Next Article in Journal
Induced Isotensor Interactions in Heavy-Ion Double-Charge-Exchange Reactions and the Role of Initial and Final State Interactions
Next Article in Special Issue
The ASTRI Mini-Array: A New Pathfinder for Imaging Cherenkov Telescope Arrays
Previous Article in Journal
On the Possibility of a Static Universe
Previous Article in Special Issue
Highlights of the Magic Florian Goebel Telescopes in the Study of Active Galactic Nuclei
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Science with the ASTRI Mini-Array: From Experiment to Open Observatory

by
Stefano Vercellone
on behalf of the ASTRI Project
INAF Osservatorio Astronomico di Brera, Via E. Bianchi 46, 23807 Merate, LC, Italy
http://www.astri.inaf.it/en/library/ (accessed on 21 January 2024).
Universe 2024, 10(2), 94; https://doi.org/10.3390/universe10020094
Submission received: 22 January 2024 / Revised: 8 February 2024 / Accepted: 13 February 2024 / Published: 16 February 2024
(This article belongs to the Special Issue Recent Advances in Gamma Ray Astrophysics and Future Perspectives)

Abstract

:
Although celestial sources emitting in the few tens of GeV up to a few TeV are being investigated by imaging atmospheric Čerenkov telescope arrays such as H.E.S.S., MAGIC, and VERITAS, at higher energies, up to PeV, more suitable instrumentation is required to detect ultra-high-energy photons, such as extensive air shower arrays, as HAWC, LHAASO, Tibet AS- γ . The Italian National Institute for Astrophysics has recently become the leader of an international project, the ASTRI Mini-Array, with the aim of installing and operating an array of nine dual-mirror Čerenkov telescopes at the Observatorio del Teide in Spain starting in 2025. The ASTRI Mini-Array is expected to span a wide range of energies (1–200 TeV), with a large field of view (about 10 degrees) and an angular and energy resolution of ∼3 arcmin and ∼10 %, respectively. The first four years of operations will be dedicated to the exploitation of Core Science, with a small and selected number of pointings with the goal of addressing some of the fundamental questions on the origin of cosmic rays, cosmology, and fundamental physics, the time-domain astrophysics and non γ -ray studies (e.g., stellar intensity interferometry and direct measurements of cosmic rays). Subsequently, four more years will be dedicated to Observatory Science, open to the scientific community through the submission of observational proposals selected on a competitive basis. In this paper, I will review the Core Science topics and provide examples of possible Observatory Science cases, taking into account the synergies with current and upcoming observational facilities.

1. Introduction

About 300 celestial sources are currently known to emit in the 0.1 < E < 30  TeV energy range (see the TeVCat Webpage1 [1]) based on their detection by the major imaging atmospheric Čerenkov telescope arrays (IACTs), such as H.E.S.S. [2], MAGIC [3], and VERITAS [4], whose energy range extends up to a few tens of TeV. Alternatively, extended air shower arrays (EAS) such as HAWC [5], LHAASO [6] and Tibet AS- γ  [7], adopt a different detection technique that allows us to investigate energies up to several hundreds of TeV, reaching the PeV limit. The sources detected by the current generation of EAS at energies E > 100  TeV, and up to a few PeVs are a factor of ten fewer.
The Čerenkov Telescope Array Observatory (CTAO [8]) will be the next large scale Čerenkov array and will cover an energy range from a few tens of GeV up to a few hundreds of TeV by means of telescopes of different sizes (see, e.g., [9]). It will be deployed in both hemispheres to observe the full sky. A few telescope prototypes were developed in recent years, among them the ASTRI-Horn dual-mirror, Schwarzschild–Couder (SC) telescope [10], currently operating on Mount Etna in Sicily (Italy), which obtained the first-light optical qualification by means of observation of Polaris, using a dedicated optical camera [11], and the first detection of very high-energy γ -ray emission from the Crab Nebula by a Čerenkov telescope in dual-mirror SC configuration [12].
In this review, I will first describe the ASTRI Mini-Array characteristics and performance in the context of currently available very high- and ultra-high energy (VHE and UHE, respectively) instrumentation (Section 2), then briefly highlight the ASTRI Mini-Array science topics that will be pursued during the first four years of operation (Section 3). In Section 4, I will describe in detail the subsequent four years of operation and the Open Observatory Phase, when the scientific investigation is mainly driven by the community.

2. The ASTRI Mini-Array

The ASTRI Mini-Array [13,14] consists of nine ASTRI dual-mirror small-sized (SSTs) Čerenkov telescopes, currently being deployed at the Observatorio del Teide (Spain), which will commence its scientific operations in late 2025. The ASTRI Mini-Array will provide a large field of view (FoV) of about 10°, a wide energy range from 1 TeV to 200 TeV, an angular resolution of ∼3 , and an energy resolution of ∼10%. Table 1 compares the ASTRI Mini-Array performance with that of the current IACTs.
A detailed description of the ASTRI Mini-Array performance is reported in [19]. Figure 1 shows the ASTRI Mini-Array differential sensitivity (turquoise points, 50 h integration time, 5 σ confidence level, C.L.) compared with those of the current major IACTs (H.E.S.S., MAGIC, and VERITAS) and of the planned CTAO. The ASTRI Mini-Array will improve the current IACT sensitivity at energies greater than a few TeVs and will be of the same order as that of CTAO North in the “alpha” (4 LSTs2 + 9 MSTs3 [9] configuration, and slightly better in the “science verification” (4 LSTs + 5 MSTs [20]) configuration, respectively, at energies greater than a few tens of TeVs.
The ASTRI Mini-Array will reserve, as we shall describe in Section 3, its first four years of operation for the investigation of a few specific science topics. This implies that most of the science operations will be performed as deep pointings, with exposures in the order of 200 h or even 500 h, towards specific sky regions. Figure 2 shows the ASTRI Mini-Array differential sensitivity curves for 200 h (turquoise squares) and 500 h (turquoise triangles) integration time, respectively. For such long integration times, the most appropriate comparison is with current EAS differential sensitivity curves, HAWC [21], Tibet AS- γ  (Takita M., priv. comm. based on [22]), and LHAASO [23].
Figure 1. ASTRI Mini-Array differential sensitivity for 50 h integration compared with those of MAGIC, H.E.S.S., VERITAS, and CTAO North. The differential sensitivity curves are drawn from [19] (ASTRI Mini-Array), [16] (MAGIC), the VERITAS official website https://veritas.sao.arizona.edu (accessed on 14 February 2024), and [24] (sensitivity curve for H.E.S.S.−I, stereo reconstruction). CTAO−N “alpha configuration” (alpha) sensitivity curve comes from [9]. The CTAO−N “science verification” (sv) sensitivity is drawn from [20].
Figure 1. ASTRI Mini-Array differential sensitivity for 50 h integration compared with those of MAGIC, H.E.S.S., VERITAS, and CTAO North. The differential sensitivity curves are drawn from [19] (ASTRI Mini-Array), [16] (MAGIC), the VERITAS official website https://veritas.sao.arizona.edu (accessed on 14 February 2024), and [24] (sensitivity curve for H.E.S.S.−I, stereo reconstruction). CTAO−N “alpha configuration” (alpha) sensitivity curve comes from [9]. The CTAO−N “science verification” (sv) sensitivity is drawn from [20].
Universe 10 00094 g001
Figure 2. ASTRI Mini-Array differential sensitivity for 200 h (turquoise squares) and 500 h (turquoise triangles) integration times compared with those of HAWC (507 d), Tibet AS- γ (1 yr), and LHAASO (1 yr). The differential sensitivity curves are drawn from ASTRI Mini-Array [19], HAWC [21], LHAASO [23], and Takita M. (priv. comm.) based on [22] (Tibet AS + MD). We note that the 507-day HAWC differential sensitivity curve corresponds to about 3000 h of acquisition on a source at a declination of 22° within its field of view [21].
Figure 2. ASTRI Mini-Array differential sensitivity for 200 h (turquoise squares) and 500 h (turquoise triangles) integration times compared with those of HAWC (507 d), Tibet AS- γ (1 yr), and LHAASO (1 yr). The differential sensitivity curves are drawn from ASTRI Mini-Array [19], HAWC [21], LHAASO [23], and Takita M. (priv. comm.) based on [22] (Tibet AS + MD). We note that the 507-day HAWC differential sensitivity curve corresponds to about 3000 h of acquisition on a source at a declination of 22° within its field of view [21].
Universe 10 00094 g002
The main advantage of EAS with respect to IACTs is the former’s 2 sr FoV and their larger duty cycle. On the other hand, as reported in Table 2, their energy and angular resolution in the same energy range as the ASTRI Mini-Array (about 10 TeV) are at least a factor of 3 to 4 times worse. Clearly, this makes the ASTRI Mini-Array extremely competitive in studying the morphology of extended sources and crowded fields and accurately monitoring multiple targets in the same pointing.
Furthermore, the ASTRI Mini-Array angular resolution will allow us to investigate the LHAASO uncertainty error box of Galactic sources, which is of the order of one degree [28] and study the different sources possibly associated with the PeV emission, in order to unambiguously identify them, when in synergy with GeV and X-ray facilities.
Figure 3 shows the ASTRI Mini-Array angular (left panel) and energy (right panel) resolution as a function of the energy as reported in [19].

