Open access peer-reviewed chapter

Extragalactic Gamma‐Ray Background

Written By

Houdun Zeng and Li Zhang

Reviewed: 22 December 2016 Published: 24 May 2017

DOI: 10.5772/67335

From the Edited Volume

New Insights on Gamma Rays

Edited by Ahmed M. Maghraby

Chapter metrics overview

1,707 Chapter Downloads

View Full Metrics

Abstract

The origin of the extragalactic gamma‐ray background (EGRB) is an important open issue in the gamma‐ray astronomy. There are many theories about the origin of EGRB: (1) some truly diffuse processes, such as dark matter (DM) annihilation or decay, which can produce gamma rays; (2) gamma rays produced by energetic particles accelerated through induced shock waves during structure formation of the universe; (3) a lot of unidentified sources, including normal galaxies, starbursts and active galactic nuclei (AGNs), contain a large number of energetic particles and can emit gamma rays. Among various extragalactic sources, blazars including flat spectral radio quasars (FSRQs) and BL Lac objects are one of the most possible sources for EGRB. As continuous accumulation of the data observed by the Fermi Gamma‐Ray Space Telescope, it is possible to directly construct gamma‐ray luminosity function (GLF) of the blazars involving evolution information. In this chapter, based on the largest clean sample of AGNs provided by Fermi Large Area Telescope (LAT), we mainly study blazar's GLFs and their contribution to EGRB. In our study, we separately construct GLFs of FSRQs and BL Lacs and then estimate the contributions to EGRB, respectively. Further, we discuss the diffuse gamma ray from other astrophysical sources and the other possible origins of the EGRB.

Keywords

  • blazars
  • gamma‐ray radiation
  • luminosity function
  • the extragalactic gamma‐rays background

1. Introduction

The large area telescope (LAT [1]) onboard Fermi gamma‐ray space telescope (Fermi) has measured the extragalactic diffuse gamma‐ray background and then provided useful information for us to study the origins of the extragalactic gamma‐ray background (EGRB) [25]. However, the origin of the EGRB is still an unsolved problem. Observationally, an isotropic component of the EGRB emission was first detected by the SAS‐2 satellite [6, 7] and subsequently measured by the energetic gamma‐ray experiment telescope (EGRET) [810]. Due to the higher sensitivity of Fermi‐LAT than that of EGRET, the observed integrated flux above 100 MeV by the LAT is (1.03±0.17)×105 photons cm2s1 [3], which is lower than (1.14±0.05)×105 photons cm2s1 measured by EGRET [11]. Recently, Fermi‐LAT has made a new measurement of the EGRB spectrum and their results shown that the EGRB energy spectrum between 0.1 and 820 GeV is to be well represented by a power law with an exponential cutoff above 300 GeV [5]. Figure 1 (left panel) shows the measured X‐ray and gamma‐ray background radiation spectra. We know that the X‐ray background spectrum has no big change with time and has been considered as the integrated light produced via the accretion process of active galactic nuclei (AGNs) [12]. However, the gamma‐ray spectrum is different from the X‐ray background spectrum due to the sensitivity of an instrument and other reasons. Before the Fermi gamma‐ray space telescope era, neither spectrum nor origin of the EGRB was well understood. In particular, the spectrum at 0.03–50 GeV reported by EGRET has a break in the several GeV. With the arrival of Fermi era, more accurate determination of the EGRB spectrum and more extragalactic source samples are provided to understand the nature of the EGRB. Note that the whole gamma‐ray sky contains diffuse galactic emission, point sources, isotropic extragalactic diffuse emission and local and solar diffuse emissions. Figure 1 (right panel) shows that the EGRB spectrum is obtained by removing the resolved point source, like as the most recent list of resolved Fermi‐LAT source (3FGL), the diffuse galactic emission determined by GALPROP, which simulates both cosmic‐ray propagation in the galaxy and the gamma‐ray flux resulting from interactions and possibly an isotropic flux of galactic, by restricting data to regions with |b| > 10° or even higher galactic latitudes.

Figure 1.

Left: The measured X‐ray and gamma‐ray background radiation spectra, which is from Ref. [5]. Right: The composition of the total gamma‐ray flux. The figure is obtained from the report of Ackermann, M. at Fermi Symposium.

Similar to the extragalactic EGRET sky, blazars are the largest source class identified by Fermi extragalactic sky and their contribution to the EGRB has been widely discussed. Typical estimated contributions of unresolved blazars to the EGRB range from 10 to 100% [1336]. Blazars are divided into two main subgroups: BL Lac objects and FSRQs [37]. Among the gamma‐ray blazar sample, the number of FSRQs detected by Fermi‐LAT is smaller than that of BL Lac objects (e.g., 2FGL, 3FGL). FSRQs generally show softer spectrum in the gamma‐ray band (e.g., [38]), which is to be detected harder than BL Lac objects at a given significant limit. On the one hand, BL Lacs are reputed as the population of extragalactic sources that show a negative or no cosmological evolution [3942], but FSRQs are regarded as those with a positive cosmological evolution, which is similar to the population of X‐ray‐selected, radio‐quiet AGNs [4345]. Ajello et al. [32] suggested that BL Lacs have a more complex evolution. At the modest redshift region, most BL Lac classes show a positive evolution with a space density peaking. Meanwhile, their results suggest that the evolution of low‐luminosity, high‐synchrotron‐peaked (HSP) BL Lac objects is strong negative with number density increasing for low redshift range (z0.5). In addition, the contributions of the EGRB from other sources or processes are very important. Those are star‐forming galaxies [46, 47], radio galaxies (e.g., [14, 46, 48]), gamma‐ray bursts (GRBs) (e.g., [49]), high galactic‐latitude pulsars (e.g., [50]), intergalactic shocks (e.g., [51, 52]), Seyferts (e.g., [53]), cascade from ultra‐high‐energy cosmic rays (e.g., [54, 55]), large galactic electron halo [56], cosmic‐ray interaction in the solar system [57] and dark matter annihilation or decay (e.g., [58]). Recently, with the assumptions and uncertainties, Ajello et al. [33] and Di Mauro and Donato [36] shown that the EGRB can be fully accounted for the sum of contributions from undetected sources including blazars and radio and star‐forming galaxies. Those results imply that little room in space is left for other processes such as shock wave or DM interactions (e.g., [33, 59]).

The extragalactic gamma‐ray sky provides an amount of gamma‐ray sources and allows us to obtain the information about the evolution of sources and estimate their contributions to the EGRB. Because the blazar's contribution is the main content of research on this chapter, the detail about how to build the gamma‐ray luminosity function (GLF) will be discussed in Section 2. In Section 3, a brief description about how to estimate different components’ contributions to the EGRB is given and finally, we give the conclusions and discussions in Section 4.

