ArticlePDF Available

Resolving the Extragalactic γ -Ray Background above 50 GeV with the Fermi Large Area Telescope

Authors:

Abstract and Figures

The Fermi Large Area Telescope (LAT) Collaboration has recently released a catalog of 360 sources detected above 50 GeV (2FHL). This catalog was obtained using 80 months of data re-processed with Pass 8, the newest event-level analysis, which significantly improves the acceptance and angular resolution of the instrument. Most of the 2FHL sources at high Galactic latitude are blazars. Using detailed Monte Carlo simulations, we measure, for the first time, the source count distribution, dN/dS, of extragalactic γ-ray sources at E>50 GeV and find that it is compatible with a Euclidean distribution down to the lowest measured source flux in the 2FHL (∼8×10-12 ph cm-2 s-1). We employ a one-point photon fluctuation analysis to constrain the behavior of dN/dS below the source detection threshold. Overall, the source count distribution is constrained over three decades in flux and found compatible with a broken power law with a break flux, Sb, in the range [8×10-12,1.5×10-11] ph cm-2 s-1 and power-law indices below and above the break of α2[1.60,1.75] and α1=2.49±0.12, respectively. Integration of dN/dS shows that point sources account for at least 86-14+16% of the total extragalactic γ-ray background. The simple form of the derived source count distribution is consistent with a single population (i.e., blazars) dominating the source counts to the minimum flux explored by this analysis. We estimate the density of sources detectable in blind surveys that will be performed in the coming years by the Cherenkov Telescope Array.
Content may be subject to copyright.
Version 00 as of November 4, 2015
To be submitted to PRL
Resolving the Extragalactic γ-ray Background above 50 GeV with Fermi-LAT
M. Ackermann,1M. Ajello,2, A. Albert,3W. B. Atwood,4L. Baldini,5, 3 J. Ballet,6G. Barbiellini,7, 8 D. Bastieri,9, 10
K. Bechtol,11 R. Bellazzini,12 E. Bissaldi,13 R. D. Blandford,3E. D. Bloom,3R. Bonino,14, 15 J. Bregeon,16
R. J. Britto,17 P. Bruel,18 R. Buehler,1G. A. Caliandro,3, 19 R. A. Cameron,3M. Caragiulo,20, 13 P. A. Caraveo,21
E. Cavazzuti,22 C. Cecchi,23, 24 E. Charles,3A. Chekhtman,25 J. Chiang,3G. Chiaro,10 S. Ciprini,22, 23
J. Cohen-Tanugi,16 L. R. Cominsky,26 F. Costanza,13 S. Cutini,22, 27, 23 F. D’Ammando,28, 29 A. de Angelis,30
F. de Palma,13, 31 R. Desiante,32, 14 S. W. Digel,3M. Di Mauro,3 , L. Di Venere,20, 13 A. Dom´ınguez,2P. S. Drell,3
C. Favuzzi,20, 13 S. J. Fegan,18 E. C. Ferrara,33 A. Franckowiak,3Y. Fukazawa,34 S. Funk,35 P. Fusco,20, 13
F. Gargano,13 D. Gasparrini,22, 23 N. Giglietto,20, 13 P. Giommi,22 F. Giordano,20, 13 M. Giroletti,28 G. Godfrey,3
D. Green,36, 33 I. A. Grenier,6S. Guiriec,33, 37 E. Hays,33 D. Horan,18 G. Iafrate,7, 38 T. Jogler,3G. J´ohannesson,39
M. Kuss,12 G. La Mura,10, 40 S. Larsson,41, 42 L. Latronico,14 J. Li,43 L. Li,41, 42 F. Longo,7, 8 F. Loparco,20, 13
B. Lott,44 M. N. Lovellette,45 P. Lubrano,23, 24 G. M. Madejski,3J. Magill,36 S. Maldera,14 A. Manfreda,12
M. Mayer,1M. N. Mazziotta,13 P. F. Michelson,3W. Mitthumsiri,46 T. Mizuno,47 A. A. Moiseev,48, 36
M. E. Monzani,3A. Morselli,49 I. V. Moskalenko,3S. Murgia,50 M. Negro,14, 15 E. Nuss,16 T. Ohsugi,47 C. Okada,34
N. Omodei,3E. Orlando,3J. F. Ormes,51 D. Paneque,52, 3 J. S. Perkins,33 M. Pesce-Rollins,12, 3 V. Petrosian,3
F. Piron,16 G. Pivato,12 T. A. Porter,3S. Rain`o,20, 13 R. Rando,9, 10 M. Razzano,12, 53 S. Razzaque,17 A. Reimer,40, 3
O. Reimer,40, 3 T. Reposeur,44 R. W. Romani,3M. S´anchez-Conde,42, 54 J. Schmid,6A. Schulz,1C. Sgr`o,12
D. Simone,13 E. J. Siskind,55 F. Spada,12 G. Spandre,12 P. Spinelli,20, 13 D. J. Suson,56 H. Takahashi,34
J. B. Thayer,3L. Tibaldo,57 D. F. Torres,43, 58 E. Troja,33, 36 G. Vianello,3M. Yassine,16 and S. Zimmer54, 42
1Deutsches Elektronen Synchrotron DESY, D-15738 Zeuthen, Germany
2Department of Physics and Astronomy, Clemson University,
Kinard Lab of Physics, Clemson, SC 29634-0978, USA
3W. W. Hansen Experimental Physics Laboratory,
Kavli Institute for Particle Astrophysics and Cosmology,
Department of Physics and SLAC National Accelerator Laboratory, Stanford University, Stanford, CA 94305, USA
4Santa Cruz Institute for Particle Physics, Department of Physics and Department of Astronomy and Astrophysics,
University of California at Santa Cruz, Santa Cruz, CA 95064, USA
5Universit`a di Pisa and Istituto Nazionale di Fisica Nucleare, Sezione di Pisa I-56127 Pisa, Italy
6Laboratoire AIM, CEA-IRFU/CNRS/Universit´e Paris Diderot,
Service d’Astrophysique, CEA Saclay, F-91191 Gif sur Yvette, France
7Istituto Nazionale di Fisica Nucleare, Sezione di Trieste, I-34127 Trieste, Italy
8Dipartimento di Fisica, Universit`a di Trieste, I-34127 Trieste, Italy
9Istituto Nazionale di Fisica Nucleare, Sezione di Padova, I-35131 Padova, Italy
10Dipartimento di Fisica e Astronomia “G. Galilei”, Universit`a di Padova, I-35131 Padova, Italy
11Dept. of Physics and Wisconsin IceCube Particle Astrophysics Center,
University of Wisconsin, Madison, WI 53706, USA
12Istituto Nazionale di Fisica Nucleare, Sezione di Pisa, I-56127 Pisa, Italy
13Istituto Nazionale di Fisica Nucleare, Sezione di Bari, I-70126 Bari, Italy
14Istituto Nazionale di Fisica Nucleare, Sezione di Torino, I-10125 Torino, Italy
15Dipartimento di Fisica Generale “Amadeo Avogadro” ,
Universit`a degli Studi di Torino, I-10125 Torino, Italy
16Laboratoire Univers et Particules de Montpellier,
Universit´e Montpellier, CNRS/IN2P3, Montpellier, France
17Department of Physics, University of Johannesburg,
PO Box 524, Auckland Park 2006, South Africa
18Laboratoire Leprince-Ringuet, ´
Ecole polytechnique, CNRS/IN2P3, Palaiseau, France
19Consorzio Interuniversitario per la Fisica Spaziale (CIFS), I-10133 Torino, Italy
20Dipartimento di Fisica “M. Merlin” dell’Universit`a e del Politecnico di Bari, I-70126 Bari, Italy
21INAF-Istituto di Astrofisica Spaziale e Fisica Cosmica, I-20133 Milano, Italy
22Agenzia Spaziale Italiana (ASI) Science Data Center, I-00133 Roma, Italy
23Istituto Nazionale di Fisica Nucleare, Sezione di Perugia, I-06123 Perugia, Italy
24Dipartimento di Fisica, Universit`a degli Studi di Perugia, I-06123 Perugia, Italy
25College of Science, George Mason University, Fairfax, VA 22030,
resident at Naval Research Laboratory, Washington, DC 20375, USA
26Department of Physics and Astronomy, Sonoma State University, Rohnert Park, CA 94928-3609, USA
arXiv:1511.00693v1 [astro-ph.CO] 2 Nov 2015