3. Core Science Topics

The ASTRI Mini-Array science program will develop in two phases. During the first four years of operations, the ASTRI Mini-Array will be run as an experiment, while in the subsequent four years, it will gradually evolve into an observatory open to the scientific community.
A graphical description of the main Core Science topics that we plan to investigate during the first four years of operations is shown in Figure 4. Our Core Science Program is based on “Main Pillars”. They are science fields in which the ASTRI Mini-Array will contribute breakthrough pieces of evidence to improve our understanding of a few key science questions.
Recently, [15] discussed the ASTRI Mini-Array Core Science, which includes the study of the: (1) origin of cosmic rays; (2) cosmology and fundamental physics; (3) GRBs and time-domain astrophysics; (4) direct measurements of cosmic rays; (5) stellar-intensity interferometry. Here, I will review some of the results on all science topics, while Section 4.1 will focus on the Observatory Science ones, presented in [29,30].

3.1. The Origin of Cosmic Rays

The LHAASO Collaboration [28] reported the discovery of twelve Galactic sources emitting γ -rays at several hundreds TeV up to 1.4 PeV. These sources are able to accelerate particles up to ∼ 10 15  eV, making them “PeVatron candidates”. We note that the majority of these sources are diffuse γ -ray structures with angular extensions up to 1°, which, together with the LHAASO limited angular resolution, make the identification of the actual sources responsible for the ultra high-energy γ -ray emission not univocal (except for the Crab Nebula). The recent publication of the First LHAASO Catalog of γ -ray Sources (1LHHASO [31]) containing 90 sources, 43 of them with emissions at energies E > 0.1  PeV, marks a fundamental step for the astrophysics at very high- and ultra high-energies. This discovery is extremely important for the ASTRI Mini-Array science, especially because of its angular resolution, which, at energies of about 100 TeV, is a factor of 3 to 4 times better in radius than the LHAASO one: 0.08° vs. 0.24–0.32°. We should also mention that both the angular resolution and the energy resolution can be improved by means of specific analysis cuts, as shown in Figure 3. The ASTRI Mini-Array will investigate these and future PeVatron sources, providing important information on their morphology above 10 TeV. The ASTRI Mini-Array wide FoV will be extremely important in the investigation of extended regions and point-like sources. A single pointing will allow us to investigate the Galactic Center or the Cygnus regions. In these regions, we can accumulate several hundreds of hours by also including the epochs of moderate Moon condition [32]. This is crucial to investigate both the (energy-dependent) morphology of the sources in these regions and their possible variability on a long time scale. The ASTRI Mini-Array will investigate the Galactic Center at a high Zenith angle (maximum culmination angle of ∼ 57 ). We expect to be able to study this region up to E∼200 TeV for an exposure time of 260 h, significantly improving the current results of other IACTs (see for further details [15]). The high-energy boundaries of the ASTRI Mini-Array will also be important to study the Crab Nebula, the only Galactic PeVatron4 currently known [33,34]. The origin of the Crab Nebula γ -ray emission detected by LHAASO does not require a hadronic contribution but cannot exclude it either. A deep ASTRI Mini-Array observation lasting about 500 h in the E > 100  TeV energy range should definitely be able to provide constraints on the proton component in this source.

3.2. Cosmology and Fundamental Physics

IACT arrays detected extra-galactic sources since the early nineties [35]. Since then, among the 280 sources listed in TeVCat, 93 are found to be extra-galactic: 55 high-peaked BL Lacs (HBLs), 10 intermediate-peaked BL Lacs (IBLs), 9 flat-spectrum radio quasars (FSRQs), 4 Blazars, 4 Fanaroff-Riley Type (FR-I) galaxies, 2 star-bursting galaxies (SBGs), 2 BL Lacertae objects with class unclear (BL Lacs), 2 unknown type AGNs, and 5 γ -ray bursts (GRBs). Extra-galactic jetted sources are excellent probes for several science cases. They can be used to investigate the extra-galactic background light, as well as to probe, by means of variability studies, the properties of the γ -ray emitting region. They can also be useful to investigate peculiar physical phenomena, such as the existence of the axion-like particle, to test the Lorentz invariance violation, and to study the intergalactic magnetic fields.
The Extra-galactic background light (EBL)—The EBL significantly affects the spectra of jetted sources at energies that can be explored by the ASTRI Mini-Array. Moreover, the EBL direct measurement in the infra-red (IR) portion of the spectrum is particularly challenging because of the dominant contribution of our Galaxy at these wavelengths. Nevertheless, the ASTRI Mini-Array can contribute to the study of the EBL IR component given the well-known relation λ max 1.24 × E TeV   μ m, between the wavelength of the target EBL photon, λ max   and the energy of the γ -ray, E TeV . The IR component in the ( 10 < λ < 100 μ m regime represents a challenge because of the dominance of local emission from both the Galaxy and our Solar system. The preferred candidates for observations with the ASTRI Mini-Array are TeV-emitting low-redshift radio-galaxies and local star-bursting galaxies. Among the sources fulfilling these criteria, we investigated the low-redshift radio-galaxies IC 310 (z∼0.0189) and M 87 (z∼0.00428). For IC 310 we assumed three different spectral states: flare [36], high and low [37]. Short (5 h) and deep (200 h) observations will allow us to detect these sources in different spectral states at energies E > 10  TeV, thus probing the IR EBL component. Similar results can be obtained for M 87 in different spectral states, as reported by [38] (low state), [39] (high state), and [40] (flaring state).
Fundamental and exotic physics—Blazar spectra above a few TeV are excellent probes of non-standard γ -ray propagation effects such as the presence of hadron beams (HB) in the jet of extreme BL Lac objects (E-HBL and references therein [41]), the existence of axion-like particles [42] (ALP), or the effects of the Lorentz invariance violation (LIV and references therein [41]) and the properties of inter-galactic magnetic fields (IGMF and references therein [43]). The most promising sources to detect spectral signatures induced by these effects are 1ES 0229+200 (E-HBL, z∼0.139) and Mrk 501 (HBL, z∼0.03298). The presence of HB implies that the spectrum of blazars extends at energies above those allowed by the standard EBL model [44]. A detection of a γ -ray photons at energies of a few tens of TeV, when the standard EBL model would imply a roll-off at a few TeV, could be the signature for the presence of the HB scenario. On the other hand, this excess at energies above a few tens of TeV could also be induced by both the ALPs and LIV effects. ALPs produce a distinctive oscillation pattern in the blazar spectrum, that could represent the unique marker for this process, but it would require a much finer energy resolution than that of the ASTRI Mini-Array to be revealed. Also, in the context of the widely studied dark matter (DM) weakly-interacting massive particles scenario, the ASTRI Mini-Array may provide interesting DM-related insights from dwarf spheroidal galaxies and Galactic center observations, particularly for the case of monochromatic γ -ray emission lines [30]. In order to address all these studies we can plan deep (in the order of 200 h) dedicated pointing, a typical exposure that could be accumulated during the first years, with the described ASTRI Mini-Array observing strategy.