Advertisement

2. The gamma‐ray luminosity function

Since the Fermi‐LAT has detected and identified more and more gamma‐ray sources and observed previously detected objects in greater detail, the method by using the gamma‐ray luminosity function (GLF) to estimate the EGRB of resolved sources has become much more reliable. In this approach, the GLF involving the evolution of redshift as well as the distribution of spectral indices of a given source class can be established for all known sources and the observed population can be extrapolated to lower fluxes.

2.1. Function derivation

As professed in Ref. [31], there is a classical approach to obtain the luminosity function, which is on account of 1/VMAX method provided by Schmidt [60] to deal with redshift bins. However, this method has a fault, which is known to introduce bias in each binning. For a small sample and/or a large span of parameters, if the bins contained significant evolution, the method would result in a loss of important information. In order to constrain the model parameters for various models of the evolving GLF, a maximum likelihood method is adopted, which is first introduced by Marshall et al. [61]. The likelihood function L is given as follows (e.g., [17, 19, 24, 62]):

L=exp(Nexp)i=1 NobsΦ(Lγ,i, zi, Γi),E1

where Nexp is the expected number of source detections:

Nexp=dΓdzdLγΦ(Lγ,i,zi,Γi), Nexp is the number of the sample of sources and Φ(Lγ,i,zi,Γi) is the distribution function of the space density of source on luminosity (Lγ), redshift (z) and photon index (Γ). The function form can be expressed as follows:

Φ(Lγ,i, zi, Γi)=d3NdLγdzdΓ=ργ(Lγ, z)×dNdΓ×dVdz×ω(Lγ, z, Γ),E2

where ργ(Lγ,z) is the γ‐ray luminosity function and dV/dz is the comoving volume element per unit redshift and unit solid angle:

dV/dz=cdL2/(H0(1+z)2ΩM(1+z)3+ΩΛ). dN/dΓ is the intrinsic photon index distribution assumed as a Gaussian exp((Γμ)2/2σ2), where μ and σ are the mean and the dispersion, respectively. ω(Lγ,z,Γ) is the detection efficiency and represents the probability of detecting an object with the γ‐ray luminosity Lγ at redshift z and photon index Γ [1, 24, 31]. The relationship between χ2 and likelihood (L) can be expressed by function χ2=2 ln (L) [63]. In this case, the function χ2=2lnL that is minimized is defined as follows:

χ2=2iNobsln(Φ(Lγ,i,zi,Γi))+2Nexp.E3

For a given GLF, the redshift distribution, luminosity distribution and photon index distribution can be divided into three intervals of size dLγdzdΓ and the three kinds of differential distributions can be expressed from GLF as follows [31]:

dNdz=ΓminΓmaxLγ,minLγ,maxd3NdLγdzdΓdLγdz,dNdz=ΓminΓmaxzminzmaxd3NdLγdzdΓdzdΓ,dNdz=ΓminΓmaxLγ,minLγ,maxd3NdLγdzdΓdLγdz,E4

The source count distribution can be derived as follows:

(>S)=ΓminΓmaxdΓzminzmaxdzLγ(z,S)Lγ, maxdLγ d3NdLγdzdΓ= ΓminΓmaxdNdΓdΓzminzmaxdVdzLγ(z,S)Lγ, maxργ(Lγ,z)ω(Lγ, z, Γ)dLγ E5

where Lγ(z,S) is the luminosity of a source at redshift z with a flux of Sγ (>100 MeV).

Through minimized Eq. (3), we can obtain the best‐fitting parameters of the models. There are multiple parameters in our various models to find the best in observational data in a multidimensional model parameter space; the MCMC technique can be employed for its high efficiency to constrain the model parameters. In this method, the Metropolis‐Hastings algorithm that generates samples from the posterior distribution using a Markov Chain is used when sampling the model parameters and the probability density distributions of the model parameters are asymptotically proportional to the number density of the sample points. For each parameter set P, one obtains the likelihood function L(P)exp(χ2(P)/2), where χ2 is obtained by comparing model predictions with observations. A new set of parameter P' is adopted to replace the existing one P with a probability of min {1,L(P)/L(P)}. The MCMC method has been reviewed by Fan et al.[64] and described in detail by Neal [65], Gamerman [66], Lewis and Bridle [67], Mackay [68].

2.2. Models description

The GLF models for different source classes are uncertainty. Currently, there are two methods for constructing the blazars’ GLF: the first method is to build the GLF by assuming a relationship between the GLF and the luminosity function in a lower energy band, for example, that the GLF relates to radio luminosity function (RLF) or to the X‐ray luminosity function (XLF) (e.g., [14, 16, 17, 19, 23, 28, 48, 6972]); the second method is to construct the GLF directly using observed gamma‐ray data of blazars (e.g., [15, 17, 22]). Before the Fermi era, constructing the GLF model indirectly was used more frequently due to the small EGRET samples, which results in blazar's contribution between the range of 10 and 100%. In next sections, we briefly review those models for directly constructing the GLF.

2.2.1. The pure density evolution

The pure density evolution (PDE) model is the simplest scenario of evolution and the GLF has a following form:

ρ(Lγ,z)=Aln(10)Lγ[(LγL)γ1+(LγL)γ2]1×e(z),E6

where e(z)=(1+z)κ is the standard power‐law evolutionary factor. In this case, there are five model parameters and other two parameters, μ and σ, are also added.

2.2.2. The pure luminosity evolution

In the pure luminosity evolution (PLE) model, the GLF can be expressed as follows:

ρ(Lγ, z)= A(1+z)κez/ξln(10)Lγ [(LγL*(1+z)κez/ξ)γ1+(LγL*(1+z)κez/ξ)γ2]1,E7

where A is a normalization factor, Li is the evolving break luminosity, γ1 is the faint‐end slope index, γ2 is the bright‐end slope index, κ and ξ represent the redshift evolution. Including the parameters μ and σ, there are 8 parameters in calculations.

2.2.3. The luminosity‐dependent density evolution

In the luminosity‐dependent density evolution (LDDE) model, the GLF evolution is decided by a redshift cutoff that depends on luminosity and the GLF can be given by

ρ(Lγ, z)= Aln(10)Lγ [(LγL*)γ1+(LγL*)γ2]1[(1+z1+zc*(Lγ/1048)α)p1+ (1+z1+zc*(Lγ/1048)α)p2],E8

where A is a normalization factor, L is evolving break luminosity, γ1 and p1 are the faint‐end slope index, γ2 and p2 are the bright‐end slope index, zc is redshift peak with a luminosity (here 1048 ergs s1) and α is power‐law index of the redshift‐peak evolution. From this, there are 10 parameters for calculation.

The detailed description about PLE and LDDE models can be found in sections 4.1 and 4.2 from Ref. [32]. These models also can be applied to X‐ray band, to determine the information of evolution of sources in X‐ray band (e.g., [62]). With the increase in the number of the detected sources, the evolutionary form of those sources becomes more complicated and the updated forms of those models can be found in Ref. [33], which allows the Gaussian mean μ of the photon index and the evolutionary factor e(z,Lγ) to change with luminosity.

2.3. The cosmological evolution

In Fermi sample, the large redshift range between z=0 and z=3.1 of gamma‐ray blazars was found. The obtained GLFs have shown that blazars have a cosmological evolution in their gamma‐ray band. We have simply discussed the redshift evolution of blazars in the “Introduction”. Ajello et al. [32] recently have presented the new results on the cosmological evolution of the BL Lac population by using the largest and most complete sample of gamma‐ray BL Lacs available in the literature and they found that for most BL Lac classes, the evolution is positive, with a space density peaking at modest redshift (z ≈ 1.2) (see Figure 2). In Figure 2, we also see that for their higher luminosity, FSRQs dominate at all redshifts z > 0.3 and the extreme growth in BL Lac numbers at low z allows them to produce ~90% of the local luminosity density. In particular, low‐luminosity, high‐synchrotron‐peaked (HSP) BL Lac objects showed different evolutionary behaviors with respect to other blazar classes (see Figure 2). They have strong negative evolution with number density increasing for z < 0.5, which confirms previous standpoints of negative evolution based on the samples of X‐ray‐selected BL Lac objects and this sample contained a large fraction of HSPs [39, 41].

Figure 2.

Left: The evolution of the luminosity density of FSRQs compared to that of BL Lac objects. Right: Number density of FSRQs, BL Lac objects and HSPs. The figures are obtained from the report of [32] and see Ref. [32] for additional details.

Advertisement

3. The extragalactic gamma‐ray background

The origin of the EGRB has been widely discussed for various gamma‐ray‐emitting sources in literature. Fermi has observed gamma‐ray emission from blazars, star‐forming galaxies, radio galaxies, GRBs and high‐latitude pulsars. Ajello et al. [33] and Di Mauro and Donato [36] suggested that blazars, star‐forming galaxies and radio galaxies are the main contributors to the EGRB. For those emitting sources, we focus on how to estimate the contribution of unresolved objects to the EGRB below, based on the best‐fitting GLF (space density of sources).

The differential intensity of the EGRB radiation can be expressed as follows:

dNdEdΩ=ΓminΓmaxdΓdNdΓzminzmaxdVdzLγ,minLγ,maxdLγΦ(Lγ,z)Fγintrinsic(E,Lγ,z,Γ)eτ(E,z)(1.0ω(Lγ,z,Γ))E9

where Φ(Lγ,z) is the GLF and eτ(E,z) is the optical depth of the extragalactic background light (EBL) for the sources at redshift z emitting gamma‐ray photon energy E. Recently, there are many studies on EBL (e.g., [21, 7375]). Generally, we adopted the model given by [73] for the EBL to calculate the optical depth. In Eq. (9), Fγintrinsic(E,Lγ,z,Γ) represents the intrinsic photon flux at energy E with γ‐ray luminosity Lγ and a power‐law spectrum at redshift z and it is expressed as follows:

Fγintrinsic(E, Lγ, z, Γ)= Lγ (1+z)2Γ4πdL2E12 {(2Γ)[(E2E1)2Γ1]1 (E100 MeV)Γ    Γ 2,1ln(E2/E1) (1E1E2)1(E100 MeV)2   Γ=2, E10

where E1=100 MeV and E2=100 GeV. Therefore, the integrated intensity between photon energy E1 and E2 (E2>E1) can be written as follows:

dNdΩ=E1E2dNdEdΩdEE11

The electrons and positrons are produced due to the interaction between very high energy (VHE) photons from TeV sources and ultraviolet‐infrared photons of EBL. The pairs could scatter the cosmic microwave background (CMB) radiation to high‐energy background radiation through the inverse Compton scattering process (e.g., [7683]). This cascading emission is regarded as a contributor to the EGRB if the flux of the cascade flux is lower than the detector's sensitivity. Now, we consider only the first generation of the electron‐positron pairs produced by the gamma‐ray absorption to obtain the cascade emission because the emission from the second generation or more than second generation of created pairs can be negligible at the GeV band [21]. The formulation of the cascade flux is given as follows [84]:

Fγcascade(E, Lγ, z, Γ)= 81 π16 λc3 εc2mec2(1+z)4UCMB 3εc/4εCMB(1+z)dγγ8exp[3εc/4γ2εCMB(1+z)1]×εmax2γdε FVHEintrinsic(5.11×105106ε, z, Lγ, Γ)[1eτ(ε,z)]E12

where λc=2.426×1010 cm is the Compton length, the dimensionless energy εc=E×106/(5.11×105), UCMB=4.0×1013 erg cm3 is the CMB energy density at z=0.0, εCMB=1.24×109mec2 is the average CMB photon energy at z=0.0 and εCMB=2.0×108 corresponding to EVHE=100TeV. FVHEintrinsic(EVHE,Lγ,z,Γ) represents the possible intrinsic TeV spectrum, which is extrapolated to the TeV energy ranges from the observed GeV spectrum Eq. (10) by assuming a power‐law spectrum. In Eq. (9), using Eq. (12) in place of Eq. (10) allows us to compute the contribution to the EGRB from the cascade emission of the source.

It is noted that there are two possible contributions for the cascade emission to the EGRB because the pairs are deflected by the extragalactic magnetic field (EGMF), which is shown in Figure 3. In case I, the cascade emission can contribute to the EGRB if the flux of the cascade emission is lower than that of the LAT sensitivity. In case II, although the flux of the cascade emission is larger than that of the LAT sensitivity, the angle between the redirected secondary gamma‐ray photons and the line of sight is larger than that of the LAT point‐spread function (PSF) (i.e., θ>θPSF). Thus, the cascade emission will not be attributed to a point source by the LAT and it then contributes to the EGRB, where θPSF=(1.7π/180)(0.001E)0.74[1+(0.001E/15)2]0.37 [85]. For more detailed information, see Refs. [81, 84].

Figure 3.

The cascade radiation processes in no or non‐zero extragalactic magnetic field (EGMF). Note that the pairs produced by the interaction between very high energy (VHE) photons and ultraviolet‐infrared photons of EBL are detected by the EGMF. The figures are obtained from the report of Marco Ajello at Fermi Symposium.

3.1. Blazars

Blazars emit gamma rays via the inverse Compton scattering processes and/or hadronic processes and dominate extragalactic gamma‐ray sources. Therefore, it is naturally expected that blazars contribute the main EGRB. However, its fraction was very uncertain in the EGRET era due to its small samples. At the same time, its fraction also severely depends on GLF. Blazars are divided into two main subgroups: BL Lac objects and FSRQs [37]. Figure 4 shows FSRQs’ EGRB spectra with LDDE model and BL Lacs’ EGRB spectra with PDE model. Compared to FSRQs, BL Lacs have lower gamma‐ray luminosities, lower redshifts and harder spectral indices in statistics (e.g., [86]). Thus, BL Lacs can provide a significant part in the contribution of blazar to the EGRB above 10 GeV. From Figure 4, we find out that the cascade emission from BL Lacs has a rather large fraction of the total EGRB energy flux and contrary to that of FSRQs, which may be caused by harder spectrum for BL Laces. Therefore, the contribution from BL Lacs cascade emission to the EGRB cannot be negligible. Based on the effect of the EGMF on the cascade contribution from blazars, Yan [84] have studied the effect of cascade radiation on the contribution to the EGRB using a simple semi‐analytical model. They suggested that if the strength of the EGMF is large enough (BEGMF>1012 G), the cascade contribution can significantly alter the spectrum of the EGRB at high energies. If the small strength of the EGMF is large enough (BEGMF<1014 G), then the cascade contribution is small, but it cannot be ignored. Recently, Ajello et al. [33] used an updated GLFs to analyze the redshift, luminosity and photon index distributions and obtained the best‐fitting evolutionary parameters of the GLFs. According to the GLFs and spectral energy distribution (SED) model consistent with the Fermi blazar observations, their result shown that blazars account for 5010+12 to the EGRB (see Figure 5).

Figure 4.

Comparison of predicted EGRB spectra from FSRQs and BL Lacs with the observed data of blazars. Note that the EGRB spectra from FSRQs and BL Lac are estimated based on LDDE and PDE models, respectively. The two figures are obtained from the report of Refs. [29, 30].

Figure 5.

The EGRB spectrum of blazars [33], star‐forming galaxies (gray band [4]) and radio galaxies (black striped band [48]) as well as summation of these three populations (yellow band), compared to the intensity of the observed ERGB [5]. The figure is obtained from the report of Ref. [33].

3.2. Radio galaxies

Radio galaxies are one of the largest subclasses of radio‐loud AGNs. It is more in number than blazars in the entire sky. Even though radio galaxies are fainter than blazars, Fermi‐LAT has just detected gamma rays from ~15 extragalactic sources, including 12 FR Is and 3 FR IIs [87]. In order to estimate the contribution to the EGRB from radio galaxies, their GLF is required. We must obtain indirectly the GLF due to the limited Fermi radio galaxy samples. Relying on a correlation between the luminosities in the radio and gamma‐ray frequencies, Inoue [48] converted the RLF [88] into the GLF and estimated about 25% of EGRB can be solved by radio galaxies (see Figure 5). This uncertainty significantly depends on the limited sample and the errors between the gamma ray and radio luminosity correlation.

3.3. Star‐forming galaxies

The Fermi‐LAT has detected gamma‐ray from ~9 star‐forming (SF) galaxies [2]. Those gamma rays are produced by interactions between cosmic rays and gas or interstellar radiation fields, including the decay of neutral pion and electron interactions (bremsstrahlung and inverse Compton scattering). Similar to radio galaxies, it is not straightforward to construct the GLF because of the limited star‐forming galaxy sample. Generally, the correlations between the IR wavelength and gamma‐ray region are used to predict the gamma‐ray diffuse emission for the unresolved SF galaxy population. Different from other types of source, the SF gamma‐ray average spectrum is difficult to firmly establish due to the paucity of statistics. Milky Way‐like SF galaxies (MW model) and an assumed power‐law spectrum (PL model) are proposed by Ackermann et al. [89] to express an average spectrum of SF Galaxies. In particular, the two predictions are different above 5 GeV, where the MW model softens significantly. Therefore, using the correlation between infrared and gamma‐ray luminosities, based on the well‐established infrared luminosity functions and the SF gamma‐ray average spectrum, the GLF of star‐forming galaxies is well built and the contribution of star‐forming galaxies to the EGRB can be estimated as 10–30% of the EGRB at >0.1 GeV [89], which can be seen in Figure 5.

It should be noted that about 95% of the EGRB can be naturally explained by blazars, star‐forming galaxies and radio galaxies in the 0.1–820 GeV range. Only modest space is left for other diffuse processes such as dark matter interactions, which suggests that other gamma‐ray‐emitting sources’ contribution can be neglected. Ajello et al. [33] also concluded that the result of their simulation gave an upper limit on DM self‐annihilation cross sections, which is similar to that from the independent types of analysis (e.g., [59]).

Advertisement

4. Conclusion and discussion

In this chapter, we reviewed the origin of EGRB and estimated the contribution of unsolved gamma‐ray‐emitting sources from Fermi‐LAT to the EGRB based on the construction of the corresponding GLFs. Since Fermi‐LAT has higher sensitivity and provides numerous gamma‐ray‐emitting sources for studies, we found two important results: (i) the redshift evolutionary information of gamma‐ray sources, particularly for blazars; HBLs show strong negative cosmological evolution, while FSRQs and luminous BL Lacs show positive evolution like as Seyferts and the cosmic star formation history. (ii) Fermi sources’ contribution to the EGRB; blazars clearly contribute to most of the EGRB (≈40-62%), as well as radio galaxies and star‐forming galaxies can occupy for the rest room of the EGRB [33, 36]. These results suggest that the contributions of other emitting sources have only little space to the EGRB. However, the uncertainties associated with these predictions from radio galaxies and star‐forming galaxies are still quite large because of the small samples. This situation is very similar to blazar studies in the early EGRET era. Therefore, further data will be required to construct the GLFs and precisely evaluate the contributions from those two populations.

Now, there are still some unresolved problems. We have not seen the signature of dark matter particles in the EGRB spectrum, although they are considered as the possible origin of EGRB. As we known, Fermi‐LAT has accurately measured the EGRB spectrum and the anisotropy of the EGRB [4] and the emission from dark matter is anisotropic and its spatial pattern is unique and predictable [90]. Therefore, we can obtain an upper limit on the annihilation cross section by comparing the expected EGRB angular power spectrum from dark matter annihilation with the measured spectrum. The work of Ajello et al. [33] shown that an analysis of the EGRB and its components can constrain diffuse emission mechanisms such as DM annihilation. Di Mauro and Donato [36] probed a possible emission coming from the annihilation of WIMP DM in the halo of our galaxy and found that the DM component can very well fit the EGRB data together with the realistic emission from a number of unresolved extragalactic sources.

The value of the EGMF has still not been determined. Since the pairs scatter CMB photons to GeV energies by Compton mechanism for cascade process around a TeV sources, Fermi‐LAT could measure those GeV photons, which would give a straight measurement of the EGMF. As continuous accumulation of the data observed and the further development of detection equipment, the imprint of the EGMF may be found in the gamma-ray spectrum and/or flux [79, 80, 91]. The EGMF imprint might also be found in the angular anisotropy of the EGRB [92]. If the effect of cascade depending on the EGMF cannot be neglected [84], the electron‐positron pairs produced in cascade process could be deflected by a high value of the EGMF, which makes GeV photons more isotropic. Therefore, the EGRB spectrum with the anisotropy could probe the strength of EGMF [87].

References

  1. 1. Atwood, W. B.; Abdo, A. A.; Ackermann, M.; Althouse, W.; Anderson, B.; et al. The large area telescope on the Fermi gamma‐ray space telescope mission. The Astrophysical Journal. 2009;697(2):1071–1102. doi:10.1088/0004‐637X/697/2/1071
  2. 2. Abdo, A. A.; Ackermann, M.; Ajello, M.; Allafort, A.; Antolini, E.; et al. The first catalog of active galactic nuclei detected by the Fermi large area telescope. The Astrophysical Journal. 2010;715(1): 429–457. doi:10.1088/0004‐637X/715/1/429
  3. 3. Abdo, A. A.; Ackermann, M.; Ajello, M.; Atwood, W. B.; Baldini, L.; Ballet, J.; et al. Spectrum of the isotropic diffuse gamma‐ray emission derived from first‐year Fermi large area telescope data. Physical Review Letters. 2010;104(10):101101. doi:10.1103/PhysRevLett.104.101101
  4. 4. Ackermann, M.; Ajello, M.; Albert, A.; Baldini, L.; Ballet, J.; Barbiellini, G.; et al. Anisotropies in the diffuse gamma‐ray background measured by the Fermi LAT. Physical Review D. 2012;85(8):083007. doi:10.1103/PhysRevD.85.083007
  5. 5. Ackermann, M.; Ajello, M.; Albert, A.; Atwood, W. B.; Baldini, L.; Ballet, J.; et al. The spectrum of isotropic diffuse gamma‐ray emission between 100 MeV and 820 GeV. The Astrophysical Journal. 2015;799(1):86, 24 pp. doi:10.1088/0004‐637X/799/1/86
  6. 6. Fichtel, C. E.; Simpson, G. A.; Thompson, D. J. Diffuse gamma radiation. The Astrophysical Journal. 1978;222(1):833–849. doi:10.1086/156202
  7. 7. Thompson, D. J.; Fichtel, C. E. Extragalactic gamma radiation—use of galaxy counts as a galactic tracer. Astronomy and Astrophysics. 1982;109(2):352–354.
  8. 8. Osborne, J. L.; Wolfendale, A. W.; Zhang, L. The diffuse flux of energetic extragalactic gamma rays. Journal of Physics G: Nuclear and Particle Physics. 1994;20(7):1089–1101. doi:10.1088/0954‐3899/20/7/010
  9. 9. Willis, T. D., Observations of the Isotropic Diffuse Gamma-Ray Background with the EGRET Telescope, Ph.D. thesis, Stanford University, Aug 1996
  10. 10. Sreekumar, P.; Bertsch, D. L.; Dingus, B. L.; Esposito, J. A.; Fichtel, C. E.; et al. EGRET observations of the extragalactic gamma‐ray emission. The Astrophysical Journal. 1998;494(2):523–534. doi:10.1086/305222
  11. 11. Strong, A. W.; Moskalenko, I. V.; Reimer, O. A new determination of the extragalactic diffuse gamma‐ray background from EGRET data. The Astrophysical Journal. 2004;613(2):956–961. doi:10.1086/423196
  12. 12. Ueda, Y.; Akiyama, M.; Hasinger, G.; Miyaji, T.; Watson, M. G. Toward the standard population synthesis model of the x‐ray background: evolution of x‐ray luminosity and absorption functions of active galactic nuclei including compton‐thick populations. The Astrophysical Journal. 2014;786(2):104, 28 pp. doi:10.1088/0004‐637X/786/2/104
  13. 13. Stecker, F. W.; Salamon, M. H.; Malkan, M. A. The high‐energy diffuse cosmic gamma‐ray background radiation from blazars. The Astrophysical Journal. 1993;410(2):L71–L74. doi:10.1086/186882
  14. 14. Padovani, P.; Ghisellini, G.; Fabian, A. C.; Celotti, A. Radio‐loud AGN and the extragalactic gamma‐ray background. Monthly Notices of the Royal Astronomical Society. 1993;260(3): L21–L24. doi:10.1093/mnras/260.1.L21
  15. 15. Chiang, J.; Fichtel, C. E.; von Montigny, C.; Nolan, P. L.; Petrosian, V. The evolution of gamma‐ray‐‐loud active galactic nuclei. The Astrophysical Journal. 1995;452:165. doi:10.1086/176287
  16. 16. Stecker, F. W.; Salamon, M. H. The gamma‐ray background from blazars: a new look. The Astrophysical Journal. 1996;464:600. doi:10.1086/177348
  17. 17. Chiang, J.; Mukherjee, R. The luminosity function of the EGRET gamma‐ray blazars. The Astrophysical Journal. 1998;496(2):752–760. doi:10.1086/305403
  18. 18. Mücke, A.; Pohl, M. The contribution of unresolved radio‐loud AGN to the extragalactic diffuse gamma‐ray background. Monthly Notices of the Royal Astronomical Society. 2000;312(1):177–193. doi:10.1046/j.1365‐8711.2000.03099.x
  19. 19. Narumoto, T.; Totani, T. Gamma‐ray luminosity function of blazars and the cosmic gamma‐ray background: evidence for the luminosity‐dependent density evolution. The Astrophysical Journal. 2006;634(1):81–91. doi:10.1086/502708
  20. 20. Dermer, C. D. Statistics of cosmological black hole jet sources: blazar predictions for the gamma‐ray large area space telescope. The Astrophysical Journal. 2007;659(2):958–975. doi:10.1086/512533
  21. 21. Kneiske, T. M.; Mannheim, K. BL Lacertae contribution to the extragalactic gamma‐ray background. Astronomy and Astrophysics. 2008;479(1):41–47. doi:10.1051/0004‐6361:20065605
  22. 22. Bhattacharya, D.; Sreekumar, P.; Mukherjee, R. Gamma‐ray luminosity function of gamma‐ray bright AGNs. Research in Astronomy and Astrophysics. 2008;9(1):85–94. doi:10.1088/1674‐4527/9/1/007
  23. 23. Inoue, Y.; Totani, T. The blazar sequence and the cosmic gamma‐ray background radiation in the Fermi era. The Astrophysical Journal. 2009;702(1):523–536. doi:10.1088/0004‐637X/702/1/523
  24. 24. Abdo, A. A.; Ackermann, M.; Ajello, M.; Antolini, E.; Baldini, L.; Ballet, J.; et al. The Fermi‐LAT high‐latitude survey: source count distributions and the origin of the extragalactic diffuse background. The Astrophysical Journal. 2010;720(1):435–453. doi:10.1088/0004‐637X/720/1/435
  25. 25. Ghirlanda, G.; Ghisellini, G.; Tavecchio, F.; Foschini, L.; Bonnoli, G. The radio‐γ‐ray connection in Fermi blazars. Monthly Notices of the Royal Astronomical Society. 2011;413(2):852–862. doi:10.1111/j.1365‐2966.2010.18173.x
  26. 26. Stecker, F. W.; Venters, T. M. Components of the extragalactic gamma‐ray background. The Astrophysical Journal. 2011;736(1):40, 13 pp. doi:10.1088/0004‐637X/736/1/40
  27. 27. Singal, J.; Petrosian, V.; Ajello, M. Flux and photon spectral index distributions of Fermi‐LAT blazars and contribution to the extragalactic gamma‐ray background. The Astrophysical Journal. 2012;753(1):45, 11 pp. doi:10.1088/0004‐637X/753/1/45
  28. 28. Zeng, H. D.; Yan, D. H.; Sun, Y. Q.; Zhang, L. γ‐Ray luminosity function and the contribution to extragalactic γ‐ray background for Fermi‐detected blazars. The Astrophysical Journal. 2012;749(2):151, 8 pp. doi:10.1088/0004‐637X/749/2/151
  29. 29. Zeng, H.; Yan, D.; Zhang, L. A revisit of gamma‐ray luminosity function and contribution to the extragalactic diffuse gamma‐ray background for Fermi FSRQs. Monthly Notices of the Royal Astronomical Society. 2013;431(1):997–1003. doi:10.1093/mnras/stt223
  30. 30. Zeng, H.; Yan, D.; Zhang, L. Gamma‐ray luminosity function of BL Lac objects. Monthly Notices of the Royal Astronomical Society. 2014;441(2):1760–1768. doi:10.1093/mnras/stu644
  31. 31. Ajello, M.; Shaw, M. S.; Romani, R. W.; Dermer, C. D.; Costamante, L.; et al. The luminosity function of Fermi‐detected flat‐spectrum radio quasars. The Astrophysical Journal. 2012;751(2):108, 20 pp. doi:10.1088/0004‐637X/751/2/108
  32. 32. Ajello, M.; Romani, R. W.; Gasparrini, D.; Shaw, M. S.; Bolmer, J.; et al. The cosmic evolution of Fermi BL lacertae objects. The Astrophysical Journal. 2014;780(1):73, 24 pp. doi:10.1088/0004‐637X/780/1/73
  33. 33. Ajello, M.; Gasparrini, D.; Sánchez‐Conde, M.; Zaharijas, G.; Gustafsson, M.; et al. The origin of the extragalactic gamma‐ray background and implications for dark matter annihilation. The Astrophysical Journal Letters. 2015;800(2): L27, 7 pp. doi:10.1088/2041‐8205/800/2/L27
  34. 34. Di Mauro, M.; Donato, F.; Lamanna, G.; Sanchez, D. A.; Serpico, P. D. Diffuse γ‐ray emission from unresolved BL lac objects. The Astrophysical Journal. 2014;786(2):129, 12 pp. doi:10.1088/0004‐637X/786/2/129
  35. 35. Di Mauro, M.; Calore, F.; Donato, F.; Ajello, M.; Latronico, L.. Diffuse γ‐ray emission from misaligned active galactic nuclei. The Astrophysical Journal. 2014;780(2):161, 14 pp. doi:10.1088/0004‐637X/780/2/161
  36. 36. Di Mauro, M.; Donato, F. Composition of the Fermi‐LAT isotropic gamma‐ray background intensity: emission from extragalactic point sources and dark matter annihilations. Physical Review D. 2015;91(12):123001. doi:10.1103/PhysRevD.91.123001
  37. 37. Urry, C. Megan; Padovani, Paolo. Unified schemes for radio‐loud active galactic nuclei. Publications of the Astronomical Society of the Pacific. 1995;107:803. doi:10.1086/133630
  38. 38. Acero, F.; Ackermann, M.; Ajello, M.; Albert, A.; Atwood, W. B.; Axelsson, M.; et al. Fermi large area telescope third source catalog. The Astrophysical Journal Supplement Series. 2015;218(2):23, 41 pp. doi:10.1088/0067‐0049/218/2/23
  39. 39. Rector, Travis A.; Stocke, John T.; Perlman, Eric S.; Morris, Simon L.; Gioia, Isabella M. The properties of the x‐ray‐selected EMSS sample of BL lacertae objects. The Astronomical Journal. 2000;120(4):1626–1647. doi:10.1086/301587
  40. 40. Caccianiga, A.; Maccacaro, T.; Wolter, A.; Della Ceca, R.; Gioia, I. M. On the cosmological evolution of BL lacertae objects. The Astrophysical Journal. 2002;566(1):181–186. doi:10.10 86/338073
  41. 41. Beckmann, V.; Engels, D.; Bade, N.; Wucknitz, O. The HRX‐BL Lac sample—evolution of BL lac objects. Astronomy and Astrophysics. 2003;401:927–938. doi:10.1051/0004‐6361: 20030184
  42. 42. Padovani, P.; Giommi, P.; Landt, H.; Perlman, Eric S. The deep X‐ray radio blazar survey. III. Radio number counts, evolutionary properties and luminosity function of blazars. The Astrophysical Journal. 2007;662(1):182–198. doi:10.1086/516815
  43. 43. Dunlop, J. S.; Peacock, J. A. The redshift cut‐off in the luminosity function of radio galaxies and quasars. Monthly Notices of the Royal Astronomical Society. 1990;247(1):19.
  44. 44. Ueda, Y.; Akiyama, M.; Ohta, K.; Miyaji, T. Cosmological evolution of the hard x‐ray active galactic nucleus luminosity function and the origin of the hard x‐ray background. The Astrophysical Journal. 2003;598(2):886–908. doi:10.1086/378940
  45. 45. Hasinger, G.; Miyaji, T.; Schmidt, M. Luminosity‐dependent evolution of soft X‐ray selected AGN. New Chandra and XMM‐Newton surveys. Astronomy and Astrophysics. 2005;441(2):417–434. doi:10.1051/0004‐6361:20042134
  46. 46. Strong, A. W.; Wolfendale, A. W.; Worrall, D. M.. Origin of the diffuse gamma ray background. Monthly Notices of the Royal Astronomical Society. 1976;175:23–27. doi:10.1093/mnras/175.1.23P
  47. 47. Pavlidou, V.; Fields, B. D. The guaranteed gamma‐ray background. The Astrophysical Journal. 2002;575(1):L5–L8. doi:10.1086/342670
  48. 48. Inoue, Y. Contribution of gamma‐ray‐loud radio galaxies’ core emissions to the cosmic MeV and GeV gamma‐ray background radiation. The Astrophysical Journal. 2011;733(1):66, 9 pp. doi:10.1088/0004‐637X/733/1/66
  49. 49. Casanova, S.; Dingus, B. L.; Zhang, B. Contribution of GRB emission to the GeV extragalactic diffuse gamma‐ray flux. The Astrophysical Journal. 2007;656(1):306–312. doi:10.1086/510613
  50. 50. Faucher‐Giguère, Claude‐André; Loeb, Abraham. The pulsar contribution to the gamma‐ray background. Journal of Cosmology and Astroparticle Physics. 2010;01(01):005. doi:10.1088/1475‐7516/2010/01/005
  51. 51. Loeb, A.; Waxman, E. Cosmic γ‐ray background from structure formation in the intergalactic medium. Nature. 2000;405(6783):156–158.
  52. 52. Totani, T.; Kitayama, T. Forming clusters of galaxies as the origin of unidentified GEV gamma‐ray sources. The Astrophysical Journal. 2000;545(2):572–577. doi:10.1086/317872
  53. 53. Inoue, Y.; Totani, T.; Ueda, Y. The cosmic MeV gamma‐ray background and hard x‐ray spectra of active galactic nuclei: implications for the origin of hot AGN coronae. The Astrophysical Journal Letters. 2008;672(1): L5. doi:10.1086/525848
  54. 54. Dar, A.; Shaviv, N. J. Origin of the high energy extragalactic diffuse gamma ray background. Physical Review Letters. 1995;75(17):3052–3055. doi:10.1103/PhysRevLett.75.3052
  55. 55. Kalashev, O. E.; Semikoz, D. V.; Sigl, G. Ultrahigh energy cosmic rays and the GeV‐TeV diffuse gamma‐ray flux. Physical Review D. 2009;79(6): 063005. doi:10.1103/PhysRevD.79.063005
  56. 56. Keshet, U.; Waxman, E.; Loeb, A. The case for a low extragalactic gamma‐ray background. Journal of Cosmology and Astroparticle Physics. 2004;04:006. doi:10.1088/1475‐7516/2004/04/006
  57. 57. Moskalenko, I. V.; Porter, T. A. Isotropic gamma‐ray background: cosmic‐ray‐induced albedo from debris in the solar system? The Astrophysical Journal Letters. 2009;692(1):L54–L57. doi:10.1088/0004‐637X/692/1/L54
  58. 58. Bergström, L.; Edsjö, J.; Ullio, P. Spectral gamma‐ray signatures of cosmological dark matter annihilations. Physical Review Letters. 2001;87(25):251301. doi:10.1103/PhysRevL ett.87.251301
  59. 59. Ackermann, M.; Albert, A.; Anderson, B.; Baldini, L.; Ballet, J.; Barbiellini, G.; Bastieri, D.; Bechtol, K.; Bellazzini, R.; Bissaldi, E.; and 112 coauthors. Dark matter constraints from observations of 25 Milky Way satellite galaxies with the Fermi large area telescope. Physical Review D. 2014;89(4):042001. doi:10.1103/PhysRevD.89.042001
  60. 60. Schmidt, M. Space distribution and luminosity functions of quasi‐stellar radio sources. The Astrophysical Journal. 1968;151:393. doi:10.1086/149446
  61. 61. Marshall, H. L.; Tananbaum, H.; Avni, Y.; Zamorani, G.. Analysis of complete quasar samples to obtain parameters of luminosity and evolution functions. The Astrophysical Journal. 1983;269(1):35–41. doi:10.1086/161016
  62. 62. Ajello, M.; Costamante, L.; Sambruna, R. M.; Gehrels, N.; Chiang, J.; Rau, A.; et al. The evolution of Swift/BAT Blazars and the origin of the MeV background. The Astrophysical Journal. 2009;699(1):603–625. doi:10.1088/0004‐637X/699/1/603
  63. 63. Kochanek, C. S. The flat‐spectrum radio luminosity function, gravitational lensing, galaxy ellipticities and cosmology. The Astrophysical Journal. 1996;473:595. doi:10.1086/178175
  64. 64. Fan, Z. H.; Liu, S. M.; Yuan, Q.; Fletcher, L. Lepton models for TeV emission from SNR RX J1713.7–3946. Astronomy and Astrophysics. 2010;517:L4, 4 pp. doi:10.1051/0004‐6361/201015169
  65. 65. Neal, R. M. Probabilistic Inference Using Markov Chain Monte Carlo Methods, Technical Report. Department of Computer Science, Univ.Toronto: 1993
  66. 66. Gamerman, D. Markov Chain Monte Carlo: Stochastic Simulation for Bayesian Inference. Chapman and Hall: London; 1997
  67. 67. Lewis, A.; Bridle, S. Cosmological parameters from CMB and other data: a Monte Carlo approach. Physical Review D. 2002;66(10):103511. doi:10.1103/PhysRevD.66.103511
  68. 68. MacKay, D. J. C. Information theory, inference and learning algorithms[M]. Cambridge university press, 2003;P 640, ISBN-13: 978-0521642989
  69. 69. Zhang, L.; Cheng, K. S.; Fan, J. H. The radio and gamma‐ray luminosities of blazars. Publications of the Astronomical Society of Japan. 2001;53(2):207–213. doi:10.1093/pasj/53.2.207
  70. 70. Narumoto, T.; Totani, T. Gamma‐ray luminosity function of blazars and the cosmic gamma‐ray background: evidence for the luminosity‐dependent density evolution. Astrophysics and Space Science.2007;309(1–4):73–79. doi:10.1007/s10509‐007‐9453‐4
  71. 71. Abazajian, K. N.; Blanchet, S.; Harding, J. P. Contribution of blazars to the extragalactic diffuse gamma‐ray background and their future spatial resolution. Physical Review D. 2011;84(10):103007. doi:10.1103/PhysRevD.84.103007
  72. 72. Li, F.; Cao, X.‐W. BL Lacertae objects and the extragalactic γ‐ray background. Research in Astronomy and Astrophysics. 2011;11(8):879–887. doi:10.1088/1674‐4527/11/8/001
  73. 73. Finke, J. D.; Razzaque, S.; Dermer, C. D. Modeling the extragalactic background light from stars and dust. The Astrophysical Journal. 2010;712(1):238–249. doi:10.1088/0004‐637X/712/1/238
  74. 74. Domínguez, A.; Primack, J. R.; Rosario, D. J.; Prada, F.; Gilmore, R. C.; Faber, S. M.; et al. Extragalactic background light inferred from AEGIS galaxy‐SED‐type fractions. Monthly Notices of the Royal Astronomical Society. 2011;410(4):2556–2578. doi:10.1111/j.1365‐2966.2010.17631.x
  75. 75. Inoue, Y.; Inoue, S.; Kobayashi, M. A. R.; Makiya, R.; Niino, Y.; Totani, T. Extragalactic Background Light from hierarchical galaxy formation: gamma‐ray attenuation up to the epoch of cosmic reionization and the first stars. The Astrophysical Journal. 2013;768(2):197, 17 pp. doi:10.1088/0004‐637X/768/2/197
  76. 76. Fan, Y. Z.; Dai, Z. G.; Wei, D. M. Strong GeV emission accompanying TeV blazar H1426+428. Astronomy and Astrophysics. 2004;415:483–486. doi:10.1051/0004‐6361:20034472
  77. 77. Murase, K.; Takahashi, K.; Inoue, S.; Ichiki, K.; Nagataki, S. Probing intergalactic magnetic fields in the GLAST era through pair echo emission from TeV blazars. The Astrophysical Journal Letters. 2008;686(2): L67. doi:10.1086/592997
  78. 78. Yang, C. Y.; Fang, J.; Lin, G. F.; Zhang, L.. Possible GeV Emission from TeV blazars. The Astrophysical Journal. 2008;682(2):767–774. doi:10.1086/589326
  79. 79. Neronov, A.; Vovk, I. Evidence for strong extragalactic magnetic fields from Fermi observations of TeV blazars. Science. 2010;328(5974):73. doi:10.1126/science.1184192
  80. 80. Tavecchio, F.; Ghisellini, G.; Foschini, L.; Bonnoli, G.; Ghirlanda, G.; Coppi, P. The intergalactic magnetic field constrained by Fermi/large area telescope observations of the TeV blazar 1ES0229+200. Monthly Notices of the Royal Astronomical Society: Letters. 2010;406(1):L70–L74. doi:10.1111/j.1745‐3933.2010.00884.x
  81. 81. Dermer, C. D.; Cavadini, M.; Razzaque, S.; Finke, J. D.; Chiang, J.; Lott, B. Time delay of cascade radiation for TeV blazars and the measurement of the intergalactic magnetic field. The Astrophysical Journal Letters. 2011;733(2):L21. doi:10.1088/2041‐8205/733/2/L21
  82. 82. Huan, H.; Weisgarber, T.; Arlen, T.; Wakely, S. P.. A new model for gamma‐ray cascades in extragalactic magnetic fields. The Astrophysical Journal Letters. 2011;735(2):L28, 5 pp. doi:10.1088/2041‐8205/735/2/L28
  83. 83. Inoue, Y.; Ioka, K. Upper limit on the cosmological gamma‐ray background. Physical Review D. 2012;86(2):023003. doi:10.1103/PhysRevD.86.023003
  84. 84. Yan, D.; Zeng, H.; Zhang, L. Contribution from blazar cascade emission to the extragalactic gamma‐ray background: what role does the extragalactic magnetic field play? Monthly Notices of the Royal Astronomical Society. 2012;422(2):1779–1784. doi:10.1111/j.1365‐2966.2012.20752.x
  85. 85. Taylor, A. M.; Vovk, I.; Neronov, A. Extragalactic magnetic fields constraints from simultaneous GeV‐TeV observations of blazars. Astronomy & Astrophysics. 2011;529:A144, 9 pp. doi:10.1051/0004‐6361/201116441
  86. 86. Ackermann, M.; Ajello, M.; Allafort, A.; Antolini, E.; Atwood, W. B.; et al. The second catalog of active galactic nuclei detected by the Fermi large area telescope. The Astrophysical Journal. 2011;743(2):171, 37 pp. doi:10.1088/0004‐637X/743/2/171
  87. 87. Massaro. F.; Thompson, D. J.; Ferrara, E. C. The extragalactic gamma‐ray sky in the Fermi era. The Astronomy and Astrophysics Review. 2016;24:1.
  88. 88. Willott, C. J.; Rawlings, S.; Blundell, K. M.; Lacy, M.; Eales, S. A. The radio luminosity function from the low‐frequency 3CRR, 6CE and 7CRS complete samples. Monthly Notices of the Royal Astronomical Society. 2001;322(3):536–552. doi:10.1046/j.1365‐8711.2001.04101.x
  89. 89. Ackermann, M.; Ajello, M.; Allafort, A.; Baldini, L.; Ballet, J.; Bastieri, D.; et al. GeV observations of star‐forming galaxies with the Fermi large area telescope. The Astrophysical Journal. 2011;755(2):164, 23 pp. doi:10.1088/0004‐637X/755/2/164
  90. 90. Ando, S.; Komatsu, E.; Narumoto, T.; Totani, T. Dark matter annihilation or unresolved astrophysical sources? Anisotropy probe of the origin of the cosmic gamma‐ray background. Physical Review D. 2007;75(6):063519. doi:10.1103/PhysRevD.75.063519
  91. 91. Tavecchio, F.; Ghisellini, G.; Bonnoli, G.; Foschini, L. Extreme TeV blazars and the intergalactic magnetic field. Monthly Notices of the Royal Astronomical Society. 2011;414(4):3566–3576. doi:10.1111/j.1365‐2966.2011.18657.x
  92. 92. Venters, T. M.; Pavlidou, V. Probing the intergalactic magnetic field with the anisotropy of the extragalactic gamma‐ray background. Monthly Notices of the Royal Astronomical Society. 2013;432(4):3485–3494. doi:10.1093/mnras/stt697

Written By

Houdun Zeng and Li Zhang

Reviewed: 22 December 2016 Published: 24 May 2017