27INAF Osservatorio Astronomico di Roma, I-00040 Monte Porzio Catone (Roma), Italy
28INAF Istituto di Radioastronomia, I-40129 Bologna, Italy
29Dipartimento di Astronomia, Universit`a di Bologna, I-40127 Bologna, Italy
30Dipartimento di Fisica, Universit`a di Udine and Istituto Nazionale di Fisica Nucleare,
Sezione di Trieste, Gruppo Col legato di Udine, I-33100 Udine
31Universit`a Telematica Pegaso, Piazza Trieste e Trento, 48, I-80132 Napoli, Italy
32Universit`a di Udine, I-33100 Udine, Italy
33NASA Goddard Space Flight Center, Greenbelt, MD 20771, USA
34Department of Physical Sciences, Hiroshima University, Higashi-Hiroshima, Hiroshima 739-8526, Japan
35Erlangen Centre for Astroparticle Physics, D-91058 Erlangen, Germany
36Department of Physics and Department of Astronomy,
University of Maryland, College Park, MD 20742, USA
37NASA Postdoctoral Program Fellow, USA
38Osservatorio Astronomico di Trieste, Istituto Nazionale di Astrofisica, I-34143 Trieste, Italy
39Science Institute, University of Iceland, IS-107 Reykjavik, Iceland
40Institut f¨ur Astro- und Teilchenphysik and Institut f¨ur Theoretische Physik,
Leopold-Franzens-Universit¨at Innsbruck, A-6020 Innsbruck, Austria
41Department of Physics, KTH Royal Institute of Technology, AlbaNova, SE-106 91 Stockholm, Sweden
42The Oskar Klein Centre for Cosmoparticle Physics, AlbaNova, SE-106 91 Stockholm, Sweden
43Institute of Space Sciences (IEEC-CSIC), Campus UAB, E-08193 Barcelona, Spain
44Centre d’ ´
Etudes Nucl´eaires de Bordeaux Gradignan, IN2P3/CNRS,
Universit´e Bordeaux 1, BP120, F-33175 Gradignan Cedex, France
45Space Science Division, Naval Research Laboratory, Washington, DC 20375-5352, USA
46Department of Physics, Faculty of Science, Mahidol University, Bangkok 10400, Thailand
47Hiroshima Astrophysical Science Center, Hiroshima University, Higashi-Hiroshima, Hiroshima 739-8526, Japan
48Center for Research and Exploration in Space Science and Technology
(CRESST) and NASA Goddard Space Flight Center, Greenbelt, MD 20771, USA
49Istituto Nazionale di Fisica Nucleare, Sezione di Roma “Tor Vergata”, I-00133 Roma, Italy
50Center for Cosmology, Physics and Astronomy Department,
University of California, Irvine, CA 92697-2575, USA
51Department of Physics and Astronomy, University of Denver, Denver, CO 80208, USA
52Max-Planck-Institut f¨ur Physik, D-80805 M¨unchen, Germany
53Funded by contract FIRB-2012-RBFR12PM1F from the Italian Ministry of Education, University and Research (MIUR)
54Department of Physics, Stockholm University, AlbaNova, SE-106 91 Stockholm, Sweden
55NYCB Real-Time Computing Inc., Lattingtown, NY 11560-1025, USA
56Department of Chemistry and Physics, Purdue University Calumet, Hammond, IN 46323-2094, USA
57Max-Planck-Institut f¨ur Kernphysik, D-69029 Heidelberg, Germany
58Instituci´o Catalana de Recerca i Estudis Avan¸cats (ICREA), Barcelona, Spain
(Dated: November 4, 2015)
The Fermi Large Area Telescope (LAT) Collaboration has recently released a catalog of 360
sources detected above 50 GeV (2FHL). This catalog was obtained using 80 months of data re-
processed with Pass 8, the newest event-level analysis, which significantly improves the acceptance
and angular resolution of the instrument. Most of the 2FHL sources at high Galactic latitude are
blazars. Using detailed Monte Carlo simulations, we measure, for the first time, the source count
distribution, dN/dS, of extragalactic γ-ray sources at E > 50 GeV and find that it is compatible
with a Euclidean distribution down to the lowest measured source flux in the 2FHL (8×1012 ph
cm2s1). We employ a one-point photon fluctuation analysis to constrain the behavior of dN/dS
below the source detection threshold. Overall the source count distribution is constrained over
three decades in flux and found compatible with a broken power law with a break flux, Sb, in the
range [8 ×1012 ,1.5×1011] ph cm2s1and power-law indices below and above the break of
α2[1.60,1.75] and α1= 2.49 ±0.12 respectively. Integration of dN/dS shows that point sources
account for at least 86+16
14% of the total extragalactic γ-ray background. The simple form of the
derived source count distribution is consistent with a single population (i.e. blazars) dominating the
source counts to the minimum flux explored by this analysis. We estimate the density of sources
detectable in blind surveys that will be performed in the coming years by the Cherenkov Telescope
Array.
PACS numbers:
The origin of the extragalactic γ-ray background
(EGB), the Universe’s glow in γrays, has been debated
since the first measurement with the SAS-2 satellite [1].
The EGB spectrum has been accurately measured, from
100 MeV to 820 GeV, by the Large Area Telescope (LAT)
on board the Fermi Gamma-Ray Space Telescope mission
[2]. Part of the EGB arises from the emission of resolved
and unresolved point sources like blazars, star-forming
and radio galaxies [e.g. 3, 4, 5], which are routinely de-
tected in γrays. A possible contribution to the EGB
may also come from diffuse processes such as annihilat-
ing/decaying dark matter particles (see [6] for a review).
Here we show for the first time that Fermi-LAT is able
to resolve the high-energy EGB into point-like sources.
Indeed, thanks to the accrual of 80 months of data (see
right panel of Fig. 1) and the increased acceptance and
improved point-spread function delivered by the new
event-level analysis dubbed Pass 8 [7], the LAT has re-
cently performed an all-sky survey at >50 GeV resulting
in the detection of 360 γ-ray sources that constitute the
second catalog of hard Fermi-LAT sources [2FHL, 8]
Blazars, mostly belonging to the BL Lacertae (BL Lac)
population, are the majority (74 %) of the sources in the
2FHL catalog. At Galactic latitudes (b) larger than 10
about 70% of the detected sources are associated with
BL Lacs. Only 7 % of the high-latitude (|b|>10)
sources are classified as something other than BL Lacs,
4% of them being Flat Spectrum Radio Quasars (FS-
RQs), while blazars of uncertain type and unassociated
sources constitute the remaining 23 % of the sample. The
median of the synchrotron peak frequencies for blazars
of uncertain type is very similar to that of BL Lacs
(log10(νS
peak/Hz) = 15.7 vs. 15.6). The same holds
for the median spectral index of unassociated sources
(Γ =3.0 vs. 3.1). This is supporting the fact that blazars
of uncertain type and unassociated sources are almost en-
tirely BL Lacs. Therefore, the fraction of likely blazars
in the high-latitude 2FHL sample is 97 % (93% BL Lacs
and 4% FSRQs).
