Fast-spinning Black Holes Inferred from Symmetrically Limb-brightened Radio Jets

, , , , and

Published 2018 November 26 © 2018. The American Astronomical Society. All rights reserved.
, , Citation Kazuya Takahashi et al 2018 ApJ 868 82 DOI 10.3847/1538-4357/aae832

Download Article PDF
DownloadArticle ePub

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

0004-637X/868/2/82

Abstract

This paper theoretically investigates the relations between the structure of relativistic jets and produced synchrotron images, by using a steady, axisymmetric force-free jet model. We especially focus on the limb-brightened jets that are largely symmetric to the jet axes and observed in some active galactic nuclei, such as M87, Mrk 501, Cyg A, and 3C84. We find that symmetrically limb-brightened images can be produced when magnetic field lines of the jet penetrate a fast-spinning black hole (BH), as motivated by the Blandford–Znajek mechanism. On the other hand, jets with magnetic field lines that pass through a slowly spinning BH or the Keplerian accretion disk produce highly asymmetric radio images. In addition, the edge of a counterjet tends to be luminous in the accretion-disk model even for rather small viewing angles, which may be problematic for some observed jets. We also suggest that the site of particle accelerations in relativistic jets can be constrained by fitting the radio images to observations. This kind of study focusing on the jet images far away from the central engine is complementary to those concentrating directly on the innermost region with upcoming data from the Event Horizon Telescope.

Export citation and abstract BibTeX RIS

1. Introduction

The launching mechanism of collimated relativistic outflows (jets) is one of the mysteries in astrophysics. They are observed in active galactic nuclei (AGNs) and microquasars and are most probably associated with gamma-ray bursts and some tidal disruption events. They are believed to be launched from a system with a black hole (BH) and an accretion disk. In particular, AGN jets are widely thought to be electromagnetically launched along globally ordered magnetic field lines from the BHs via the Blandford–Znajek (BZ) mechanism (Blandford & Znajek 1977) or from the accretion disks via a unipolar induction mechanism (Blandford & Payne 1982). While the BZ mechanism effectively works to drive a Poynting-flux-dominated jet in general relativistic magnetohydrodynamics (GRMHD) simulations (McKinney & Gammie 2004; Komissarov 2005; Barkov & Komissarov 2008; McKinney & Blandford 2009; Tchekhovskoy et al. 2011; Qian et al. 2018) and in those with radiation (GRRMHD simulations; McKinney et al. 2014; Sadowski et al. 2014), the real foot point of astrophysical jets is yet to be confirmed from observations. Observational evidence for the BZ mechanism, if any, would support the existence of an ergosphere (Komissarov 2004; Toma & Takahara 2014, 2016; Kinoshita & Igata 2017).

Radio observations with very long baseline interferometry (VLBI) techniques can now resolve a jet at the very vicinity of the central BH (Junor et al. 1999). Recently, Hada et al. (2016) revealed an evident limb-brightened feature of the jet of M87 at ∼0.5 mas from the BH, which corresponds to 140–280 Schwarzschild radii (${r}_{g}=2{{GM}}_{\mathrm{BH}}/{c}^{2}$) for the distance to M87 (D = 16.7 Mpc; Blakeslee et al. 2009) and the BH mass (${M}_{\mathrm{BH}}\,\sim (3\mbox{--}6)\times {10}^{9}\,{M}_{\odot };$ Gebhardt et al. 2011; Walsh et al. 2013). The limb-brightened feature is largely symmetric to the jet axis and is observed with VLBI in the downstream at least up to $\sim {10}^{4}{r}_{g}$ (projected) from the center (Kovalev et al. 2007; Walker et al. 2008), while the feature is still less clear for the faint counterjet. We note that largely symmetric limb-brightened jets are also observed in other AGNs such as Mrk 501 (Giroletti et al. 2004, and references therein), Cyg A (e.g., Boccardi et al. 2016), and 3C84 (Nagai et al. 2014; Giovannini et al. 2018), whereas their spatial structures have been less resolved.

Theoretically, Broderick & Loeb (2009, hereafter BL09) proposed a steady, axisymmetric jet model to synthesize radio images of the M87 jet. They supposed a paraboloid-shaped force-free magnetic field that corotates with a Keplerian accretion disk at the equator. Their model succeeded in reproducing a jet length similar to observations and a dim counterjet for an assumed spatial distribution of the nonthermal electrons. However, the produced images do not show limb-brightened features but do illuminate the jet axis. While BL09 focused more on images of the BH shadow that will be detected by the Event Horizon Telescope (EHT; Doeleman et al. 2012; Akiyama et al. 2017), it will be important to ensure the consistency of the model with the downstream observations.

In this paper, we investigate the relations between the structure of relativistic jets and observed radio images. We employ the force-free paraboloidal jet model of BL09, which will be suitable at least for the M87 jet since the force-free approximation would be reasonable, especially in the base of the M87 jet (Kino et al. 2014, 2015), and the shape of the M87 jet can be reasonably fit by a parabola (Asada & Nakamura 2012; Hada et al. 2013; Nakamura et al. 2018). We introduce some new physics to the model of BL09. Motivated by the BZ mechanism, we newly consider jets with a rigidly rotating magnetic field, as well as those with Keplerian rotation. It is found that the difference in the jet launching point qualitatively changes the whole jet structure and leads to qualitatively and quantitatively different radio images even for the same distribution of emitting particles. We also try more general patterns of the distribution of nonthermal electrons, since it is not well constrained where and how particles are accelerated in relativistic jets. As shown later in this paper, we find that symmetrically limb-brightened features can be synthesized when the magnetic field lines penetrate a fast-spinning BH. Depending on the viewing angle, the counterjet becomes either luminous or dim. It is also shown, on the other hand, that symmetrically limb-brightened features cannot be produced when the magnetic field lines corotate with the Keplerian accretion disk, even if the nonthermal electrons are distributed on the jet edge. Since the jet model and the distribution of the nonthermal electrons are critical to producing BH shadows (Dexter et al. 2012; Mościbrodzka et al. 2016), this kind of study to constrain the jet base structure from the observational jet images at the far zone must be complementary to those employing the upcoming EHT data.

The paper is organized as follows. We briefly introduce our steady, axisymmetric force-free model in the next section, while the details are explained in Appendix A. Section 3 presents our calculated radio images for various parameter sets, where we fix some quantities to our fiducial values. The dependence on some of the fixed parameters is separately studied in Appendix B, though it does not affect our conclusions. We pay close attention to the difference between our force-free model and more realistic models by discussing in Section 4 how our synthesized radio images can change in a cold, ideal MHD treatment. Effects of the viewing angle are also discussed in the latter part of Section 4. We finally summarize and conclude our study in Section 5.

2. Method

To simulate radio emissions from relativistic AGN jets, we use an analytic model. Section 2.1 introduces the jet model including the magnetic and velocity fields, as well as the distribution of the nonthermal electrons. Section 2.2 explains the method to calculate a radio intensity map produced by synchrotron radiation. Section 2.3 is devoted to the strategy to choose our model parameters.

2.1. Force-free Jet Model

As shown below, our force-free model is essentially the same as in BL09. Although we employ a flat spacetime outside the BH, the magnetic field configuration is not much different from that with a general relativistic treatment even near the hole (McKinney & Narayan 2007). The resultant radio images will not be significantly changed as long as we focus on the limb-brightened features seen far from the central warped region. We put the detailed formulation in Appendix A and briefly explain the salient results below.

2.1.1. Electromagnetic Field

In a steady, axisymmetric force-free field, a stream function Ψ gives the electromagnetic field. Following BL09, we assume a parametrically controlled paraboloid-like jet instead of an exact solution of a force-free field. The stream function is given by

Equation (1)

In the above expression, $(r,\theta ,\phi )$ denote the standard spherical coordinates, and the minus and plus signatures are for $z\geqslant 0$ and z < 0, respectively. Here, ν is the parameter to control the jet shape, where ν = 1 gives paraboloidal jets, and A is a constant that has the dimension of $[{r}^{2-\nu }B]$, with B being the magnetic field. The electromagnetic field is then given by

Equation (2)

Equation (3)

Equation (4)

where B and E denote the magnetic and electric fields, respectively. Here, $(R,\phi ,z)$ are the standard cylindrical coordinates, the subscript p is assigned for the poloidal component, $\hat{{\boldsymbol{\phi }}}$ is the azimuthal unit vector, and ${{\rm{\Omega }}}_{{\rm{F}}}={{\rm{\Omega }}}_{{\rm{F}}}({\rm{\Psi }})$ corresponds to the rotational frequency of magnetic fields. It should be noted that the magnetic field is wound up and toroidal dominant, $B\sim | {B}_{\phi }| $, for $R| {{\rm{\Omega }}}_{{\rm{F}}}| /c\gg 1$, while it is poloidal dominant, $B\sim | {{\boldsymbol{B}}}_{p}| $, for $R| {{\rm{\Omega }}}_{{\rm{F}}}| /c\ll 1$.

2.1.2. Fluid Velocity

The fluid velocity cannot be determined in the force-free limit in principle, since the inertia is totally neglected. In this paper, we use the following drift velocity, v, as the fluid velocity by following BL09:

Equation (5)

The above velocity holds the following conditions for the electromagnetic field given by Equations (2)–(4): (1) the velocity does not exceed the speed of light in the entire region for $\nu \leqslant \sqrt{2}$, (2) the electric field vanishes in the fluid rest frame (the frozen-in condition), and (3) the velocity is asymptotically the same as in cold, ideal MHD when $R| {{\rm{\Omega }}}_{{\rm{F}}}| /c\gg 1$ (see Appendix A.3). Figure 1 sketches an example of the twisted magnetic and velocity field lines in a paraboloid-shaped jet (ν = 1). We note that the velocity is perpendicular to the magnetic field, while their poloidal components, ${{\boldsymbol{v}}}_{p}$ and Bp, are parallel to each other. The asymptotic relations of the velocity for $R| {{\rm{\Omega }}}_{{\rm{F}}}| /c\gg 1$ are given by

Equation (6)

Equation (7)

where β denotes the speed normalized by c, and $g(\theta ,\nu )$ is a factor that is an order of tenth and approaches unity toward the jet axis (θ = 0, π; see Figure 12). That is, the fluid velocity is dominated by the poloidal component and becomes relativistic around the jet axis if $R| {{\rm{\Omega }}}_{{\rm{F}}}| /c\gg 1$. For $R| {{\rm{\Omega }}}_{{\rm{F}}}| /c\ll 1$, on the other hand, the following relations are obtained:

Equation (8)

Equation (9)

That is, the fluid velocity is not relativistic and is dominated by the toroidal component.