3.3. Multi-Messenger and Time-Domain Astrophysics

Transients and multi-messenger studies such as γ -ray bursts (GRBs), gravitational waves (GWs), and neutrino emission ( ν s ) from VHE sources are indeed the new frontiers of high-energy astrophysics.
γ -ray bursts—GRBs have only been detected by IACTs starting from 2018 and, at the time of writing, we only count six5 GRBs detected at energies in excess of 0.1 TeV: GRB 160821B ( z = 0.162 , MAGIC] [45]), GRB 180720B ( z = 0.653 , H.E.S.S. [46]), GRB 190114C ( z = 0.424 , MAGIC [47]), GRB 190829A ( z = 0.078 , H.E.S.S. [48]), GRB 201015A ( z = 0.42 , MAGIC [49]), GRB 201216C ( z = 1.1 , MAGIC [50]), GRB 221009A ( z = 0.151 , LHAASO] [51]). Two of them are particularly relevant for the ASTRI Mini-Array. GRB 190114C is the first GRB detected at VHE within one minute from the T 0 , up to an energy of about 1 TeV. GRB 221009A, described as “to be a once-in-10,000-year event” [52,53] was detected by LHAASO up to 13 TeV [54], challenging the standard emission scenario for the canonical EBL absorption [55,56]. Taking into account the energetics observed in GRB 221009A, we can estimate that this event, placed at a different redshift ( z = 0.078 , 0.25 , 0.42 ), can be detected up and above 10 TeV by the ASTRI Mini-Array within a few minutes from the event (L. Nava, Priv. Comm.).
Neutrinos—AGNs can be sources of extra-galactic ν s . The IceCube data [57] seem to indicate that there could be an association between ν s emission and a few AGNs: a Seyfert-2 galaxy (NGC 1068, D = 14.4  Mpc), and two BL Lac objects (TXS 0506 + 056, z = 0.3365 ; PKS 1424 + 240, z = 0.16 ). Although the two latter sources are known TeV emitters, NGC 1068 shows prominent emission only in the 0.1–300 GeV energy band [58,59]. NGC 1068 could show emission above 10 TeV under the assumption of a dominant contribution of relativistic particles accelerated by the AGN-driven wind, as discussed in Section 4.4.

3.4. Non γ -ray Astrophysics

Stellar intensity interferometry—Stellar intensity interferometry (SII) is based on the second-order coherence of light, which allows imaging sources at the level of 100  μ as. This means that it is possible to reveal details on the surface and of the environment surrounding bright stars in the sky, which typically have angular diameters of 1–10 mas. The SII observing mode will take advantage of an additional, dedicated instrument that is being designed and will be installed on the ASTRI Mini-Array telescopes [60].
Direct measurements of cosmic rays—More than 99% of the signal acquired by the ASTRI Mini-Array is hadronic in nature. In particular, this hadronic component could be useful to investigate the cosmic ray composition in the TeV–PeV energy range and the measurement of the cosmic ray spectrum at energies characteristics of its “knee”.

3.5. Synergies with Other Facilities

The ASTRI Mini-Array will be operating during a period when several facilities will cover the whole electromagnetic spectrum, from radio to PeV. The Sardinia radio telescope (SRT) will complement VHE observation with radio data at different frequencies, both for Galactic and extra-galactic objects. Galactic sources already observed with SRT are W 44, IC 433, and Tycho [61,62]. In the optical energy band, the Telescopio Nazionale Galileo [63] and the GASP/WEBT Consortium [64] can provide excellent coverage, as well as several facilities managed by Instituto de Astrofísica de Canarias (IAC) at the Canary Island. In the X-ray energy band, in addition to the long-standing ESA and NASA legacy Observatories, we can now exploit the eROSITA [65] surveys and, in particular, the IXPE [66] X-ray polarimetric data. At the extreme energy boundary, LHAASO, HAWC, and Tibet AS- γ will extend data at energies of a few PeV.
The ASTRI Mini-Array location at the Observatorio del Teide and its collaboration with the IAC will allow us to investigate sources synergically with both the MAGIC and CTAO−N arrays. In particular, they will be of paramount importance for their capability to investigate not only the local Universe but also to reach redshifts well beyond one and perform cosmological studies on extra-galactic sources. Moreover, both MAGIC and CTAO−N will allow us to extend the ASTRI Mini-Array spectral performance in the sub-TeV regime, with almost no breaks from a few tens of GeV up to hundreds of TeV.

4. The Observatory Phase

The ASTRI Mini-Array science program will gradually evolve from an experiment towards an Observatory Phase, built on the experience and results from the Core Science phase, and open to observational proposals from the scientific community at large. We foresee important synergies with the above-mentioned facilities to yield the best scientific return from the proposed observations.
An example of such synergies is illustrated in Figure 5, where we show, in Galactic coordinates and Aitoff projection, the First LHAASO Catalog of γ -ray Sources (1LHHASO, [31]) plotted as orange dots and, superimposed, the ASTRI Mini-Array Core Science target regions for both Pillar-1 (blue circles) and Pillar-2 (red circles) targets. Some 1LHAASO sources already overlap the ASTRI Mini-Array selected Pillar regions, in particular along the Galactic Plane. We also note that Mrk 501 and Mrk 421 are natural candidates for common variability studies.
We can now discuss a few examples of foreseeable investigations that can be performed during the Observatory Phase. Since this phase will be open to the scientific community through competitive proposals, these examples represent ideas on how to best employ the ASTRI Mini-Array capabilities.

4.1. Cygnus Region Mini-Survey

The ASTRI Mini-Array wide FoV is well suited to perform mini-surveys of selected sky regions. One of the most important ones, as shown in Figure 5, is the Cygnus Region ( 60 < l < 90 ), which contains several sources emitting above a few TeV, as reported in the First LHAASO Catalog. Figure 6 shows the results of a possible ASTRI Mini-Array mini-survey of this region with different exposure times, 50 h, 100 h, and 200 h from top to bottom, respectively.
The simulations, discussed in [29], combined fifty different pointings, at the same Galactic latitude and spaced by 0.4° in Galactic longitude, from (l, b) = (64, 0) to (l, b) = (84, 0), and lasting 1 h, 2 h, and 4 h hours each, respectively. The very high-energy simulated sources were drawn from the Third HAWC Catalog of very high-energy γ -ray sources (3HWC [67]). Thirteen of them fall inside the area considered and were simulated according to their published spectral parameters. Ten of these very high-energy sources are always significantly detected by the ASTRI Mini-Array, even at the shortest (50 h) exposure time. Recently, [68] reported the detection of an extend (about 6 in diameter) γ -ray emission centered on Cygnus-X (l,b)≈(80 , 0 ) with 66 photon-like events with energies greater than 400 TeV (see Figure 1 in [68]). The ASTRI Mini-Array wide FoV (∼ 10 in diameter) and the stable off-axis performance (see Section 2) will allow us to investigate this region with a single pointing and prolonged exposure, performing a more accurate morphological measurement on the core region of the bubble discovered by LHAASO.

4.2. Gamma-Ray Binaries—LS 5039

A recent study [69] discusses the properties of γ -ray binaries. We currently know ten non-transient γ -ray binaries: seven have a compact source (six are located in our Galaxy and one, LMC P3, in the Large Magellanic Cloud), while three are colliding-wind binaries. We discuss the results of ASTRI Mini-Array simulations of LS 5039 to show the ASTRI Mini-Array capabilities of reproducing both the folded light-curve and the spectrum in different orbital phases for this source. We simulated 300 h of total exposure, 250 h in the low state, and 50 h in the high state (see [70] for a detailed description of the different flux levels). The orbit-averaged spectrum6, described in [71], is a cut-off power-law with Γ = 2.06 ± 0.05 and E cut = 13.0 ± 4.1  TeV. We simulated a fixed exposure time of 10 h for each phase bin. Figure 7 shows the simulation results, the flux (left panel), and the 1  σ uncertainty ( δ Γ ) on the spectral index (right panel) as a function of the orbital phase. The simulated source flux is fully consistent with the flux expected from the model. Moreover, for 90% of the orbital phase, the uncertainty on the photon index, δ Γ , is between 0.1 and 0.25, while only in the case of the lowest-flux bin (phase range 0.1–0.2) its value rises up to about 0.4.

4.3. Spectral Features—Mrk 501

Mrk 501 ( z = 0.032983 ± 0.00005 ) is the second extra-galactic source detected at VHE [72]. It is classified as a high-synchrotron-peaked BL Lac object, which means that the synchrotron peak of the usual double-humped blazar spectral energy distribution reaches the ultra-violet energy band or even higher frequencies ( ν peak 10 15  Hz). This source is extremely variable, at almost all frequencies. Recently MAGIC detected a peculiar spectral feature during the highest X-ray ( E > 0.3  KeV) flux ever recorded from this source [73].
The spectral feature, emerging during the highest X-ray flux state at about 3 TeV with a significance of ≈ 4 σ , can be modeled both as a curved narrow-band log-parabola or a Gaussian function superimposed to a broad-band simple log-parabola, respectively. The physical interpretation is still debated and three possible scenarios can be invoked: a two-zone emitting region model, a pile-up in the electron energy distribution, or a pair cascade from electrons accelerated in a black hole magnetospheric vacuum gap. Figure 8 shows the ASTRI Mini-Array simulations performed to investigate its capabilities in terms of energy resolution to detect such spectral features. Simulations (see [30] for a detailed discussion) were performed to investigate the percentage of number of detections of the spectral feature with respect to a broad-band log-parabola above 5 σ confidence level for 200 realizations. In order to have at least a ≈50% probability of detection of the feature, 1.5 h of observation time would be required, increasing up to ≈80% probability for 2 h of exposure.

