On the Jet Properties of γ-Ray-loud Active Galactic Nuclei

Published 2018 April 19 © 2018. The American Astronomical Society. All rights reserved.
, , Citation Liang Chen 2018 ApJS 235 39 DOI 10.3847/1538-4365/aab8fb

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

0067-0049/235/2/39

Abstract

Based on broadband spectral energy distributions (SEDs), we estimate the jet physical parameters of 1392 γ-ray-loud active galactic nuclei (AGNs), the largest sample so far. The (SED) jet power and magnetization parameter are derived for these AGNs. Out of these sources, the accretion disk luminosity of 232 sources and (extended) kinetic jet powers of 159 sources are compiled from archived papers. We find the following. (1) Flat-spectrum radio quasars (FSRQs) and BL Lacs are well separated by ${\rm{\Gamma }}=-0.127\mathrm{log}{L}_{\gamma }+8.18$ in the γ-ray luminosity versus photon index plane with a success rate of 88.6%. (2) Most FSRQs present a (SED) jet power larger than the accretion power, which suggests that the relativistic jet-launching mechanism is dominated by the Blandford–Znajek process. This result confirms previous findings. (3) There is a significant anticorrelation between jet magnetization and a ratio of the (SED) jet power to the (extended) kinetic jet power, which, for the first time, provides supporting evidence for the jet energy transportation theory: a high-magnetization jet may more easily transport energy to a large scale than a low-magnetization jet.

Export citation and abstract BibTeX RIS

1. Introduction

The gravitational potential of the supermassive black hole (BH) at the center of an active galactic nucleus (AGN) is believed to be the ultimate energy source of the AGN. An accretion disk can be formed by way of matter falling onto the BH, and the angular momentum can be lost through viscosity or turbulence (e.g., Rees 1984) or via outflow or magnetic field processes (e.g., Cao 2011; Cao & Spruit 2013). About 10% of AGNs have relative stronger radio emissions compared with their optical emissions (i.e., radio-loud AGNs3 ), which are believed to host relativistic jets launched from the central accreting system (Urry & Padovani 1995; Yuan et al. 2008; Cao 2016). The jet launching can be related to the Blandford–Znajek (BZ) process (Blandford & Znajek 1977) through extracting the rational energy of BHs and/or the Blandford–Payne (BP) process (Blandford & Payne 1982) by releasing the gravitational energy of the accretion disk. As an extreme subclass of radio-loud AGNs, blazars show broadband emissions (radio through γ-ray), rapid variability, high and variable polarization, superluminal motions, and core-dominated nonthermal continua, which are believed to be caused by a relativistic Doppler-beaming effect due to a small viewing angle between the relativistic jet and the line of sight (see, e.g., Blandford & Rees 1978; Urry & Padovani 1995; Falomo et al. 2014; Madejski & Sikora 2016). These properties provide the blazar an ideal laboratory to study AGN jet physics. Broadband spectral energy distributions (SEDs) of blazars usually show two significant bumps: one peaks at the infrared (IR) to X-ray bands, which is believed to be the synchrotron emission of energetic electrons within the jet, and the second peaks at the γ-ray band, which may be the inverse Compton (IC) emission of the same electron population emitting the synchrotron bump (e.g., Dermer & Schlickeiser 1993; Sikora et al. 1994; Bloom & Marscher 1996; see also the hadronic model, Mannheim 1993; Aharonian 2000). According to the rest-frame equivalent width (EW) of the broad emission lines, blazars are classified as flat-spectrum radio quasars (FSRQs) with EW ≥ 5 Å and BL Lacertae objects (BL Lacs) with EW < 5 Å; see Scarpa & Falomo 1997.

Indisputably, the jet property is basically important in studying jet physics, including jet launching, energy transportation, energy transform and conversion, radiative processes, etc. Limited by the spatial resolution of modern telescopes, only the very long baseline interferometry (VLBI) radio technique can resolve subparsec-scale jets (even down to several Schwarzschild radii for very near and massive BHs, e.g., M87; Doeleman et al. 2012). Superluminal motion is a common phenomenon in blazar VLBI observations, which set a strong constraint that these jets should move very fast (apparent velocity on the order of ${\beta }_{\mathrm{app}}\equiv {v}_{\mathrm{app}}/c\gtrsim 10$ and can reach ∼30 for some special sources; see, e.g., Lister et al. 2013). The rapid variability at the γ-ray band can also set a lower limit of jet velocity due to the fact that these γ-ray photons overcome the possible absorption by the soft photons through the photon annihilation process and escape from the emission region (e.g., Dondi & Ghisellini 1995). The jet properties at the launching and energy dispassion regions can also be constrained through broadband SED modeling (e.g., Ghisellini et al. 1998, 2014; Ghisellini & Tavecchio 2015; Chen 2017). This powerful method can offer a constraint on the main jet parameters, including, within the one-zone leptonic model, the Doppler factor (the bulk Lorentz factor sometimes), the strength of the magnetic field, the energy density of the relativistic electrons, the jet power, the energy distribution of the electrons, and the location of the emission region (e.g., Ghisellini et al. 2014; Kang et al. 2014; Ghisellini & Tavecchio 2015; Chen 2017). Benefiting from a better knowledge of the high-frequency radio, millimeter, far-IR, and γ-ray continuum given by the WMAP, Planck, WISE, and Fermi Large Area Telescope (LAT) satellites, Ghisellini et al. (2014) modeled the broadband SED of the largest blazar sample yet (the number of blazars reaches 217; see also Ghisellini & Tavecchio 2015) and got the jet physical parameters. Note that within the one-zone leptonic model, the estimation of the jet parameters mainly depends on the peak frequency and luminosity of the two SED bumps (see, e.g., Tavecchio et al. 1998; Chen 2017).

Since its launch on 2008 June 11, the LAT onboard the Fermi satellite has revolutionized our knowledge of γ-ray AGNs above 100 MeV. A large number of AGNs are detected by Fermi/LAT and are compiled as the LAT Bright AGN Sample (LBAS; Abdo et al. 2009) and the First/Second/Third LAT AGN Catalogs (1LAC/2LAC/3LAC; Abdo et al. 2010b; Ackermann et al. 2011, 2015). The 3LAC includes 1591 AGNs located at high Galactic latitudes ($| b| \gt 10^\circ $) and detected at ≳100 MeV with a test statistic greater than 25 between 2008 August 4 and 2012 July 31. The number of γ-ray AGNs is even beyond the previous estimation (e.g., Cao & Bai 2008). Most of these AGNs are blazars (98%) and can be obtained with radio, optical, and X-ray data. In this paper, we will investigate the AGN jet physics based on broadband SED. Section 2 describes the sample. Section 3 presents the method to calculate the jet parameters. The results and discussion will be given in Section 4. Our conclusion will be drawn in Section 5. A ΛCDM cosmology with values from the Planck results is used in our calculation; in particular, Ωm = 0.32, ΩΛ = 0.68, and the Hubble constant H0 = 67 km s−1 Mpc−1 (Planck Collaboration et al. 2014).

2. Sample

The 3LAC contains 1591 AGNs, of which 1559 are blazars (467 FSRQs, 632 BL Lacs, and 460 BCUs4 ). Due to the uniform all-sky exposure of the Fermi/LAT, these sources form a γ-ray flux-limited sample. Note that since the blazars are rapidly variable, it is better to use the simultaneous multifrequency data, which seems impossible for the 3LAC, given the very large number of sources. Through collecting the multifrequency data (radio to X-ray bands) from the NASA/IPAC Extragalactic Database (NED), Fan et al. (2016) successfully obtained the broadband SED of 1392 blazars (461 FSRQs, 620 BL Lacs, and 311 BCUs). They employed a log-parabolic function, $\mathrm{log}(\nu {F}_{\nu })=-{P}_{1}$ ${(\mathrm{log}\nu -\mathrm{log}{\nu }_{p})}^{2}+\mathrm{log}({\nu }_{p}{F}_{{\nu }_{p}})$, to fit the SED; therefore, the synchrotron peak frequency (νp), spectral curvature (P1), and peak flux (${\nu }_{p}{F}_{{\nu }_{p}}$) are obtained. Of the total number of blazars, 999 have measured redshifts. In this paper, we will study the jet physical properties of the 1392 blazars, based on the data of Fan et al. (2016). The data regarding the γ-ray (i.e., photon index and luminosity) are compiled from Acero et al. (2015) and Ackermann et al. (2015).

3. Methods

We directly collect the peak frequency and flux of the synchrotron bump from Fan et al. (2016). For the IC bump, the observed γ-ray luminosity and photon index are used to estimate the peak luminosity and frequency. Abdo et al. (2010a) collected quasi-simultaneous broadband SEDs of 48 blazars and employed a third-degree polynomial to fit the synchrotron and IC bumps. The obtained peak frequency and luminosity are found to be significantly correlated with the γ-ray photon index and luminosity, respectively. They found a tight relation between the photon index and IC peak frequency: $\mathrm{log}{\nu }_{\mathrm{IC}}^{p}=-4{\rm{\Gamma }}+31.6$ (see Equation (5) in Abdo et al. 2010a). For the same sample, we plot γ-ray luminosity versus IC peak luminosity in Figure 1. The linear fitting shows $\mathrm{log}{L}_{\mathrm{IC}}^{p}\,=(0.946\pm 0.018)\mathrm{log}{L}_{\gamma }+(2.18\pm 0.88)$, and the Pearson test yields the chance probability p = 8.72 × 10−39. Because there is not enough data to construct the IC bump for this large sample, the above two formulae are used to estimate the IC peak frequency and luminosity.