Figure 1.

Figure 1. Example of field lines in a paraboloidal jet with ΩF > 0 (z ≥ 0). The red and green lines represent magnetic field lines and stream lines, respectively. The jet axis coincides with the z axis. The thick field lines originate from a point on the (x, y) plane, while the dashed lines stem from the centrosymmetric point with respect to the origin. For visibility in the figure, we omitted the field lines in the counterjet (z < 0), which has a symmetric structure with respect to the equatorial plane except in the direction of the poloidal magnetic field. Also plotted are (X, Y, Z) coordinates, where the X axis coincides with the x axis and the Z axis is inclined toward the −y direction. The observer is assumed to be in the Z direction, and the angle between the z and Z axes corresponds to the viewing angle Θ. Thus (X, Y) give the coordinates on the sky as viewed by the observer.

Standard image High-resolution image

2.1.3. Nonthermal Electrons

Motivated by the limb-brightened jets, we consider the case where the nonthermal electrons are distributed away from the jet axis, in contrast to BL09, who assumed the nonthermal electrons are clinging to the axis. Such a spatial distribution concentrated away from the axis could be realized for jets launched from an accretion disk and even for those launched from the BH. The particles are supplied from a disk at the jet foot point for the former case, while several options for the particle injection can be considered for the latter case. As shown in MHD simulations (McKinney & Gammie 2004; Komissarov 2005; Komissarov et al. 2007; Barkov & Komissarov 2008; McKinney & Blandford 2009; Tchekhovskoy et al. 2011; McKinney et al. 2014; Sadowski et al. 2014; Qian et al. 2018), the jets driven by the BZ mechanism are confined by the external pressure of the ambient matter, that is, a geometrically thick disk with an advection-dominated accretion flow (ADAF; Narayan & Yi 1994), or the disk wind. Thermal charged particles are prevented by the globally ordered magnetic field in the funnel region from diffusing into there from the ADAF, but high-energy hadrons can diffuse into there (Toma & Takahara 2012; Kimura et al. 2014, 2015), and high-energy photons can annihilate and supply ${e}^{-}{e}^{+}$ pairs there (Levinson & Rieger 2011; Mościbrodzka et al. 2011). The particles in a jet flow outward from the separation surface (the stagnation surface), which is much closer to the BH and the hottest part of the disk farther away from the jet axis (Takahashi et al. 1990; McKinney & Gammie 2004; Pu et al. 2015; Nakamura et al. 2018).7 Thus, the particle injection for the outflow can be dominated at the jet edge. A pair creation gap created around the separation surface could also be a particle supplier (Levinson & Rieger 2011; Broderick & Tchekhovskoy 2015). The fluid instability or magnetic reconnection at the layer between the jet and disk wind may also produce nonthermal particles (see Matsumoto & Masada 2013; Parfrey et al. 2015; Toma et al. 2017).

It is beyond the scope of this paper to discuss in detail the above injection and acceleration mechanisms (upcoming EHT data would shed light on those mechanisms). In this study, we simply assume that the spatial distribution of the nonthermal is described in a parametric way and the energy spectrum is given by a single power law. The spatial distribution is characterized by the cross sections at z = ±z1 for simplicity, where the electrons are assumed to be in a ring shape and the number density is given by

Equation (10)

where Rp is the radius where n peaks on the plane, and Δ gives the width of the ring; n0 is a normalization constant. Our prescription is identical to that in BL09 when Rp = 0 and ${\rm{\Delta }}={z}_{1}=5{r}_{g}$. We also set ${z}_{1}=5{r}_{g}$ hereafter while Rp and Δ remain as free parameters. In the vertical direction, n is assumed to obey the continuity equation (BL09), which is given as follows by using Equation (5):8

Equation (11)

We assume that the nonthermal electrons are isotropic in the fluid rest frame and obey an energy distribution of a single power law given by an index p:

Equation (12)

where γ' is the Lorentz factor of an electron measured in the proper frame, which has lower and higher cutoffs at ${\gamma }_{\min }^{{\prime} }$ and ${\gamma }_{\max }^{{\prime} }$, respectively. The synchrotron emissivity does not depend on ${\gamma }_{\max }^{{\prime} }$ but only on ${\gamma }_{\min }^{{\prime} }$ provided ${\gamma }_{\min }^{{\prime} }$ and ${\gamma }_{\max }^{{\prime} }$ are sufficiently small and large, respectively (see Appendix A.5). We can, hence, set ${\gamma }_{\max }^{{\prime} }=\infty $ for a large higher cutoff, while we use ${\gamma }_{\min }^{{\prime} }=100$ for the lower cutoff by following BL09. As in BL09, the energy distribution is fixed to Equation (12) everywhere, which means that some energy supplier is assumed to replenish high-energy electrons to compensate for cooling processes such as synchrotron and adiabatic cooling.

2.2. Synchrotron Radio Images

The quantities given by Equations (1)–(11) give the synchrotron emissivity at each location in a jet that is received by the observer at a frequency ω as follows (Rybicki & Lightman 1985; Shibata et al. 2003):

Equation (13)

where the quantities with prime symbols are evaluated in the fluid rest frame.9 Here, ${\boldsymbol{n}}$ is a unit vector that is directed to the observer at infinity, μ is the cosine of the angle between ${\boldsymbol{n}}$ and v, and ${\rm{\Gamma }}=1/\sqrt{1-{\beta }^{2}}$ is the Lorentz factor. The factor in the right-hand side is attributed to relativistic effects that are due to the bulk fluid motion. Note that ${j}_{\omega ^{\prime} }^{{\prime} }({\boldsymbol{n}}^{\prime} )$ is given by Equation (55).

In higher frequencies, radio jets in AGNs are optically thin to synchrotron emissions. The intensity of radio images observed on the sky is, then, calculated by integrating Equation (13) along the line of sight after fixing the viewing angle Θ:

Equation (14)

where (X, Y) are the coordinates of the sky and ${dZ}$ is the line element parallel to the line of sight. In the following, the X axis is chosen to coincide with the x axis, and the Z axis is inclined toward the −y direction so that the angle between the z and Z axes becomes Θ (see Figure 1). A simulated VLBI image is obtained after the convolution with a beam kernel, which is introduced in the next subsection.

2.3. Model Parameters

In our force-free model, 10 parameters remain: ΩF, Rp, Θ, Δ, ν, p, MBH, A, n0, and ω. We systematically change them and investigate the effects on our synthetic radio images. The first two parameters (ΩF and Rp) are especially important, since they can drastically change radio images, as shown in Section 3. The viewing angle Θ is found to be less important for limb-brightened features, while it can be important for the brightness ratio between the jet and counterjet (see Section 4.2). The choices of the other parameters do not qualitatively alter the synthetic images (see Appendix B).

We consider two patterns of ΩF. The first choice of ΩF is the same as in BL09, where the magnetic field is threaded through a razor-thin accretion disk at the equatorial plane. Since the field rotates with the disk, ΩF is given by

Equation (15)

where $\tilde{R}$ is the foot point radius of a given magnetic field line measured on the equatorial plane, and RISCO is the radius of the innermost stable circular orbit (ISCO) for prograde rotations. Here, ${{\rm{\Omega }}}_{\mathrm{Kep}}$ is the Keplerian angular frequency given by the dimensionless Kerr parameter, a, as follows (Bardeen et al. 1972):

Equation (16)

where ${r}_{G}:= {r}_{g}/2={{GM}}_{\mathrm{BH}}/{c}^{2}$ is the gravitational radius.

The other choice of ΩF is motivated by the BZ process, which was not considered in BL09. In the BZ process, ΩF is nearly a constant given by

Equation (17)

where ${{\rm{\Omega }}}_{\mathrm{BH}}$ is the rotational frequency of the Kerr BH and ${r}_{+}=(1+\sqrt{1-{a}^{2}}){r}_{G}$ is the horizon radius (Blandford & Znajek 1977; McKinney & Gammie 2004). We assume that the shape of the magnetic field lines changes above the equator so that the field lines penetrate the event horizon, while the shape far away from the equator is given by Equation (1). Such a field configuration may be possible, depending on the profile of the external pressure of the disk wind or corona, which collimates the jet and is responsible for the global jet shape (McKinney 2006; Nakamura & Asada 2013). Since the jet structure near the central region is not important for the limb-brightened feature observed far from the central BH, we use Equation (1) in the entire region for simplicity.

The ring radius (Rp) is systematically changed from zero to some sufficiently large value. For comparison to BL09, the other parameters are fixed to the fiducial values in BL09: ${M}_{\mathrm{BH}}=3.4\times {10}^{9}{M}_{\odot }$, ${\rm{\Delta }}=5{r}_{g}$, ν = 1, p = 1.1, and Θ =25°, which were chosen for the M87 jet. Accordingly, we henceforth consider M87, which is an example of an AGN that shows a symmetrically limb-brightened jet with a dim counterjet.10 We calculate jet images on the scale of several milliarcseconds, where the limb-brightened feature of the jet is observed with VLBI (Ly et al. 2007; Walker et al. 2008; Hada et al. 2011, 2013, 2016; Mertens et al. 2016). For the fiducial mass of the BH, 1 mas corresponds to ∼0.08 pc ∼250 rg for D = 16.7 Mpc (Blakeslee et al. 2009). We use the beam kernel for VLBA given in Walker et al. (2008) and assume that the M87 jet is inclined toward the northeast by 20° as measured from the east direction on the sky. We note that A and n0 are related only to the normalization of the intensity. We adopt the following values throughout the paper: ${{Az}}_{1}^{2-\nu }=100$ G, which corresponds to the strength of the magnetic field at (R, z) = (0, ±z1) (Kino et al. 2015); and n0 = 1 cm${}^{-3}$, which produces a peak intensity that is roughly consistent with observations of M87 in order of magnitude for our best model described in Section 3.2.1.11 A typical value of the magnetization factor $\sigma ={B}^{2}/(4\pi {\rm{\Gamma }}{{nm}}_{p}{c}^{2})$, where mp is the proton mass, is given by

Equation (18)

In fact, the force-free approximation σ ≫ 1 consistently holds for our jet models, as shown in the next section. We use the observed frequency of ω/(2π) = 44 GHz to synthesize the intensity maps, while the choice of ω does not affect the shape of radio contour maps under the optically thin assumption.12 We summarize the model parameters in Table 1. In Appendix B, some of the above parameters are varied around our fiducial values to study the effects on radio images; they do not change our conclusions.