4.4. Disentangling Spectral Models in Misaligned Jetted Sources—NGC 1068

NGC 1068 ( z = 0.00379 ± 0.00001 ) is a powerful γ -ray Seyfert-2 galaxy detected by Fermi-LAT. It also hosts starburst activity in its central region and AGN-driven winds. The origin of the γ -ray emission is still debated because of the presence of different particle acceleration sites, such as the starburst ring, the circum-nuclear disk, and the jet [58,74]. Moreover, it has recently been associated with a possible source of neutrino emission [57]. While the canonical jet model does not extend above 10 TeV (see [74]), the AGN wind model predicts a hard spectrum that extends in the very high energy band. Figure 9 shows the results of a simulated deep observation (200 h) to test if we can detect VHE emission expected by the AGN wind model. The ASTRI Mini-Array is able to measure the source spectrum in the energy bins ∼2–5 TeV and ∼5–13 TeV at about 5  σ level.

5. Conclusions

The ASTRI Mini-Array will commence scientific observations at the end of 2025 from the Observatorio del Teide, collecting data that will create a natural connection between current and future VHE facilities and other multi-wavelength observatories by providing light-curves, spectra, and high-resolution images of point-like and extended sources. Its 10° field of view will allow us to investigate both extended sources (e.g., supernova remnants) and crowded/rich fields (e.g., the Galactic Center) with a single pointing, while its 3 angular resolution at 10 TeV will allow us to perform detailed morphological studies of extended sources. Moreover, its sensitivity, extending above 100 TeV with a moderate degradation (about a factor of 2) up to the edge of the FoV, will make it the most sensitive IACT in the 5–200 TeV energy range in the Northern Hemisphere before the advent of CTAO−N. The ASTRI Mini-Array will join the energy domain typical of EASs with the precision domain (excellent angular and energy resolutions) typical of IACTs, allowing several synergies with LHAASO, HAWC, and Tibet AS- γ , investigating PeV-only sources, obtaining broad-band spectra, and detailed source morphology. For the first four years, the ASTRI Mini-Array will be run as an experiment with dedicated pointings in order to address specific Core Science Topics. Afterward, we expect a smooth transition toward an Observatory Phase open to observational proposals from the scientific community.

Funding

This work received no funding.

Data Availability Statement

This research has made use of the ASTRI Mini-Array Instrument Response Functions (IRFs) provided by the ASTRI Project [75] which are publicly available. All physical models are available through the cited literature. The ASTRI Mini-Array simulations were performed using the publicly available software ctools [v1.6.3] [76] and gammapy [v0.17] [77].

Acknowledgments

This work was conducted in the context of the ASTRI Project thanks to the support of the Italian Ministry of University and Research (MUR) as well as the Ministry for Economic Development (MISE), with funds explicitly assigned to the Italian National Institute of Astrophysics (INAF). We acknowledge the support of the Brazilian Funding Agency FAPESP (Grant 2013/10559-5) and the South African Department of Science and Technology through Funding Agreement 0227/2014 for the South African Gamma-Ray Astronomy Program. IAC is supported by the Spanish Ministry of Science and Innovation (MICIU). They are partially supported by H2020-ASTERICS, a project funded by the European Commission Framework Programme Horizon 2020 Research and Innovation action under grant agreement n. 653477. The ASTRI project is becoming a reality thanks to Giovanni “Nanni” Bignami, Nicolò “Nichi” D’Amico, two outstanding scientists who, in their capability as INAF Presidents, provided continuous support and invaluable guidance. Although Nanni was instrumental in starting the ASTRI telescope, Nichi transformed it into the Mini-Array in Tenerife. Now, the project is being built owing to the unfaltering support of Marco Tavani, the current INAF President. Paolo Vettolani and Filippo Zerbi, the past and current INAF Science Directors, and Massimo Cappi, the Coordinator of the High Energy branch of INAF, have been also very supportive of our work. We are very grateful to all of them. Unfortunately, Nanni and Nichi passed away, but their vision still guides us. This review went through the internal ASTRI review process. SV also wishes to thank the Universe Editor of the Special Issue “Recent Advances in Gamma Ray Astrophysics and Future Perspectives”, P. Romano, for inviting him to write a review, and the referees for their comments that helped improve the manuscript.

Conflicts of Interest

The author declares no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
ALPAxion-like particles
AS- γ Air shower γ -ray array
ASTRIAstrofisica con specchi a tecnologia replicante italiana
C.L.Confidence Limit
CTAOČerenkov telescope array Observatory
DMDark matter
EASExtended air showers arrays
ESAEuropean space agency
EBLExtra-galactic background light
E-HBLExtreme high-peaked BL Lacs
eROSITAExtended Roentgen survey with an imaging telescope array
FOVField of view
FRFanaroff–Riley galaxies
FSRQFlat-spectrum radio quasar
GASPGLAST-AGILE support programme
GRBGamma-ray burst
GWGravitational wave
HAWCHigh-altitude water Čerenkov observatory
HBHadron Beam
HBLHigh-peaked BL Lacs
HEHigh-energy
H.E.S.S.High-energy stereoscopic system
IACInstituto de Astrofísica de Canarias
IACTImaging atmospheric Čerenkov telescope arrays
IBLIntermediate-peaked BL Lacs
IGMFInter-galactic magnetic field
IRInfra-red
IXPEImaging X-ray polarimetry explorer
LHAASOLarge high-altitude air shower observatory
LIVLorentz invariance violation
LSTLarge-sized telescope
MAGICMajor atmospheric gamma-ray imaging Čerenkov telescopes
MSTMedium-sized telescope
NASANational aeronautics and space administration
SBGStar-bursting galaxies
SCSchwarzschild-Couder
SIIStellar intensity interferometry
SRTSardinia radio telescope
SSTSmall-sized telescope
TNGTelescopio Nazionale Galileo
VERITASVery energetic radiation imaging telescope array system
VHEVery high-energy
WEBTWhole-Earth blazar telescope

Notes

1
http://tevcat.uchicago.edu/, accessed on 14 February 2024.
2
Large-sized telescopes.
3
Medium-sized telescopes.
4
As noted in [33], a 1.1 PeV photon requires a parent electron of energy E∼2.3 PeV.
5
GRB 160821B and GRB 201015A have been detected at a significance of ∼3.1  σ and ∼3.5  σ , respectively, contrary to all the other GRBs whose detection significance exceeds ∼5  σ .
6
More detailed, phase-resolved spectra could not be investigated given the short exposure times in each phase bin.