Figure 1.

Figure 1. Relation between γ-ray luminosity and IC peak luminosity for blazars from Abdo et al. (2010a), with the best linear fit $\mathrm{log}{L}_{\mathrm{IC}}^{p}=(0.946\pm 0.018)\mathrm{log}{L}_{\gamma }\,+(2.18\pm 0.88)$ and the chance probability p = 8.72 × 10−39 (Pearson test).

Standard image High-resolution image

For BL Lac objects, the optical emission lines are very weak or even missed. Their γ-ray emission is believed to be synchrotron self-Compton (SSC) emission. The γ-ray emission of FSRQs, with strong optical emission lines, is believed to be emission from IC external seed photons produced (external inverse Compton, "EC") from, e.g., a broad emission-line region (BLR) or dust torus. As suggested by Ghisellini et al. (2009), BL Lacs and FSRQs are neatly separated in the γ-ray photon index versus γ-ray luminosity plane (see also Abdo et al. 2009; Ackermann et al. 2015). For confirmed FSRQs and BL Lacs in the 3LAC (Ackermann et al. 2015), we plot the γ-ray photon index versus γ-ray luminosity in Figure 2. It can be seen that FSRQs have larger γ-ray luminosity and softer γ-ray spectra (red squares, mostly in the upper right corner) compared with BL Lacs (blue squares, mostly in the lower left corner). The physical difference between these two subclasses may be related to the different accretion model of the central BH; i.e., a standard cold accretion disk exists in FSRQs, while advection-dominated accretion flow (ADAF; Yuan & Narayan 2014) exists in BL Lacs (e.g., Wang et al. 2002; Cao 2003; Ghisellini et al. 2009; Xu et al. 2009; Sbarrato et al. 2014). Despite the different physical origin, we choose a phenomenological criterion (a line in the γ-ray photon index versus luminosity plane) to roughly separate these two subclasses. This criterion/line should satisfy two criteria (see Figure 2): (1) the fraction of FSRQs above the line and the fraction of BL Lacs below the line should be the same, and (2) this fraction value must reach maximum. Finally, we get the criterion (the black line in Figure 2) ${\rm{\Gamma }}=-0.127\mathrm{log}{L}_{\gamma }\,+8.18$. In this case, the fraction of FSRQs above the line and the fraction of BL Lacs below the line reaches the maximum value: 88.6%. As mentioned above, 311 out of 1932 blazars are BCUs. In this paper, this criterion is employed to classify these BCUs into FSRQs or BL Lacs. In this case, the class type is labeled "UF" for BCUs classified as FSRQs and "UB" for BCUs classified as BL Lacs, while the class types "CF" and "CB" are for these confirmed FSRQs and BL Lacs, respectively (see the class type in Table 1).

Figure 2.

Figure 2. The γ-ray photon index vs. γ-ray luminosity. The FSRQs and BL Lacs are well separated by ${\rm{\Gamma }}=-0.127\mathrm{log}{L}_{\gamma }+8.18$ with a success rate of 88.6%.

Standard image High-resolution image

Table 1.  AGN Jet Physical Parameters

Name z Cl. $\mathrm{log}{\gamma }_{0}$ b $\mathrm{log}R$ δ $\mathrm{log}B$ $\mathrm{log}{P}_{\mathrm{jet}}$ $\mathrm{log}\sigma $ $\mathrm{log}\eta $ $\mathrm{log}M$ $\mathrm{log}{L}_{\mathrm{disk}}$ Model
(1) (2) (3) (4) (5) (6) (7) (8) (9) (10) (11) (12) (13) (14)
J0001.2-0748 CB 4.25 0.60 17.2 56.1 −2.46 46.6 −3.36 ST

Note. The source names labeled "∗" are blazars having extreme Doppler factors ($\delta \lt 1$ or $\delta \gt 100$), and therefore we instead use the median values (${\delta }_{{\rm{m}}}\simeq 14.3$ for BL Lacs and ${\delta }_{{\rm{m}}}\simeq 10.7$ for FSRQs) to calculate other jet parameters (see text for details). Column (1) gives the 3FGL name. Column (2) gives the redshift. Column (3) is the class of the source: "UF" for BCUs classified as FSRQs, "UB" for BCUs classified as BL Lacs, and "CF" and "CB" for confirmed FSRQs and BL Lacs, respectively. Column (4) is the electron peak energy. Column (5) is the curvature b of the electron energy distribution. Column (6) is the radius of the emission sphere in units of cm. Column (7) is the Doppler factor. Column (8) is the strength of the magnetic field in units of Gs (strength of magnetic field). Column (9) is the jet (SED) power in units of erg s−1. Column (10) is the magnetization parameter $\sigma ={U}_{{\rm{B}}}/{U}_{{\rm{e}}}$. Column (11) is the ratio of the jet SED power to (extended) jet kinetic power, $\eta ={P}_{\mathrm{jet}}/{P}_{\mathrm{jet},\mathrm{ext}}$. Column (12) is the available BH mass in units of M. Column (13) is the available disk luminosity in units of erg s−1. Column (14) is the model used in the calculations: "ST" for SSC/Thomson, "SK" for SSC/KN, "ET" for EC/Thomson, and "EK" for EC/KN. (This table is available in its entirety in ASCII text form.)

Only a portion of this table is shown here to demonstrate its form and content. A machine-readable version of the full table is available.

Download table as:  DataTypeset image

The SSC emission of FSRQs usually peaks at the X-ray band and dominates its X-ray emission (see, e.g., Kang et al. 2014; Ghisellini & Tavecchio 2015; Zhang et al. 2015). Therefore, the X-ray luminosity can be set as the upper limit of the SSC luminosity in FSRQs. Fan et al. (2016) presented X-ray luminosity at 1 keV for 275 out of 486 FSRQs and found a tight relation between X-ray and γ-ray luminosity, L1GeV = 0.91L1keV + 5.1 (p < 10−4; see Table 3 in Fan et al. 2016), which is used to estimate the X-ray luminosity for FSRQs without X-ray measurement.

In our sample, 393 sources have no measured redshift (157 BL Lacs and 236 BCUs). For these sources, we use the median values of the known redshifts of confirmed BL Lacs (z = 0.359) and FSRQs (z = 1.108) for unknown-redshift BL Lacs (including BCUs divided into BL Lacs) and BCUs divided into FSRQs, respectively. Finally, we have 803 BL Lacs and 589 FSRQs, of which the class type, redshift, and other parameters are presented in Table 1.

We note that there is only a blazar having extremely large synchrotron peak frequency above >1020 Hz in the sample: 3FGL J1504.5–8242 (1RXS J150537.1–824233) having $\mathrm{log}{\nu }_{{\rm{s}}}^{{\rm{p}}}(\mathrm{Hz})=20.12$, which is a BCU (Acero et al. 2015; Fan et al. 2016). It has a γ-ray photon index Γγ = 2.3 (Acero et al. 2015), which is a typical value of FSRQs. According to the above criterion, this source is classified as an FSRQ. Its synchrotron peak frequency may be overestimated because the quality of its synchrotron broadband SED is bad.5 For this source, the synchrotron peak frequency is assumed to be an average value of FSRQs: $\mathrm{log}{\nu }_{{\rm{s}}}^{{\rm{p}}}(\mathrm{Hz})=13.94$ (see Fan et al. 2016).

A one-zone synchrotron + IC model is adopted in this paper to estimate the jet physical parameters. This model assumes a homogeneous and isotropic emission region, which is a sphere with radius R, a uniform magnetic field with strength B, and a uniform electron energy distribution N(γ). The emission region moves relativistically with a Lorentz factor ${\rm{\Gamma }}=1/\sqrt{1-{\beta }^{2}}$ and a viewing angle θ, which forms the Doppler factor $\delta \,=1/[{\rm{\Gamma }}(1-\beta \cos \theta )]$. The frequency and luminosity transform from jet to AGN frames as6 ν = δν' and $\nu L(\nu )={\delta }^{4}\nu ^{\prime} L^{\prime} (\nu ^{\prime} )$, respectively. As stressed in Tavecchio et al. (1998), the jet parameters specifying the one-zone model are uniquely determined once the basic observables and variability timescale are known. Therefore, along the lines of the analytical treatment in Tavecchio et al. (1998) for the one-zone SSC model, it is possible to derive a useful approximate analytical expression for the jet physical parameters as a function of the observed SED quantities (e.g., the peak frequency and luminosity of the synchrotron and IC components). Because the synchrotron emission at the peak frequency is usually optical thin, we have

Equation (1)

where the emitting efficient is

Equation (2)

Here ${P}_{\mathrm{sy}}(\nu ^{\prime} ,\gamma )$ is the synchrotron emission power of a single electron. Because the SED of the synchrotron emission of a single electron is very narrow in the frequency space (see, e.g., Blumenthal & Gould 1970; Rybicki & Lightman 1979), one sometimes assumes that all emission is produced at a particular frequency, i.e., monochromatic approximation (see Chen 2017 for details), using the following equation as an approximation of ${P}_{\mathrm{sy}}(\nu ^{\prime} ,\gamma )$:

Equation (3)

where νL = eB/2πmec is the Lamer frequency and UB = B2/8π is the magnetic field energy density. In this case, the synchrotron luminosity (Equation (1)) will be reduced to

Equation (4)

where $\nu ^{\prime} =(4/3){\nu }_{L}{\gamma }^{2}$.