Table 1.  Model Parameters

Quantity Symbol Fiducial value for Figures 210
Rotational frequency of the magnetic field ΩF Equation (15) (Keplerian) or Equation (17) (rigid)
Radius where n peaks on z = ±z1 Rp Varied in $[0,100{r}_{g}]$
Viewing angle Θ 25°
Width of the Gaussian ring Δ 5rg
Jet shape ν 1 (paraboloidal jet)
Energy spectral index of the nonthermal electrons p 1.1
Mass of the BH MBH $3.4\times {10}^{9}{M}_{\odot }$
Strength of the magnetic field at (R,z) = (0,±z1) ${{Az}}_{1}^{2-\nu }$ 100 G
Number density of the nonthermal electrons at $(R,z)=({R}_{p},\pm {z}_{1})$ n0 1 cm${}^{-3}$
Dimensionless Kerr parameter of the BH a Varied in $[0,0.998]$
Height of the plane where n is given in a ring shape by Equation (10) z1 5rg
Minimal Lorentz factor of the nonthermal electrons ${\gamma }_{\min }^{{\prime} }$ 100
Maximal Lorentz factor of the nonthermal electrons ${\gamma }_{\max }^{{\prime} }$ $\infty $
Observational frequency ω/(2π) 44 GHz
Luminosity distance to the jet D 16.7 Mpc
Inclination of the projected jet axis measured from the east direction   20° toward northeast
Beam kernel   Walker et al. (2008) (VLBA)

Download table as:  ASCIITypeset image

3. Results

3.1. Case 1: Disk-threaded Model

First, we examine the case of the disk-threaded model, in which the magnetic fields penetrate the Keplerian accretion disk and ΩF is given by Equation (15). We here show the results for a = 0.998, which is the fiducial value in BL09, since those for smaller a are qualitatively the same. This model has ${R}_{\mathrm{ISCO}}\sim 1.2{r}_{+}\sim 0.62{r}_{g}$ and ${{\rm{\Omega }}}_{\mathrm{Kep}}({R}_{\mathrm{ISCO}})\sim 2.5\times {10}^{-5}$ s−1.

The jet structure of this model is presented in Figure 2.13 The upper left panel shows the distribution of an important quantity, $R{{\rm{\Omega }}}_{{\rm{F}}}/c$ (see also Figure 18 for the close-up around the origin). As shown by the white lines, the so-called light "cylinder," where $R{{\rm{\Omega }}}_{{\rm{F}}}/c=1$ is satisfied, forms not only a vertical surface around the jet axis but also a curved one far from the jet axis (Blandford 1976). The former truncated cylinder is formed at $R={R}_{\mathrm{lc},1}:= c/{{\rm{\Omega }}}_{\mathrm{Kep}}({R}_{\mathrm{ISCO}})\sim 1.2{r}_{g}\sim 4.8\times {10}^{-3}$ mas due to the uniform rotation of the magnetic field passing through inside the ISCO. The latter curved surface ($z\propto {R}^{4/3}$ at $R\gg {r}_{G};$ see Appendix C) is, on the other hand, attributed to the differential rotation of the magnetic field lines anchored to the accretion disk. As a result, $R{{\rm{\Omega }}}_{{\rm{F}}}/c$ exceeds unity only in a limited region bound by these two surfaces. This means that the jet edge part is dominated by a poloidal magnetic field and is not efficiently accelerated to poloidal directions, as shown in the lower left and upper middle panels in Figure 2 (see Equations (6) and (8) for asymptotic relations between $| {{\boldsymbol{v}}}_{p}| $ and $R{{\rm{\Omega }}}_{{\rm{F}}}/c$). The highly relativistic poloidal speed is realized, on the other hand, only at $R\gtrsim {R}_{\mathrm{lc},1}$, which is near the jet axis. The jet rotational speed, vϕ, is shown in the upper right panel in Figure 2. Note that vϕ peaks ∼0.5c around the light cylinder and reduces away from it, as indicated in Equations (7) and (9). The lower middle panel in Figure 2 shows the Lorentz factor, which manifestly shows the jet is relativistic only near the jet axis, as explained above. The lower right panel shows the number density of the nonthermal electrons for Rp = 40rg (one of the fiducial cases) in the logarithmic scale. As is evident, the nonthermal electrons are concentrated on the magnetic field lines that pass through $R=40{r}_{g}$ at z = 5rg (${\rm{\Psi }}/A\sim 35.3{r}_{g};$ drawn as black lines in Figure 2), while the number density rapidly decreases away from the lines. Note that n is also reduced along a field line upward. We note that the magnetization factor σ given by Equation (18) is low at the dense region around the black lines but is much larger than unity in the displayed region (σ ≳ 8 × 103), which ensures the use of the force-free approximation for this model.

Figure 2.

Figure 2. Physical quantities in the jet (y = 0, z > 0) for Case 1. The jet is axisymmetric around the z axis, and the BH exists at the origin. The jet structure is symmetric with respect to the equatorial plane (z = 0). Upper left panel: color map of $R{{\rm{\Omega }}}_{{\rm{F}}}/c$. Upper middle: poloidal speed normalized by the speed of light, $| {{\boldsymbol{v}}}_{p}| /c$. Upper right: azimuthal speed normalized by the speed of light, ${v}_{\phi }/c$. Lower left: ratio of the toroidal and poloidal magnetic field strengths, $| {B}_{\phi }| /{B}_{p}$. Lower middle: the Lorentz factor, Γ. Lower right: number density of the nonthermal electrons for Rp = 40rg in the logarithmic scale, where the region with $\mathrm{log}n\lt -6$ is filled with the same color for visibility of the dense region. In these panels, the white lines are the light cylinder, and the black ones are the poloidal magnetic field lines that pass through $R=40{r}_{g}$ at z = 5rg. The jet was cut out along the magnetic field surface that goes through $R=150{r}_{g}$ at $z=5{r}_{g}$. Note that 1 mas corresponds to ∼250rg (i.e., $1{r}_{g}\sim 4\times {10}^{-3}$ mas).

Standard image High-resolution image

Figure 3 shows the beaming factor $\delta := 1/[{\rm{\Gamma }}(1-\beta \mu )]$ on some horizontal slices of the jet. The top and middle left panels are for the counterjet (z < 0), and the others are for the jet (z > 0) that is directed toward the observer with a viewing angle of Θ = 25°. The beaming effect becomes remarkable in the region with $R{{\rm{\Omega }}}_{{\rm{F}}}/c\gg 1$, where the poloidal speed becomes relativistic and the Lorentz factor is large. In the approaching jet side, the distribution of δ is highly asymmetric because of the jet rotation, which reaches ∼0.5c around the curved light "cylinder," as presented in the upper right panel in Figure 2. Through μ in Equation (13), jet rotations lead to the opposite effects of relativistic beaming in the left and right sides of the jet. The left side of the jet is coming toward the observer and, as a result, strongly beams light toward the observer, whereas the right side of the jet is going away from the observer and, hence, does not efficiently beam light to the observer. We note that the peaks of δ and Γ in each slice do not necessarily coincide, due to the misalignment of the observer and flow directions. In the counterjet side, on the other hand, δ is suppressed below unity almost in the entire region. The suppression is especially strong in the region with $R{{\rm{\Omega }}}_{{\rm{F}}}/c\gg 1$, and the asymmetry of δ due to the jet rotation is also seen as in the approaching jet side.

Figure 3.

Figure 3. Beaming factor δ = 1/[Γ(1−βμ)] for the observer with Θ = 25° on some horizontal slices of the approaching jets (z > 0) and counterjets (z < 0) of Case 1. The sliced plane is designated in the upper right corner in each panel. Same as in Figure 2, the white and black lines indicate the light cylinder and the magnetic field surface that passes through $R=40{r}_{g}$ at z = 5rg. Note that 1 mas corresponds to ∼250rg (i.e., $1{r}_{g}\sim 4\times {10}^{-3}$ mas).

Standard image High-resolution image

The left panel in Figure 4 shows the calculated radio image for Rp = 0. This model is essentially the same as the standard (M0) model in BL09, where the nonthermal electrons cling to the jet axis. As expected, the jet axis is the brightest, due to the concentration of electrons, and any limb-brightened feature is not seen. The counterjet is not seen in the radio map, due to the relativistic beaming to the direction opposite to the observer. We also note that the radio intensity is larger in the left-hand side of the jet in the figure because of the asymmetric beaming effect shown in Figure 3.

Figure 4.

Figure 4. Radio intensity maps for Case 1, where the magnetic field penetrates the Keplerian accretion disk. The unit of the intensity is milli-Jansky per beam. The contours are drawn as follows. The inner 20 contours are for ${\sqrt{2}}^{-k}$ ($k=0,\,\cdots ,\,19$), while the outermost two are for ${\sqrt{2}}^{-21}$ and $0.1{\sqrt{2}}^{-21}$, respectively. The Y axis coincides with the projected jet axis, and the origin is the projected location of the BH. The particle distributions are given by Rp = 0 and 40rg in the left and right panels, respectively, as designated above each panel. The beam shape is also plotted in gray at the top right corner in each panel. Note that 1 mas corresponds to ∼250rg (i.e., $1{r}_{g}\sim 4\times {10}^{-3}$ mas).