References

  1. Wakely, S.P.; Horan, D. TeVCat: An online catalog for Very High Energy Gamma-Ray Astronomy. Int. Cosm. Ray Conf. 2008, 3, 1341–1344. [Google Scholar]
  2. Aharonian, F.; Akhperjanian, A.G.; Bazer-Bachi, A.R.; Beilicke, M.; Benbow, W.; Berge, D.; Bernlöhr, K.; Boisson, C.; Bolz, O.; Borrel, V.; et al. Observations of the Crab nebula with HESS. A&A 2006, 457, 899–915. [Google Scholar] [CrossRef]
  3. Aleksić, J.; Alvarez, E.A.; Antonelli, L.A.; Antoranz, P.; Asensio, M.; Backes, M.; Barrio, J.A.; Bastieri, D.; Becerra González, J.; Bednarek, W.; et al. Performance of the MAGIC stereo system obtained with Crab Nebula data. Astropart. Phys. 2012, 35, 435–448. [Google Scholar] [CrossRef]
  4. Weekes, T.C.; Badran, H.; Biller, S.D.; Bond, I.; Bradbury, S.; Buckley, J.; Carter-Lewis, D.; Catanese, M.; Criswell, S.; Cui, W.; et al. VERITAS: The Very Energetic Radiation Imaging Telescope Array System. Astropart. Phys. 2002, 17, 221–243. [Google Scholar] [CrossRef]
  5. Abeysekara, A.U.; Albert, A.; Alfaro, R.; Alvarez, C.; Álvarez, J.D.; Arceo, R.; Arteaga-Velázquez, J.C.; Ayala Solares, H.A.; Barber, A.S.; Bautista-Elivar, N.; et al. Observation of the Crab Nebula with the HAWC Gamma-Ray Observatory. Astrophys. J. 2017, 843, 39. [Google Scholar] [CrossRef]
  6. Cao, Z. A future project at tibet: The large high altitude air shower observatory (LHAASO). Chin. Phys. C 2010, 34, 249–252. [Google Scholar] [CrossRef]
  7. Amenomori, M.; Bi, X.J.; Chen, D.; Chen, T.L.; Chen, W.Y.; Cui, S.W.; Danzengluobu; Ding, L.K.; Feng, C.F.; Feng, Z.; et al. Search for Gamma Rays above 100 TeV from the Crab Nebula with the Tibet Air Shower Array and the 100 m2 muon Detector. Astrophys. J. 2015, 813, 98. [Google Scholar] [CrossRef]
  8. Cherenkov Telescope Array Consortium; Acharya, B.S.; Agudo, I.; Al Samarai, I.; Alfaro, R.; Alfaro, J.; Alispach, C.; Alves Batista, R.; Amans, J.P.; Amato, E.; et al. Science with the Cherenkov Telescope Array; World Scientific: Singapore, 2019. [Google Scholar] [CrossRef]
  9. Gueta, O. The Cherenkov Telescope Array: Layout, design and performance. In Proceedings of the 37th International Cosmic Ray Conference, Virtual, 12–23 July 2022; p. 885. [Google Scholar] [CrossRef]
  10. Scuderi, S. The ASTRI Program. Eur. Phys. J. Web Conf. 2019, 209, 01001. [Google Scholar] [CrossRef]
  11. Giro, E.; Canestrari, R.; Sironi, G.; Antolini, E.; Conconi, P.; Fermino, C.E.; Gargano, C.; Rodeghiero, G.; Russo, F.; Scuderi, S.; et al. First optical validation of a Schwarzschild Couder telescope: The ASTRI SST-2M Cherenkov telescope. A&A 2017, 608, A86. [Google Scholar] [CrossRef]
  12. Lombardi, S.; Catalano, O.; Scuderi, S.; Antonelli, L.A.; Pareschi, G.; Antolini, E.; Arrabito, L.; Bellassai, G.; Bernlöhr, K.; Bigongiari, C.; et al. First detection of the Crab Nebula at TeV energies with a Cherenkov telescope in a dual-mirror Schwarzschild-Couder configuration: The ASTRI-Horn telescope. A&A 2020, 634, A22. [Google Scholar] [CrossRef]
  13. Scuderi, S.; Giuliani, A.; Pareschi, G.; Tosti, G.; Catalano, O.; Amato, E.; Antonelli, L.A.; Becerra Gonzàles, J.; Bellassai, G.; Bigongiari, C.; et al. The ASTRI Mini-Array of Cherenkov telescopes at the Observatorio del Teide. J. High Energy Astrophys. 2022, 35, 52–68. [Google Scholar] [CrossRef]
  14. Scuderi, S. The ASTRI Mini-Array: A new pathfinder for Imaging Cherenkov Telescope Arrays. Universe, 2024; submitted. [Google Scholar]
  15. Vercellone, S.; Bigongiari, C.; Burtovoi, A.; Cardillo, M.; Catalano, O.; Franceschini, A.; Lombardi, S.; Nava, L.; Pintore, F.; Stamerra, A.; et al. ASTRI Mini-Array core science at the Observatorio del Teide. J. High Energy Astrophys. 2022, 35, 1–42. [Google Scholar] [CrossRef]
  16. Aleksić, J.; Ansoldi, S.; Antonelli, L.A.; Antoranz, P.; Babic, A.; Bangale, P.; Barceló, M.; Barrio, J.A.; Becerra González, J.; Bednarek, W.; et al. The major upgrade of the MAGIC telescopes, Part II: A performance study using observations of the Crab Nebula. Astropart. Phys. 2016, 72, 76–94. [Google Scholar] [CrossRef]
  17. Holder, J.; Atkins, R.W.; Badran, H.M.; Blaylock, G.; Bradbury, S.M.; Buckley, J.H.; Byrum, K.L.; Carter-Lewis, D.A.; Celik, O.; Chow, Y.C.K.; et al. The first VERITAS telescope. Astropart. Phys. 2006, 25, 391–401. [Google Scholar] [CrossRef]
  18. De Naurois, M. H.E.S.S.-II—Gamma ray astronomy from 20 GeV to hundreds of TeV’s. Eur. Phys. J. Web Conf. 2017, 136, 03001. [Google Scholar] [CrossRef]
  19. Lombardi, S.; Antonelli, L.A.; Bigongiari, C.; Cardillo, M.; Gallozzi, S.; Green, J.G.; Lucarelli, F.; Saturni, F.G. Performance of the ASTRI Mini-Array at the Observatorio del Teide. In Proceedings of the 37th International Cosmic Ray Conference, Virtual, 12–23 July 2022; p. 884. [Google Scholar]
  20. Zanin, R.; (CTAO Observatory, Bologna, Italy) CTAO Project Scientist. Personal Communication, 2021.
  21. Abeysekara, A.U.; Alfaro, R.; Alvarez, C.; Álvarez, J.D.; Arceo, R.; Arteaga-Velázquez, J.C.; Avila Rojas, D.; Ayala Solares, H.A.; Barber, A.S.; Bautista-Elivar, N.; et al. The HAWC Real-time Flare Monitor for Rapid Detection of Transient Events. Astrophys. J. 2017, 843, 116. [Google Scholar] [CrossRef]
  22. Amenomori, M.; Bao, Y.W.; Bi, X.J.; Chen, D.; Chen, T.L.; Chen, W.Y.; Chen, X.; Chen, Y.; Cirennima; Cui, S.W.; et al. First Detection of Photons with Energy beyond 100 TeV from an Astrophysical Source. Phys. Rev. Lett. 2019, 123, 051101. [Google Scholar] [CrossRef]
  23. di Sciascio, G.; Lhaaso Collaboration. The LHAASO experiment: From Gamma-Ray Astronomy to Cosmic Rays. Nucl. Part. Phys. Proc. 2016, 279-281, 166–173. [Google Scholar] [CrossRef]
  24. Holler, M.; de Naurois, M.; Zaborov, D.; Balzer, A.; Chalmé-Calvet, R. Photon Reconstruction for H.E.S.S. Using a Semi-Analytical Model. Int. Cosm. Ray Conf. 2015, 34, 980. [Google Scholar] [CrossRef]
  25. Abeysekara, A.U.; Albert, A.; Alfaro, R.; Alvarez, C.; Álvarez, J.D.; Arceo, R.; Arteaga-Velázquez, J.C.; Ayala Solares, H.A.; Barber, A.S.; Baughman, B.; et al. The 2HWC HAWC Observatory Gamma-Ray Catalog. Astrophys. J. 2017, 843, 40. [Google Scholar] [CrossRef]
  26. Kawata, K.; Sako, T.K.; Ohnishi, M.; Takita, M.; Nakamura, Y.; Munakata, K. Energy determination of gamma-ray induced air showers observed by an extensive air shower array. Exp. Astron. 2017, 44, 1–9. [Google Scholar] [CrossRef]
  27. Aharonian, F.; An, Q.; Axikegu; Bai, L.X.; Bai, Y.X.; Bao, Y.W.; Bastieri, D.; Bi, X.J.; Bi, Y.J.; Cai, H.; et al. Observation of the Crab Nebula with LHAASO-KM2A—A performance study. Chin. Phys. C 2021, 45, 025002. [Google Scholar] [CrossRef]
  28. Cao, Z.; Aharonian, F.A.; An, Q.; Axikegu; Bai, Y.X.; Bao, Y.W.; Bastieri, D.; Bi, X.J.; Bi, Y.J.; Cai, H.; et al. Ultrahigh-energy photons up to 1.4 petaelectronvolts from 12 γ-ray Galactic sources. Nature 2021, 594, 33–36. [Google Scholar] [CrossRef]