In this paper, we derive the source detection efficiency
of the 2FHL catalog analysis using accurate Monte Carlo
simulations of the γ-ray sky. We then infer the intrinsic
flux distribution dN/dS of sources located at a latitude
|b|>10, where Sis the photon flux (ph cm2s1) mea-
sured in the 50 GeV–2 TeV energy band.
The simulations were performed using the gtobssim
tool, which is part of the Fermi ScienceTools distribu-
tion, and using the same pointing and live time his-
tory and event selection as used in the 2FHL cata-
log. We have employed the P8R2 SOURCE V6 in-
strument response function for the simulations and
analysis and the Galactic and isotropic diffuse emis-
sion were simulated using the gll iem v06.fits and
iso P8R2 SOURCE V6 v06.txt templates 1. The last in-
gredient of the simulations is an isotropic population
of point sources that has the characteristics of blazars
(fluxes and spectra) as detected in 2FHL. The simula-
tions described here were produced iteratively and ul-
1See http://fermi.gsfc.nasa.gov/ssc/
timately rely on the source count distribution dN/dS
Sαas determined at the end of photon fluctuation anal-
ysis (see later), which is, a broken power law with a break
flux Sb= 1 ×1011 ph cm2s1and a Euclidean slope
above the break, α1= 5/2, while below Sbthe slope is
α2= 1.65. Sources were generated with fluxes in the
range [Smin, Smax ] = [1014,109] ph cm2s1and with
power-law spectra of the form dN/dE EΓ. For each
source the photon index Γ is drawn from a Gaussian dis-
tribution with average value 3.2 and standard deviation
0.7 (this reproduces the observed distribution as shown
on the right panel of Fig. 2). Galactic sources are not
considered in the simulations since we are interested in
the flux distribution of blazars at |b|>10. We pro-
duced 10 simulations of the γ-ray sky following these
prescriptions and in Fig. 1 the sky map of one simula-
tion is shown together with the real one. Clearly visible
in both maps are the diffuse emission along the Galac-
tic plane, the Fermi bubbles [9], the emission from point
sources and the isotropic diffuse emission.
The energy spectrum of the simulations is consistent
within 10 %, at all energies of interest and for photons
detected at |b|>10, with that of the LAT observations.
As clearly visible in Fig. 1, the spatial distribution of
gamma rays of the real map is also correctly reproduced.
The 10 simulations are analyzed exactly as the real data
were for the 2FHL catalog. This starts from detecting
source candidates using a sliding-cell algorithm and a
wavelet analysis [10] then analyzing each with the stan-
dard Fermi Science Tools, in order to derive the γ-ray
properties of detectable sources (see [8] for more details).
As in the 2FHL catalog, detected sources are those with
a test statistic (TS)>25 and at least 3 associated photons
predicted by the likelihood fit. This leads to the detec-
tion, in the simulations, of 271 ±18 sources at |b|>10,
which is in good agreement with the 253 sources detected
in the 2FHL. Moreover, the simulations show that the
2FHL catalog contains at most 1% of false detections.
In order to further validate our analysis we have per-
formed two consistency checks on the simulations. The
first compares the input source fluxes Strue with the
fluxes Smeas measured with the Fermi Science Tools in
the simulations. The result displayed in the left panel of
Fig. 2 shows that for bright sources this ratio converges
to 1 as expected in the absence of biases or errors. On
the other hand Smeas/Strue for faint sources deviates sys-
tematically from 1. This effect is readily understood as
caused by the Eddington bias, which is the statistical
fluctuations of sources with a simulated flux below the
threshold to a flux above the detection threshold [11].
Our second check compares of the average photon index
distribution (dN/dΓ), as derived from the simulations,
with the same distribution as derived from the 2FHL
catalog. This is reported in the right panel of Fig. 2 and
it shows that our description of the γ-ray sky and of the
blazar population is faithful to the real one.
FIG. 1: In the left (right) panel the adaptively smoothed count map of one simulation (real sky) in the energy range 50 GeV-2
TeV is represented in Galactic coordinates and Hammer-Aitoff projection. The two maps contain about 60000 γ-ray events.
The results from analyzing the sources in the simu-
lated data can be used to measure the detection effi-
ciency ω(S), which is a weighting factor that takes into
account the probability to detect a source as a function
of flux. The detection efficiency is simply derived from
the simulations measuring the ratio between the number
of detected sources and the number of simulated ones
as a function of measured source flux. The result re-
ported in Fig. 3 shows that the LAT detects any source
in the |b|>10sky for fluxes larger than 2×1011 ph
cm2s1, but misses 80–90 % of the sources with fluxes
of 1×1011 ph cm2s1and many more below this
flux. The peak (ω(S)>1) clearly visible at a flux of
2×1011 ph cm2s1is due to the Eddington bias.
A reliable estimate of the detection efficiency is funda-
mental in order to correct the observed flux distribution
of the 2FHL catalog and in turn to derive the intrinsic
source count distribution, which is obtained as:
dN
dS (Si) = 1
Ω∆Si
Ni
ω(Si)[cm2s deg2],(1)
where Ω is the solid angle of the |b|>10sky, ∆Siis
the width of the flux bin, Niis the number of sources in
each flux bin and Siis the flux at the center of a given
bin i. We verified through simulations that this method
allows us to retrieve the correct source count distribution
as long as the distribution used in the simulations is a
faithful representation of the real one.
This is found to be consistent, down to the sensitivity
of the 2FHL catalog (8×1012 ph cm2s1), with a
power-law function with slope α1= 2.49 ±0.12 (see right
panel of Fig. 3). This best-fit value is consistent with
the Euclidean expectation and motivated us to choose
α1= 2.5 in the simulations.
Fig. 4 shows the cumulative source count distribution
that is defined as:
N(> S) = ZSmax
S
dN
dS0dS0[deg2],(2)
where Smax is fixed to be 108ph cm2s1.
In order to infer the shape of the dN/dS below the flux
threshold for detecting point sources we have performed
a photon fluctuation analysis. This helps us to probe the
source count distribution to the level where sources con-
tribute on average 0.5photons each. The analysis is per-
formed by comparing the histogram of the pixel counts
of the real sky with the ones obtained via Monte Carlo
simulations and allows us to constrain the slope of the
differential flux distribution below the threshold of the
survey [15, 16]. We consider a differential flux distribu-
tion described as a broken power law where the slope
above the break is α1= 2.5 as determined in this work
while below the break the slope varies in different sim-
ulations between α2[1.3,2.7]. For each value of the
slope we derive the model pixel count distribution av-
eraging over the pixel count distributions obtained from
20 simulations. The simulated and real maps have been
pixelized using the HEALPix tool 2[17]. We have used a
resolution of order 9, which translates into 3145728 pixels
and an pixel size of about 0.11. Consistent results are
obtained when using a resolution of order 8. We consider
a single energy bin from 50 GeV to 2 TeV.
The model (averaged) pixel count distributions are
compared to the real data using a χ2analysis to deter-
mine the most likely scenario. As expected, there is a
degeneracy between the best-fit value of the slope α2and
the choice of the break flux, Sb. The result of the analy-
sis is that the break flux is limited to the range between
Sb[8 ×1012,1.5×1011 ] ph cm2s1while the index
below the break is in the range α2[1.60,1.75]. The
best configuration, which we refer to as our benchmark
model, has a break flux at 1 ×1011 ph cm2s1and
a slope α2= 1.65 with a χ2= 12.4 (for 12 degrees of
freedom). This implies that the source count distribu-
tion must display a hard break |α1α2| ≈ 0.9 from the
Euclidean behavior measured at bright fluxes. We show
in Fig. 5, for the best-fit configuration, the comparison
2See http://healpix.sourceforge.net
FIG. 2: Left Panel: ratio of the measured-to-simulated source
flux (as derived from the analysis of the simulations described
in the text) as a function of simulated source flux. Right
Panel: comparison between the photon index distributions of
sources detected in 2FHL (blue points) and the average of the
simulations (red points).
between the pixel count distribution evaluated for the av-
erage of 20 simulations, and the same quantity as derived
from the real data. The figure also shows the differences
between these two distributions.
