ANTARES detector ANTARES is the first undersea neutrino telescope and the largest of its kind in the Northern Hemisphere [4]. It is located between 2475 m (seabed) and 2025 m below the Mediterranean Sea level, 40 km offshore from Toulon (France). The telescope consists of 12 detection lines with 25 storeys each. A standard storey includes three optical modules (OMs) [5] each housing a 10-inch photomultiplier [6] and a local control module that contains the electronics [7, 8]. The OMs are orientated 45 downwards in order to optimize their acceptance to upgoing light and to avoid the effect of sedimentation and biofouling [9]. The length of a line is 450 m and the horizontal distance between neighboring lines is 60-75 m. In one of the lines, the upper storeys are dedicated to a test system for acoustic neutrino detection [10]. Similar acoustic devices are also installed in an additional line that contains instrumentation aimed to measure environmental parameters [11]. The location of the active components of the lines is known better than 10 cm by a combination of tiltmeters and compasses in each storey and a series of acoustic transceivers (emitters and receivers) in certain storeys along the line and surrounding the telescope [12]. A charged current interaction of a ν μ (ν μ ) with the matter below or around the detector produces a relativistic muon that can travel hundreds of meters and cross the detector or pass nearby. This muon induces Cherenkov light when travelling through water, and some of the emitted photons produce a signal in the PMTs (the hits) with the corresponding charge and time information. The hits are used to reconstruct the direction of the track that is well correlated to the neutrino direction. The signature due to ν e (ν e ) and to neutral current interactions are particle showers, i.e. very localized energy deposits and light emission. Different reconstruction strategies are used, which provides worse angular resolution with respect to the ν μ channel and a better energy resolution for contained events. Data taking started with the first 5 lines of the detector installed in 2007. The full detector was completed in May 2008 and has been operating continuously ever since, except for some periods in which repair and maintenance operations have taken place. 3. Searches for sources of cosmic neutrinos The searches for significant excesses of events from particular small regions of the sky need the best possible determination of the incoming neutrino direction. Only the ν μ are suited for this method. The track-finding algorithm consists of a multi-step procedure to fit the muon direction by maximizing a likelihood parameter Λ, which describes the quality of the reconstruction. Neutrino candidates are selected by

452 M. Spurio / Physics Procedia 61 (2015) 450 458 Fig. 1. a) Cumulative distribution of the track reconstruction quality parameter, Λ, for tracks reconstructed as upgoing (cos θ < 0.1). The bottom panel shows the ratio between data and simulation. The green (red) distribution corresponds to the simulated atmospheric muons (neutrinos), where a 50% [13] (30% [14]) relative error was assigned. Data errors correspond to statistical errors only. b) 90% C.L. flux upper limits and sensitivities on the νμ flux for six years of ANTARES data. IceCube results are also shown for comparison. The light-blue markers show the upper limit for any point source located in the ANTARES visible sky in declination bands of 1. The solid blue (red) line indicates the ANTARES (IceCube) sensitivity for a point-source with an E 2 spectrum as a function of the declination. The blue (red) squares represent the upper limits for the ANTARES (IceCube) candidate sources. The dashed dark blue (red) line indicates the ANTARES (IceCube) sensitivity for a point-source and for neutrino energies lower than 100 TeV. From [15]. the requirement of upgoing direction, and a value of Λ sufficiently good to reject additional atmospheric muons which are poorly reconstructed as upgoing. Fig. 1a) shows the cumulative distribution of the track reconstruction quality parameter for data (points), atmospheric neutrinos (the red band) and atmospheric muons (the green band) wrongly-reconstructed as upgoing. The atmospheric neutrinos are simulated using the Bartol flux [16], and the atmospheric muons using the MUPAGE generator [17]. The application of a cut on selected events of Λ > 5.2 keeps a large fraction of really upgoing events, retaining only 10% of misreconstructed downgoing muons. The latest search for point-like sources uses a period corresponding to a total livetime of 1338 days [15]. After the selection cuts, the final data sample consists of 5516 events. The estimated median neutrino angular resolution is 0.38 for a neutrino spectrum E 2. Different blind analysis have been performed, both over the full sky, or in the direction of a-priori selected candidate source locations. For the latter, the locations of known γ-ray emitters, in particular in the Galaxy, are well-suited to look for steady, point-like sources. Fig. 1b) summarizes the results of these searches. The