Standard image High-resolution image

We here pick up the result for Rp = 40rg, while the results for Rp > 0 are qualitatively the same, as mentioned later. In this example, the nonthermal electrons are nearly on the curved light-cylinder surface for $| z| \leqslant 10$ mas, as indicated by the magnetic field lines with ${\rm{\Psi }}/A\sim 35.3{r}_{g}$ (the black lines in Figures 2 and 3), which reasonably trace the dense region of the nonthermal electrons. The right panel in Figure 4 shows the synthesized radio image for Rp = 40rg. One of the most striking features is the strongly asymmetric limb brightening: the left-hand side of the jet axis in the figure (i.e., the northern part on the sky) is more luminous than the counterpart in the right-hand side (the southern part). The limb brightening is understood just as a reflection of the assumed Rp. The large asymmetry is, on the other hand, due to the rotation of the jet. As seen in Figure 3, the asymmetry of the beaming factor in the approaching jet side is relatively large near the magnetic field line ${\rm{\Psi }}/A\sim 35.3{r}_{g}$, where vϕ reaches (∼0.5c), as plotted in the upper right panel in Figure 2. We note that the synchrotron emission is intrinsically asymmetric even in the fluid rest frame, since the pitch angle of the relativistic electrons that are directed toward the observer is different between the right and left sides of the jet because of winding magnetic field lines, which is included in Equation (55) through $\sin \psi ^{\prime} ({\boldsymbol{n}}^{\prime} )$. The intrinsic asymmetry is, however, found to be minor compared to the asymmetry induced by the relativistic beaming.

The luminous counterjet is another notable feature for Rp = 40rg, as seen in the right panel in Figure 4. The counterjet becomes apparent, in contrast to the observations of the M87 jet (e.g., Hada et al. 2016), since the relativistic boost to poloidal directions is so weak. As shown in the upper middle panel in Figure 2, the poloidal speed on the magnetic field line ${\rm{\Psi }}/A\sim 35.3{r}_{g}$ is relatively small ($| {{\boldsymbol{v}}}_{p}| \lesssim \mathrm{0.7c}$). As a result of the decrease of $| {{\boldsymbol{v}}}_{p}| $ toward the jet edge, δ increases to unity toward the jet edge in the counterjet side, as shown in Figure 3. Note that δ is ∼0.5 on the magnetic field line ${\rm{\Psi }}/A\sim 35.3{r}_{g}$, which is not sufficient to darken the counterjet.

The results for other Rp > 0 are qualitatively the same. We confirmed that the radio images still keep the strong asymmetry for $0\lt {R}_{p}\lt 40{r}_{g}$, as indicated by the asymmetric candle-flame-like image with the brighter northern edge for Rp = 0 in the left panel in Figure 4. The counterjet for $0\lt {R}_{p}\lt 40{r}_{g}$ becomes less luminous than for Rp = 40rg owing to larger $| {{\boldsymbol{v}}}_{p}| $. We also found for ${R}_{p}\gt 40{r}_{g}$ that the asymmetry of the limb brightening can be weaker thanks to the smaller asymmetry of δ between the right and left sides of the jet, which is due to smaller vϕ, but the counterjet becomes more luminous due to smaller $| {{\boldsymbol{v}}}_{p}| $.

3.2. Case 2: BH-threaded Model

We investigate the other case, where ΩF is a constant given by Equation (17), motivated by the magnetic field lines penetrating the BH. The Kerr parameter is crucial in this case, since it directly controls ΩF. We thus systematically study the dependence of the radio image on the Kerr parameter as well as the effect of Rp. We pick two extreme cases of a = 0.998 and a = 0.1 as the best examples.

3.2.1. Fast-spinning BH

First, we show the results for a = 0.998, for which ΩF is $1.4\times {10}^{-5}$ s−1. The upper left panel in Figure 5 shows the color map of $R{{\rm{\Omega }}}_{{\rm{F}}}/c$ in the jet. Since the magnetic field lines rigidly rotate, $R{{\rm{\Omega }}}_{{\rm{F}}}/c$ monotonically increases with R, and the light cylinder is given by $R={R}_{\mathrm{lc},2{\rm{f}}}:= {{\rm{\Omega }}}_{{\rm{F}}}/c\sim 2.1{r}_{g}\sim 8.6\,\times {10}^{-3}$ mas. We emphasize here that the jet structure is qualitatively different from those in the disk-threaded model, in which another curved surface of the light cylinder exists. The magnetic field is thus toroidally dominated in almost the entire region in the jet except for the inside of the thin, light cylinder, as depicted in the lower left panel in Figure 5, which is a sharp contrast to the previous case. As a result, the velocity field is also qualitatively different away from the jet axis: $| {{\boldsymbol{v}}}_{p}| $ becomes highly relativistic $(\sim c)$, and vϕ is suppressed to nonrelativistic speed $(\lesssim 0.1c)$, as presented in the upper middle and right panels in Figure 5. Around the jet axis ($R\lesssim {R}_{\mathrm{lc},2{\rm{f}}}$), on the other hand, the velocity field is not much different from that in the previous disk-threaded model, since the magnetic field lines near the jet axis rigidly rotate with comparable angular frequencies in these models (see ${{\rm{\Omega }}}_{{\rm{F}}}\sim 1.4\times {10}^{-5}$ s−1 for this model and ${{\rm{\Omega }}}_{\mathrm{Kep}}({R}_{\mathrm{ISCO}})\sim 2.5\times {10}^{-5}$ s−1 for the previous disk-threaded model). The Lorentz factor is shown in the lower middle panel in Figure 5. The lower right panel exhibits $\mathrm{log}n$ for Rp = 40rg as an example. The nonthermal electrons are concentrated on the magnetic field lines ${\rm{\Psi }}/A\sim 35.3{r}_{g}$ (black lines), on which the magnetization factor σ is minimized to ∼8 × 105 in the presented region but still holds a sufficiently large value for the force-free approximation.

Figure 5.

Figure 5. Same as Figure 2 but for Case 2 with a = 0.998.

Standard image High-resolution image

Figure 6 shows δ in the jet. As expected from the slow vϕ, the difference in δ is small between the right and left sides with respect to the observer. Due to the large $| {{\boldsymbol{v}}}_{p}| $, δ in the counterjet is suppressed and the radiation is strongly debeamed for the observer. In the approaching jet side, on the other hand, a part of the front side of the jet strongly beams the light to the observer (δ ∼ 5), whereas the back side does not, due to the misalignment of the highly relativistic velocity and the observer direction. We also note that the asymmetry that is due to the anisotropic synchrotron radiation in the fluid rest frame is again found to be negligible.

Figure 6.

Figure 6. Same as Figure 3 but for Case 2 with a = 0.998.

Standard image High-resolution image

The left panel in Figure 7 displays the synthesized radio map for Rp = 0. Neither a limb-brightened feature nor the counterjet is seen in the radio map as in Case 1 with Rp = 0. This result is again attributed to the strong beaming effect to polar directions that is due to the velocity field near the jet axis.

Figure 7.

Figure 7. Same as Figure 4 but for Case 2 with a = 0.998.

Standard image High-resolution image

The radio maps for Rp > 0 can successfully show a symmetrically limb-brightened jet without a luminous counterjet, as displayed in the right panel in Figure 7, where the result for Rp = 40rg is shown for comparison to the counterpart in Figure 4. The symmetry of the jet image is recovered thanks to the small vϕ in the outer part of the jet away from the axis, which suppresses the beaming/debeaming asymmetry in the jet northern/southern sides, as displayed in Figure 6. The counterjet is less luminous due to the highly relativistic poloidal speed, which beams the emission to the direction opposite to the observer.

For larger ring radii (${R}_{p}\gt 40{r}_{g}$), the results are qualitatively the same as for Rp = 40rg, while the width of the jet image becomes wider. For smaller Rp ($0\lt {R}_{p}\lt 40{r}_{g}$), the jet width becomes smaller while keeping the symmetrically limb-brightened feature and gradually approaches the result for Rp = 0.

3.2.2. Slowly Spinning BH

We here show the results for the slowly spinning BH with a = 0.1. The Kerr parameter results in ${{\rm{\Omega }}}_{{\rm{F}}}=7.5\times {10}^{-7}$ s−1, which is $5.3\times {10}^{-2}$ times as large as that for a = 0.998 and shifts the light cylinder outward to the ∼19 times larger radius, $R={R}_{\mathrm{lc},2{\rm{s}}}\sim 40{r}_{g}\sim 0.16$ mas, as well as the other contour lines of $R{{\rm{\Omega }}}_{{\rm{F}}}/c$, as presented in the upper left panel in Figure 8. The change is also reflected in the distribution of the ratio of the toroidal to poloidal magnetic field strengths, as visible in the lower left panel in Figure 8. As a result, the region with slow poloidal speeds and fast azimuthal ones is extended from the jet axis to $R\lesssim {R}_{\mathrm{lc},2{\rm{s}}}$, as visible in the upper middle and right panels in Figure 8. In the outer part of the jet, $R\gg {R}_{\mathrm{lc},2{\rm{s}}}$, vϕ is increased by ∼19 times, compared to the case of a = 0.998 at the same radius, since vϕ is inversely proportional to $R{{\rm{\Omega }}}_{{\rm{F}}}/c$, as given by Equation (7). The poloidal speeds for $R\gg {R}_{\mathrm{lc},2{\rm{s}}}$ are not much different from those in the previous case of a = 0.998, on the other hand, since it is asymptotically determined by the angle from the jet axis, as given by Equation (6), where $R{{\rm{\Omega }}}_{{\rm{F}}}/c$ appears in the higher order corrections. The resultant Lorentz factor is displayed in the lower middle panel in Figure 8, which has the asymptotically same structure as for a = 0.998 in the jet edge part, due to the dominance of the poloidal speed. The lower right panel presents the density profile of the nonthermal electrons for Rp = 40rg, which are concentrated on the magnetic field lines ${\rm{\Psi }}/A\sim 35.3{r}_{g}$ (black lines). The minimal value of σ ∼ 5 × 103 in the displayed area is consistent with the force-free assumption.

Figure 8.

Figure 8. Same as Figure 2 but for Case 2 with a = 0.1.