The lowest flux that the photon fluctuation analysis
is sensitive to can be estimated by adding to the source
count distribution one more break flux below that of the
benchmark model. We fixed the slope below this second
break to α3=1.80, which is at the edge of the derived
range for α2, while the break flux is varied in the range
Slim [5 ×1013,5×1012 ] ph cm2s1to register
when a worsening of the χ2(with respect to the best-fit
one) is observed. The result of this analysis is that the fit
worsened by more than 3σfor Slim &1.3×1012 ph cm2
s1. The results of the photon fluctuation analysis are
reported in Figs. 3 and 4, which show that this technique
FIG. 3: Left Panel: detection efficiency ω(S) (blue points) as
a function of source flux and normalized distribution of source
fluxes detected in 2FHL (grey shaded histogram). Right
Panel: intrinsic S2dN/dS distribution (black points). The
black solid line shows our best-fit model, while the grey and
cyan bands show the 1σand 3σuncertainty bands from the
photon fluctuation analysis. The vertical brown dotted line
represents the sensitivity of the photon fluctuation analysis.
The orange and red curves indicate where 85% and 100% of
the EGB intensity above 50 GeV [2] would be produced when
extrapolating the flux distribution below the break with dif-
ferent values of faint-end slope, α2.
allows us to measure the source count distribution over
almost three decades in flux.
We have tested also the possibility that a new source
population could emerge in the flux distribution with a
Euclidean distribution, as might be expected, for exam-
ple, from star-forming galaxies [18]. In this test we set
α3= 2.50 and follow the method described above to de-
rive the maximum flux at which a possible re-steepening
of the source counts might occur. This is found to be
Slim 7×1013 ph cm2s1and the integrated emission
of such a population would exceed at fluxes of 7×1014
ph cm2s1the totality of the EGB intensity.
FIG. 4: Cumulative source count distribution N(> S) with
the uncertainty bands as in Fig. 3 together with the theo-
retical predictions from Ref. [12] (blue dashed line), [4] (red
dashed line) and [13] (green band). The vertical dotted brown
line shows the 5 mCrab flux reachable by CTA in 240 hrs of
exposure [14].
Our best-fit model for the flux distribution dN/dS is
therefore, for S&1012 ph cm2s1, a broken power-
law with break flux in the range Sb[0.8,1.5] ×1011,
slopes above and below the break of α1= 2.49 ±0.12
and α2[1.60,1.75], respectively and a normalization
K= (4.60 ±0.35) ×1019 deg2ph1cm2s. We believe
this describes the source counts of a single population
(blazars), because no re-steepening of the source count
distribution is observed and because the large majority
(97 %) of the detected sources are likely blazars.
Fig. 4 reports the theoretical expectations for the
source count distribution given by blazars [4, 13] and BL
Lacs [12]. These models are consistent with the obser-
vations at bright fluxes, but are above the experimental
N(> S) by about a factor of 2 at S= 1012 ph cm2
s1. We include in the same figure also the predicted
5 mCrab sensitivity reachable by CTA in 240 hours in
the most sensitive pointing strategy [14]. At these fluxes
the source density is 0.0194 ±0.0044 deg2, which trans-
lates to the serendipitous detection of 200 ±45 blazars in
one quarter of the full sky. It is also interesting to note
that our analysis constrains the source count distribution
to fluxes that are much fainter than those reachable by
CTA in short exposures.
Once known, the source count distribution can be used
to estimate the contribution of point sources to the EGB.
This is performed by integrating the flux distribution
dN/dS as follows:
I=ZSmax
0
S0dN
dS0dS0[ph cm2s1sr1].(3)
Choosing Smax = 108ph cm2s1we find that the
FIG. 5: Comparison between the pixel count distribu-
tion from the average of 20 simulations (blue points), and
the distribution from the real sky (red points). The
green points show the difference between the two distribu-
tions. In each number of photon bin Nphotons ranging be-
tween [Nphoton,1, Nphoton,2] we display Npixel with Nphotons
[Nphoton,1, Nphoton,2).
total integrated flux from point sources is 2.07+0.40
0.34 ×
109ph cm2s1sr1which constitutes 86+16
14% of the
EGB above 50 GeV estimated in [2]. This validates the
predictions of models [3, 4, 12]. Point sources with fluxes
S > 1.3×1012 ph cm2s1produce 1.47+0.20
0.24 ×109ph
cm2s1sr1, while 6.0+2.0
1.0×1010 ph cm2s1sr1
is produced by sources below that flux.
The Fermi-LAT has measured the angular power spec-
trum of the diffuse γ-ray background at |b|>30and in
four energy bins spanning the 1-50 GeV energy range [19].
For multipoles l155 the angular power CPis found to
be almost constant, suggesting that the anisotropy is pro-
duced by an unclustered population of unresolved point
sources. Indeed, Refs. [20, 21, 22] argue that most of
the angular power measured by the Fermi-LAT is due to
unresolved emission of radio-loud active galactic nuclei.
The angular power due to unresolved sources at
>50 GeV can be readily predicted from the source count
distribution as:
CP=ZSmax
0
(1 ω(S0)) S02dN
dS0dS0[(ph cm2s1)2sr1],
(4)
The angular power evaluates to CP(E > 50 GeV) =
9.4+1.0
1.6×1022 (ph/cm2/s)2sr1. This is the first
observationally-based prediction of the angular power at
>50 GeV. Our estimation for CP(E > 50GeV ) is in good
agreement with the extrapolation of the Fermi-LAT an-
gular power measurements [19] above 50 GeV and is con-
sistent with the calculated anisotropy due to radio loud
active galactic nuclei made in Refs. [20, 21].
In conclusion, the Fermi-LAT collaboration has used
the new event-level analysis Pass 8 to conduct an all-sky
survey above 50 GeV. The resulting 2FHL catalog con-
tains 253 sources at |b|>10and closes the energy gap
between the LAT and Cherenkov telescopes. We have
thoroughly studied the properties of both resolved and
unresolved sources in the 50 GeV–2 TeV band using de-
tailed Monte Carlo simulations and a photon fluctuation
analysis. This allowed us to characterize, for the first
time, the source count distribution above 50GeV, which
is found to be compatible at &1012 ph cm2s1with
a broken power-law model with a break flux in the range
Sb[0.8,1.5] ×1011 ph cm2s1, and slopes above
and below the break of, respectively, α1= 2.49 ±0.12
and α2[1.60,1.75]. A photon fluctuation analysis con-
strains a possible re-steepening of the flux distribution
to a Euclidean behavior (α3= 2.50) to occur at fluxes
lower than 7×1013 ph cm2s1. Our analysis per-
mits us to estimate that point sources, and in particu-
lar blazars, explain almost the totality (86+16
14 %) of the
>50 GeV EGB.
This might have a number of important consequences,
since any other contribution, exotic or not, must nec-
essarily be small. This bound might imply strong con-
straints for the annihilation cross section or decay time of
high-mass dark matter particles producing γ-rays [3, 4].