  29. D’Aì, A.; Amato, E.; Burtovoi, A.; Compagnino, A.A.; Fiori, M.; Giuliani, A.; La Palombara, N.; Paizis, A.; Piano, G.; Saturni, F.G.; et al. Galactic observatory science with the ASTRI Mini-Array at the Observatorio del Teide. J. High Energy Astrophys. 2022, 35, 139–175. [Google Scholar] [CrossRef]
  30. Saturni, F.G.; Arcaro, C.H.E.; Balmaverde, B.; Becerra González, J.; Caccianiga, A.; Capalbi, M.; Lamastra, A.; Lombardi, S.; Lucarelli, F.; Alves Batista, R.; et al. Extragalactic observatory science with the ASTRI mini-array at the Observatorio del Teide. J. High Energy Astrophys. 2022, 35, 91–111. [Google Scholar] [CrossRef]
  31. Cao, Z.; Aharonian, F.; An, Q.; Axikegu; Bai, Y.X.; Bao, Y.W.; Bastieri, D.; Bi, X.J.; Bi, Y.J.; Cai, J.T.; et al. The First LHAASO Catalog of Gamma-Ray Sources. arXiv 2023, arXiv:2305.17030. [Google Scholar] [CrossRef]
  32. Saturni, F.G.; Lombardi, S.; Antonelli, L.A.; Bigongiari, C.; Fedorova, E.; Giuliani, A.; Lucarelli, F.; Mastropietro, M.; Pareschi, G. Performance of the ASTRI Mini-Array with different levels of the Night Sky Background. In Proceedings of the 38th International Cosmic Ray Conference—PoS(ICRC2023), Nagoya, Japan, 26 July–3 August 2023; Volume 444, p. 717. [Google Scholar] [CrossRef]
  33. Cao, Z.; Aharonian, F.A.; An, Q.; Axikegu; Bai, Y.X.; Bao, Y.W.; Bastieri, D.; Bi, X.J.; Bi, Y.J.; Cai, H.; et al. Peta‚Äìelectron volt gamma-ray emission from the Crab Nebula. Science 2021, 373, 425–430. [Google Scholar] [CrossRef] [PubMed]
  34. Aharonian, F.; An, Q.; Axikegu; Bai, Y.X.; Bao, Y.W.; Bastieri, D.; Bi, X.J.; Bi, Y.J.; Cai, H.; Tian, W.W.; et al. Performance of LHAASO-WCDA and observation of the Crab Nebula as a standard candle. Chin. Phys. C 2021, 45, 085002. [Google Scholar] [CrossRef]
  35. Punch, M.; Akerlof, C.W.; Cawley, M.F.; Chantell, M.; Fegan, D.J.; Fennell, S.; Gaidos, J.A.; Hagan, J.; Hillas, A.M.; Jiang, Y.; et al. Detection of TeV photons from the active galaxy Markarian 421. Nature 1992, 358, 477–478. [Google Scholar] [CrossRef]
  36. Ahnen, M.L.; Ansoldi, S.; Antonelli, L.A.; Arcaro, C.; Babić, A.; Banerjee, B.; Bangale, P.; Barres de Almeida, U.; Barrio, J.A.; Becerra González, J.; et al. First multi-wavelength campaign on the gamma-ray-loud active galaxy IC 310. A&A 2017, 603, A25. [Google Scholar] [CrossRef]
  37. Aleksić, J.; Antonelli, L.A.; Antoranz, P.; Babic, A.; Barres de Almeida, U.; Barrio, J.A.; Becerra González, J.; Bednarek, W.; Berger, K.; Bernardini, E.; et al. Rapid and multiband variability of the TeV bright active nucleus of the galaxy IC 310. A&A 2014, 563, A91. [Google Scholar] [CrossRef]
  38. MAGIC Collaboration; Acciari, V.A.; Ansoldi, S.; Antonelli, L.A.; Arbet Engels, A.; Arcaro, C.; Baack, D.; Babić, A.; Banerjee, B.; Bangale, P.; et al. Monitoring of the radio galaxy M 87 during a low-emission state from 2012 to 2015 with MAGIC. MNRAS 2020, 492, 5354–5365. [Google Scholar] [CrossRef]
  39. Aharonian, F.; Akhperjanian, A.G.; Bazer-Bachi, A.R.; Beilicke, M.; Benbow, W.; Berge, D.; Bernlöhr, K.; Boisson, C.; Bolz, O.; Borrel, V.; et al. Fast Variability of Tera-Electron Volt γ Rays from the Radio Galaxy M87. Science 2006, 314, 1424–1427. [Google Scholar] [CrossRef] [PubMed]
  40. Aliu, E.; Arlen, T.; Aune, T.; Beilicke, M.; Benbow, W.; Bouvier, A.; Bradbury, S.M.; Buckley, J.H.; Bugaev, V.; Byrum, K.; et al. VERITAS Observations of Day-scale Flaring of M 87 in 2010 April. Astrophys. J. 2012, 746, 141. [Google Scholar] [CrossRef]
  41. Biteau, J.; Prandini, E.; Costamante, L.; Lemoine, M.; Padovani, P.; Pueschel, E.; Resconi, E.; Tavecchio, F.; Taylor, A.; Zech, A. Progress in unveiling extreme particle acceleration in persistent astrophysical jets. Nat. Astron. 2020, 4, 124–131. [Google Scholar] [CrossRef]
  42. Galanti, G.; Roncadelli, M. Behavior of axionlike particles in smoothed out domainlike magnetic fields. Phys. Rev. D 2018, 98, 043018. [Google Scholar] [CrossRef]
  43. Alves Batista, R.; Saveliev, A. The Gamma-Ray Window to Intergalactic Magnetism. Universe 2021, 7, 223. [Google Scholar] [CrossRef]
  44. Tavecchio, F.; Romano, P.; Landoni, M.; Vercellone, S. Putting the hadron beam scenario for extreme blazars to the test with the Cherenkov Telescope Array. MNRAS 2019, 483, 1802–1807. [Google Scholar] [CrossRef]
  45. Acciari, V.A.; Ansoldi, S.; Antonelli, L.A.; Arbet Engels, A.; Asano, K.; Baack, D.; Babić, A.; Baquero, A.; Barres de Almeida, U.; Barrio, J.A.; et al. MAGIC Observations of the Nearby Short Gamma-Ray Burst GRB 160821B. Astrophys. J. 2021, 908, 90. [Google Scholar] [CrossRef]
  46. Abdalla, H.; Adam, R.; Aharonian, F.; Ait Benkhali, F.; Angüner, E.O.; Arakawa, M.; Arcaro, C.; Armand, C.; Ashkar, H.; Backes, M.; et al. A very-high-energy component deep in the γ-ray burst afterglow. Nature 2019, 575, 464–467. [Google Scholar] [CrossRef]
  47. MAGIC Collaboration; Acciari, V.A.; Ansoldi, S.; Antonelli, L.A.; Arbet Engels, A.; Baack, D.; Babić, A.; Banerjee, B.; Barres de Almeida, U.; Barrio, J.A.; et al. Teraelectronvolt emission from the γ-ray burst GRB 190114C. Nature 2019, 575, 455–458. [Google Scholar] [CrossRef]
  48. H. E. S. S. Collaboration; Abdalla, H.; Aharonian, F.; Ait Benkhali, F.; Angüner, E.O.; Arcaro, C.; Armand, C.; Armstrong, T.; Ashkar, H.; Backes, M.; et al. Revealing X-ray and gamma ray temporal and spectral similarities in the GRB 190829A afterglow. Science 2021, 372, 1081–1085. [Google Scholar] [CrossRef]
  49. Suda, Y.; Artero, M.; Asano, K.; Berti, A.; Nava, L.; Noda, K.; Terauchi, K.; Acciari, V.A.; Ansoldi, S.; Antonelli, L.A.; et al. Observation of a relatively low luminosity long duration GRB 201015A by the MAGIC telescopes. In Proceedings of the 37th International Cosmic Ray Conference, Virtual, 12–23 July 2022; p. 797. [Google Scholar] [CrossRef]
  50. Fukami, S.; Berti, A.; Loporchio, S.; Suda, Y.; Nava, L.; Noda, K.; Bošnjak, Ž.; Asano, K.; Longo, F.; The MAGIC Collaboration; et al. Very-high-energy gamma-ray emission from GRB 201216C detected by MAGIC. In Proceedings of the 37th International Cosmic Ray Conference, Virtual, 12–23 July 2022; p. 788. [Google Scholar] [CrossRef]
  51. Huang, Y.; Hu, S.; Chen, S.; Zha, M.; Liu, C.; Yao, Z.; Cao, Z.; Experiment, T.L. LHAASO observed GRB 221009A with more than 5000 VHE photons up to around 18 TeV. GRB Coord. Netw. 2022, 32677, 1. [Google Scholar]
  52. Burns, E.; Svinkin, D.; Fenimore, E.; Kann, D.A.; Agüí Fernández, J.F.; Frederiks, D.; Hamburg, R.; Lesage, S.; Temiraev, Y.; Tsvetkova, A.; et al. GRB 221009A: The Boat. Astrophys. J. Lett. 2023, 946, L31. [Google Scholar] [CrossRef]
  53. LHAASO Collaboration; Cao, Z.; Aharonian, F.; An, Q.; Axikegu; Bai, L.X.; Bai, Y.X.; Bao, Y.W.; Bastieri, D.; Bi, X.J.; et al. A tera-electron volt afterglow from a narrow jet in an extremely bright gamma-ray burst. Science 2023, 380, 1390–1396. [Google Scholar] [CrossRef] [PubMed]
  54. Cao, Z.; Aharonian, F.; An, Q.; Axikegu; Bai, Y.X.; Bao, Y.W.; Bastieri, D.; Bi, X.J.; Bi, Y.J.; Cai, J.T.; et al. Very high-energy gamma-ray emission beyond 10 TeV from GRB 221009A. Sci. Adv. 2023, 9, eadj2778. [Google Scholar] [CrossRef]