Standard image High-resolution image

The plots of δ in Figure 9 clearly exhibit different patterns compared to the case for a = 0.998. In the approaching jet side, δ is more asymmetric between the right and left sides, due to the larger vϕ. In the counterjet side, δ is still suppressed almost in the jet edge region, due to highly relativistic $| {{\boldsymbol{v}}}_{p}| $, whereas δ is close to unity inside the light cylinder because of nonrelativistic $| {{\boldsymbol{v}}}_{p}| $.

Figure 9.

Figure 9. Same as Figure 3 but for Case 2 with a = 0.1.

Standard image High-resolution image

Figure 10 shows the produced radio images for Rp = 0 (left) and Rp = 40rg (right). Most importantly, the radio image becomes highly asymmetric between the northern and southern parts because of the enhanced relativistic beaming by larger vϕ, which is incompatible with the M87 jet. We also note that the counterjet becomes more luminous, which is clearer in the case of Rp = 0 (see the left panels in Figures 7 and 10), due to smaller $| {{\boldsymbol{v}}}_{p}| $ around the jet axis, which relaxes the relativistic beaming to the antidirection to the observer.

Figure 10.

Figure 10. Same as Figure 4 but for Case 2 with a = 0.1.

Standard image High-resolution image

We also confirmed that the results for $0\lt {R}_{p}\lt 40{r}_{g}$ present extremely asymmetric jets, as inferred from the results for Rp = 0 and 40rg. The large asymmetry is also maintained for larger Rp in our search up to ${R}_{p}=100{r}_{g}$, which produces a sufficiently wide jet image for M87.

4. Discussion

4.1. Jet Images in Cold, Ideal MHD Treatment

In Section 3, it was found that the disk-threaded model would not produce the symmetrically limb-brightened jets in our model and seems to be inappropriate for the M87 jet. We here discuss whether this result changes or not if we give another velocity different from that given by Equation (5) as the jet velocity. This is worth considering, since the jet edge part in the disk-threaded model corresponds to the region with $\zeta \leqslant 1$, where the drift velocity, Equation (5), may not approach the velocity in cold, ideal MHD, which will be the next simplest approximation. Comparing the drift velocity with the velocity in cold, ideal MHD jets, we argue for expected changes of limb-brightened radio images in the MHD treatment through the modified beaming effects for the observer.

We first focus on the azimuthal speed, since it is critical to the asymmetry in radio images. The toroidal speed in cold, ideal MHD outflows under the steady and axisymmetric assumptions is given as follows (e.g., Toma & Takahara 2013):

Equation (19)

where $\zeta := R{{\rm{\Omega }}}_{{\rm{F}}}/c$, and the letters with tilde symbols denote the quantities at the inlet. Here, $\tilde{{\rm{\Gamma }}}\sim 1$ is the initial Lorentz factor at the inlet. The ratio to the azimuthal speed in our force-free model, the second term in Equation (5), is then given by

Equation (20)

where the letters with MHD and FF are evaluated in a cold, ideal MHD model and our force-free one, respectively. As long as the force-free approximation is reasonable, the shapes of the poloidal magnetic fields are the same in ours and in MHD. This assumption of the same-shaped field yields ${\zeta }_{\mathrm{FF}}={\zeta }_{\mathrm{MHD}}=\zeta $. Since ${g}_{\mathrm{FF}}(\theta ,\nu )$ is also determined by the shape of Bp, we can omit the subscript, FF, hereafter: gFF = g.

If the ratio given by Equation (20) exceeds unity (i.e., ${v}_{\phi ,\mathrm{MHD}}\geqslant {v}_{\phi ,\mathrm{FF}}$), the disk-threaded model (Case 1) will not be preferred even in cold, ideal MHD models, due to more asymmetric limb-brightened features (see Appendix D for the proof that faster rotational speeds always lead to more asymmetric images). From Equation (20), the inequality ${v}_{\phi ,\mathrm{MHD}}\geqslant {v}_{\phi ,\mathrm{FF}}$ holds for

Equation (21)

We can put $1-{\tilde{\zeta }}^{2}\sim 1$, since we are now interested in the outer jet with $\zeta \lesssim 1$ in Case 1, which roughly corresponds to ${\rm{\Psi }}/A\gtrsim 35.3{r}_{g}$ for $| z| \lt 10$ mas (the black curve in Figure 2) and ${\tilde{\zeta }}^{2}\lesssim 0.01$. Since $g\sim 1$ at high latitudes where the limb brightens (see Figure 12), we can reduce Equation (21) to

Equation (22)

That is, if the above condition holds, the simulated radio images for ${R}_{p}\geqslant 40{r}_{g}$ would be more asymmetric in the disk-threaded model with cold, ideal MHD treatment.

Although the actual value of ${{\rm{\Gamma }}}_{\mathrm{MHD}}$ in cold, ideal MHD treatment could be obtained with a detailed model, it is beyond the scope of this paper. Instead, we here consider whether the limb brightening of the M87 jet can emanate from the region with $\zeta \leqslant 1$ by assuming that the pattern speed observed in the M87 jet corresponds to ${{\rm{\Gamma }}}_{\mathrm{MHD}}$. Mertens et al. (2016) reports that the Lorentz factor of the fast component of the M87 jet exceeds ∼2 at $z\gtrsim 3$ mas, which means from Equation (22) that the azimuthal speed should be larger than that in our model provided the limb brightening originates from the region with $\zeta \leqslant 1$. A larger vϕ enhances the asymmetry of the radio images, which is not consistent with observations.

A possible change of the poloidal speed would always produce problematic jet images: the counterjet becomes more luminous for smaller $| {{\boldsymbol{v}}}_{p}| $, while the asymmetry of the emission from "the coming quadrisection" of the jet is enhanced for larger $| {{\boldsymbol{v}}}_{p}| $ (see Appendix D).

From the above discussions, the disk-threaded model would not be suitable for the M87 jet even in the cold, ideal MHD treatment, while MHD numerical simulations should be incorporated for more quantitative discussions. It is noted, on the other hand, that Mertens et al. (2016) conjectured a jet launched from a Keplerian accretion disk, based on analyses of observed pattern speeds in the M87 jet with cold, ideal MHD treatment. The reason for this discordance should be pursued, although it is beyond the scope of this paper.

4.2. Effects of the Viewing Angle

The viewing angle Θ will be another important parameter, as well as ΩF and Rp, for producing radio images, since it changes the line-of-sight speed, which strongly beams or debeams the synchrotron emission to the observer. While the viewing angle of the M87 jet is thought to be in the range of ∼10° to 45° based on optical observations of superluminal motion around the HST-1 (Biretta et al. 1999) and radio observations of proper motion and brightness ratio of the jet and counterjet (Ly et al. 2007; Hada et al. 2016; Mertens et al. 2016), it will be interesting to study whether the limb-brightened features are kept if the viewing angle is much larger or smaller than the above constraint. Below we set Θ = 5° and 75°, for example, while the other parameters are the same as in Case 2 with a = 0.998 and Rp = 40rg.

The left panel in Figure 11 presents the result for Θ = 5°, which still shows a limb-brightening feature. As Θ decreases, the jet becomes more luminous, while the counterjet becomes dimmer because of stronger beaming and debeaming effects to the observer. At the same time, the jet image is expanded in the transverse direction by the projection effect.

Figure 11.

Figure 11. Same as the right panel in Figure 7 but for Θ = 5° (left) and 75° (right).

Standard image High-resolution image

The right panel in Figure 11 displays the case for Θ = 75°. The large viewing angle softens the relativistic beaming to the observer. As a result, the jet becomes less luminous while the counterjet becomes more luminous. The limb-brightened feature is still visible in the jet side while it is also apparent in the counterjet.

As presented above, limb-brightened features are observed even for viewing angles much different from our fiducial value. This fact suggests that the limb-brightened features observed in other objects such as Mrk 501, which has a viewing angle of ${\rm{\Theta }}\sim 5-15^\circ $ (Giroletti et al. 2004, 2008), and Cyg A, which has ${\rm{\Theta }}\sim 75^\circ $ (Boccardi et al. 2016), are also attributed to the jet structure with magnetic field lines penetrating a fast-spinning BH and nonthermal electrons away from the jet axis. It is interesting, however, that Boccardi et al. (2017) came to another conclusion that the jet base of Cyg A is widely extended and appears to be anchored to the accretion disk. We may need more detailed models with parameters tuned for these objects to make credible conclusions, which will be studied in a forthcoming paper.

5. Summary and Conclusions

This paper investigated the relations between the jet structure of AGNs and observed radio images of the jet. We focused on the limb-brightened features observed in some AGNs such as M87 that appear to be largely symmetric to the jet axis. We employed the basically same steady, axisymmetric force-free jet model as in BL09 but introduced new points of view to produce limb-brightened jets. We compared paraboloidal jets launched from the Keplerian accretion disk and from the central BH. The latter was not investigated in BL09. It was found that they have qualitatively different jet structures, including the jet rotation pattern and speed, which produce qualitatively and quantitatively different radio images even for the same distribution of emitting particles. We treated the spatial distribution of the nonthermal electrons as a parameter, instead of linking it to some physical processes or just concentrating the particles around the jet axis as in BL09. Simulating radio maps produced by synchrotron radiation, we constrained several important jet parameters for symmetrically limb-brightened jet images.

We demonstrated that symmetrically limb-brightened jets may be launched from a fast-spinning BH with the nonthermal electrons distributed away from the jet axis: we assumed that the magnetic field lines penetrate the BH and the magnetic field lines rigidly rotate with the half angular frequency of the BH. Far away from the jet axis, the jet is sufficiently accelerated to poloidal directions and the jet rotation relatively slows down, which occurs more effectively for larger Kerr parameters. Such a velocity field leads to symmetric jet images with low-luminosity counterjets. Slowly spinning BHs and the particle distribution concentrated near the jet axis are disfavored: the former results in extremely asymmetric radio emissions because of faster jet rotations, while the latter never brightens the edge but ends in a candle-flame-like pattern.