Tight constraints could also be inferred on other γ-ray
emission mechanisms due to other diffusive processes
such as UHECRs [23, 24]. Finally, if the neutrinos
detected by IceCube have been generated in hadronic
cosmic-ray interactions, then the same sources producing
the neutrino background will produce part of the sub-
TeV γ-ray background [25]. Because blazars were found
not to be responsible for the majority of the neutrino flux
[26], the fact that the 50 GeV–2 TeV γ-ray background is
almost all due to blazars constrains the contribution of
other source classes to the neutrino background. Such
constraints will be presented in a dedicated paper.
The Fermi -LAT Collaboration acknowledges support
for LAT development, operation and data analysis
from NASA and DOE (United States), CEA/Irfu and
IN2P3/CNRS (France), ASI and INFN (Italy), MEXT,
KEK, and JAXA (Japan), and the K.A. Wallenberg
Foundation, the Swedish Research Council and the Na-
tional Space Board (Sweden). Science analysis support
in the operations phase from INAF (Italy) and CNES
(France) is also gratefully acknowledged.
Electronic address: ma jello@slac.stanford.edu
Electronic address: mattia.dimauro@to.infn.it
[1] C. E. Fichtel, R. C. Hartman, D. A. Kniffen, D. J.
Thompson, H. Ogelman, M. E. Ozel, T. Tumer, and G. F.
Bignami, ApJ 198, 163 (1975).
[2] M. Ackermann et al. (Fermi-LAT), Astrophys.J. 799, 86
(2015), 1410.3696.
[3] M. Di Mauro and F. Donato, Phys.Rev. D91, 123001
(2015), 1501.05316.
[4] M. Ajello, D. Gasparrini, M. S´anchez-Conde, G. Zahari-
jas, M. Gustafsson, et al., Astrophys.J. 800, L27 (2015),
1501.05301.
[5] C. D. Dermer, ApJ 659, 958 (2007), arXiv:astro-
ph/0605402.
[6] M. Fornasa and M. A. S´anchez-Conde (2015),
1502.02866.
[7] W. Atwood, L. Baldini, J. Bregeon, P. Bruel, A. Chekht-
man, et al., Astrophys.J. 774, 76 (2013), 1307.3037.
[8] The Fermi-LAT Collaboration, ArXiv e-prints (2015),
1508.04449.
[9] M. Ackermann et al. (Fermi-LAT), Astrophys. J. 793, 64
(2014), 1407.7905.
[10] S. Ciprini, G. Tosti, F. Marcucci, C. Cecchi, G. Discepoli,
E. Bonamente, S. Germani, D. Impiombato, P. Lubrano,
and M. Pepe, in The First GLAST Symposium, edited
by S. Ritz, P. Michelson, and C. A. Meegan (2007), vol.
921 of American Institute of Physics Conference Series,
pp. 546–547.
[11] A. S. Eddington, MNRAS 73, 359 (1913).
[12] M. Di Mauro, F. Donato, G. Lamanna, D. Sanchez, and
P. Serpico, Astrophys.J. 786, 129 (2014), 1311.5708.
[13] P. Giommi and P. Padovani, MNRAS 450, 2404 (2015),
1504.01978.
[14] G. Dubus, J. L. Contreras, S. Funk, Y. Gallant, T. Has-
san, J. Hinton, Y. Inoue, J. Kn¨odlseder, P. Martin,
N. Mirabal, et al., Astroparticle Physics 43, 317 (2013),
1208.5686.
[15] G. Hasinger, R. Burg, R. Giacconi, G. Hartner,
M. Schmidt, J. Trumper, and G. Zamorani, A&A 275, 1
(1993).
[16] D. Malyshev and D. W. Hogg, ApJ 738, 181 (2011),
1104.0010.
[17] K. M. G´orski, E. Hivon, A. J. Banday, B. D. Wandelt,
F. K. Hansen, M. Reinecke, and M. Bartelmann, ApJ
622, 759 (2005), astro-ph/0409513.
[18] M. B´ethermin, E. Daddi, G. Magdis, M. T. Sargent,
Y. Hezaveh, D. Elbaz, D. Le Borgne, J. Mullaney,
M. Pannella, V. Buat, et al., ApJL 757, L23 (2012),
1208.6512.
[19] M. Ackermann, M. Ajello, A. Albert, et al., Phy. Rev. D
85, 083007 (2012), 1202.2856.
[20] A. Cuoco, E. Komatsu, and J. Siegal-Gaskins, Phys.Rev.
D86, 063004 (2012), 1202.5309.
[21] M. Di Mauro, A. Cuoco, F. Donato, and J. M. Siegal-
Gaskins, JCAP 1411, 021 (2014), 1407.3275.
[22] A. E. Broderick, C. Pfrommer, E. Puchwein, K. M.
Smith, and P. Chang (Heidelberg Institute for Theoreti-
cal Studies, Perimeter Institute for Theoretical Physics),
Astrophys. J. 796, 12 (2014), 1308.0015.
[23] M. Ahlers and J. Salvado, Phy. Rev. D 84, 085019 (2011),
1105.5113.
[24] G. B. Gelmini, O. Kalashev, and D. V. Semikoz, JCAP
1, 044 (2012), 1107.1672.
[25] M. Aartsen et al. (IceCube), Phys.Rev.Lett. 113, 101101
(2014), 1405.5303.
[26] T. Gl¨usenkamp and for the IceCube Collaboration
(2015), 1502.03104.
... the efficiency and use it as a correction to reconstruct the true underlying dN/dS [3][4][5][6]. For all those sources which are too faint to be resolved, their cumulative distribution of photons in the sky defines an almost isotropic cosmic field, conventionally called the isotropic diffuse gamma-ray background (IGRB) or, more precisely, the unresolved gamma-ray background (UGRB). ...
... We thank M. Negro for valuable insight on Fermi-LAT data analysis and data reduction and for the numerical tool Xgam. 4 We thank G. Zaharijas for serving as LAT internal referee and providing useful comments on the manuscript. We acknowledge support from: Additional support for science analysis during the operations phase is gratefully acknowledged from the Istituto Nazionale di Astrofisica in Italy and the Centre National d'Études Spatiales in France. ...
Article
Full-text available
We reconstruct the extra-galactic gamma-ray source-count distribution, or dN/dS , of resolved and unresolved sources by adopting machine learning techniques. Specifically, we train a convolutional neural network on synthetic 2-dimensional sky-maps, which are built by varying parameters of underlying source-counts models and incorporate the Fermi -LAT instrumental response functions. The trained neural network is then applied to the Fermi -LAT data, from which we estimate the source count distribution down to flux levels a factor of 50 below the Fermi -LAT threshold. We perform our analysis using 14 years of data collected in the (1,10) GeV energy range. The results we obtain show a source count distribution which, in the resolved regime, is in excellent agreement with the one derived from cataloged sources, and then extends as dN/dS ∼ S ⁻² in the unresolved regime, down to fluxes of 5 · 10 ⁻¹² cm ⁻² s ⁻¹ . The neural network architecture and the devised methodology have the flexibility to enable future analyses to study the energy dependence of the source-count distribution.
... The interactions of these CRs produce γ-rays and neutrinos (e.g., [566,567]) which contribute to the extragalactic diffuse backgrounds. These backgrounds have been measured with IceCube [568][569][570] and Fermi-LAT [571]. LIRGs have recently emerged as a viable contributor to these backgrounds following a neutrino detection from the LIRG NGC 1068 [381] (see also Section 4.1). ...
... Combined multi-messenger observations of neutrinos and γ-rays suggest that some neutrino sources may be opaque to γ-rays [572,573]. If true, it would alleviate the observational constraints on source populations that can generate neutrino backgrounds without violating the observed γ-ray background [571]. Dusty LIRGs may experience suppressed γ-ray emission through pair production interactions within their intense IR radiation fields (e.g., [249]) or dense gas [574]. ...