In leptonic blazar jet models, the synchrotron and IC emission components are radiations of nonthermal electron populations that are assumed to be isotropic in the jet fluid frame. One technique is to fit the data by injecting power-law electron distribution and allowing the electrons to evolve in response to radiative and adiabatic energy losses (e.g., Böttcher & Chiang 2002; Katarzyński et al. 2003; Moderski et al. 2003). In this case, many parameters must be specified, including the cutoff energies, injection indexes, and power, but this method is potentially useful to follow the dynamic spectral behavior of blazars. Contrary to this approach, we abandon any preconceptions about particle acceleration and employ the simplest functional form that is able to provide reasonably good fits to the SED data (see, e.g., Finke et al. 2008 for SSC modeling of TeV blazars with power-law electron distributions). For this purpose, and with the goal of minimizing the number of free parameters, a three-parameter log-parabolic function is employed to describe electron energy distribution in this paper (see Dermer et al. 2014; Fan et al. 2016):

Equation (5)

This electron energy distribution is only phenomenologically assumed to follow the log-parabola, without taking self-consistently into account the evolution due to injection and cooling effects (e.g., Kirk et al. 1998; Tramacere et al. 2011). Despite this, the log-parabolic model can successfully represent the broadband SED of blazars in both observation (single-source and sample studies; e.g., Massaro et al. 2004; Chen 2014; Dermer et al. 2015; Fan et al. 2016; Krauß et al. 2016; Xue et al. 2016) and theoretical SED modeling (e.g., Tramacere et al. 2011; Cerruti et al. 2013; Dermer et al. 2014; Yan et al. 2015; Ding et al. 2017; Hu et al. 2017). As for why there is a term (γ/γ0)−3, it is to guarantee that the electrons of γ0 emit at the peak frequency in the $\mathrm{log}\nu ^{\prime} -\mathrm{log}\nu ^{\prime} {L}_{\nu ^{\prime} }^{{\prime} }$ frame (see below). The entire description of the electron energy distribution is then given by three parameters: the normalization N0, the peak Lorentz factor γ0 of nonthermal electrons, and the spectral curvature parameter b.

The curvature, b, of the electron energy distribution has a relation with that of the SED (synchrotron bump), b ≈ 5P1 (see, e.g., Massaro et al. 2006; Chen 2014). In this case, the synchrotron peak frequency in the $\mathrm{log}\nu ^{\prime} -\mathrm{log}\nu ^{\prime} {L}_{\nu ^{\prime} }^{{\prime} }$ frame is

Equation (6)

and the corresponding peak luminosity is

Equation (7)

For the jet opening angle θj, the causality condition requires θjΓ ≲ 1 (Clausen-Brown et al. 2013), which is also supported by numerical simulations of axisymmetric, magnetically driven outflows (Komissarov et al. 2009), while constraints imposed by the SSC process (Nalewajko et al. 2014a) and observations of radio cores (e.g., Jorstad et al. 2005; Pushkarev et al. 2009) indicate θjΓ ≳ 0.1–0.7. In the calculations presented, we adopted a value of θjΓ = 1, and in this case, we have δ ≲ Γ (with a viewing angle of roughly θθj), of which the exact values will not significantly affect our conclusions.

The SSC emission may account for the emission at the peak of the IC component of BL Lacs, while the EC emission may account for that of FSRQs. The emissions at the IC peak can be at the Thomson or Klein–Nishina (KN) regimes. We will discuss these cases separately. The variability timescale can set an upper limit on the size of the emission region due to the causality, RδcΔt/(1 + z). The variability timescales are not required to be the same in various sources (see Ulrich et al. 1997 for a review; e.g., Bonnoli et al. 2011 and Abdo et al. 2011a for the well-studied blazars 3C 454.3 and Mrk 421; and Nalewajko 2013 for a systematic study indicating a typical variability timescale in the source frame in the Fermi/LAT band of ≈1 day; see also Hu et al. 2014), while for simplicity, the average value Δt/(1 + z) ≈ 1 day (in the source frame) is assumed in our calculation (e.g., Kang et al. 2014; Zhang et al. 2015).

3.1. SSC at the Thomson Regime

In the case of Thomson scattering, the SSC peak frequency follows

Equation (8)

Integrating the monochromatic luminosity (see Equations (4) and (5)), one can get the total synchrotron luminosity,

Equation (9)

Within the one-zone model, the average synchrotron energy density is (see Chen 2017)

Equation (10)

Therefore, the SSC peak luminosity follows

Equation (11)

Combining Equations (6), (9), (8), (11), and

Equation (12)

we have

Equation (13)

3.2. SSC at the KN Regime

In the case of KN scattering (${\gamma }_{0}h{\nu }_{\mathrm{sy}}^{{\prime} p}\gtrsim {m}_{e}{c}^{2}$), the SSC peak frequency follows

Equation (14)

The effective SSC seed photon energy density can be easily derived through integrating from the minimum frequency to the critical frequency ${\nu }_{\mathrm{sy}-c}^{{\prime} }=(\sqrt{3}/2)({m}_{e}{c}^{2}/h{\gamma }_{0})$,

Equation (15)

where

Equation (16)

In the case of the KN regime, the SSC peak luminosity follows

Equation (17)

Because Equation (17) is actually an integral equation (tc is a function of δ through ${\nu }_{\mathrm{sy},\mathrm{ssc}}^{{\prime} p}={\nu }_{\mathrm{sy},\mathrm{ssc}}^{p}/\delta $), the final estimated jet parameters cannot be analytically expressed as in Equation (13). The parameters (γ0, N0, B, δ, and R) can be derived through numerically solving the set of Equations (6), (9), (12), (14), and (17).

3.3. EC at the Thomson Regime

The FSRQs usually have very strong emission from the BLR and dust torus, and these photons can be IC scattered by relativistic electrons in the jet. In the frame of the jet, the external photon energy density will be enhanced and the frequency will be amplified: ${U}_{\mathrm{ext} \mbox{-} \mathrm{Th}}^{{\prime} }\approx (17/12){{\rm{\Gamma }}}^{2}{U}_{\mathrm{ext}}$ and ${\nu }_{\mathrm{ext}}^{{\prime} }\approx {\rm{\Gamma }}{\nu }_{\mathrm{ext}}$. In this case, the EC peak luminosity follows

Equation (18)

and the EC peak frequency,

Equation (19)

Combining Equations (6), (12), (18), and (19), we have (see also Chen & Bai 2011)

Equation (20)

This means that the external photon properties (i.e., ${U}_{\mathrm{ext}}/{\nu }_{\mathrm{ext}}^{2}$) will be determined when given the synchrotron/EC peak frequency and luminosity. For the BLR, the emission is mainly contributed by Lyα lines, νext-BLR ≈ 2 × 1015 Hz (Ghisellini & Tavecchio 2009). The dust torus reprocesses the disk emission into the IR band. As indicated by Spitzer observations, the typical peak frequency of IR dust torus emission is around νext-IR ≈ 3 × 1013 (Cleary et al. 2007; Ghisellini & Tavecchio 2009; Gu 2013). Following Ghisellini & Tavecchio (2009), the radiation from the reprocessed dust torus (IR) or BLR is described as a blackbody spectrum. The issue of whether the radiation regions are inside or outside the BLRs is highly debated. Considering the γ-ray photons can be absorbed via photon–photon pair production above ≳10 GeV of BLR emission (e.g., Liu & Bai 2006; Bai et al. 2009), some authors suggested that the emission region should be outside the BLR (e.g., Sikora et al. 2009; Tavecchio & Mazin 2009; Tavecchio et al. 2013). The broadband SED modeling of a large sample also suggests that the emission region may be outside the BLR for most sources (e.g., Kang et al. 2014). In this case, the IC/dust process will dominate the γ-ray emission (Ghisellini & Tavecchio 2009). In this paper, the IC/dust torus models are considered in our calculation.

We assume a Doppler factor of $\delta =1/{\rm{\Gamma }}(1-\beta \cos \theta )\approx {\rm{\Gamma }}$ for the relativistic jet close to the line of sight in blazars with a viewing angle θ ≲ 1/Γ (Jorstad et al. 2005; Pushkarev et al. 2009). Combining Equations (6), (9), (11), (12), (18) (or (20)), and (19), the jet parameters can be derived (the SSC emission in this case is within the Thomson regime),

Equation (21)

3.4. EC at the KN Regime

In the case of the KN regime (γ0Γextmec2), the EC peak frequency is

Equation (22)

The effective EC seed photon energy density can be easily calculated through integrating from the minimum frequency to the critical frequency ${\nu }_{\mathrm{ext} \mbox{-} {\rm{c}}}^{{\prime} }=(\sqrt{3}/2)({m}_{e}{c}^{2}/h{\gamma }_{0})$,

Equation (23)

where7

Equation (24)

In the case of the KN regime, the EC peak luminosity follows

Equation (25)

Combining Equations (6), (9), (11), (12), and (22), we have (the SSC in this case is within the Thomson regime)

Equation (26)

After getting these parameters, the external energy density, Uext, can be easily derived through combining Equations (12) and (25).

As discussed above, we use the SSC model for all BL Lacs and the EC model for all FSRQs. The input parameters include ${\nu }_{\mathrm{sy}}^{{\rm{p}}}$, ${\nu }_{\mathrm{ssc}/\mathrm{ec}}^{{\rm{p}}}$, ${\nu }_{\mathrm{sy}}^{p}L({\nu }_{\mathrm{sy}}^{p})$, ${\nu }_{\mathrm{ssc}}^{p}L({\nu }_{\mathrm{ssc}}^{p})$, and P1. The criterion from the Thomson to KN regimes in the SSC model is ${\gamma }_{0}h{\nu }_{\mathrm{sy}}^{{\prime} p}\gtrsim {m}_{e}{c}^{2}$ (see Section 3.2), and that in the EC model is ${\gamma }_{0}{\rm{\Gamma }}h{\nu }_{\mathrm{ext}}\gtrsim {m}_{e}{c}^{2}$ (see Section 3.4).

4. Results and Discussion

The COS B satellite first detected 3C 273 as a γ-ray blazar (Swanenburg et al. 1978). After that, nearly 100 blazars were discovered by the Energetic Gamma-Ray Experiment Telescope (EGRET) onboard the Compton Gamma Ray Observatory (Hartman et al. 1999; Nandikotkur et al. 2007). The 20-fold (sensitivity) improvement of Fermi/LAT has now detected more than 1000 blazars, which allows us to do population studies. For all 1392 γ-ray blazars in our sample, we estimate the jet physical parameters.

We first calculate the Doppler factor (δ) and the size of the emission region (R), for which (histogram) distributions are presented in Figure 3. It can be seen that R and δ follow the same distributions, which is due to the assumptions RδcΔt/(1 + z) and Δt/(1 + z) ≈ 1 day in our calculation. The median values of the Doppler factors of FSRQs, BL Lacs, and total blazars are δ ≃ 10.7, 22.3, and 13.1, respectively. The significantly larger δ of BL Lacs relative to that of FSRQs seems to be inconsistent with estimations from other methods. The independent methods found that the Doppler factors of BL Lacs are comparable with or even smaller than those of FSRQs, such as from radio variability (Lähteenmäki & Valtaoja 1999; Fan et al. 2009; Savolainen et al. 2010; Liodakis et al. 2017), SED modeling (Ghisellini et al. 1998, 1993), pair production (Mattox et al. 1993; Fan et al. 2014), or apparent superluminal motions (Jorstad et al. 2005; Hovatta et al. 2009). From Figure 3, we find that some BL Lacs present extremely large values of δ, even reaching ∼104, which may be incorrect for blazars. The typical values of the Doppler factor in blazars range from a few to ∼50 based on various methods as mentioned above (e.g., Mattox et al. 1993; Ghisellini et al. 1993, 1998, 2014; Lähteenmäki & Valtaoja 1999; Jorstad et al. 2005; Fan et al. 2009; Hovatta et al. 2009; Savolainen et al. 2010; Fan et al. 2014; Ghisellini & Tavecchio 2015; Liodakis et al. 2017). There is also a small fraction of BL Lacs having very low values of the Doppler factor (δ < 1). In our model, the Doppler factor is assumed to be equal to the jet bulk Lorentz factor δ = Γ, which must be larger than 1. For these sources with extreme Doppler factors, the estimations of other jet parameters may also be incorrect. We calculate the median value of the Doppler factors of all BL Lacs with 1 < δ < 100, which gives δm ≃ 14.3. This median value is used for 268 BL Lacs with extreme Doppler factors (δ < 1 or δ > 100) and determining other jet parameters. In this case, δ is a new input parameter other than ${\nu }_{\mathrm{sy}}^{{\rm{p}}}$, ${\nu }_{\mathrm{ssc}}^{{\rm{p}}}$, ${\nu }_{\mathrm{sy}}^{p}L({\nu }_{\mathrm{sy}}^{p})$, ${\nu }_{\mathrm{ssc}}^{p}L({\nu }_{\mathrm{ssc}}^{p})$, and P1. Due to the large uncertainty of the estimation of ${\nu }_{\mathrm{ssc}}^{{\rm{p}}}$ (see Abdo et al. 2010a and discussion below), we use δ instead of ${\nu }_{\mathrm{ssc}}^{{\rm{p}}}$ as the input parameter to calculate other jet parameters. There are no FSRQs having δ > 100 and only three FSRQs having δ < 1. We use the same method to determine the other jet parameters of these three FSRQs (using δm ≃ 10.7, the median value of δ of all FSRQs with δ > 1). We calculate all jet parameters and present them in Table 1, where sources with δ < 1 or δ > 100 are labeled with an asterisk. The full version of Table 1 is available in online ASCII form, which can be downloaded publicly. In Figure 4, we present the (histogram) distributions of some jet parameters, and the median values of these are listed in Table 2.

Figure 3.

Figure 3. Distributions of δ and R. Left panel: The median values of the Doppler factor δ of FSRQs, BL Lacs, and total blazars are 10.7, 22.4, and 13.1, respectively. For sources with 1 < δ < 100 (i.e., within the two green dotted lines), the median values of δ of FSRQs and BL Lacs are 10.7 and 14.3, respectively. Right panel: The median values of the size of the emission region R of FSRQs, BL Lacs, and total blazars are 2.76, 5.79, and 3.38 × 1016 cm, respectively. For sources with 1 < δ < 100 (i.e., within the two green dotted lines), the median values of R of FSRQs and BL Lacs are 2.78 and 3.70 × 1016 cm, respectively. See also Table 2.

Standard image High-resolution image
Figure 4.

Figure 4. Distributions of some jet parameters. Upper left panel: The median values of the electron peak energy γ0 of FSRQs, BL Lacs, and total blazars are 1167.8, 12077, and 3646.1, respectively. Upper right panel: The median values of the strength of the magnetic field B of FSRQs, BL Lacs, and total blazars are 1.56, 0.119, and 0.446, respectively. Bottom left panel: The median values of the curvature b of FSRQs, BL Lacs, and total blazars are 0.6, 0.55, and 0.55, respectively. Bottom right panel: The median values of the jet power Pjet of FSRQs, BL Lacs, and total blazars are 20.0, 6.3, and 12.0 × 1045 erg s−1, respectively. See also Table 2.

Standard image High-resolution image

Table 2.  Median Values of Some Jet Physical Parameters

  ${\gamma }_{0}$ b R δ B ${P}_{\mathrm{jet}}$ η σ ${P}_{\mathrm{jet}}/{L}_{\mathrm{acc}}$ ${P}_{\mathrm{jet}}/{L}_{\mathrm{Edd}}$ ${L}_{\mathrm{disk}}/{L}_{\mathrm{Edd}}$
  (1) (2) (3) (4) (5) (6) (7) (8) (9) (10) (11)
FSRQ(t) 1167.8 0.6 2.78 10.7 1.56 20.0 57.2 6.42 0.768 0.382 0.148
BL Lac(t) 12077 0.55 3.70 14.3 0.119 6.3 230 0.0285
Total(t) 3646.1 0.55 3.70 14.3 0.446 12.0 73.9 0.640
FSRQ(z) 1075.9 0.65 2.93 11.3 1.54 22.2 57.2 6.42 0.768 0.382 0.148
BL Lac(z) 13888 0.55 3.70 14.3 0.0748 5.08 230 0.0121
Total(z) 2713.3 0.60 3.70 14.3 0.528 14.8 73.9 0.743
FSRQ(cz) 1050.8 0.65 2.96 11.4 1.54 22.2 57.2 6.43 0.768 0.382 0.148
BL Lac(cz) 14036 0.55 3.70 14.3 0.0768 5.11 213 0.0131
Total(cz) 2614.6 0.60 3.70 14.3 0.554 15.3 73.4 0.862

Note. Case (t) is for all sources in our sample. Case (z) is for sources having known redshifts and including BCUs, while case (cz) is for sources having known redshifts and excluding BCUs. Column (1) is the electron peak energy. Column (2) is the curvature of the electron energy distribution. Column (3) is the radius of the emission sphere in units of 1016 cm. Column (4) is the Doppler factor. Column (5) is the strength of the magnetic field in units of Gs. Column (6) is the jet (SED) power in units of 1045 erg s−1. Column (7) is the ratio of the jet SED power to the (extended) jet kinetic power, $\eta ={P}_{\mathrm{jet}}/{P}_{\mathrm{jet},\mathrm{ext}}$. Column (8) is the magnetization parameter $\sigma ={U}_{{\rm{B}}}/{U}_{{\rm{e}}}$. Column (9) is the jet power in units of accretion power. Column (10) is the jet power in units of Eddington power. Column (11) is the disk luminosity in units of Eddington power.

Download table as:  ASCIITypeset image

The upper left panel in Figure 4 is the distribution of the electron peak energy γ0. The red line refers to FSRQs, the blue line indicates BL Lacs, and the black dashed line is for total blazars. It clearly shows two separated populations/dichotomic distributions for FSRQs and BL Lacs, with median values γ0 ≃ 1167.8 and 12077, respectively, which are comparable to previous studies8 (e.g., Ghisellini & Tavecchio 2015; Zhang et al. 2015; Qin et al. 2018). The median value for total blazars is γ0 ≃ 3646.1. These separated populations between FSRQs and BL Lacs are actually results from the separated distribution of synchrotron peak frequency (also for the IC component) between FSRQs and BL Lacs, which are presented in Figure 5. These separated populations may reflect that FSRQs and BL Lacs are different in nature. As discussed above, the high accretion rate makes FSRQs usually luminous AGNs having very strong line/dust torus emission. Therefore, nonthermal electrons in the jets of FSRQs will suffer efficient cooling and attain a smaller typical energy. The low accretion rate makes BL Lacs low-luminosity AGNs, with no or very weak emission lines. Therefore, the electrons in BL Lac jets suffer less cooling, which leads to a larger typical energy. These results are consistent with the so-called blazar sequence (see, e.g., Fossati et al. 1998; Ghisellini et al. 1998; Chen & Bai 2011). Blazars were classified as low-, intermediate-, and high-synchrotron-peaked (LSP, ISP, and HSP) blazars based on their synchrotron peak frequency by Abdo et al. (2010b) and later by Fan et al. (2016) based on their Bayesian classification of peak frequency for a larger sample of Fermi/LAT-detected blazars. From LSP to ISP to HSP blazars, the luminosity forms a gradual decreasing trend, which is called the blazar sequence (Fossati et al. 1998; Ghisellini et al. 1998; Chen & Bai 2011). However, some authors claimed that this sequence was from a selection effect, because after being corrected by the Doppler-beaming effect, the negative correlation between the peak frequency and luminosity disappeared (see Nieppola et al. 2008; Wu et al. 2009; Fan et al. 2017). The IC emission of smaller-energy electrons in FSRQs corresponds to softer γ-ray emission, while the higher-energy electrons in BL Lacs produce harder γ-ray emission. This is consistent with the fact that FSRQs and BL Lacs are distributed at two separate areas in the Lγ–Γ plane, as shown in Figure 2 (see also Abdo et al. 2009; Ghisellini et al. 2009; Ackermann et al. 2015).

Figure 5.

Figure 5. Distribution of synchrotron (left panel) and IC (right panel) peak frequency (measured in the AGN frame). The median values are νs = (0.689, 8.13, 2.48) × 1014 Hz and νIC = (0.692, 33.1, 5.75) × 1022 Hz for FSRQs, BL Lacs, and total blazars, respectively.

Standard image High-resolution image

The distribution of the magnetic field is shown in the upper right panel of Figure 4. The median values of FSRQs, BL Lacs, and total blazars are B ≃ 1.56, 0.119, and 0.446 Gs, respectively. These values are roughly consistent with the blazar sample SED modeling results (Ghisellini & Tavecchio 2015), in which the typical values of the magnetic field strength of FSRQs are found to range from 1 to 10 Gs and BL Lacs widely range from 0.001 to 10 Gs (see, e.g., Tavecchio et al. 2010; Zhang et al. 2012; Costamante et al. 2017).

The distribution of the curvature b of the electron energy distribution (see Equation (5)) is presented in the bottom left panel of Figure 4. Note that we calculate this parameter b = 5P1, where P1 is the curvature of the observed SED of the synchrotron component, whose values are given by Fan et al. (2016). The median value of FSRQs (b ≃ 0.6) is only slightly larger than that of BL Lacs (b ≃ 0.55), with the median of total blazars b ≃ 0.55 (see Chen 2014).

The uncertainty of the estimation of these parameters comes from the input parameters. An important contribution comes from the IC peak frequency ${\nu }_{\mathrm{IC}}^{p}$, which is derived from the γ-ray photon index through $\mathrm{log}{\nu }_{\mathrm{IC}}^{p}=-4{\rm{\Gamma }}+31.6$ (Equation (5) in Abdo et al. 2010a). As proposed by Abdo et al. (2010a), the error of the estimated $\mathrm{log}{\nu }_{\mathrm{IC}}^{p}$ is about ∼0.7, which leads to uncertainty of the estimated jet parameters. For the SSC model, we have a corresponding uncertainty about ${\rm{\Delta }}\mathrm{log}{\gamma }_{0}\sim 0.35$, ${\rm{\Delta }}\mathrm{log}\delta \sim 0.35$, ${\rm{\Delta }}\mathrm{log}B\sim 1.05$, and ${\rm{\Delta }}\mathrm{log}R\sim 0.35$ (simply for Thomson scattering; see Equation (13)). For the EC model, we have a corresponding uncertainty about ${\rm{\Delta }}\mathrm{log}{\gamma }_{0}\sim 0.175$, ${\rm{\Delta }}\mathrm{log}\delta \sim 0.175$, ${\rm{\Delta }}\mathrm{log}B\sim 0.525$, and ${\rm{\Delta }}\mathrm{log}R\sim 0.175$ (simply for Thomson scattering; see Equation (21)).

4.1. Disk–Jet Connection and Jet Launching

The most promising scenario for launching astrophysical relativistic jets involves large-scale magnetic fields anchored in rapidly rotating compact objects. The idea of driving outflows by rotating magnetic fields, originally invented by Weber & Davis (1967) to explain the spindown of young stars, was successfully applied to pulsar winds (Michel 1969) and became a dominant mechanism in theories of relativistic jets in AGNs (Lovelace et al. 1987; Li et al. 1992; Vlahakis & Königl 2004) and γ-ray bursts (e.g., Spruit et al. 2001; Vlahakis & Königl 2001). Powerful jets in AGNs can be powered by the innermost portions of accretion disks and/or by rapidly rotating BHs (Blandford & Znajek 1977; Blandford & Payne 1982; Spruit 2010). The correlation between the jet and the accretion disk/BLR has been presented in many studies (e.g., Cao & Jiang 1999; Ho & Peng 2001; Wang et al. 2003; Ghisellini et al. 2014; Liu et al. 2014; Du et al. 2016; Sbarrato et al. 2016). Because the disk continuum and line emission are usually missing in BL Lacs, we study the jet–disk connection only for FSRQs through comparing the accretion (e.g., accretion disk/BLR/BH mass) with the jet properties.

The disk bolometric luminosity Lbol can be derived from a continuum luminosity based on the empirical relation (Shen et al. 2011). In order to avoid contamination by the nonthermal continuum, the accretion disk luminosity can be derived from the total luminosity of broad emission lines as a proxy of Ldisk ≈ 10LBLR (Calderone et al. 2013), and the total luminosity of broad emission lines can be reconstructed through visible broad lines (see, e.g., Francis et al. 1991; Vanden Berk et al. 2001). Hβ, Mg ii, and C iv (Shaw et al. 2012) are widely used to scale to the quasar template spectrum of Francis et al. (1991) to calculate the total BLR luminosity LBLR. In Francis et al. (1991), Lyα is used as a reference of 100, and the total relative BLR flux is 555.77, of which Hβ is 22, Mg ii is 34, and CIV is 63 (see also, e.g., Vanden Berk et al. 2001; Ghisellini et al. 2014). The BH mass and accretion rate are fundamental parameters of AGNs. The virial method is now the most widely used to estimate the BH mass (e.g., Gu et al. 2001; Wu et al. 2004; Greene & Ho 2005). The uncertainty of a BH mass derived in this way is a factor of ∼3–4 (Vestergaard & Peterson 2006). Other methods include using the relation between BH mass and stellar velocity dispersion (e.g., Wu et al. 2002) and between BH mass and host-galaxy bulge luminosity (e.g., Laor 2001). Shaw et al. (2012) reported the optical spectroscopy of a large sample of γ-ray-detected blazars and estimated their BH mass through the virial method. We collect their available data (BH mass and emission lines) for our sample and finally get 144 FSRQs. As in the method discussed above, we use their optical emission lines to scale to the quasar template spectrum to calculate the total BLR luminosity. Because the BLR luminosities derived from various emission lines are consistent with each other (e.g., Francis et al. 1991; Shen et al. 2011; Shaw et al. 2012), we use the average value of the BLR luminosity if more than one emission line is available. The disk luminosity is then derived by Ldisk ≈ 10LBLR. Shen et al. (2011) studied the properties of quasars in the Sloan Digital Sky Survey Data Release 7 catalog. They calculated the virial BH mass and bolometric luminosity Lbol through the correlation between Lbol and continuum luminosity as presented in Richards et al. (2006). The disk luminosity can be calculated Ldisk = Lbol/2 (see Calderone et al. 2013). We further compile available data from Shen et al. (2011) and get another 46 FSRQs with disk luminosity and 44 FSRQs with BH mass. We note that some famous FSRQs are not included in these samples, e.g., 3C 273 and 3C 454.3. We then search the literature for FSRQs with well-measured disk luminosity and BH mass (Xiong & Zhang 2014). Finally, we have 232 FSRQs with measured disk luminosity and 229 with measured BH mass; 222 FSRQs have both values. The three FSRQs with Doppler factors δ < 1 are not included in these sources.

Integrating the observed broadband SED, including both synchrotron and IC components, we can calculate the jet bolometric luminosity Lobs. Figure 6 shows Lobs as a function of Ldisk for the 232 FSRQs with well-measured disk luminosity. These show a robust correlation9 $\mathrm{log}{L}_{\mathrm{disk},46}=(0.788\,\pm 0.076)\mathrm{log}{L}_{\mathrm{obs},46}-(1.44\pm 0.14)$ (solid blue line) with a chance probability p = 4.76 × 10−15 (Pearson test). The black dotted line represents Lobs = Ldisk and the black dashed line Lobs = 100Ldisk, which indicates that the jet bolometric luminosity is of the order of ∼100 times the disk luminosity Lobs ≈ 100Ldisk. The entire radiation power of the jet can be estimated through Pr = 2LobsΓ2/δ4 ≈ 2Lobs/δ2, which is the power that the jet expends in producing the nonthermal radiation. Based on VLBI-detected superluminal motion, or the broadband SED model, the jet Doppler factor is found to be an order of δ ≳ 10 (e.g., Lister et al. 2013; Ghisellini & Tavecchio 2015). Our estimation gives a median value of δ = 10.7 for FSRQs (see the discussion above and/or the upper left panel of Figure 4). In this case, the entire jet radiation power will be on the order of the disk luminosity Pr ≈ Ldisk. The jet radiative efficiency is believed to be on the order of Pr/Pjet ∼ 10%, which holds for AGNs, γ-ray bursts, and even BH X-ray binaries (Nemmen et al. 2012; Zhang et al. 2013; Ma et al. 2014), which gives an inevitable consequence that the jet power, Pjet ≈ 10Pr, is larger than the disk luminosity, Pjet ≈ 10Ldisk. This suggests that the jet-launching processes and the way of transporting energy from the vicinity of the BH to infinity must be very efficient.

Figure 6.

Figure 6. Jet bolometric luminosity vs. disk luminosity. The solid blue line shows a robust correlation, $\mathrm{log}{L}_{\mathrm{disk},46}=(0.788\pm 0.076)\mathrm{log}{L}_{\mathrm{obs},46}-(1.44\pm 0.14)$, with a chance probability p = 4.76 × 10−15 through a Pearson test. The black dotted line represents Lobs = Ldisk, and the black dashed line represents Lobs = 100Ldisk.

Standard image High-resolution image

The relativistic jet may be driven from the rapidly rotating BH. Increasing the spin of the BH shrinks the innermost stable orbit while increasing the accretion disk radiative efficiency ${\eta }_{\mathrm{acc}}={L}_{\mathrm{disk}}/\dot{M}{c}^{2}$ to a maximum value ηacc ≈ 0.3 (Thorne 1974), where $\dot{M}$ is the mass accretion rate. Assuming the accretion disk radiative efficiency ηacc = 0.3, we calculate the accretion power ${L}_{\mathrm{acc}}=\dot{M}{c}^{2}={L}_{\mathrm{disk}}/{\eta }_{\mathrm{acc}}$ (same as that in, e.g., Ghisellini et al. 2014).

With the jet physical parameters, we calculate the jet power through10 ${P}_{\mathrm{jet}}=2\pi {R}^{2}c{{\rm{\Gamma }}}^{2}({U}_{{\rm{B}}}+{U}_{{\rm{e}}}+{U}_{{\rm{p}}})$, where Ue and Up are the electron and proton energy density in the jet comoving frame (see, e.g., Celotti & Ghisellini 2008; Ghisellini et al. 2014). We adopt the conventional assumption that jet power is carried by electrons and protons (with one cold proton per emitting electron). In this case, Ue and Up can be easily derived for the log-parabolic distribution of electrons (see Equation (5)): ${U}_{{\rm{e}}}=\int N(\gamma )\gamma {m}_{{\rm{e}}}{c}^{2}d\gamma ={N}_{0}{\gamma }_{0}^{2}{m}_{{\rm{e}}}{c}^{2}\sqrt{\pi \mathrm{log}10/b}{10}^{1/4b}$ and ${U}_{{\rm{p}}}\,={m}_{{\rm{p}}}{c}^{2}\int N(\gamma )d\gamma ={N}_{0}{\gamma }_{0}{m}_{{\rm{p}}}{c}^{2}\sqrt{\pi \mathrm{log}10/b}{10}^{1/b}$. The existence of electron–positron pairs would reduce the jet power. Note that the median value of γ0 = 1167.8 is near the mass ratio of proton to electron, mp/me, implying that the jet power will not reduce significantly when considering electron–positron pairs. In addition, pairs cannot largely outnumber protons, because otherwise the Compton rocket effect would stop the jet (see, e.g., Ghisellini & Tavecchio 2010).

In order to study the relations among the jet and accretion disk, we plot the accretion power as a function of the jet power in Figure 7 (upper panel). The solid blue line shows the best-fit relation11 : $\mathrm{log}{L}_{\mathrm{acc},46}$ $=\,(2.25\pm 0.27)\mathrm{log}{P}_{\mathrm{jet},46}-(0.34\pm 0.12)$ (the Pearson test shows a chance probability p = 2.85 × 10−11). The dashed black line is the equality line. This result shows that a significant number of FSRQs have jet power greater than the accretion power, supported by the distribution of the ratio of jet power to accretion power (Pjet/Lacc) as shown in the left panel of Figure 8, implying that accretion power is not sufficient to launch the jets. The production of such a large-power jet cannot be treated as a marginal by-product of the accretion disk flow and most likely is governed by the BZ mechanism (Blandford & Znajek 1977). The gravitational energy released from accretion cannot only be transformed into heat and radiation but also powers jets through the BP process (Blandford & Payne 1982). In addition, the gravitational energy can also amplify the magnetic field, allowing the field to access the rotational energy of the BH and transform part of it into powerful jets through the BZ process (Blandford & Znajek 1977). As revealed by recently general relativistic magnetohydrodynamical (MHD) numerical simulations (e.g., Tchekhovskoy et al. 2011; McKinney et al. 2012; Li 2014), the average outflowing jet/wind power can even exceed the total accretion power for the case of spin value a = 0.99 when the BZ process dominates. The magnetic flux required to explain the production of the most powerful jets has been found to agree with the maximum magnetic flux that can be confined on BHs by the ram pressure of "magnetically arrested disks" (MAD; Narayan et al. 2003). In recent years, the MAD scenario has been thoroughly investigated and is now considered to be the likely remedy for the production of very powerful jets (Tchekhovskoy et al. 2011; McKinney et al. 2012; Sikora & Begelman 2013). However, as numerical simulations suggest, the powers of the jets launched in the MAD scenario depend not only on the spin and magnetic flux but also on the disk's geometrical thickness (Avara et al. 2016), with the jet power scaling approximately quadratically with all of these quantities (Sikora 2016). This large power of the relativistic jet in some FSRQs suggests that the BZ process will dominate the jet launching, which confirms some previous studies (see, e.g., Ghisellini et al. 2014; Zamaninasab et al. 2014).

Figure 7.

Figure 7. Jet power Pjet vs. accretion power Lacc (upper panel), Eddington luminosity LEdd (middle panel), and Eddington ratio λ (bottom panel). The solid blue lines show robust correlations, $\mathrm{log}{L}_{\mathrm{acc},46}=(2.25\pm 0.27)\mathrm{log}{P}_{\mathrm{jet},46}-(0.34\pm 0.12)$, with a chance probability p = 2.76 × 10−11 (Pearson test) for upper panel and $\mathrm{log}{L}_{\mathrm{Edd},46}=(1.38\pm 0.14)\mathrm{log}{P}_{\mathrm{jet},46}+(0.215\pm 0.064)$, with a chance probability p = 2.98 × 10−14 (Pearson test) for the middle panel. The black dashed line is an equality line. There is almost no correlation between the jet power and Eddington ratio (chance probability p = 0.0354).

Standard image High-resolution image
Figure 8.

Figure 8. Distributions of Eddington ratio (left panel) and jet power in units of accretion power (Pjet/Pacc; middle panel) and Eddington power (Pjet/LEdd; right panel), of which the median values are 0.148, 0.768, and 0.382, respectively.

Standard image High-resolution image

The middle panel of Figure 7 indicates jet power as a function of Eddington luminosity, LEdd = 1.26 × 1038(M/M) erg s−1, for 229 sources with measured BH mass, which presents a significant correlation with the best linear fitting12 $\mathrm{log}{L}_{\mathrm{Edd},46}\,=(1.38\pm 0.14)\mathrm{log}{P}_{\mathrm{jet},46}+(0.220\pm 0.064)$ (solid blue line) and a chance probability p = 4.0 × 10−14 (the Pearson test). The black dashed line is the equality line, which shows that the jet power in units of Eddington luminosity is less than 1 for most sources. This is supported by the distribution of the ratio of jet power to Eddington luminosity, as shown in the right panel of Figure 8, with a median value of 0.382.

The bottom panel of Figure 7 shows a relation between the Eddington ratio (defined as λ = Ldisk/LEdd) and jet power, suggesting almost no correlation between these two parameters (the Pearson test gives r = 0.144 and p = 0.0325; see, e.g., Zhang et al. 2015 for similar results of FSRQs). This suggests that the accretion rate may not be very important for driving the jet. As discussed above, this is consistent with the idea that these jets may not be launched through the BP process, related to the accretion process, but rather through the BZ process extracting the rotational energy of the BH. Actually, BL Lacs are believed to have very weak disk continuum emission and low accretion rates compared with FSRQs (Cao 2002; Ghisellini et al. 2009; Xu et al. 2009; Falomo et al. 2014). BL Lacs have comparable (very slightly lower) jet powers with FSRQs (the median values Pjet = 20.0, 6.30, and 12.0 × 1045 erg s−1 for FSRQs, BL Lacs, and total blazars, respectively; see discussion next subsection, Table 2, and/or the bottom right panel of Figure 4). Therefore, one may expect that the noncorrelation between Eddington ratio and jet power will extend down to the low Eddington ratio tail when considering BL Lacs.

The distribution of the Eddington ratio (Ldisk/LEdd) is presented in the left panel of Figure 8 with a median value of 0.148, consistent with previous results (e.g., Ghisellini et al. 2014). This implies that FSRQs have the standard geometrically thin, optically thick accretion disk (Shakura & Sunyaev 1973), which can launch powerful jets, contrary to some expectations (e.g., Livio et al. 1999; Meier 2002).

4.2. Jet Magnetization and Energy Transportation

The most diverse opinions about the nature of AGN relativistic jets concern their magnetic field and magnetization σ = PB/Pkin, where PB and Pkin are the magnetic and kinetic power, respectively. As discussed above, the current picture describing the powerful jet launching is that the jet can be produced from the central accretion system through the BZ mechanism by extracting BH rotation energy. This process is attributed to a key role of the dynamically important magnetic field, by means of which the BH spin energy is extracted and channeled into a Poynting/magnetic flux (Blandford & Znajek 1977; Tchekhovskoy et al. 2009, 2011). This magnetic flux cannot be developed by dynamo mechanisms in standard radiation-dominated accretion disks (Ghosh & Abramowicz 1997). However, such a flux is expected to be accumulated in the inner regions of the accretion flow and/or on the BH by the advection of magnetic fields from external regions (see Cao & Spruit 2013; Sikora & Begelman 2013 and references therein). In fact, the magnetic flux close to the BH horizon is so large that accretion likely occurs through an MAD flow, as discussed above (Narayan et al. 2003; Tchekhovskoy et al. 2011; McKinney et al. 2012). Starting as Poynting flux-dominated outflows, the jets are smoothly accelerated as magnetic power is being progressively converted to kinetic power and their magnetization drops, until a substantial equipartition between the magnetic and kinetic power is established (σ ≈ 1; see, e.g., Komissarov et al. 2007; Tchekhovskoy et al. 2009; Vlahakis 2015 and references therein). Therefore, at the end of this acceleration phase, the jet should still carry a substantial fraction (≈half) of its power in the form of a Poynting flux. The jet acceleration becomes inefficient when σ ≲ 1, from which point on, in the ideal MHD picture, σ will decrease logarithmically (Lyubarsky 2010).

However, even though blazars can have jets originating from MADs and be powered by the BZ mechanism, at the same time, they can have σ ≪ 1 at the radio core and/or the blazar zone (i.e., the region where most of the radiation is produced, as indicated by blazar models with jet opening angles 1/Γ; Nalewajko et al. 2014b; Janiak et al. 2015; Zdziarski et al. 2015). Zdziarski et al. (2015) suggested that the magnetic-to-kinetic energy flux conversion is assumed to result from the differential collimation of poloidal magnetic surfaces, which is the only currently known conversion mechanism in steady-state, axisymmetric, and nondissipative jets13 (the conversion process can initially proceed quite efficiently; see also Tchekhovskoy et al. 2009; Lyubarsky 2010). Hence, it is likely that other mechanisms are involved in the conversion process working at σ ≲ 1, such as MHD instabilities (see Komissarov 2011 and references therein), randomization of magnetic fields (Heinz & Begelman 2000), reconnection of magnetic fields (Drenkhahn & Spruit 2002; Lyubarsky 2010), and/or impulsive modulation of jet production (Lyutikov & Lister 2010; Granot et al. 2011). Little is known about the feasibility and efficiency of these processes in the context of AGNs. The MHD instabilities can develop when σ drops to unity or even earlier if stimulated by high-amplitude fluctuations of the jet power and direction, which are predicted in the MAD model (e.g., McKinney et al. 2012). In this case, the magnetization can reach σ ≪ 1 even prior to the radio core and/or the blazar zone.

Studies of the σ parameter are important, not only for better understanding of the dynamical structure and evolution of relativistic jets but also because its value determines the dominant particle acceleration mechanism and its efficiency. Dissipation of part of the kinetic (through shocks) and/or magnetic (through reconnection) power leads to the acceleration of particles up to ultrarelativistic energies, producing the nonthermal emission we observe from blazars. Both versions (the first/second order) of Fermi acceleration were, and still are, commonly invoked as responsible for the ultrarelativistic electrons in AGN jets (see, e.g., Hoshino et al. 1992; Sironi & Spitkovsky 2011; Marscher 2014). Based on so-called particle-in-cell simulations, it is shown that electrons can be efficiently heated in very low magnetized jets (σ ≪ 1; Madejski & Sikora 2016). The reconnection of magnetic energy is also a promising mechanism for particle acceleration in AGN jets (e.g., Giannios et al. 2009; Uzdensky 2011; Cerutti et al. 2012). A prediction of this scenario, resulting from detailed particle-in-cell simulations, is the substantial equipartition between the magnetic field and the accelerated electrons downstream of the reconnection site, where particles cool and emit the radiation we observe. For example, relativistic reconnection 2D models predict a lower limit of the order of σ ∼ 0.3 in the dissipation region (Sironi et al. 2015). Needless to say, the magnetic reconnection scenario requires that jets carry a sizable fraction of their power in the magnetic form up to the emission regions.

Based on the jet parameters estimated in this paper, the magnetization properties can be studied. Here we define a magnetization parameter as the ratio of energy densities between the magnetic field and relativistic electrons14 σ =UB/Ue (also called the equipartition parameter). The distribution of σ is shown in Figure 9. It can be seen that FSRQs have a narrower distribution than BL Lacs, with median values σ = 6.42 and 0.0285 (and 0.640 for total blazars), respectively (see also Table 2). The Kolmogorov–Smirnov (KS) test yields the significance level probability for the null hypothesis that FSRQs and BL Lacs are drawn from the same distribution p = 4.67 × 10−76 and the statistic DKS = 0.503 (the maximum separation of the two cumulative fractions). Similar results have been presented on single sources (Abdo et al. 2011b; Acciari et al. 2011; Aleksić et al. 2015) and through modeling the SEDs of the large sample blazars (Zhang et al. 2013; Tavecchio & Ghisellini 2016), which demonstrate that the magnetization parameter of the large majority of BL Lacs is commonly at ∼10−2. Costamante et al. (2017) studied six hard-TeV BL Lacs and found that the equipartition parameters range from 10−3 to 10−5. Kharb et al. (2012) presented deep Chandra/ACIS observations and Hubble Space Telescope (HST) Advanced Camera for Surveys observations of two FSRQs, PKS B0106+013 and 3C 345, and found the shocked jet regions upstream of the radio hot spots, the kiloparsec-scale jet wiggles, and a "nose cone"-like jet structure in PKS B0106+013, as well as the V-shaped radio structure in 3C 345, which suggest that the jet still has a large σ at these large scales. Modeling of FSRQ SEDs is also in agreement with this result (e.g., Ghisellini & Tavecchio 2015).

Figure 9.

Figure 9. Distribution of the magnetization parameters σ = UB/Ue of a total of 1392 blazars with median values of FSRQs, BL Lacs, and total blazars of 6.42, 0.0285, and 0.640, respectively. The KS test yields the significance level probability for the null hypothesis that FSRQs and BL Lacs are drawn from the same distribution P = 4.67 × 10−76 and the statistic DKS = 0.503 (the maximum separation of the two cumulative fractions).

Standard image High-resolution image

The smaller σ in BL Lacs relative to FSRQs implies that BL Lacs jets would suffer deceleration more easily than those of FSRQs (Piner et al. 2008; Homan et al. 2015), although both present similar violent γ-ray emission from the central region. This is consistent with the idea that some BL Lacs show almost no superluminal motion in the Very Long Baseline Array (VLBA) scale (e.g., Mrk 501 and Mrk 421; Giroletti et al. 2004, 2006), while the multiwavelength SED and variability imply a highly Doppler-booting emission, which indicates that the jet has already been decelerated at that scale (e.g., Giroletti et al. 2004; Ghisellini et al. 2005). For some FSRQs (e.g., 3C 279 and 3C 345; Lobanov & Zensus 1999; Piner et al. 2003), the Monitoring of Jets in AGN with VLBA Experiments (MOJAVE) project shows that jets will still be accelerated at the VLBA scale (see, Homan et al. 2015). The low magnetization in BL Lacs is essential in order to avoid efficient cooling of relativistic electrons, which would lead to much larger values of γ0 of BL Lacs relative to FSRQs (see Table 2 and/or Figure 4). This also seems to imply that the magnetic reconnection cannot be strong in these emission regions. The particle acceleration in situ may originate from the energy dissipated in shocks and/or in boundary shear layers. The spine/layer jet structure has been revealed through many observational techniques and can contribute significant X-ray emission due to the enhancement of the seed photon energy density by relative opposite motions between the spine and layer (see Chen 2017 for a recent review).

So far, there is no observational evidence supporting the theoretical prediction that a high-magnetization jet will more easily transport energy to a large scale than a low-magnetization jet. As found above, FSRQs have relatively larger values of σ than BL Lacs, which simply predicts that the jet may transport energy more efficiently to a large scale in FSRQs than BL Lacs. Therefore, we may expect a larger extended jet kinetic power (relative to the core SED jet power) in FSRQs than BL Lacs. To answer this question, we should estimate the (extended) jet kinetic power of these sources.

The jet power is basically important for understanding jet formation, the jet–disk relation (Blandford & Znajek 1977; Blandford & Payne 1982; Gu et al. 2009; Spruit 2010; Li & Cao 2012; Narayan & McClintock 2012; Wu et al. 2013; Cao 2014), and the AGN feedback on structure formation (Best et al. 2007; Magliocchetti & Brüggen 2007). It is generally believed that the X-ray cavities are direct evidence of AGN feedback and provide a direct measurement of the mechanical energy released by the AGN jets (jet kinetic power) through the work done on the hot, gaseous halos surrounding them (e.g., Bîrzan et al. 2004; Allen et al. 2006; Cavagnolo et al. 2010). It is also found that the cavity kinetic power is correlated with the radio extended power (Rawlings & Saunders 1991; Bîrzan et al. 2004; Cavagnolo et al. 2010), and the scaling relationship is roughly consistent with the theoretical relation (e.g., Willott et al. 1999; Cavagnolo et al. 2010; Meyer et al. 2011). Therefore, in systems where the X-ray cavities are lacking or not observed, the radio extended luminosity is used to estimate the jet kinetic power. The relations presented in Willott et al. (1999) are commonly used to do this estimation (Cavagnolo et al. 2010). Based on this method, Meyer et al. (2011) presented the largest blazar sample with measured (extended) jet kinetic power, until now.

In this paper, we collect the (extended) jet kinetic power, Pjet,ext, from Meyer et al. (2011), resulting in a total of 159 blazars, including 76 BL Lacs and 83 FSRQs, whose data are presented in Table 1. To study the jet energy transportation, we define η as a ratio of the jet SED power (from the central nuclei, Pjet) to the (extended) jet kinetic power (Pjet,ext), η = Pjet/Pjet,ext. The distribution of η is shown in Figure 10 with median values η = 57.2, 230, and 73.9 for FSRQs, BL Lacs, and total blazars, respectively, which implies that the jet SED power is significantly larger than the (extended) jet kinetic power. Two reasons may account for this: (1) the (SED) jet power may represent jet power measured in AGN active state, while the (extended) jet kinetic power is the historically average power, and (2) most of the energy is dissipated at the central region before being transported to a large scale. We note that FSRQs have an average smaller η than BL Lacs (by about one order of magnitude). The KS test gives the significance level probability that FSRQs and BL Lacs are drawn from the same distribution P = 2.58 × 10−10 and the statistic DKS = 0.524. This suggests that relatively more energy is transported to a large scale in FSRQs than in BL Lacs. As discussed above, the central jets of FSRQs also have relatively higher magnetization (σ) than those of BL Lacs, suggesting that a higher-magnetization jet can more easily transport energy to a large scale. As discussed above, some BL Lacs have extremely estimated Doppler factors. We instead use the median value of δ to determine the other jet parameters for these sources. This uncertainty may affect our results. Therefore, we exclude all BL Lacs with δ > 100 or δ < 1 and find that the remaining 56 BL Lacs have median values of η = 230 and σ = 0.00389, which is also consistent with the above result that the higher-magnetization jet can more easily transport energy to a large scale. For further research, we plot η versus σ in Figure 11 for sources with 1 < δ < 100, which presents a significant anticorrelation. The solid black line shows the best linear fit, $\mathrm{log}\sigma =(8.54\,\pm 0.69)-(4.44\pm 0.33)\mathrm{log}\eta $, with a chance probability p = 1.23 × 10−26 (Pearson test). This result confirms the above result, suggesting that a higher-magnetization jet can more easily transport energy to a larger scale, which, for the first time, offers supporting evidence for the jet energy transportation theory.

Figure 10.

Figure 10. Distribution of η = Pjet/Pjet,ext for blazars having measured extended radio emission (159 blazars, including 83 FSRQs and 76 BL Lacs), with median values of 57.2, 230, and 73.9 for FSRQs, BL Lacs, and total blazars, respectively. The KS test gives the significance level probability that FSRQs and BL Lacs are drawn from the same distribution P = 2.58 × 10−10 and the statistic DKS = 0.524.

Standard image High-resolution image
Figure 11.

Figure 11. Jet magnetization parameters (σ = UB/Ue) vs. the ratio of (SED) jet power to (extended) kinetic jet power (η = Pjet/Pjet,ext). The solid black line presents a best linear fit for the total sources, $\mathrm{log}\sigma =-(4.44\pm 0.33)\mathrm{log}\eta +(8.54\pm 0.69)$, with a chance probability $p=1.23\times {10}^{-26}$ (Pearson test).

Standard image High-resolution image

Our sample include 393 unknown-redshift sources (157 BL Lacs and 236 BCUs), which is about 28% of the sample. We use the median values of known redshifts of blazars for these sources. In order to explore the possible effect of this replacement on our results, we recalculate the median values of all jet parameters of another two cases: case 1 uses known-redshift sources including BCUs, and case 2 uses known-redshift sources excluding BCUs. All of these values are presented in Table 2, where the label (z) is for case 1 and the label (cz) is for case 2. It can be seen that both cases are consistent with our above results.

5. Conclusion

In this paper, based on broadband SEDs, we estimate the jet physical parameters of 1392 γ-ray-loud AGNs (Ackermann et al. 2015; Fan et al. 2016). These values are presented in Table 1, and the median values are shown in Table 2, which are roughly consistent with previous studies. Out of these sources, the accretion disk luminosities of 232 sources and the (extended) kinetic jet powers of 159 sources are compiled from archived papers. The full version of the data is available online (publicly). The main results are summarized here.

  • 1.  
    We show that γ-ray FSRQs and BL Lacs are well separated by ${\rm{\Gamma }}=-0.127\mathrm{log}{L}_{\gamma }+8.18$ in the γ-ray luminosity versus photon index plane with a success rate of 88.6%. This criterion is employed to divide 311 BCUs into FSRQs or BL Lacs.
  • 2.  
    The peak electron energy γ0 forms two distinct distributions between FSRQs and BL Lacs (also for νs/νC distributions), implying that the electrons in FSRQ jets suffer strong cooling, while they suffer less cooling in BL Lac jets.
  • 3.  
    A significant number of FSRQs present a jet bolometric luminosity larger than the disk luminosity by ∼2 orders of magnitude, which implies a (SED) jet power (assuming typical values of the Doppler factor δ ∼ 10 and jet radiative efficiency ∼10%) larger than the disk luminosity by ∼10 times. This suggests that the jet-launching processes and the way of transporting energy from the vicinity of the BH to infinity must be very efficient.
  • 4.  
    Most FSRQs present a (SED) jet power larger than the accretion power, which suggests that the relativistic jet-launching mechanism is dominated by the BZ process, at least in these FSRQs.
  • 5.  
    The magnetization of jets in BL Lacs is significantly lower than that in FSRQs, which is consistent with the idea that BL Lac jets may be more easily decelerated than FSRQ jets.
  • 6.  
    The ratio of the (extended) kinetic jet power to the (SED) jet power in FSRQs is significantly larger than that in BL Lacs. There is a significant anticorrelation between the jet magnetization parameter and the ratio of the SED jet power to the (extended) kinetic jet power. These results, for the first time, provide supporting evidence for the jet energy transportation theory: a high-magnetization jet can more easily transport energy to a large scale than a low-magnetization jet.

It should be noted that the nonsimultaneity of the SED does not always have an optimal impact on the derived parameters (and therefore large uncertainties) for single sources, but this is compensated for by the large number of sources ensuring that, from the statistical point of view, they should be only statistically meaningful.

We thank the anonymous referee for insightful comments and constructive suggestions. We are grateful for help from Xinwu Cao, Shiyin Shen, and Zhaoming Gan. This work is supported by the CAS grant (QYZDJ-SSW-SYS023).

Footnotes

  • The radio-loudness parameter, the ratio of the radio flux at 5 GHz to the optical flux at the B band, $R\equiv {f}_{5\mathrm{GHz}}/{f}_{{\rm{B}}}\gt 10$ for radio-loud AGNs (Kellermann et al. 1989).

  • BCU refers to blazar candidates of uncertain type (Ackermann et al. 2015).

  • The prime refers to values measured in the jet frame.

  • The factor 2.82 accounts for the difference between the peak frequency of the blackbody spectrum and the kT/h.

  • The median value of γ0 ≃ 1167.8 of FSRQs in this paper is slightly larger than the previous result, where γ0 ranges from 100 to 1000 (e.g., Ghisellini & Tavecchio 2015; Zhang et al. 2015).

  • Ldisk,46 = (Ldisk/1046) erg s−1 and Lobs,46 = (Lobs/1046) erg s−1.

  • 10 

    The factor of 2 accounts for the two jets.

  • 11 

    Lacc,46 = (Lacc/1046) erg s−1 and Pjet,46 = (Pjet/1046) erg s−1.

  • 12 

    LEdd,46 = (LEdd/1046) erg s−1.

  • 13 

    If there is a dissipative mechanism, the magnetic energy could be converted to thermal and kinetic energy.

  • 14 

    Note that the definition of σ = UB/Ue here is different from PB/Pkin =UB/(Ue + Up). For typical jet parameters (median values), Up/Ue = 103/4b(mp/me)/γ0 ≈ 28.0, 3.51, and 11.6 for FSRQs, BL Lacs, and total blazars.

Please wait… references are loading.
10.3847/1538-4365/aab8fb