ANTARES upper limits (at 90% C.L.) on a E 2 flux from 50 selected candidate sources are indicated with blue dots. The corresponding upper limits from sources in the opposite hemisphere and seen by the IceCube [18] are also reported as red dots. The dashed dark blue (red) line indicates the ANTARES (IceCube) sensitivity for a point-source and for neutrino energies lower than 100 TeV, which shows that the IceCube sensitivity for sources in the Southern hemisphere is mostly due to events of higher energy. A second analysis concerns a time-integrated full-sky search looking for an excess of events over the atmospheric neutrino background in the declination band [ 90 ; +48 ]. The search algorithm is based on an unbinned maximum likelihood which includes the information on the point-spread function and on the

M. Spurio / Physics Procedia 61 (2015 ) 450 458 453 expected background rate as a function of the declination and of the number of hits used in the track reconstruction. Upper limits at the 90% C.L. on the muon neutrino flux from point sources located anywhere in the visible ANTARES sky are given by small blue dots in Fig. 1b). Each value corresponds to the highest upper-limit obtained in declination bands of 1. Following the recent evidence of high energy neutrinos by IceCube [19], a point source close to the Galactic Centre has been proposed to explain the accumulation of seven events in its neighborhood with a flux normalization corresponding to 6 10 8 GeVcm 2 s 1 [20]. Due to the large uncertainty of the direction estimates of these IceCube events, this hypothetical source might be located at a different point in the sky. The ANTARES data set was analyzed with the full sky algorithm restricted to a region of 20 around the proposed location near the Galactic centre. The trial factor related to this analysis is smaller than in the full sky search because of the smaller size of the region, providing a better sensitivity. No significant cluster has been found and the presence of a point source with a flux normalization of 6 10 8 GeVcm 2 s 1 anywhere in the region is excluded. Therefore, the excess found by IceCube in this region cannot be caused by a single point source. Furthermore, a source width of 0.5 for declinations lower than 11 is also excluded [15]. For an E 2 spectrum, neutrinos with E > 2 PeV contribute only 7% to the event rate, hence these results are hardly affected by a cutoff at energies on the order of PeV. 4. Searches for diffuse fluxes To date, no cosmic high-energy neutrino sources have been identified. This motivates the complementary approach of a search for a diffuse flux of astrophysical neutrinos. A cumulative flux is composed of the integrated flux of all neutrino sources and could be detected even if the individual sources are too faints. This applies in particular to extragalactic sources, as Active Galactic Nuclei, which are among the candidate sources of ultra high-energy cosmic rays and could produce a detectable high energy neutrino signal. In ANTARES, the measurement in the highest energy region of the atmospheric neutrino spectrum (see Sect. 6.2) constraints the diffuse flux of cosmic neutrinos. The details of this analysis, updating the published results [21], are reported in a dedicated contribution [22]. 4.1. Neutrino from the direction of galactic Fermi Bubbles Analysis of data collected by the Fermi-LAT [23] experiment has revealed two large circular structures near the Galactic Centre, above and below the galactic plane, the so-called Fermi bubbles (FBs). A possible explanation of the observed γ-rays (which have an almost uniform E 2 spectrum, with a normalization uncertainty of a factor of 2) is the acceleration process of cosmic rays [24]. In the interaction with the interstellar medium they produce neutral and charged pions, whose decay yields γ-rays and neutrinos, respectively. A dedicated search for an excess of neutrinos from the FB direction using 806 livedays has been performed by comparing the rate of events observed in the region of the FBs (ON zone) to that observed in equivalent areas of the Galaxy (OFF zones). No statistically significant excess of events above the chosen energy cut was observed in the ON zone with respect to the average expected number from the OFF zones. Corresponding upper limits on the neutrino flux from the FBs have therefore been derived for different assumptions on the energy cutoff at the source, as shown in Fig. 2a). Some of these limits are relatively close to the expected fluxes, suggesting that at least part of the phase space for hadronic models of γ-ray emission in the FBs could be probed with the full ANTARES dataset, or after about one year of operation of the next-generation KM3NeT neutrino telescope [26]. 5. Multi-messenger neutrino searches ANTARES has developed specific strategies to look for neutrinos with timing and/or directional correlations with known (or potential) sources of other cosmic messengers. The restricted space-time search window allows a sensible reduction of the background from atmospheric neutrinos, therefore enhancing the sensitivity to faint signals.