Article
Full-text available
Galaxy evolution is an important topic, and our physical understanding must be complete to establish a correct picture. This includes a thorough treatment of feedback. The effects of thermal–mechanical and radiative feedback have been widely considered; however, cosmic rays (CRs) are also powerful energy carriers in galactic ecosystems. Resolving the capability of CRs to operate as a feedback agent is therefore essential to advance our understanding of the processes regulating galaxies. The effects of CRs are yet to be fully understood, and their complex multi-channel feedback mechanisms operating across the hierarchy of galaxy structures pose a significant technical challenge. This review examines the role of CRs in galaxies, from the scale of molecular clouds to the circumgalactic medium. An overview of their interaction processes, their implications for galaxy evolution, and their observable signatures is provided and their capability to modify the thermal and hydrodynamic configuration of galactic ecosystems is discussed. We present recent advancements in our understanding of CR processes and interpretation of their signatures, and highlight where technical challenges and unresolved questions persist. We discuss how these may be addressed with upcoming opportunities.
... Since the launch of the Fermi-Large Area Telescope (Fermi-LAT) in 2008, high-energy astrophysics has undergone a transformative Fermi era marked by profound discoveries (Abdo, Ackermann, Ajello, Allafort, Antolini, Atwood, Axelsson, Baldini, Ballet, Barbiellini, Bastieri, Baughman, Bechtol, Bellazzini, Belli, et al. 2010;Abdo, Ackermann, Ajello, Allafort, Antolini, Atwood, Axelsson, Baldini, Ballet, Barbiellini, Bastieri, Baughman, Bechtol, Bellazzini, Berenji, et al. 2010;. Though nearly 20% LSPs were out of detection, it was found that the diffuse extra-galactic γ-ray background is significantly dominated by emission from blazars Ackermann et al. 2016;Arsioli and Polenta 2018). However, the exact location of γ-ray emission region is still on debate (Agudo et al. 2012;W. ...
Preprint
Full-text available
The discovery that blazars dominate the extra-galactic {\gamma}-ray sky is a triumph in the Fermi era. However, the exact location of {\gamma}-ray emission region still remains in debate. Low-synchrotron-peaked blazars (LSPs) are estimated to produce high-energy radiation through the external Compton process, thus their emission regions are closely related to the external photon fields. We employed the seed factor approach proposed by Georganopoulos et al. It directly matches the observed seed factor of each LSP with the characteristic seed factors of external photon fields to locate the {\gamma}-ray emission region. A sample of 1138 LSPs with peak frequencies and peak luminosities was adopted to plot a histogram distribution of observed seed factors. We also collected some spectral energy distributions (SEDs) of historical flare states to investigate the variation of {\gamma}-ray emission region. Those SEDs were fitted by both quadratic and cubic functions using the Markov-chain Monte Carlo method. Furthermore, we derived some physical parameters of blazars and compared them with the constraint of internal {\gamma}{\gamma}-absorption. We find that dusty torus dominates the soft photon fields of LSPs and most {\gamma}-ray emission regions of LSPs are located at 1-10 pc. The soft photon fields could also transition from dusty torus to broad line region and cosmic microwave background in different flare states. Our results suggest that the cubic function is better than the quadratic function to fit the SEDs.
... This implies that the afterglow contribution can be compared to the prompt phase in several GeV bands, but it still holds a nondominant position. Considering suggestions that blazars (Ackermann et al. 2016) or starforming galaxies (Roth et al. 2021) dominate the IGRB, GRB afterglows may occupy a considerable fraction of the rest part of the IGRB. If the major component of the IGRB from blazars or star-forming galaxies can be accurately determined, the rest part of the IGRB may then put useful constraints on the properties of the GRB afterglow in turn. ...
Article
Full-text available
The isotropic diffuse γ -ray background (IGRB) serves as a fundamental probe of the evolution of the extreme Universe. Although various astrophysical sources have been proposed as potential contributors to the IGRB, the dominant population is still under debate. γ -ray bursts (GRBs) are among candidate contributors of IGRB, although they are not as frequently discussed as blazars or starburst galaxies. Recent observations of TeV emission from GRB afterglows have provided fresh insights into this subject. This work aims to investigate the potential contribution of GRB afterglows to the IGRB under the standard afterglow model. We carefully examine the influence of each microphysical parameter of the afterglow model on this contribution, with a particular emphasis on the significant role played by the initial kinematic energy. To determine the energy and quantify the contribution of GRB afterglow to IGRBs, we utilize the observed GRB afterglow energy emissions from the Swift X-ray Telescope and Fermi Large Area Telescope instruments. Our calculations, considering the synchrotron self-Compton emission, suggest that GRB afterglows make up less than 10% of the IGRBs. To enhance the precision of our findings, it is crucial to further constrain these parameters by conducting additional ground-based observations of GRB afterglows.
... to gamma rays. Unsurprisingly, given that most extragalactic Fermi LAT sources are blazars, also the IGRB above a few tens of GeV is dominated by emission from unresolved blazars [29]. Evidence for TeV neutrino emission from blazars has also been found, in particular in relation to the source TXS 0506 + 056 [30,31]. ...
Article
Full-text available
Many instruments for astroparticle physics are primarily geared towards multi-messenger astrophysics, to study the origin of cosmic rays and to understand high-energy astrophysical processes. Since these instruments observe the Universe at extreme energies and in kinematic ranges not accessible at accelerators these experiments provide also unique and complementary opportunities to search for particles and physics beyond the standard model of particle physics. In particular, the reach of IceCube, Fermi and KATRIN to search for and constrain Dark Matter, Axions, heavy Big Bang relics, sterile neutrinos and Lorentz invariance violation will be discussed. The contents of this article are based on material presented at the Humboldt–Kolleg ‘Clues to a mysterious Universe—exploring the interface of particle, gravity and quantum physics’ in June 2022. This article is part of the theme issue ‘The particle-gravity frontier’.
Article
The astrophysical γ -ray photons carry the signatures of the violent phenomena happening on various astronomical scales in our Universe. This includes supernova remnants, pulsars, and pulsar wind nebulae in the Galactic environment and extragalactic relativistic jets associated with active galactic nuclei (AGN). However, ∼30% of the γ -ray sources detected with the Fermi Large Area Telescope lack multiwavelength counterpart association, precluding us from characterizing their origin. Here we report, for the first time, the association of a collisional ring galaxy system in our Galactic neighborhood (distance ∼10 Mpc), formed as a consequence of a smaller “bullet” galaxy piercing through a larger galaxy, as the multifrequency counterpart of an unassociated γ -ray source 4FGL J1647.5−5724. The system, also known as “ Kathryn ’ s Wheel ,” contains two dwarf irregular galaxies and an edge-on, late-type, spiral galaxy surrounded by a ring of star-forming knots. We utilized observations taken from the Neil Gehrels Swift observatory, Rapid ASKAP Continuum Survey, SuperCOSMOS H α Survey, Dark Energy Survey, and VIsible MultiObject Spectrograph at Very Large Telescope to ascertain the association with 4FGL J1647.5−5724 and to explore the connection between the star-forming activities and the observed γ -ray emission. We found that star formation alone cannot explain the observed γ -ray emission, and additional contribution likely from the pulsars/supernova remnants or buried AGN is required. We conclude that arcsecond/subarcsecond-scale observations of this extraordinary γ -ray-emitting galaxy collision will be needed to resolve the environment and explore the origin of cosmic rays.
Article
Understanding the origin of the diffuse flux of high-energy astrophysical neutrinos detected by IceCube has become a challenging issue within present High Energy Astrophysics. In this work, we present a model to explore the potential neutrino emission of starburst galaxies (SBG) by considering three different neutrino production zones that can be associated to a typical single SBG. The first zone is the starburst nucleus, where due to the high rate of supernova explosions, a significant amount of protons can be accelerated to high energies and undergo pp interactions with cold protons of the interstellar medium. The second zone we consider is the corresponding to the starburst wind, which is formed by the hot gas that emerges from the nucleus and interacts with the intergalactic medium generating shocks. Protons accelerated there can undergo pp interactions with the ambient matter. The third neutrino production zone we consider, is an external one, where we account for the possibility that protons escaping from the whole system interact with the cosmic microwave background. Finally, adding the neutrino contributions of the three zones, we calculate the diffuse neutrino flux and the diffuse photon flux by integration on the redshift range appropriate for SBG. We find that the model behaves well applied to nearby galaxies such as M82 and NGC 253. The contributions made to the diffuse neutrino flux are able to explain part of the data provided by IceCube if typical parameters are considered.
Article
Full-text available
Understanding the physical mechanisms that control galaxy formation is a fundamental challenge in contemporary astrophysics. Recent advances in the field of astrophysical feedback strongly suggest that cosmic rays (CRs) may be crucially important for our understanding of cosmological galaxy formation and evolution. The appealing features of CRs are their relatively long cooling times and relatively strong dynamical coupling to the gas. In galaxies, CRs can be close to equipartition with the thermal, magnetic, and turbulent energy density in the interstellar medium, and can be dynamically very important in driving large-scale galactic winds. Similarly, CRs may provide a significant contribution to the pressure in the circumgalactic medium. In galaxy clusters, CRs may play a key role in addressing the classic cooling flow problem by facilitating efficient heating of the intracluster medium and preventing excessive star formation. Overall, the underlying physics of CR interactions with plasmas exhibit broad parallels across the entire range of scales characteristic of the interstellar, circumgalactic, and intracluster media. Here we present a review of the state-of-the-art of this field and provide a pedagogical introduction to cosmic ray plasma physics, including the physics of wave–particle interactions, acceleration processes, CR spatial and spectral transport, and important cooling processes. The field is ripe for discovery and will remain the subject of intense theoretical, computational, and observational research over the next decade with profound implications for the interpretation of the observations of stellar and supermassive black hole feedback spanning the entire width of the electromagnetic spectrum and multi-messenger data.
Article
We present a phenomenological framework for starburst-driven neutrino production via proton-proton collisions and apply it to (ultra)luminous infrared galaxies (U/LIRGs) in the Great Observatories All-Sky LIRG Survey (GOALS). The framework relates the infrared luminosity of a GOALS galaxy, derived from consistently available Herschel Space Observatory data, to the expected starburst-driven neutrino flux. The model parameters that define this relation can be estimated from multiwavelength data. We apply the framework in a case study to the LIRG NGC 3690 (Arp 299, Mrk 171) and compare the obtained neutrino fluxes to the current sensitivity of the IceCube Neutrino Observatory. Using our framework, we conclude that the neutrino emission in the LIRG NGC 1068, recently presented as the first steady IceCube neutrino point source, cannot be explained by a starburst-driven scenario and is therefore likely dominated by the active galactic nucleus in this galaxy. In addition to the single-source investigations, we also estimate the diffuse starburst-driven neutrino flux from GOALS galaxies and the total LIRG population over cosmic history.
Article
Full-text available
The origin of the extragalactic $\gamma$-ray background (EGB) has been debated for some time. { The EGB comprises the $\gamma$-ray emission from resolved and unresolved extragalactic sources, such as blazars, star-forming galaxies and radio galaxies, as well as radiation from truly diffuse processes.} This letter focuses on the blazar source class, the most numerous detected population, and presents an updated luminosity function and spectral energy distribution model consistent with the blazar observations performed by the {\it Fermi} Large Area Telescope (LAT). We show that blazars account for 50$^{+12}_{-11}$\,\% of the EGB photons ($>$0.1\,GeV), and that {\it Fermi}-LAT has already resolved $\sim$70\,\% of this contribution. Blazars, and in particular low-luminosity hard-spectrum nearby sources like BL Lacs, are responsible for most of the EGB emission above 100\,GeV. We find that the extragalactic background light, which attenuates blazars' high-energy emission, is responsible for the high-energy cut-off observed in the EGB spectrum. Finally, we show that blazars, star-forming galaxies and radio galaxies can naturally account for the amplitude and spectral shape of the background in the 0.1--820\,GeV range, leaving only modest room for other contributions. This allows us to set competitive constraints on the dark-matter annihilation cross section.
Article
Full-text available
The γ-ray sky can be decomposed into individually detected sources, diffuse emis- sion attributed to the interactions of Galactic cosmic rays with gas and radiation fields, and a residual all-sky emission component commonly called the isotropic diffuse γ-ray background (IGRB). The IGRB comprises all extragalactic emissions too faint or too diffuse to be resolved in a given survey, as well as any residual Galactic foregrounds that are approximately isotropic. The first IGRB measurement with the Large Area Tele- scope (LAT) on board the Fermi Gamma-ray Space Telescope (Fermi) used 10 months of sky-survey data and considered an energy range between 200 MeV and 100 GeV. Im- provements in event selection and characterization of cosmic-ray backgrounds, better understanding of the diffuse Galactic emission, and a longer data accumulation of 50 months, allow for a refinement and extension of the IGRB measurement with the LAT, now covering the energy range from 100 MeV to 820 GeV. The IGRB spectrum shows a significant high-energy cutoff feature, and can be well described over nearly four decades in energy by a power law with exponential cutoff having a spectral index of 2.32 ± 0.02 and a break energy of (279±52) GeV using our baseline diffuse Galactic emission model. The total intensity attributed to the IGRB is (7.2 ± 0.6) × 10−6 cm−2 s−1 sr−1 above 100 MeV, with an additional +15%/−30% systematic uncertainty due to the Galactic diffuse foregrounds.
Article
Full-text available
A search for high-energy neutrinos interacting within the IceCube detector between 2010 and 2012 provided the first evidence for a high-energy neutrino flux of extraterrestrial origin. Results from an analysis using the same methods with a third year (2012–2013) of data from the complete IceCube detector are consistent with the previously reported astrophysical flux in the 100 TeV–PeV range at the level of 10−8 GeV cm−2 s−1 sr−1 per flavor and reject a purely atmospheric explanation for the combined three-year data at 5.7σ. The data are consistent with expectations for equal fluxes of all three neutrino flavors and with isotropic arrival directions, suggesting either numerous or spatially extended sources. The three-year data set, with a live time of 988 days, contains a total of 37 neutrino candidate events with deposited energies ranging from 30 to 2000 TeV. The 2000-TeV event is the highest-energy neutrino interaction ever observed.
Article
Full-text available
The Fermi bubbles are two large structures in the gamma-ray sky extending to $55^\circ$ above and below the Galactic center. We analyze 50 months of Fermi Large Area Telescope data between 100 MeV and 500 GeV above $10^\circ$ in Galactic latitude to derive the spectrum and morphology of the Fermi bubbles. We thoroughly explore the systematic uncertainties that arise when modeling the Galactic diffuse emission through two separate approaches. The gamma-ray spectrum is well described by either a log parabola or a power law with an exponential cutoff. We exclude a simple power law with more than 7$\sigma$ significance. The power law with an exponential cutoff has an index of $1.9 \pm 0.2$ and a cutoff energy of $110\pm 50$ GeV. We find that the gamma-ray luminosity of the bubbles is $4.4^{+2.4}_{-0.9} \times 10^{37}$ erg s$^{-1}$. We confirm a significant enhancement of gamma-ray emission in the south-eastern part of the bubbles, but we do not find significant evidence for a jet. No significant variation of the spectrum across the bubbles is detected. The width of the boundary of the bubbles is estimated to be $3.4^{+3.7}_{-2.6}$ deg. Both inverse Compton (IC) models and hadronic models including IC emission from secondary leptons fit the gamma-ray data well. In the IC scenario, the synchrotron emission from the same population of electrons can also explain the WMAP and Planck microwave haze with a magnetic field between 5 and 20 $\mu$G.
Article
Full-text available
Based on the experience gained during the four and a half years of the mission, the Fermi -LAT collaboration has undertaken a comprehensive revision of the event-level analysis going under the name of Pass 8. Although it is not yet finalized, we can test the improvements in the new event reconstruction with the special case of the prompt phase of bright Gamma-Ray Bursts (GRBs), where the signal to noise ratio is large enough that loose selection cuts are sufficient to identify gamma- rays associated with the source. Using the new event reconstruction, we have re-analyzed ten GRBs previously detected by the LAT for which an x-ray/optical follow-up was possible and found four new gamma rays with energies greater than 10 GeV in addition to the seven previously known. Among these four is a 27.4 GeV gamma-ray from GRB 080916C, which has a redshift of 4.35, thus making it the gamma ray with the highest intrinsic energy (147 GeV) detected from a GRB. We present here the salient aspects of the new event reconstruction and discuss the scientific implications of these new high-energy gamma rays, such as constraining extragalactic background light models, Lorentz invariance violation (LIV) tests, the prompt emission mechanism and the bulk Lorentz factor of the emitting region.
Article
The blazar-simplified view is a new paradigm that explains well the diverse statistical properties of blazars observed over the entire electromagnetic spectrum on the basis of minimal assumptions on blazars' physical and geometrical properties. In this paper, the fourth in a series, we extend the predictions of this paradigm below the sensitivity of existing surveys and estimate the contribution of blazars to the X-ray and γ-ray extragalactic backgrounds. We find that the integrated light from blazars can explain up to 100 per cent of the cosmic background at energies larger than ̃10 GeV, and contribute ≈40-70 per cent of the γ-ray diffuse radiation between 100 MeV and 10 GeV. The contribution of blazars to the X-ray background, between 1 and 50 keV, is approximately constant and of the order of 4-5 per cent. On the basis of an interpolation between the estimated flux at X-ray and γ-ray energies, we can expect that the contribution of blazars raises to ̃10 per cent at 100 keV, and continues to increase with energy until it becomes the dominant component at a few MeV. Finally, we show that a strong dependence of the synchrotron peak frequency on luminosity, as postulated by the blazar sequence, is ruled out by the observational data as it predicts a γ-ray background above a few GeV that is far in excess of the observed value.
Article
A new estimation of the isotropic diffuse gamma-ray background (IGRB) observed by the Large Area Telescope (LAT) on board the Fermi Gamma-ray Space Telescope (Fermi) has been presented for 50 months of data, in the energy range 100 MeV-820 GeV and for different modelings of the Galactic foreground. We attempt here the interpretation of the Fermi-LAT IGRB data in terms of the gamma-ray unresolved emission from different extragalactic populations. We find very good fits to the experimental IGRB, obtained with theoretical predictions for the emission from active galactic nuclei and star forming galaxies. In addition, we probe a possible emission coming from the annihilation of weakly interacting dark matter (DM) particles in the halo of our Galaxy. We set stringent limits on its annihilation cross section into gamma-rays, which are about the thermal relic value for a wide range of DM masses. We also identify regions in the DM mass and annihilation cross section parameter space which can significantly improve the fit to the IGRB data. Our analysis is conducted within the different IGRB data sets obtained from different models for the Galactic emission, which is shown to add a significant ambiguity on the IGRB interpretation.
Article
Fermi has been instrumental in constraining the luminosity function and redshift evolution of gamma-ray bright blazars. This includes limits upon the spectrum and anisotropy of the extragalactic gamma-ray background (EGRB), redshift distribution of nearby Fermi active galactic nuclei (AGN), and the construction of a log(N)-log(S) relation. Based upon these, it has been argued that the evolution of the gamma-ray bright blazar population must be much less dramatic than that of other AGN. However, critical to such claims is the assumption that inverse Compton cascades reprocess emission above a TeV into the Fermi energy range, substantially enhancing the strength of the observed limits. Here we demonstrate that in the absence of such a process, due, e.g., to the presence of virulent plasma beam instabilities that preempt the cascade, a population of TeV-bright blazars that evolve similarly to quasars is consistent with the population of hard gamma-ray blazars observed by Fermi. Specifically, we show that a simple model for the properties and luminosity function is simultaneously able to reproduce their log(N)-log(S) relation, local redshift distribution, and contribution to the EGRB and its anisotropy without any free parameters. Insofar the naturalness of a picture in which the hard gamma-ray blazar population exhibits the strong redshift evolution observed in other tracers of the cosmological history of accretion onto halos is desirable, this lends support for the absence of the inverse Compton cascades and the existence of the beam plasma instabilities.
Article
In principle, the angular anisotropy in the extragalactic gamma-ray background (EGRB) places severe constraints upon putative populations of unresolved gamma-ray point sources. Existing estimates of the EGRB anisotropy have been constructed by excising known point sources, e.g., taken from the First or 2 Year Fermi-LAT Source Catalog (1FGL or 2FGL, respectively) and statistically analyzing the residual gamma-ray sky maps. We perform an independent check of the EGRB anisotropy limits by comparing the values obtained from the 1FGL-masked sky maps to the signal implied by sources that lie below the 1FGL detection threshold in the more-sensitive 2FGL and 1FHL (First Fermi-LAT catalog of >10 GeV sources). Based upon this, we find evidence for substantially larger anisotropies than those previously reported at energies above 5 GeV, where BL Lacs are likely to provide the bulk of their contribution to the EGRB. This uncertainty in the EGRB anisotropy cautions against using it as an independent constraint for the high-redshift gamma-ray Universe. Moreover, this would suggest that contrary to previous claims, smooth extensions of the resolved point source population may be able to simultaneously explain both the isotropic and anisotropic components of the EGRB.
Article
A study of the statistics of cosmological black hole jet sources is applied to EGRET blazar data and predictions are made for GLAST. Black hole jet sources are modeled as collimated relativistic plasma outflows with radiation beamed along the jet axis due to strong Doppler boosting. The comoving rate density of blazar flares is assumed to follow a blazar formation rate (BFR), modeled by analytic functions based on astronomical observations and fits to EGRET data. The redshift and size distributions of γ-ray blazars observed with EGRET, separated into BL Lac objects (BLs) and flat spectrum radio quasar (FSRQ) distributions, are fit with monoparametric functions for the distributions of the jet Lorentz factor Γ, comoving directional power l, and spectral slope. A BFR factor ≈10× greater at z 1 than at present is found to fit the FSRQ data. A smaller comoving rate density and greater luminosity of BL flares at early times compared to the present epoch fits the BL data. Based on the EGRET observations, ≈1000 blazars consisting of ≈800 FSRQs and FR 2 radio galaxies and ≈200 BL Lac objects and FR 1 radio galaxies will be detected with GLAST during the first year of the mission. Additional AGN classes, such as hard-spectrum BL Lac objects that were mostly missed with EGRET, could add more GLAST sources. The FSRQ and BL contributions to the EGRET γ-ray background at 1 GeV are estimated at the level of ≈10%-15% and ≈2%-4%, respectively. EGRET and GLAST sensitivities to blazar flares are considered in the optimal case, and a GLAST analysis method for blazar detection is outlined.