  55. Galanti, G.; Roncadelli, M.; Tavecchio, F. Assessment of ALP scenarios for GRB 221009A. arXiv 2022, arXiv:2211.06935. [Google Scholar] [CrossRef]
  56. Galanti, G.; Nava, L.; Roncadelli, M.; Tavecchio, F.; Bonnoli, G. Observability of the Very-High-Energy Emission from GRB 221009A. Phys. Rev. Lett. 2023, 131, 251001. [Google Scholar] [CrossRef]
  57. IceCube Collaboration; Abbasi, R.; Ackermann, M.; Adams, J.; Aguilar, J.A.; Ahlers, M.; Ahrens, M.; Alameddine, J.M.; Alispach, C.; Alves, A.A., Jr.; et al. Evidence for neutrino emission from the nearby active galaxy NGC 1068. Science 2022, 378, 538–543. [Google Scholar] [CrossRef]
  58. Lamastra, A.; Fiore, F.; Guetta, D.; Antonelli, L.A.; Colafrancesco, S.; Menci, N.; Puccetti, S.; Stamerra, A.; Zappacosta, L. Galactic outflow driven by the active nucleus and the origin of the gamma-ray emission in NGC 1068. A&A 2016, 596, A68. [Google Scholar] [CrossRef]
  59. Acciari, V.A.; Ansoldi, S.; Antonelli, L.A.; Arbet Engels, A.; Baack, D.; Babić, A.; Banerjee, B.; Barres de Almeida, U.; Barrio, J.A.; Becerra González, J.; et al. Constraints on Gamma-Ray and Neutrino Emission from NGC 1068 with the MAGIC Telescopes. Astrophys. J. 2019, 883, 135. [Google Scholar] [CrossRef]
  60. Zampieri, L.; Bonanno, G.; Bruno, P.; Gargano, C.; Lessio, L.; Naletto, G.; Paoletti, L.; Rodeghiero, G.; Romeo, G.; Bulgarelli, A.; et al. A stellar intensity interferometry instrument for the ASTRI Mini-Array telescopes. In Optical and Infrared Interferometry and Imaging VIII; Mérand, A., Sallum, S., Sanchez-Bermudez, J., Eds.; Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series; SPIE: Bellingham, WA, USA, 2022; Volume 12183, p. 121830F. [Google Scholar] [CrossRef]
  61. Egron, E.; Pellizzoni, A.; Iacolina, M.N.; Loru, S.; Marongiu, M.; Righini, S.; Cardillo, M.; Giuliani, A.; Mulas, S.; Murtas, G.; et al. Imaging of SNR IC443 and W44 with the Sardinia Radio Telescope at 1.5 and 7 GHz. MNRAS 2017, 470, 1329–1341. [Google Scholar] [CrossRef]
  62. Loru, S.; Pellizzoni, A.; Egron, E.; Righini, S.; Iacolina, M.N.; Mulas, S.; Cardillo, M.; Marongiu, M.; Ricci, R.; Bachetti, M.; et al. Investigating the high-frequency spectral features of SNRs Tycho, W44, and IC443 with the Sardinia Radio Telescope. MNRAS 2019, 482, 3857–3867. [Google Scholar] [CrossRef]
  63. Bortoletto, F.; Bonoli, C.; D’Alessandro, M.; Ragazzoni, R.; Conconi, P.; Mancini, D.; Pucillo, M. Commissioning of the Italian National Telescope Galileo. In Advanced Technology Optical/IR Telescopes VI; Stepp, L.M., Ed.; Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series; SPIE: Bellingham, WA, USA, 1998; Volume 3352, pp. 91–101. [Google Scholar] [CrossRef]
  64. Villata, M.; Raiteri, C.M.; Larionov, V.M.; Kurtanidze, O.M.; Nilsson, K.; Aller, M.F.; Tornikoski, M.; Volvach, A.; Aller, H.D.; Arkharov, A.A.; et al. Multifrequency monitoring of the blazar 0716+714 during the GASP-WEBT-AGILE campaign of 2007. A&A 2008, 481, L79–L82. [Google Scholar] [CrossRef]
  65. Predehl, P.; Andritschke, R.; Arefiev, V.; Babyshkin, V.; Batanov, O.; Becker, W.; Böhringer, H.; Bogomolov, A.; Boller, T.; Borm, K.; et al. The eROSITA X-ray telescope on SRG. A&A 2021, 647, A1. [Google Scholar] [CrossRef]
  66. Weisskopf, M.C.; Ramsey, B.; O’Dell, S.; Tennant, A.; Elsner, R.; Soffitta, P.; Bellazzini, R.; Costa, E.; Kolodziejczak, J.; Kaspi, V.; et al. The Imaging X-ray Polarimetry Explorer (IXPE). In Space Telescopes and Instrumentation 2016: Ultraviolet to Gamma Ray; den Herder, J.W.A., Takahashi, T., Bautz, M., Eds.; SPIE: Bellingham, WA, USA, 2016; Society of Photo-Optical Instrumentation 72 Engineers (SPIE) Conference Series; Volume 9905, p. 990517. [Google Scholar] [CrossRef]
  67. Albert, A.; Alfaro, R.; Alvarez, C.; Camacho, J.R.A.; Arteaga-Velázquez, J.C.; Arunbabu, K.P.; Avila Rojas, D.; Ayala Solares, H.A.; Baghmanyan, V.; Belmont-Moreno, E.; et al. 3HWC: The Third HAWC Catalog of Very-high-energy Gamma-Ray Sources. Astrophys. J. 2020, 905, 76. [Google Scholar] [CrossRef]
  68. LHAASO Collaboration. An ultrahigh-energy γ-ray bubble powered by a super PeVatron. Sci. Bull. 2024, 69, 449–457. [Google Scholar] [CrossRef] [PubMed]
  69. Chernyakova, M.; Malyshev, D.; Paizis, A.; La Palombara, N.; Balbo, M.; Walter, R.; Hnatyk, B.; van Soelen, B.; Romano, P.; Munar-Adrover, P.; et al. Overview of non-transient γ-ray binaries and prospects for the Cherenkov Telescope Array. A&A 2019, 631, A177. [Google Scholar] [CrossRef]
  70. Pintore, F.; Giuliani, A.; Belfiore, A.; Paizis, A.; Mereghetti, S.; La Palombara, N.; Crestan, S.; Sidoli, L.; Lombardi, S.; D’Aì, A.; et al. Scientific prospects for a mini-array of ASTRI telescopes: A γ-ray TeV data challenge. J. High Energy Astrophys. 2020, 26, 83–94. [Google Scholar] [CrossRef]
  71. Aharonian, F.; Akhperjanian, A.G.; Bazer-Bachi, A.R.; Beilicke, M.; Benbow, W.; Berge, D.; Bernlöhr, K.; Boisson, C.; Bolz, O.; Borrel, V.; et al. 3.9 day orbital modulation in the TeV γ-ray flux and spectrum from the X-ray binary LS 5039. A&A 2006, 460, 743–749. [Google Scholar] [CrossRef]
  72. Quinn, J.; Akerlof, C.W.; Biller, S.; Buckley, J.; Carter-Lewis, D.A.; Cawley, M.F.; Catanese, M.; Connaughton, V.; Fegan, D.J.; Finley, J.P.; et al. Detection of Gamma Rays with E > 300 GeV from Markarian 501. Astrophys. J. 1996, 456, L83. [Google Scholar] [CrossRef]
  73. MAGIC Collaboration; Acciari, V.A.; Ansoldi, S.; Antonelli, L.A.; Babić, A.; Banerjee, B.; Barres de Almeida, U.; Barrio, J.A.; Becerra González, J.; Bednarek, W.; et al. Study of the variable broadband emission of Markarian 501 during the most extreme Swift X-ray activity. A&A 2020, 637, A86. [Google Scholar] [CrossRef]
  74. Lamastra, A.; Tavecchio, F.; Romano, P.; Landoni, M.; Vercellone, S. Unveiling the origin of the gamma-ray emission in NGC 1068 with the Cherenkov Telescope Array. Astropart. Phys. 2019, 112, 16–23. [Google Scholar] [CrossRef]
  75. ASTRI Project, ASTRI Mini-Array Instrument Response Functions (Prod2, v1.0). 2022. Available online: https://zenodo.org/records/6827882 (accessed on 21 January 2024). [CrossRef]
  76. Knödlseder, J.; Mayer, M.; Deil, C.; Cayrou, J.B.; Owen, E.; Kelley-Hoskins, N.; Lu, C.C.; Buehler, R.; Forest, F.; Louge, T.; et al. GammaLib and ctools. A software framework for the analysis of astronomical gamma-ray data. A&A 2016, 593, A1. [Google Scholar] [CrossRef]
  77. Donath, A.; Terrier, R.; Remy, Q.; Sinha, A.; Nigro, C.; Pintore, F.; Khélifi, B.; Olivera-Nieto, L.; Ruiz, J.E.; Brügge, K.; et al. Gammapy: A Python package for gamma-ray astronomy. A&A 2023, 678, A157. [Google Scholar] [CrossRef]
Figure 3. ASTRI Mini-Array angular (left panel) and energy (right panel) resolution as a function of the energy. The angular resolution is defined as the 68% γ -ray event containment radius (in degrees). The black points were computed with analysis cuts optimizing the differential sensitivity in 50 h; the long-dashed, dark-green lines were instead derived with analysis cuts taking into account also the angular/energy resolution in the optimization; the short-dashed, light-green lines mark the 0.05° and the 0.1% threshold for the angular and energy resolution, respectively. Adapted from [19].