We also suggested that symmetrically limb-brightened jets are not launched from a geometrically thin accretion disk with Keplerian rotation, which was assumed in BL09. Reasonably, the jet edge is not illuminated unless the nonthermal electrons exist there. The nonthermal electrons away from the jet axis, however, produce strongly asymmetric radio images. This is because the fast jet rotation enhances the difference in the relativistic beaming to the observer between the northern and southern sides of the jet. The luminous counterjet is also a problem of this model in the case of the M87 jet, which is not dimmed because of the slow poloidal speeds in the jet edge. We also pointed out that the disk-threaded model would not be appropriate for the M87 jet even in cold, ideal MHD treatment, since the asymmetry of radio images would be enhanced and the counterjet could be more prominent. This challenges the interpretation that the jet is launched from an accretion disk (e.g., Mertens et al. 2016).

We cannot exclude, however, the magnetic field lines converging to a narrow ring region on the accretion disk instead of penetrating the BH horizon, which may cause an almost rigidly rotating magnetic field, whereas it should be debatable whether such a concentrated configuration can be realized. It is also noted that the disk-threaded model might relax the asymmetry of jet images and veil the counterjet by assuming an accretion disk rotating with another law that has a weaker dependence on R than for ${{\rm{\Omega }}}_{\mathrm{Kep}}$ or by finely tuning all of the parameters in our model, whereas only slowing down the rotation speed is insufficient to solve the problems (see Appendix B.2). We need a more detailed fit to observations in order to totally reject the disk-threaded model.

In our BH-threaded model, the symmetry of radio images is dependent on the Kerr parameter: the symmetric pattern is gradually recovered as the Kerr parameter increases. Therefore, the spin of the central BH could be constrained by fitting the calculated jet image to the observations. Such detailed studies are complementary to those concentrating directly on the innermost region with upcoming EHT data (Dexter et al. 2012; Mościbrodzka et al. 2016), since the size of observed BH shadows only has a weak dependence on the BH spin (Psaltis et al. 2015 and references therein). Furthermore, in addition to M87, the limb-brightened jet structures in other AGNs such as Mrk 501 and Cyg A might be also explained in the same manner with the BH-threaded model. A detailed study on these specific objects will be presented in a forthcoming paper.

It is worth noting again that our results indicate the existence of nonthermal electrons away from the jet axis, which is inevitable in order to produce limb-brightened images. This constraint is important, since the distribution of the nonthermal particles is one of the most ambiguous points even in more elaborate models using global GRMHD simulations (e.g., Mościbrodzka et al. 2016). While the distribution of nonthermal electrons should be given by microscopic processes, our findings might be a hint to search for the site of particle accelerations in relativistic jets. Other sophisticated numerical simulations of relativistic jets, such as Broderick & McKinney (2010) and Porth et al. (2011), also do not show limb-brightened features because of the assumed spatial distribution of the nonthermal electrons, although their distributions are based on physically motivated models. We also note that Porth et al. (2011) assumed jets launched from an accretion disk, so their simulations would not produce a symmetrically limb-brightened jet with a dim counterpart even if they had employed other spatial distributions of emitting particles.

While our simple treatment of relativistic jets leads to suggestive results, a comprehensive treatment with an accretion disk with funnel flows handled in a more detailed way, for example in general relativistic radiation MHD, must be incorporated in future work and is necessary for a consistent understanding of the jet–disk system of AGNs.

We thank the participants in the Mizusawa Project Meetings in 2016 and 2017 for fruitful discussions on the M87 jet from various points of view. K.T. and K.T. thank Taiki Ogihara for daily discussions on relativistic jets. The first author thanks Prof. Hiroshi Nagai for his comments on the use of terminology and references. We also thank the anonymous referee for his or her fruitful comments and suggestions. Numerical calculations were performed on Draco, a computer cluster at the Frontier Research Institute for Interdisciplinary Sciences in Tohoku University. This work is partly supported by JSPS Grants-in-Aid for Scientific Research, JP17H06362 (K. Takahashi), 15H05437 (K. Toma), JP18K03656 (M.K.), and JP18H03721 (M.K., K.H.), and also a JST grant "Building of Consortia for the Development of Human Resources in Science and Technology."

Appendix A: Force-free Jet Model

A.1. Steady, Axisymmetric Force-free Field

Steady, axisymmetric electromagnetic fields have been widely considered in the literature (Mestel 1961; Okamoto 1974; Bekenstein & Oron 1978; Camenzind 1986; Tomimatsu & Takahashi 2003; Vlahakis & Königl 2003; Beskin 2009). We review here such fields with the force-free approximation. The basic equations consist of the Maxwell equations and the conservation laws of a fluid coupled with an electromagnetic field.

We start from the relations that are derived only from the steady, axisymmetric condition before imposing the force-free approximation. Analogous to the two-dimensional incompressible flows, a stream function exists for the poloidal magnetic field, by which each magnetic surface is labeled, because of the divergence-free condition of the magnetic field in an axisymmetric geometry. The poloidal magnetic field, Bp, is then given as follows (Narayan et al. 2007; Broderick & Loeb 2009; Toma & Takahara 2013):

Equation (23)

where ${\rm{\Psi }}:= {{RA}}_{\phi }$ is a stream function with Aϕ being the toroidal component of the magnetic vector potential. The vectors with a hat are the unit coordinate bases. We note that the stream function Ψ(R,z) is essentially the total magnetic flux penetrating within radius R except for a factor of 2π: that is, Φ = 2πΨ is satisfied for any magnetic flux Φ (Narayan et al. 2007).

The steady, axisymmetric condition reduces the poloidal component of Faraday's law, ${\boldsymbol{\nabla }}\times {\boldsymbol{E}}=0$, to the relation ${E}_{\phi }\equiv 0$, where Eϕ stands for the toroidal component of the electric field.

In the force-free approximation, the plasma inertia and thermal pressure are neglected in dynamics (Narayan et al. 2007; Broderick & Loeb 2009). In this prescription, the fluid contributes only as the charge and current sources. The equation of motion is, hence, reduced to

Equation (24)

where ρe and j are charge and current densities, respectively.

The projection of both sides of the force-free condition, Equation (24), to the direction of B yields the condition that the magnetic and electric fields are orthogonal to each other: ${\boldsymbol{E}}\cdot {\boldsymbol{B}}=0$. Due to the absence of the toroidal electric field, the orthogonal condition gives the electric field as follows (Lyubarsky 2009; Toma & Takahara 2013):

Equation (25)

where ${{\rm{\Omega }}}_{{\rm{F}}}(R,z)$ is a scalar function. That is, the surface of Ψ =const. is also an equipotential surface. Substituting Equation (25) into Faraday's law, we obtain the following conservation law from the toroidal component:

Equation (26)

which means that ΩF is conserved along a magnetic field line and, hence, is a function of Ψ: ${{\rm{\Omega }}}_{{\rm{F}}}={{\rm{\Omega }}}_{{\rm{F}}}({\rm{\Psi }})$ (Toma & Takahara 2013).

The other Maxwell equations recover the corresponding charge and current sources for a given electromagnetic field. The charge density is obtained by Gauss's law (Narayan et al. 2007):

Equation (27)

while the current density is given by Ampère's law as follows (Narayan et al. 2007):

Equation (28)

Equation (29)

The toroidal component of the force-free condition, Equation (24), gives a conservation law for the total poloidal current passing through a toroidal loop of radius R, $I\propto {{RB}}_{\phi }$. In fact, the equation gives (${\boldsymbol{j}}\times {\boldsymbol{B}}{)}_{\phi }=0$ because ${E}_{\phi }\equiv 0$, which is satisfied only if jp is parallel to Bp. Comparing these poloidal vectors given by Equations (23) and (29), one notices that RBϕ should be a function of Ψ (Narayan et al. 2007). That is, RBϕ is conserved along a magnetic field line:

Equation (30)

We already projected Equation (24) to the directions of B and $\hat{{\boldsymbol{\phi }}}$. Because of the orthogonal relations, ${\boldsymbol{E}}\cdot {\boldsymbol{B}}={\boldsymbol{E}}\cdot \hat{{\boldsymbol{\phi }}}=0$, the projection onto E gives a relation independent of the former ones. The last equation determines Ψ for a given ΩF and Bϕ as follows (Narayan et al. 2007):

Equation (31)

which gives the shape of the magnetic field that satisfies the force balance in the transfield direction.

A.2. Magnetic Field

Equation (31) becomes singular at the critical surface RΩF/c =1, and a regular solution is found only for an appropriate choice of the functional form of Bϕ for a given ΩF. Otherwise, the solution cannot be continuous beyond the singular surface (Fendt 1997; Contopoulos et al. 1999; Beskin 2009; Takamori et al. 2014). Since it is generally a tough task to find such a fully consistent regular solution, we use a stream function that approximately describes the force-free numerical solution obtained by Tchekhovskoy et al. (2008), which gives a paraboloid-shaped jet and was also adopted in BL09. The stream function is given by

Equation (32)

where A is a constant that has the dimension of $[{r}^{2-\nu }B]$, and ν is the parameter that determines the jet shape. The minus and plus signs are for z ≥ 0 and z < 0, respectively, and the function is symmetric with respect to the equatorial plane, z = 0. We note that Equation (32) is a good approximation to the exact solution of the steady, axisymmetric force-free field, as well as results in numerical simulations (Tchekhovskoy et al. 2008). As special cases, Equation (32) gives a split-monopole field for ν = 0 and a paraboloidal field for ν = 1. Since we are interested in collimated jets, we assume ν > 0 hereafter. The components of the poloidal magnetic field are given by

Equation (33)

Equation (34)

which yield

Equation (35)

where

Equation (36)

We henceforth assume $\nu \leqslant \sqrt{2}$ (for the drift speed less than c; see Appendix A.3). Then the function $g(\theta ,\nu )$ satisfies $\sqrt{1+{\nu }^{2}}/2\leqslant g(\theta ,\nu )\leqslant 1$, as shown in Figure 12. We note that $g(\theta ,\nu )$ is reduced to $\cos (\theta /2)$ and $\sin (\theta /2)$ for z ≥ 0 and z < 0, respectively, in the case of ν = 1.

Figure 12.

Figure 12. Plots of $g(\theta ,\nu )$ for $\nu \leqslant \sqrt{2}$.

Standard image High-resolution image

Corresponding to the given shape of the jet, Equation (32), Bϕ is given by (Tchekhovskoy et al. 2008)

Equation (37)

We note that BL09 also use the same prescription for Bϕ.

The magnitude of the magnetic field is given by Equations (35) and (37) as follows:

Equation (38)

which gives the following asymptotic relation:

Equation (39)

A.3. Fluid Velocity

The force-free approximation does not give the fluid velocity, since the fluid inertia is totally neglected and, hence, the motion along a magnetic field cannot be determined. Following BL09, we use the so-called drift velocity as the fluid velocity (Narayan et al. 2007):

Equation (40)

This prescription ensures that (1) the fluid speed does not exceed the speed of light for $\nu \leqslant \sqrt{2}$, (2) the electric field vanishes in the proper frame, which is consistent with the infinite conductivity, and (3) the velocity asymptotically approaches the fluid velocity in cold, ideal MHD as relativistically accelerated to poloidal directions.

The first and second statements are straightforwardly confirmed by calculation. In fact, the normalized speed of the fluid is given by

Equation (41)

where the equality holds for $g(\theta ,\nu )=1$ and $R| {{\rm{\Omega }}}_{{\rm{F}}}| /c=\infty $. We also note that the azimuthal speed is bound by $c/2$, which can be shown in the same manner.

Equations (39) and (40) give the asymptotic relations of the fluid velocity for $R| {{\rm{\Omega }}}_{{\rm{F}}}| /c\ll 1$ as follows:

Equation (42)

Equation (43)

Equation (44)

where ${\beta }_{p}:= | {{\boldsymbol{v}}}_{p}| /c$ and ${\beta }_{\phi }=| {v}_{\phi }| /c$ are the normalized poloidal and toroidal speeds, respectively. That is, the fluid velocity is nonrelativistic and dominated by the toroidal component. For $R| {{\rm{\Omega }}}_{{\rm{F}}}| /c\gg 1$, on the other hand, the following relations are obtained:

Equation (45)

Equation (46)

Equation (47)

That is, the fluid velocity is dominated by the poloidal component, which becomes relativistic as $g(\theta ,\nu )$ approaches unity. We note here that, as $g(\theta ,\nu )\to 1$, the leading terms in Equations (45) and (46) approach those in the asymptotic relations in steady, axisymmetric cold outflows in ideal MHD (Toma & Takahara 2013):

Equation (48)

Equation (49)

which holds for $R| {{\rm{\Omega }}}_{{\rm{F}}}| /c\gg 1$ and ${\rm{\Gamma }}\gg \tilde{{\rm{\Gamma }}}\sim 1$, where the letters with tilde symbols denote quantities at the inlet.

A.4. Nonthermal Electrons

The number density of the nonthermal electrons, n, is assumed to be given by the continuity equation for a fluid, ${\boldsymbol{\nabla }}\cdot (n{\boldsymbol{v}})=0$, by following BL09, although it is not so obvious whether the nonthermal electrons obey the equation. For ${{RB}}_{\phi }{{\rm{\Omega }}}_{{\rm{F}}}\ne 0$, the continuity equation is reduced to

Equation (50)

which means that n scales with B2 along a given magnetic field. We also note that the continuity equation also derives the conservations of the ratio of the mass flux to the magnetic flux as in ideal MHD: ${\boldsymbol{B}}\cdot {\boldsymbol{\nabla }}(n| {{\boldsymbol{v}}}_{p}| /{B}_{p})=0$ by using Equation (40). In this paper, we assume the following ring-shaped distribution of the nonthermal electrons on the planes z = ±z1 (z1 ≥ 0):

Equation (51)

where Rp is the radius where n has the peak on the plane and Δ gives the width of the ring, while n0 is the number density at the peak. We note that BL09 considered only Rp = 0, where the nonthermal electrons are concentrated on the jet axis at z = ±z1.

Equations (50) and (51) give the number density of the nonthermal electrons at a given point on a magnetic field labeled by ${\rm{\Psi }}^{\prime} $ as follows:

Equation (52)

where ${R}_{1}({\rm{\Psi }}^{\prime} )$ denotes the radial coordinate of the intersections of Ψ = Ψ' and z = ±z1. We omit an artificial factor of $(1-\exp [-{r}^{2}/{z}_{1}^{2}])$ in Equation (52) that was introduced in BL09 to reduce plasma in the innermost region, $r\lt {z}_{1}$. Our results are not qualitatively different, however, even if the factor is taken into account.

We assume that the distribution of the nonthermal electrons is isotropic in the fluid rest frame and the energy distribution is described by a single power law with an index p:

Equation (53)

where and hereafter quantities with a prime are evaluated in the fluid rest frame. Here, γ' is the Lorentz factor of an electron, ${\gamma }_{\min }^{{\prime} }$ and ${\gamma }_{\max }^{{\prime} }$ are the minimal and maximal Lorentz factors, respectively, and C is a normalization constant, which is given for $p\ne 1$ by (Shibata et al. 2003)

Equation (54)

We assume that the energy distribution is given by Equation (53) in the entire region; that is, we assume some energy supplier that compensates for the energy loss that is due to cooling processes such as the synchrotron cooling.

A.5. Synchrotron Emissivity in the Fluid Rest Frame

Since we consider highly relativistic electrons, the synchrotron emission is highly beamed in the direction of the electron motion. In this case, the synchrotron emissivity in the fluid rest frame, ${j}_{\omega ^{\prime} }^{{\prime} }({\boldsymbol{n}}^{\prime} )$, is given by (Rybicki & Lightman 1985; Shibata et al. 2003)

Equation (55)

where e, me, and $\bar{{\rm{\Gamma }}}(\ldots )$ are the elementary charge, the mass of an electron, and the gamma function, respectively. Here, $\psi ^{\prime} ({\boldsymbol{n}}^{\prime} )$ is the pitch angle of the electrons that are directed toward the observer, which are most responsible for producing radio images because of relativistic beaming effects (Shibata et al. 2003):

Equation (56)

In the derivation of Equation (55), we used the approximation that ${\gamma }_{\min }^{{\prime} }$ and ${\gamma }_{\max }^{{\prime} }$ are sufficiently small and large, respectively, to evaluate an energy integral (Rybicki & Lightman 1985). In this case, the energy cutoffs affect the synchrotron emissivity only through the normalization constant C.

Appendix B: Parameter Dependence

B.1. Fast-spinning BH-threaded Models

We study the dependence of radio images on the parameters that were fixed in the main text. It is important to note that our conclusions in the main text are not changed even if these parameters are altered, whereas the radio images are slightly modified. We use the fast-spinning BH-threaded model (Case 2 with a = 0.998) with Rp = 40rg shown in the right panel in Figure 7 as a fiducial model, since it resembles the observed images better than the other models. We change four parameters, Δ (ring width), ν (jet shape), p (power index of the energy distribution of electrons), and MBH (BH mass) around the fiducial model as in Table 2 while fixing the other parameters such as ΩF and a as well as Rp. Comparing the produced radio images, we discuss the effects of each parameter below.

Table 2.  Fast-spinning BH-threaded Models

Name Δ ν p MBH
  (rg)     (109 M)
A (fiducial) 5 1 1.1 3.4
B 1 1 1.1 3.4
C 10 1 1.1 3.4
D 5 0.75 1.1 3.4
E 5 1.25 1.1 3.4
F 5 1 2 3.4
G 5 1 3 3.4
H 5 1 1.1 6.6

Download table as:  ASCIITypeset image

The dependence on Δ is displayed in Figure 13. As naturally expected, the larger Δ makes radio images wider in the north–south direction, since the electrons are more distributed to the edge region, although the effect is rather limited within this range of Δ.

Figure 13.

Figure 13. Same as Figure 7 but for models B, A, and C from left to right.

Standard image High-resolution image

The dependence on the jet shape is shown in Figure 14, where the jet is less (more) collimated in the left (right) panel. We note that the jet shape is expressed by $R\propto {z}^{\xi }$ far from the BH ($\theta \ll 1$), where ξ is defined by ν = 2–2ξ (Tchekhovskoy et al. 2008). That is, ν = 0.75, 1, and 1.25 (i.e., ξ = 0.625, 0.5, and 0.375) give the asymptotic jet shape of $z\propto {R}^{8/5}$, R2, ${R}^{8/3}$, respectively. The jet shape is clearly reflected in the radio image as tightly collimated jets produce narrower radio images.

Figure 14.

Figure 14. Same as Figure 13 but for models D, A, and E from left to right.

Standard image High-resolution image

Figure 15 manifests that the harder energy distribution of electrons leads to more compact radio images. That is, the contrast of intensity is enhanced for larger p, since the difference in the magnetic and velocity fields at different locations is enhanced by $(p-1)/2$, as given in Equation (55). The limb-brightened feature becomes discrete, as a result, for large p, while it is still discernible in Figure 15. We note that Hada et al. (2016) reported $p\sim 2.2\mbox{--}2.6$ for the M87 jet.

Figure 15.

Figure 15. Same as Figure 13 but for models A, F, and G from left to right.

Standard image High-resolution image

Massive BHs produce "larger" radio images, as shown in Figure 16. One should be careful in interpreting this result, since the Schwarzschild radius changes as ${r}_{g}\propto {M}_{\mathrm{BH}}$, while we used ${R}_{p}=40{r}_{g}$ in both models. That is, the electrons are distributed more far away from the jet axis in model H with a more massive BH, which directly makes the radio image wider in the X direction. We also note that the BH mass changes ΩF, which is proportional to ${M}_{\mathrm{BH}}^{-1}$ in the BH-threaded model for a fixed Kerr parameter. Thus, the increase in MBH for a fixed a has effects similar to the decrease in a for a fixed MBH (note: ${{\rm{\Omega }}}_{{\rm{F}}}\propto a/(1+\sqrt{1-{a}^{2}})$).

Figure 16.

Figure 16. Same as Figure 13 but for models A (left) and H (right). Note that 1 mas corresponds to ∼250rg (i.e., $1{r}_{g}\sim 4\times {10}^{-3}$ mas) for the left panel, while 1 mas ∼130rg (i.e., 1rg ∼ 8 × 10−3 mas) for the right one.

Standard image High-resolution image

B.2. Sub-Keplerian Disk-threaded Models

We study here the parameter dependence of radio intensity maps of the disk-threaded model (Case 1). We focus on the disk rotation, which characterizes disk-threaded models, and consider sub-Keplerian motion. Introducing a factor q ($0\lt q\leqslant 1$), we modify Equation (15) as follows:

Equation (57)

where $0\lt q\lt 1$ gives a sub-Keplerian disk while q = 1 coincides with Case 1. We pick the cases with q = 0.1 and q = 0.5 for example, while keeping the other parameters the same as in Case 1. The former is an extreme case of slowly rotating disks, and the latter corresponds to ADAFs. Figure 17 shows the radio intensity maps for q = 0.1, 0.5, and 1 for reference. As the disk rotation slows down, the radio image recovers the symmetry. However, the limb feature becomes less prominent and the counterjet keeps the brightness, which are inconsistent with observations of M87.

Figure 17.

Figure 17. Same as the right panel in Figure 4 but for q = 0.1, 0.5, and 1 from left to right.

Standard image High-resolution image

These changes in the radio image are explained as follows. As the disk rotational speed decreases, the light "cylinder" surfaces, both the curved and vertical ones, shrink inward. The shrinkage of the light cylinder decreases vϕ on ${\rm{\Psi }}/A\sim 35.3{r}_{g}$, where most of the emitting particles exist (black lines in Figure 2), since vϕ peaks around the light "cylinder" and decreases toward the jet edge part (see the upper right panel in Figure 2). This is why the asymmetry of radio images is weakened for smaller q. The shrinkage of the light cylinder, at the same time, slows down the poloidal speed, since $| {{\boldsymbol{v}}}_{p}| $ becomes smaller away from the curved light-cylinder surface. Thus, the light emanating from ${\rm{\Psi }}/A\sim 35.3{r}_{g}$ is less beamed as q decreases. As a result, the counterjet keeps the feature and the limb becomes less prominent with respect to the central core, which weakens the limb-brightening feature.

Appendix C: Asymptotic Shape of the Light Cylinder in Case 1

We derive here the asymptotic shape of the light "cylinder" in our disk-threaded model (Case 1), which has a curved surface as shown in Figure 2. We consider the far zone where each magnetic field line is anchored to the accretion disk far from the gravitational radius, that is, $\tilde{R}\gg {r}_{G}$. Thus, each magnetic field line rotates with ${{\rm{\Omega }}}_{{\rm{F}}}={{\rm{\Omega }}}_{\mathrm{Kep}}(\tilde{R})\sim \sqrt{{{GM}}_{\mathrm{BH}}/{\tilde{R}}^{3}}$. The condition for the light cylinder, $R{{\rm{\Omega }}}_{{\rm{F}}}/c=1$, is then reduced to the following cubic equation for z:

Equation (58)

where we used the relation $\tilde{R}={\rm{\Psi }}/A=\sqrt{{R}^{2}+{z}^{2}}\mp z$. The real root of Equation (58) is given by

Equation (59)

Equation (60)

We note that the deviation of the surface given by the above asymptotic relation, Equation (59), from the curved surface of the light cylinder is rather small even at $R\gtrsim {r}_{G}$, as shown in Figure 18.

Figure 18.

Figure 18. Close-up of the left panel in Figure 2 around the origin. The newly drawn dashed curve shows the line given by Equation (59).

Standard image High-resolution image

Appendix D: Ratio of the Beaming Factors on the Left and Right Sides of the Jet

In our model, asymmetric radio images are produced mainly by the difference in the beaming factors, $\delta := 1/[{\rm{\Gamma }}(1-\beta \mu )]$, on the left and right sides of the jet with respect to the observer, which is caused by the jet rotation. We here discuss how the asymmetric feature changes when the jet speed changes while the other quantities remain the same values by studying the dependence of the ratio of δ between the two sides of the jet. We note that the following discussion is quite general and uses only the assumption of an axisymmetric flow.

We use Cartesian coordinates $(x,y,z)$ where the flow is axisymmetric around the z axis and the observer direction is given by ${\boldsymbol{n}}=(0,\sin {\rm{\Theta }},\cos {\rm{\Theta }})$ with $0\leqslant {\rm{\Theta }}\leqslant \pi /2$ being the viewing angle. We consider two points $P=(\cos \chi ,\sin \chi ,z)$ and $Q=(-\cos \chi ,\sin \chi ,z)$, where $-\pi /2\lt \chi \lt \pi /2$ is the azimuthal angle of P measured from the x axis. Note that P and Q are symmetric positions with respect to the yz plane (see Figure 19). Let ${{\boldsymbol{\beta }}}_{P}$ and ${{\boldsymbol{\beta }}}_{Q}$ be the velocities normalized by c at P and Q, respectively. Due to the axisymmetry, they are generally given by

Equation (61)

Equation (62)

where ${\beta }_{R}$ and ${\beta }_{\phi }$ are, respectively, the radial and azimuthal velocities at P (or equivalently at Q). We can assume ${\beta }_{\phi }\geqslant 0$ without loss of generality. Equations (61) and (62) yield

Equation (63)

Equation (64)

The ratio of the beaming factors at P and Q is given by $\varepsilon := {\delta }_{P}/{\delta }_{Q}=(1-\beta {\mu }_{Q})/(1-\beta {\mu }_{P})$, where $\beta =| {{\boldsymbol{\beta }}}_{P}| =| {{\boldsymbol{\beta }}}_{Q}| $. Rotations with ${\beta }_{\phi }\gt 0$ lead to $\delta \gt 1$, which means that light emitted from P is more beamed to the observer than from Q. Without rotations (${\beta }_{\phi }=0$) or viewed from the z axis (Θ = 0), P and Q become equivalent for the observer and, hence, ε is unity.

Figure 19.

Figure 19. Considered two points, P and Q, which are on symmetric positions with respect to the yz plane.

Standard image High-resolution image

We first note that faster rotations always enhance the difference in the relativistic beaming to the observer between P and Q unless Θ = 0. That is, the ratio ε monotonically increases with ${\beta }_{\phi }$, which follows from

Equation (65)

where the inequality holds, since $\alpha := {\beta }^{2}[1-\beta ({\beta }_{R}\sin \chi \sin {\rm{\Theta }}\,+{\beta }_{z}\cos {\rm{\Theta }})]$ turns out to be nonnegative as follows:

Equation (66)

Equation (67)

Equation (68)

where the plus and minus signs are for ${\beta }_{R}\gt 0$ (expanding flows) and ${\beta }_{R}\leqslant 0$ (converging flows). Here, ${\beta }_{p}=\sqrt{{\beta }_{R}^{2}+{\beta }_{z}^{2}}$ is the poloidal speed, and ${\varphi }_{\pm }$ is given by $\cos {\varphi }_{\pm }=\pm {\beta }_{R}/{\beta }_{p}$ and $\sin {\varphi }_{\pm }={\beta }_{z}/{\beta }_{p}$.

The behavior of the ratio ε is more complicated for the change in the poloidal velocities, ${\beta }_{R}$ and ${\beta }_{z}$, as shown below. The differential of ε with respect to ${\beta }_{R}$ is given by

Equation (69)

The signature of $\partial \varepsilon /\partial {\beta }_{R}$ depends on the signature of ${\beta }_{R}+{\beta }^{3}\sin \chi \sin {\rm{\Theta }}$, which can be either positive or negative. It should be noted, however, that $\partial \varepsilon /\partial {\beta }_{R}$ is nonnegative for χ > 0 and ${\beta }_{R}\gt 0$. That is, the increase of the radial speed amplifies the difference of the beaming effects between P and Q in the half of the expanding flow that is near the observer when divided by the xz plane, unless Θ = 0 or βϕ = 0. This is the case for our jet model.

The differential of ε with respect to ${\beta }_{z}$ is given by

Equation (70)

which can be positive or negative, depending on the sign of ${\beta }_{z}+{\beta }^{3}\cos {\rm{\Theta }}$. It is, however, worth noting that $\partial \varepsilon /\partial {\beta }_{z}$ is nonnegative for ${\beta }_{z}\gt 0$, that is, when the outflow comes toward the observer. If applied to our jet model, it means that the increase of ${\beta }_{z}$ enhances the difference in the beaming effects between the left and right sides of the jet, whereas it is not always the case for the counterjet.

Footnotes

  • The separation surface is the separatrix between outflowing matter that is launched as a jet and inflowing matter that is swallowed into the BH. Note that it is not taken into account in the flow velocity given by Equation (5), since we only model jet outflows by neglecting general relativistic effects.

  • Note that BL09 further multiplied the number density by an artificial factor of $(1-\exp [-{r}^{2}/{z}_{1}^{2}])$ that works to reduce n in the innermost region $r\lt {z}_{1}$, which is an ad hoc treatment of gravitational effects. We do not introduce this factor, though it does not change our conclusions.

  • We excise the spherical region inside the horizon, where the emissivity is set to zero.

  • 10 

    As a first step, we investigate the relations between the jet images and the important jet parameters (ΩF and Rp) while fixing other parameters to the fiducial values, and we try to produce radio images with a symmetrically limb-brightened jet and a dim counterjet. We do not try to find the best-fit parameters for the M87 jet images.

  • 11 

    As seen in the right panel in Figure 7, the peak intensity ∼103 mJy per beam for Rp = 40rg is roughly consistent with those observed in M87 (∼5 × 102 mJy per beam; Hada et al. 2016) in order of magnitude. The peak intensity would be reduced for a more realistic model, since our model assumes optically thin jets, while the central core of the M87 jet is actually optically thick for 44 GHz.

  • 12 

    The optically thin assumption holds well for this frequency for the jet models in the main body of this paper, which correspond to Figures 4, 7, 10, and 11.

  • 13 

    The intrinsic jet length of 10 mas corresponds to the projected length of ∼4.2 mas for the viewing angle of Θ = 25°.

Please wait… references are loading.
10.3847/1538-4357/aae832