454 M. Spurio / Physics Procedia 61 ( 2015 ) 450 458 Fig. 2. a) Upper limits on the neutrino flux from the Fermi bubble regions. Four different neutrino spectra have been assumed: E 2 with no cutoff (black solid), 500 TeV (red dashed), 100 TeV (green dot-dashed), 50 TeV (blue dotted) together with the theoretical predictions for the case of a purely hadronic model (the same colours, areas filled with dots, inclined lines, vertical lines and horizontal lines respectively). The limits are drawn for the energy range where 90% of the signal is expected. From [25]. b) Sum of the 296 individual γ-ray burst muon neutrino spectra (red and blue solid lines) and limits set by ANTARES on the total flux (red and blue dashed lines) assuming two different theoretical models [27]. The IceCube IC 40+IC 59 limit [30] on the neutrino emission from 300 GRBs is also shown in black. This method has been used to put upper limits on the neutrino flux from known γ-ray transients, using a set of 296 long GRBs [27]. The expected neutrino fluxes have been calculated for each burst individually using a fully numerical neutrino emission model (NeuCosmA, [28]) which includes Monte Carlo simulations of the full underlying photo-hadronic interaction processes in the GRB. No coincident neutrino event is found within 10 of any of the GRBs in the sample, and 90% C.L. upper limits are placed on the total expected flux according to the NeuCosmA model as well as to the commonly used model [29], as shown in Fig. 2a). Although not competitive with the IceCube limits, the ANTARES results concern a different sample of GRBs (only 10% of them being also in IceCube field of view), thus offering complementary sky coverage and energy range. Specific searches have been performed during periods of intense activity of some sources, as reported by X-ray and/or γ-ray observatories. In particular, a search for neutrinos in coincidence with 10 flaring blazars chosen on basis of high state periods of the light curve reported by Fermi-LAT was performed using data from 2008. A similar analysis is ongoing on the 2008-2012 period using data from Fermi and from the Cherenkov Telescopes HESS, MAGIC and VERITAS. A third search for neutrinos in coincidence with X-ray or γ-ray outbursts of 6 microquasars was performed using 2007-2010 data [31]. No coincident neutrino events were detected in all studies for flaring objects, and in some cases the inferred limits are close to predictions which may be tested by ANTARES, in particular for the two microquasars GX339-4 and CygX-3. Coincident searches of neutrinos and gravitational waves are yet another example of a multi-messenger strategy that could help reveal hidden sources, so far undetected with conventional photon astronomy. A dedicated paper on the searches for possible coincidences among ANTARES and the last-generation GW interferometers VIRGO and LIGO using 2007 data was published [32]. Analysis is ongoing for a larger data set. Finally, the occurrence of a special event in a neutrino telescope (such as the near-simultaneous arrival of two or more neutrinos from the same direction) could indicate that a highly energetic burst has occurred and may be used as a trigger for optical, X-ray, and γ-ray follow-ups. Since 2009, ANTARES alerts (about 25/year) have been sent on a regular basis to a network of fast-response, wide-field robotic optical telescopes (TAROT, ROTSE and ZADKO) and more recently also to the SWIFT/XRT telescope [33]. Due to the repositioning time of the telescope, the first optical image can typically be obtained after a few seconds.

M. Spurio / Physics Procedia 61 (2015 ) 450 458 455 Fig. 3. a) The atmospheric neutrino energy spectrum E 3.5 dφ ν /de ν measured by ANTARES (black full squares) [37]. The full line represents the Bartol group ν μ flux. The red and blue dashed lines include two prompt neutrino production models. Data and predictions are averaged over the zenith angle region θ>90. The results of the AMANDA-II unfolding [38] averaged in the region θ>100 (red circles) and that of IceCube40 [39] zenith-averaged above 97 (blue triangles) are shown. b) 90% C.L. contour of the neutrino oscillation parameters as derived from the fit of the E R / cos θ distribution. Results from MINOS and Super-Kamiokande are shown for comparison, as well as recent results from the IceCube [36]. Successive sets of images can then be recorded minutes, hours and/or days after the alert, in order to follow the time profile of different transient sources (such as GRB afterglows or core-collapse supernovae) on appropriate time scales. No optical counterpart has been observed so far in association with one of the ANTARES neutrino alerts, and limits on the magnitude of a possible GRB afterglow are in preparation. 6. Particle physics studies 6.1. Neutrino oscillation Atmospheric muons and atmospheric neutrinos are produced by cosmic ray interactions with atmospheric nuclei and represent the irreducible background for cosmic neutrinos. Up to 100 TeV, muons and neutrinos are produced mainly by decays of charged pions and kaons in the cascade and their spectra are related by the kinematics of the π μν and K μν decays, yielding an energy spectrum Φ ν E 3.7 (conventional atmospheric neutrinos). Rarely, also charmed mesons are produced; their immediate decay yields a harder neutrino energy spectrum (prompt neutrino flux), which should exceed that of the conventional neutrinos above 100 TeV. ANTARES can detect neutrino events with energy as low as few tens of GeV. In this energy range, the ν μ events suffer for the effect of oscillations [34, 35] when arriving from the nadir. Low energy events have been used for neutrino oscillation studies, providing the first determination of oscillation parameters by neutrino telescope using an overall livetime of 863 days. The region of allowed parameters is shown in Fig. 3b), which includes also the more recent IceCube results. 6.2. Measurement of the atmospheric ν μ spectrum The measurement of the atmospheric ν μ spectrum up to the highest accessible energies has been performed using a data sample of 855 days live time. The cuts used to select a pure sample of atmospheric neutrinos were optimized using a run-by-run Monte Carlo simulation which reproduces the detail of the acquisition conditions of the detector.

456 M. Spurio / Physics Procedia 61 ( 2015 ) 450 458 The measurement of ν μ energy is challenging, because only the induced muon is observed, which loses energy in a stochastic way, and the muon travel in the detector only for a limited fraction of its range. For this reason, event-by-event energy reconstruction is not sufficient to derive the energy spectrum of the detected atmospheric neutrinos. The methods used to reconstruct the muon energy are thus based on the amount of detected light on the OMs. The muon estimated energy was determined for each event; the parent neutrino distribution was derived with unfolding procedures, following the methods described in [37]. The obtained unfolded neutrino energy spectrum, averaged over zenith angles θ>90, is shown in Fig. 3a) as a function of the energy in the range 100 GeV-200 TeV. 6.3. Dark Matter searches ANTARES can exploit the very precise determination of the ν μ candidate direction to indirect searches of dark matter (DM) in the popularly advocated form of WIMPs (Weakly Interacting Massive Particles). Such particles could get gravitationally trapped in massive bodies such as the Earth, the Sun or the Galactic Center, where they can subsequently self-annihilate into different channels (as χχ ll ; qq ; W + W ; Z 0 Z 0 ; γγ; hh, where l represents a charged lepton, q a quark, W ±, Z 0 the gauge bosons, γ the photon and h the Higgs scalar boson). If the WIMP is not the neutralino or another Majorana particle, neutrinos could also be directly produced by χχ ν f ν f annihilations. As only neutrinos escape from dense regions, the subsequent decay of some final-state particles could originate a neutrino flux. An excess of neutrinos in the 10GeV-1TeVenergy range and in the direction of one such body could therefore provide an indirect signature of the existence of dark matter. A search in the direction of the Sun has been performed using about 300 livedays [40]. Conventionally, two reference channels are used: the hard channel assumes that all WIMPs annihilations produce a τ + τ (or W + W ) pair. The so-called soft channel assumes on the other hand 100% production of bb pairs. The former assumption induces a harder neutrino energy spectrum. For each WIMP mass and annihilation channel considered, the quality cuts were designed as to minimize the average 90% C.L. upper limit on the WIMP-induced neutrino flux. No excess above the atmospheric background was found, yielding upper limits on the spin-dependent and spin-independent WIMP-proton cross-section. An improved analysis has been recently unblinded, using 1321 effective livedays. Improvements arise at low energy from the addition of events reconstructed with only one line. Also the new analysis shows no significant number of events in excess to the background. Indirect constrains are particularly competitive with respect to direct experiment results for the spin-dependent WIMP-nucleus scattering. Fig. 4 shows the 90% C.L. upper limits in terms of spin-dependent (SD) WIMP-proton cross-sections derived from the analysis and compared to predictions of the MSSM-7 [41] model, a simplified version of the Minimal SuperSymmetric Model containing a neutralino as lightest stable particle. 6.4. Searches for exotics Magnetic monopoles (MM) are predicted by various gauge theories to have been produced in the early universe. Theory does not provide definite predictions on the magnetic monopole abundance. However, by requiring that MMs do not short-circuit the galactic magnetic field faster than the dynamo mechanism can regenerate it, a flux upper limit can be obtained. This is the so-called Parker bound (Φ MM 10 15 cm 2 s 1 sr 1 ), whose value sets the scale of the detector exposure for MMs search. The most stringent limit in the widest interval of velocities arises from the MACRO experiment [43]. The signature for relativistic MMs in a neutrino telescope is quite visible, as the intensity of the light they emit in the detector is O(10 4 ) times greater than the Cherenkov light. The analysis of data up to 2008 yields one observed event, compatible with the expected background. ANTARES limit on relativistic MMs [42], shown in Fig. 4b), is at present the most restrictive. 7. Conclusions The ANTARES neutrino telescope is the first and only deep sea neutrino telescope currently in operation. Since the connection of its first detection line in 2006, it has been continuously monitoring the declination band [ 90 ; +48 ], which includes the galactic center, with unprecedented sensitivity. ANTARES

M. Spurio / Physics Procedia 61 (2015 ) 450 458 457 Fig. 4. a) ANTARES 1321 days upper limits (90% C.L.) on the spin-dependent WIMP-proton cross-sections as a function of the WIMP mass, for the three self-annihilation channels: bb (green), W + W (blue) and τ + τ (red). Expected values of a grid scan of the MSSM-7 [41] are included (light grey shaded areas respectively). These limits improve those published in [40], that contains the references to other experiments whose limits are drown in the figure. b) 90% C.L. upper limit on flux of relativistic upgoing magnetic monopole. The reference Parker bound and that of other experiments is also shown [42]. demonstrated to accurately measure atmospheric ν μ from energy regions around 20 GeV, where neutrino oscillations reduce the number of upgoing events, up to 200 TeV. Besides atmospheric neutrino studies, several searches for neutrinos of astrophysical origin have been performed. An extended multi-messenger program complements and expands its astrophysics reach. Neutrino telescopes cover also different particle physics topics, as indirect searches for dark matter, searches for exotic or fossile particles. ANTARES paves the way to a multi-km3 neutrino telescope in the Northern Hemisphere, KM3NeT, to complement the Ice- Cube detector deployed at the South Pole. More competitive results are expected in the future as ANTARES will continue taking data and improving the results at least until the end of 2016, when it gets eventually superseded by the next-generation detector in the Mediterranean Sea. References [1] L. A. Anchordoqui et al., Journal of High Energy Astrophysics 1-2 (2014) [2] C. Spiering, arxiv:1402.2096; U.F. Katz, C. Spiering. Prog.Part. Nucl. Phys. 67 (2012) 651. [3] T. Chiarusi, M. Spurio Eur. Phys.J. C65 (2010) 649. [4] M. Ageron et al., Nucl. Instrum. Meth. A 656 (2011) 11. [5] P. Amram et al., Nucl. Instrum. Meth. A 484 (2002) 369. [6] J.A. Aguilar et al., Nucl. Instrum. Meth. A 555 (2005) 132. [7] J.A. Aguilar et al., Nucl. Instrum. Meth. A 622 (2010) 59. [8] J.A. Aguilar et al., Nucl. Instrum. Meth. A 570 (2007) 107. [9] P. Amram et al., Astropart. Phys. 19 (2003) 253. [10] J.A. Aguilar et al., Nucl. Instrum. Meth. A 626-627 (2011) 128. [11] J.A. Aguilar et al., Astropart. Phys. 26 (2006) 314. [12] S. Adrián-Martínez et al., JINST 7 (2012) T08002. [13] J. A. Aguilar et al., Astropart. Phys. 34 (2010) 179. [14] G. D. Barr, T. K. Gaisser, S. Robbins and T. Stanev, Phys. Rev. D 74 (2006) 094009. [15] S. Adrián-Martínez et al. arxiv:1402.6182. [16] G. D. Barr, T. K. Gaisser, P. Lipari, S. Robbins and T. Stanev, Phys. Rev. D 70 (2004) 023006. [17] G. Carminati, M. Bazzotti, A. Margiotta and M. Spurio, Comput. Phys. Commun. 179 (2008) 915. [18] M. G. Aartsen et al. [IceCube Collaboration], Astrophys. J. 779 (2013) 132. [19] M. G. Aartsen et al. [IceCube Collaboration], Phys. Rev. Lett. 111 (2013) 021103. [20] M.C. Gonzalez-Garcia et al., arxiv:1310.7194. [21] J.A. Aguilar et al., Phys. Lett B 696 (2011) 16. [22] J. Schnabel, contribution to this conference. [23] M. Su, T. Slatyer, D. Finkbeiner, Astrophys. J. 724 (2010) 1044. [24] R. M. Crocker and F. Aharonian, Phys. Rev. Lett. 106 (2011) 101102. [25] S. Adrián-Martínez et al., Eur. Phys. J. C 74 (2014) 2701.

458 M. Spurio / Physics Procedia 61 ( 2015 ) 450 458 [26] S. Adrián-Martínez et al. [KM3NeT Collaboration], Astropart. Phys. 42 (2013) 7. [27] S. Adrián-Martínez et al. A&A 559 (2013) A9. [28] S. Hummer, P. Baerwald and W. Winter, Phys. Rev. Lett. 108 (2012) 231101 [arxiv:1112.1076]. [29] D. Guetta, D. Hooper, J. Alvarez-Muniz, F. Halzen and E. Reuveni, Astropart. Phys. 20 (2004) 429. [30] R. Abbasi et. al (IceCube coll.) Nature, 484 (2012) 351. [31] S. Adrián-Martínez et al. arxiv:1402.1600 [32] S. Adrián-Martínez et al. [ANTARES, LIGO Scientific and Virgo Collaborations], JCAP 1306 (2013) 008. [33] M. Ageron et al., Astropart. Phys. 35 (2012) 530. [34] Y. Fukuda et al. [Super-Kamiokande Collaboration], Phys. Rev. Lett. 81 (1998) 1562. [35] M. Ambrosio et al. [MACRO Collaboration], Physics Letters B 434 (1998) 451. [36] Ch. Wiebusch for the IceCube Collaboration, 33rd ICRC, id 848. [37] S. Adrián-Martínez et al., Eur. Phys. J. C 73 (2013) 2606. [38] R. Abbasi et al., Astropart. Phys. 34 (2010) 48. [39] R. Abbasi et al., Phys. Rev. D 83 (2011) 012001. [40] S. Adrián-Martínez et al., JCAP11 (2013) 032. [41] L. Bergstrom, P. Gondolo, Astropart. Phys. 5 (1996) 263. [42] S. Adrián-Martínez et al., Astropart. Phys. 35 (2012) 634640. [43] M. Ambrosio et al. [MACRO Collaboration] Eur. Phys. J. C 25 (2002) 511.