Figure 3. ASTRI Mini-Array angular (left panel) and energy (right panel) resolution as a function of the energy. The angular resolution is defined as the 68% γ -ray event containment radius (in degrees). The black points were computed with analysis cuts optimizing the differential sensitivity in 50 h; the long-dashed, dark-green lines were instead derived with analysis cuts taking into account also the angular/energy resolution in the optimization; the short-dashed, light-green lines mark the 0.05° and the 0.1% threshold for the angular and energy resolution, respectively. Adapted from [19].
Universe 10 00094 g003
Figure 4. Graphical description of the ASTRI Mini-Array Core Science program.
Figure 4. Graphical description of the ASTRI Mini-Array Core Science program.
Universe 10 00094 g004
Figure 5. Pillar-1 and Pillar-2 target regions (red and blue circles, ≈ 10 in diameter) and 1LHAASO sources (orange dots) on the sky in Galactic coordinates (Aitoff projection). The black solid line shows the declination limit for the ASTRI Mini-Array pointings.
Figure 5. Pillar-1 and Pillar-2 target regions (red and blue circles, ≈ 10 in diameter) and 1LHAASO sources (orange dots) on the sky in Galactic coordinates (Aitoff projection). The black solid line shows the declination limit for the ASTRI Mini-Array pointings.
Universe 10 00094 g005
Figure 6. ASTRI Mini-Array simulations of the Cygnus region mini-survey. The count maps were produced assuming for each pointing an exposure of 1 h (top panel), 2 h (middle panel), and 4 h (bottom panel), respectively. Sky map units are counts/pixels. From [29].
Figure 6. ASTRI Mini-Array simulations of the Cygnus region mini-survey. The count maps were produced assuming for each pointing an exposure of 1 h (top panel), 2 h (middle panel), and 4 h (bottom panel), respectively. Sky map units are counts/pixels. From [29].
Universe 10 00094 g006
Figure 7. ASTRI Mini-Array simulations of LS 5039. Left panel: orbital modulation obtained with 10 h-long simulations per orbital phase bin. The open squares are the expected fluxes from the models, while the filled circles are the simulated fluxes in 0.8–200 TeV. Error bars are at 1 σ C.L. Right panel: 1 σ uncertainty ( δ Γ ) on the spectral index obtained for 10 h-long simulations per orbital bin. From [29].
Figure 7. ASTRI Mini-Array simulations of LS 5039. Left panel: orbital modulation obtained with 10 h-long simulations per orbital phase bin. The open squares are the expected fluxes from the models, while the filled circles are the simulated fluxes in 0.8–200 TeV. Error bars are at 1 σ C.L. Right panel: 1 σ uncertainty ( δ Γ ) on the spectral index obtained for 10 h-long simulations per orbital bin. From [29].
Universe 10 00094 g007
Figure 8. ASTRI Mini-Array simulations of the spectral feature emerging at ≈3 TeV in the spectrum on Mrk 501 during its highest ever-recorded X-ray flux state. LP = log-parabola; eplogpar = curved log-parabola; EBL = extra-galactic background light. From [30].
Figure 8. ASTRI Mini-Array simulations of the spectral feature emerging at ≈3 TeV in the spectrum on Mrk 501 during its highest ever-recorded X-ray flux state. LP = log-parabola; eplogpar = curved log-parabola; EBL = extra-galactic background light. From [30].
Universe 10 00094 g008
Figure 9. ASTRI Mini-Array simulations of the NGC 1068 VHE spectral energy distribution. Different emission models (see [74] for a detailed description) have been considered. From [30].
Figure 9. ASTRI Mini-Array simulations of the NGC 1068 VHE spectral energy distribution. Different emission models (see [74] for a detailed description) have been considered. From [30].
Universe 10 00094 g009
Table 1. Performance of the ASTRI Mini-Array compared with the main current IACT arrays. References: ASTRI Mini-array [15], MAGIC [16], VERITAS [17] and https://veritas.sao.arizona.edu (accessed on 14 February 2024), H.E.S.S. [2].
Table 1. Performance of the ASTRI Mini-Array compared with the main current IACT arrays. References: ASTRI Mini-array [15], MAGIC [16], VERITAS [17] and https://veritas.sao.arizona.edu (accessed on 14 February 2024), H.E.S.S. [2].
QuantityASTRI Mini-ArrayMAGICVERITASH.E.S.S.
Location28° 18 04 N28° 45 22 N31° 40 30 N23° 16 18 S
16° 30 38 W17° 53 30 W110° 57 7.8 W16° 30 00 E
Altitude [m]2390239612681800
FoV 10 ° 3.5 ° 3.5 ° 5 °
Angular Res. 0.05 ° (10 TeV)0.07° (1 TeV)0.07° (1 TeV)0.06° (1 TeV)
Energy Res.10% (10 TeV)16% (1 TeV)17% (1 TeV)15% (1 TeV)
Energy Range(0.5–200) TeV(0.05–20) TeV(0.08–30) TeV(0.02–30) TeV ( a )
Notes: ( a ) : considering the contribution of H.E.S.S. II telescope unit [18].
Table 2. Summary of the performance of the current main particle sampling arrays compared with those of the ASTRI Mini-Array. References: ASTRI Mini-array [15], HAWC [5,25], LHAASO [6], Tibet AS- γ  [22,26].
Table 2. Summary of the performance of the current main particle sampling arrays compared with those of the ASTRI Mini-Array. References: ASTRI Mini-array [15], HAWC [5,25], LHAASO [6], Tibet AS- γ  [22,26].
QuantityASTRI Mini-ArrayHAWCLHAASOTibet AS- γ
Location28° 18 04 N18° 59 41 N29° 21 31 N30° 05 00 N
16° 30 38 W97° 18 27 W100° 08 15 E90° 33 00 E
Altitude [m]2390410044104300
FoV∼0.024 sr2 sr2 sr2 sr
Angular Res. 0.05 ° (10 TeV)0.15° ( a ) (10 TeV)(0.24–0.32)° ( b ) (100 TeV)0.2° ( c ) (100 TeV)
Energy Res.10% (10 TeV)30% (10 TeV)(13–36)% (100 TeV) ( b ) 20% ( c ) (100 TeV)
Energy Range(0.5–200) TeV(0.1–1000) TeV(0.1–1000) TeV(0.1–1000) TeV
Notes: ( a ) : (0.15–1)° as a function of the event size. ( b ) : angular resolution is (0.70–0.94)° at 10 TeV; (0.24–0.32)° at 100 TeV; 0.15° at 1000 TeV. Energy resolution is (30–45)% at 10 TeV; (13–36)% at 100 TeV; (8–20)% at 1000 TeV [27]. ( c ) : angular resolution is 0.5 ° at 10 TeV and ∼0.2° at 10 TeV at 50% containment radius [22]. Energy resolution is ∼40% at 10 TeV and ∼20% at 100 TeV [26]. The different values of the LHAASO angular and energy resolution performance at a given energy have been computed at different Zenith angles, 0 < θ < 20 , 20 < θ < 35 , and 35 < θ < 50 degrees, respectively. At lower Zenith angles, the performance is better.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Vercellone, S., on behalf of the ASTRI Project. Science with the ASTRI Mini-Array: From Experiment to Open Observatory. Universe 2024, 10, 94. https://doi.org/10.3390/universe10020094

AMA Style

Vercellone S on behalf of the ASTRI Project. Science with the ASTRI Mini-Array: From Experiment to Open Observatory. Universe. 2024; 10(2):94. https://doi.org/10.3390/universe10020094

Chicago/Turabian Style

Vercellone, Stefano on behalf of the ASTRI Project. 2024. "Science with the ASTRI Mini-Array: From Experiment to Open Observatory" Universe 10, no. 2: 94. https://doi.org/10.3390/universe10020094

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop