A Mechanism for the Triple-ridge Emission Structure of AGN Jets

, , and

Published 2019 May 20 © 2019. The American Astronomical Society. All rights reserved.
, , Citation Taiki Ogihara et al 2019 ApJ 877 19 DOI 10.3847/1538-4357/ab1909

Download Article PDF
DownloadArticle ePub

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

0004-637X/877/1/19

Abstract

Recent radio very long baseline interferometry observations of the relativistic jet in the M87 radio galaxy have shown a triple-ridge structure that consists of the conventional limb-brightened feature and a central narrow ridge. Motivated by these observations, we examine a steady axisymmetric force-free model of a jet driven by the central black hole (BH) with its electromagnetic structure being consistent with general relativistic magnetohydrodynamic simulations, and find that it can produce triple-ridge images even if we assume a simple Gaussian distribution of emitting electrons at the base of the jet. We show that the fluid velocity field associated with the electromagnetic field produces the central ridge component due to the relativistic beaming effect, while the limb-brightened feature arises due to a strong magnetic field around the jet edge that also induces the electrons to be dense there. We also show that the computed image strongly depends on the electromagnetic field structure, viewing angle, and parameters related to the electrons' spatial distribution at the jet base. This study will help constrain the non-thermal electron injection mechanism of BH jets and be complementary to theoretical analyses of the upcoming data of the Event Horizon Telescope.

Export citation and abstract BibTeX RIS

1. Introduction

Relativistic collimated outflows (or jets) have been observed in active galactic nuclei (AGNs). It is widely thought that they form systems composed of a black hole (BH) at the center of a galaxy and surrounding plasma. Their plausible formation mechanism is electromagnetic extraction of the rotational energy of the BH and/or its accretion disk (Blandford & Znajek 1977; Blandford & Payne 1982; Uchida & Shibata 1985; Lovelace et al. 1987; Meier 2001; Komissarov 2004; Beskin 2010). The former is called the Blandford–Znajek (BZ) process, and the latter is called the Blandford–Payne (BP) process. General relativistic magnetohydrodynamic (GRMHD) simulations show that the globally ordered magnetic field is realized only in the funnel region around the rotation axis, where a relativistic jet appears to form via the BZ process, and the disk wind is non-relativistic (e.g., McKinney & Gammie 2004; McKinney & Blandford 2009; Tchekhovskoy et al. 2011; Sądowski et al. 2013; Nakamura et al. 2018). In recent studies, the radiative transfer calculations have been performed based on the GRMHD simulation results (e.g., Moscibrodzka et al. 2007; Dolence et al. 2009; Broderick & McKinney 2010; Porth et al. 2011; Dexter et al. 2012; Mościbrodzka et al. 2016; Pu et al. 2016). These studies enable us to compare the numerical results to observations. If the BZ process is observationally confirmed, the existence of an ergosphere is indirectly supported. (Komissarov 2004; Toma & Takahara 2014, 2016; Kinoshita & Igata 2018).

Despite those sophisticated computations, there remains the so-called "mass-loading problem" for relativistic jets. No particle is injected into the jet from the BH, of course, and the globally ordered magnetic field prevents the surrounding thermal plasma particles from diffusing into the jet. The origin of particles in jets is still unclear. Electron–positron pair creation by ambient photons (Levinson & Rieger 2011; Mościbrodzka et al. 2011) and by high-energy photons emitted by electrons accelerated in the MHD-violated region or gap (Beskin et al. 1992; Hirotani & Okamoto 1998; Levinson & Rieger 2011; Broderick & Tchekhovskoy 2015; Hirotani et al. 2016; Ptitsyna & Neronov 2016; Levinson & Segev 2017) have been discussed for the particle injection mechanism.4 High-energy hadron injection from the surrounding plasma (Toma & Takahara 2012; Kimura et al. 2014, 2015) and magnetic reconnection induced by the accretion of fields with alternating polarity (Parfrey et al. 2015) might also be relevant. In the GRMHD simulations, these non-thermal processes are not taken into account. The density of the jet usually becomes very low and is replaced by a density floor. Although this treatment does not affect the electromagnetic field structure because the particle energy density is much smaller than the electromagnetic field density in the funnel region, the terminal Lorentz factor and the emission of the jet depend directly on the particle density distribution (McKinney 2006; Mościbrodzka et al. 2016; Jeter et al. 2018; Takahashi et al. 2018, hereafter, T18). The spatial distribution of emitting particles is still a serious ambiguity when one compares the simulation results with observations.

Radio very long baseline interferometry (VLBI) observations can resolve AGN jets and have reported the limb-brightened structure in jets of M87 (Kovalev et al. 2007; Walker et al. 2008; Hada et al. 2011, 2016; Mertens et al. 2016; Kim et al. 2018), Mrk 501 (Piner et al. 2008), Mrk 421 (Piner et al. 2010), Cyg A (Boccardi et al. 2016), and 3C84 (Nagai et al. 2014; Giovannini et al. 2018). M87's jet has been actively observed because of its proximity (D ≃ 16.7 Mpc, Mei et al. 2007; Blakeslee et al. 2009) and its length (∼10 kpc). Its central BH mass MBH is estimated as $(3.3\mbox{--}6.6)\times {10}^{9}\,{M}_{\odot }$ (Macchetto et al. 1997; Gebhardt & Thomas 2009; Gebhardt et al. 2011; Walsh et al. 2013), and then the angular size of the Schwarzschild radius ${R}_{{\rm{S}}}=2{{GM}}_{\mathrm{BH}}/{c}^{2}$ is ≈3.9–7.8 μas. The BH shadow and the jet-launching region of M87 with this size is expected to be resolved by the Global VLBI observation project Event Horizon Telescope (EHT) (Doeleman et al. 2012; Akiyama et al. 2017). The limb-brightened structure of the M87 jet is confirmed between ∼40 and 105 RS from the center, which are de-projected lengths under the assumption of the viewing angle Θ = 15° (see Wang & Zhou 2009).

Among many theoretical papers that compute the synchrotron images of MHD jets, only a few discussed the origin of the limb-brightened structure (Zakamska et al. 2008; Gracia et al. 2009; Mościbrodzka et al. 2016). T18 showed limb-brightened images using an analytic force-free model consistent with GRMHD simulation results. They showed that the toroidal velocity of the flow (and its effect on relativistic beaming) is relevant for interpreting the image structure and that a jet from a rapidly spinning BH is favored. Slowly spinning BHs and Keplerian disks do not efficiently accelerate the flow, leading to large toroidal velocities and highly asymmetric images that are not consistent with the observations of the M87 jet. It was also shown that a sufficient amount of electrons has to be injected on the magnetic field lines apart from the jet axis. This implies that the characteristic image structure can potentially constrain the electron spatial distribution and the mass-loading mechanism.

In this paper, we focus on the newly discovered characteristic feature. High-sensitivity observations with VLBA + phased-VLA (Hada 2017), recent analysis of VSOP data (Asada et al. 2016), and stacked Very Long Baseline Array images (Walker et al. 2018) have revealed that the M87 jet image has a "triple-ridge structure," i.e., a narrow central ridge in addition to the conventional limbs. Hereafter, we call the central component the "inner ridge" and the limbs "outer ridges." This feature might indicate that two different formation mechanisms work in a jet, unlike what is seen in the GRMHD simulation results, e.g., the BZ process for the inner ridge and the BP process for the outer ridges (Sob'yanin 2017). However, we show that a BZ jet can solely produce the triple-ridge images by following the formulation in T18. Thus, two different formation mechanisms are not necessarily required. We find that the MHD flow velocity structure of the BZ jets produces a distinct component near the axis due to the relativistic beaming effect, which corresponds to the inner ridge, while the outer ridges may arise mainly because of stronger magnetic field and denser electrons around the jet edge. Note that we will simply illustrate this novel idea on the triple-ridge emission structure, not trying to fit the observed data of the M87 jet.

This paper is organized as follows. We briefly introduce our model in Section 2, and show the computed triple-ridge image and its parameter dependence, along with their physical causes in Section 3. We discuss some detailed properties of the observed triple-ridge structure in Section 4, and summarize our findings in Section 5.

2. Model

To produce the synchrotron jet image, we use an analytic model that is essentially the same as that used in T18. Here, we explain this model briefly. Readers can refer to T18 as well as Broderick & Loeb (2009) for more details. Section 2.1 introduces the modeling of electromagnetic field, velocity field, and density field structure, and Section 2.2 treats electrons' energy distribution and their radiation. We set up our parameter values in Section 2.3. The quantities with prime denote those measured in the fluid rest frame.

2.1. Force-free Jet Model

We assume that the force-free condition is valid in the funnel region, where particle inertia and pressure are not important for dynamics, as shown in GRMHD simulations and implied from the current observational data of the M87 jet (Kino et al. 2014, 2015). We then activate the steady axisymmetric condition, and then the poloidal magnetic field and electric field measured in the lab frame are described as

Equation (1)

Equation (2)

(R, ϕ, z) are the standard cylindrical coordinates, with the z axis set to be the jet axis. $\hat{\phi }$ is the azimuthal unit vector. Ψ is the magnetic flux function and ΩF corresponds to the angular velocity. We use a parametrically controlled form of Ψ,

Equation (3)

where (r, θ, ϕ) are the standard spherical coordinates. A is a constant that normalizes the magnetic field strength, and ν controls the jet shape. The plus and minus signs stand for z < 0 and z > 0, respectively. Note that we focus on the emission structure far from the central BH, where we can neglect the GR effects. For this form of Ψ, the toroidal magnetic field is approximately given by (Tchekhovskoy et al. 2008)

Equation (4)

Equations (3) and (4) are not exact solutions of the Maxwell equations under steady axisymmetric condition (i.e., Grad-Shafranov equation), but they fit well to GRMHD simulation results (Tchekhovskoy et al. 2010; Nakamura et al. 2018). Ψ and ΩF are quantities conserved along each field line and the electromagnetic field is described by their distribution. The cases of ν = 0 and 1 correspond to the radial and parabolic shapes of ${{\boldsymbol{B}}}_{{\rm{p}}}$, respectively. For ν = 2, the cylindrical structure is obtained in the far zone. As explained in Section 1, only the model with field lines penetrating the rapidly rotating BH can produce the nearly symmetric limb-brightened images as observed (T18). We focus on this case, in which ΩF(Ψ) is given as ΩF(Ψ) ≈ 0.5 ΩH, where ΩH = ac/2r+, a is a spin parameter of the BH, and r+ is the horizon radius of the rotating BH. We set a = 0.998.

We may not define the fluid velocity in principle in the force-free model, but practically we can use the drift velocity (Tchekhovskoy et al. 2008; Beskin 2010),

Equation (5)

This is consistent with the velocity in the Poynting-dominated cold ideal MHD formalism at the zone where RΩFc (see Equations (45) and (46) of T18). For RΩFc, the drift velocity has ${v}_{\phi }\sim R{{\rm{\Omega }}}_{{\rm{F}}}\gg {v}_{{\rm{p}}}$ (Equations (42) and (43) of T18) for the electromagnetic field that we consider (as discussed later in Section 3.1), which is also consistent with the cold ideal MHD velocity. Cold ideal GRMHD calculations also show vϕ ≫ vp within the outer light cylinder (Pu et al. 2015), which supports our treatment of the fluid velocity. We do not treat the case in which the jet particles are injected along the magnetic field lines with a large Lorentz factor at the outflow base. We briefly discuss this case in Section 4.3.

We give the density distribution from the sourceless equation of continuity,

Equation (6)

where n is the particle number density measured in the lab frame. This equation, combined with Equation (5), leads to another conserved quantity,

Equation (7)

When we give the density value at a certain point on each field line, we can obtain the density distribution everywhere. We note that Equation (7) is consistent with nvp/Bp = const., that is, the mass flux per unit magnetic flux conserved in the cold ideal MHD formalism. Indeed, one can obtain Equation (7) by substituting Equation (5) to nvp/Bp = const.

2.2. Non-thermal Electron Energy Distribution and Radiation

We set the fluid density at z = z1 just for simplicity with

Equation (8)

where z1, Rp, Δ are free parameters. We assume that a constant fraction of the electrons are non-thermal5 and their energy distribution is described as

Equation (9)

where γ' is the Lorentz factor of electrons measured in the fluid frame, ${\gamma }_{m}^{{\prime} }$ is the minimum value of γ', and p is the power-law index. p and ${\gamma }_{m}^{{\prime} }$ are set to 1.1 and 100, respectively, throughout this paper. We set the non-thermal electron density as zero along the field lines that do not penetrate the horizon, i.e., for Ψ > Ψ(r = r+, θ = π/2).

The synchrotron emissivity in the fluid frame jω'' is written by (Rybicki & Lightman 1986)

Equation (10)

where e, me, and $\bar{{\rm{\Gamma }}}(x)$ are the elementary charge, the mass of the electron, and the gamma function, respectively. $B^{\prime} \equiv | {\boldsymbol{B}}^{\prime} | $, and α' is the pitch angle of the electrons defined by $\cos \alpha ^{\prime} =(\hat{{\boldsymbol{n}}}^{\prime} \cdot {\boldsymbol{B}}^{\prime} )/| {\boldsymbol{B}}^{\prime} | $, where $\hat{{\boldsymbol{n}}}^{\prime} $ is the unit vector directing toward the observer in the fluid frame. Then, the emissivity at each point in the jet that is received by the observer is given by (Rybicki & Lightman 1986; Shibata et al. 2003)

Equation (11)

where β is the bulk speed normalized by the speed of light, and Γ is its corresponding Lorentz factor. μ is the cosine of the angle between the direction of the bulk flow and the direction of the line of sight in the lab frame.

The jet is optically thin for synchrotron radiation in the region where the triple-ridge structure is observed (∼10–30 mas from the radio core) and at the wavebands of those observations (Asada et al. 2016; Hada 2017). The observed intensity is then calculated by

Equation (12)

where dZ is the line element along the line of sight. (X, Y) are the coordinates of the sky with the Y axis being the projected jet axis. The viewing angle Θ is defined as the angle between the line of sight (Z axis) and the jet axis (z axis). We first obtain the intensity map on the XY plane by the computation of Equation (12) with the spatial resolution of 3 RS, and obtain the simulated VLBI image after the convolution with a Gaussian beam kernel.

2.3. Model Parameters

Our model parameters are summarized in Table 1. a, ${\gamma }_{m}^{{\prime} }$, and p are set to their fiducial values for which T18 obtained limb-brightened images. While T18 used the same fiducial values for ν, Θ, and MBH as in Broderick & Loeb (2009) to compare their calculation results, we replace them with the values roughly consistent with recent observations of the M87 jet. Nakamura et al. (2018) show that the width of the observed jet of M87 can be fitted by z ∝ R1.6 (see also Asada & Nakamura 2012; Hada et al. 2013; Nakamura & Asada 2013). Since Equation (3) indicates that the shape of each field line obeys z ∝ R2/(2−ν) asymptotically, we adopt ν = 0.75, assuming that the observed jet width reflects the magnetic field structure. The BH mass and the viewing angle are set by MBH = 6.2 × 109 M (Gebhardt et al. 2011, rescaled for D = 16.7 Mpc) and Θ = 15° (Wang & Zhou 2009), respectively. Then, 1 mas corresponds to ≈136 RS for D = 16.7 Mpc. We use the beam size of 1.14 mas × 0.55 mas, for which a triple-ridge image of the M87 jet was obtained (Hada 2017). Note that the beam size is larger than that in T18.

Table 1.  Model Parameters

Quality Symbol Value
Mass of the BH MBH 6.2 × 109M
Dimensionless Kerr parameter of the BH a 0.998
Rotational frequency of the magnetic field ΩF 0.5ΩH
Viewing angle Θ 15°
Jet Shape (Equation (3)) ν 0.75
Radius where n peaks at z = ± z1 Rp 0
Height of the plane where n is given by Equation (8) z1 5 RS
Width of n distribution (Equation (8)) Δ 2 RS
Energy spectral index of the non-thermal electrons p 1.1
Minimal Lorentz factor of the non-thermal electrons ${\gamma }_{m}^{{\prime} }$ 100
Luminosity distance to the jet D 16.7 Mpc
Inclination between the jet axis and the major axis of the beam kernel   16°
Beam size   1.14 mas × 0.55 mas

Download table as:  ASCIITypeset image

For Rp, Δ, and z1, T18 showed the dependence of the image on Rp with fixed values of Δ and z1. In the case of Rp = 0, the jet image has one component around the axis, while in the case of Rp > 0 (e.g., ${R}_{{\rm{p}}}=40\,{R}_{{\rm{S}}}$, which is the fiducial value in their paper), limb-brightened images are obtained. They focused on the region ∼1–4 mas from the radio core, which is nearer than the ∼10–30 mas where the triple-ridge structure of the M87 jet is observed. Even though we compute images with the same parameters as T18 for the far region, the triple-ridge structure cannot be obtained, as shown in the Appendix. However, we found that the intensity map before the convolution for Rp = 0 has a distinct bright component along the jet axis that may correspond to the observed inner ridge (see Figure 7). Then we fixed Rp = 0, while setting large Δ to have a brighter jet edge, and found that the triple-ridge image can be obtained with Δ = 2 RS and z1 = 5 RS. We show the calculation results with these parameter values in the next section.

Note that A and n0 should be specified to obtain absolute intensity. T18 showed that reasonable values for them can lead to the observed level of intensity. However, the absolute intensity also depends on other uncertain parameters (p and ${\gamma }_{m}^{{\prime} }$), as well as the physical processes that we do not take into account in our current model, such as the radiative cooling and reacceleration of electrons. In this paper, we treat A and n0 arbitrarily and focus on the relative intensity along the transverse direction of jets in order to simply illustrate our novel idea on the mechanism for the triple-ridge images.

3. Results

In Section 3.1, we show the resultant image of the triple-ridge structure, and figure out its physical origin. In Section 3.2, we demonstrate the parameter dependence of the image on the free parameters on the geometry of the jet Θ and ν and those on the electrons' spatial distribution Δ and z1.

3.1. Triple-ridge Structure

Figure 1 shows the resultant image with the computational resolution (left panel) and the one convolved with the Gaussian beam (right panel). The Gaussian beam size is represented as the gray circle in the bottom left corner of the right panel. For both of the images, the intensity is normalized by each maximum value and the contours represent the intensity at 2k (k = 1, 2, 3, ...). In addition to the nearly symmetric two outer ridges, the inner ridge emerges in both images. In the left panel we find that the inner ridge extends along the jet axis and has an asymmetric shape. The right panel shows that the triple-ridge structure is highly smoothed by the convolution but still remains in Y > 15 mas. The counter jet appears in Y < 0 in the left panel and is overwhelmed by the bright core in the right panel.

Figure 1.

Figure 1. Left panel: the computed image with the computational resolution of 3 RS. Right panel: the image convolved with the beam size (1.14 mas × 0.55 mas), where the beam size is plotted as the gray circle in the bottom left corner. The intensity is normalized by each maximum value, and the contours represent the intensity at 2k (k = 1, 2, 3, ..., 27 for the left panel and k = 1, 2, 3, ..., 21 for the right panel). The number of contour lines is not the same as in T18. Note that 1 mas corresponds to ≈136 RS.

Standard image High-resolution image

The physical reason for this triple-ridge image can be explained by focusing on the dependence of jω on n', B' and the relativistic boosting factor 1/Γ2(1 − βμ)3 (see Equations (10) and (11)):

Equation (13)

We show the distribution of the three factors, as well as vp/vϕ, in Figure 2. Note that the distribution of $\sin \alpha ^{\prime} $ is not relevant to the image structure in this problem (but see Shibata et al. 2003, for the problem of images of pulsar wind nebulae). We confirmed that the shapes of computed images when artificially setting $\sin \alpha ^{\prime} =1$ do not change from those shown in Figure 1.

Figure 2.

Figure 2. Spatial distribution of B', n', vp/vϕ, and $1/{{\rm{\Gamma }}}^{2}{(1-\beta \mu )}^{3}$ for Θ = 15° in the northern hemisphere. B' and n' are normalized by each maximum value. The white solid lines are the z axis, and the gray dashed lines represent the field lines passing through the BH horizon at the equatorial plane. The dotted lines are the contours of 10−2, 10−2.5, 10−3, and 10−3.5 in the upper left panel, 10−4, 10−5, and 10−6 in the upper right panel, 1 and 10 in the lower left panel, and 1000, 2000, 3000, and 4000 in the lower right panel. Note that 1 RS corresponds to ≈7.3μas.

Standard image High-resolution image

The upper left panel of Figure 2 shows the profile of B', which is roughly understood by writing down three magnetic field components (for z > 0) from Equations (1) and (4):

Equation (14)

For a given R, each component is roughly proportional to Ψ, so it decreases for larger z. The Lorentz transformation from the lab frame to the fluid frame does not significantly change the profile. The upper right panel of Figure 2 shows the density distribution. At high z, the density around the axis becomes much smaller than the jet edge even though we set the density distribution concentrated around the axis at z = z1. This is because the distribution along the field line is derived from Equation (7) and the magnetic field is stronger around the jet edge. Therefore, one can see that the bright outer ridges are produced by the strong magnetic field and high density around the jet edge.

The magnetic field is weaker and the electron density is lower around the axis, but the strong relativistic beaming makes the bright inner ridge. As seen in Equation (14) the magnetic field is poloidal-dominant at R ≪ cF and toroidal-dominant at R ≫ cF ($c/{{\rm{\Omega }}}_{{\rm{F}}}\simeq 2.13\,{R}_{{\rm{S}}}$). Correspondingly, the fluid velocity is toroidal-dominant at R ≪ cF and poloidal-dominant at R ≫ cF (see the lower left panel of Figure 2). In between, there is a region where the fluid velocity is almost parallel to the line of sight, i.e., μ ≈ 1. This enhances the factor $1/{{\rm{\Gamma }}}^{2}{(1-\beta \mu )}^{3}$, as shown in the lower right panel of Figure 2. Although the magnetic field, electron density, and Lorentz factor are small near the jet axis, the beaming effect produces the bright sharp inner ridge.

The transverse width of the relativistically beamed area, ∼10 RS ∼ 0.1 mas, shown in the bottom right panel of Figure 2, corresponds to the width of the inner ridge of the left panel of Figure 1. We expect that observations with better resolution can illustrate the narrower inner ridge. Besides, the inner ridge appears only on one side of the jet axis. Which side the inner ridge appears on depends on the sign of ΩF and tells us the direction of the BH rotation.

3.2. Parameter Dependence

We calculate the images with different Θ, ν, Δ, and z1 to examine the parameter dependence on the image structure. For illustration, we take Y = 25 mas, for which we plot the resultant transverse intensity profiles in Figure 3. Note that we normalize the intensity by the value at X = 0 of each line in Figure 3. The upper left panel represents the intensity profiles in the case of Θ = 7°, 10°, 15°, 20°, and 75°, with the other parameters unchanged. For smaller Θ, the relation $z=Y/\sin {\rm{\Theta }}$ for a fixed value of Y means that the line of sight passes points farther from the BH, and the jet width appears to be larger. For larger Θ, the outer ridges are debeamed because the line of sight and the beaming cone of the outer ridges are misaligned, while the inner ridge remains.

Figure 3.

Figure 3. Dependence of the transverse intensity profile at Y = 25 mas on Θ (upper left), ν (upper right), Δ (lower left), and z1 (lower right). The solid line in each panel is identical to the case of Figure 1. The dashed, solid, dotted, dashed–dotted, and dashed double-dotted lines represent the results calculated when changing the parameter values from the case of the solid line as shown in each panel. The vertical axis is normalized by the value at X = 0 of each line.

Standard image High-resolution image

The upper right panel of Figure 3 shows the intensity profiles in the case of ν = 0.7, 0.75, and 0.8 with the other parameters unchanged. When ν increases, the field lines and density concentrate on the axis. This leads to an image shaped like a candle flame. On the other hand, when ν decreases, the field lines and density distribution expand. This results in the bright outer ridges, which dominate the inner-ridge component.

As for the dependence on the parameters related to the density distribution, we examine only the dependence on Δ and z1 because the dependence on Rp was discussed in T18. As Δ increases or z1 decreases, the density in the jet edge increases and the outer ridges become brighter, as shown in the lower panels in Figure 3. The positions X of the outer ridge peaks do not change between the cases of Δ = 2 RS and Δ = 3 RS, and between the cases of z1 = 2.5 RS and z1 = 5 RS because the jet width is limited by the outermost magnetic field line threading the BH. Although our setup for the electron density distribution is a toy model, this analysis indicates that the observed image structure can strongly constrain the spatial electron density distribution when Θ and ν are estimated by other observational information such as the blob pattern speed, the brightness ratio between the approaching and counter jets, and the width profile of the jet.

4. Discussion

4.1. Valleys between Ridges

As seen in Figures 1 and 3, the inner ridge and outer ridges produced by our model are not so clearly separate as reported in Hada (2017). This is because the jet edges have a sheath-like three-dimensional structure and the emission from the sheath enhances the brightness of the valleys between the ridges.

For more details, we show jω distributions along lines of sight passing (X, Y) = (±0.5 mas, 25 mas), (±1 mas, 25 mas), and (±2 mas, 25 mas) as functions of z in Figure 4 to examine which parts of the jet contribute to the transverse intensity profile ${I}_{\omega }(X,Y=25\,\mathrm{mas})$ (the black lines in Figure 3). The jet emission is composed of the sheath-like jet-edge component and the beamed inner-ridge component. For the line of sight of (X, Y) = (−0.5 mas, 25 mas), the inner-ridge component appears in addition to the jet-edge component, while for the line of sight of (X, Y) = (0.5 mas, 25 mas), the emission around the axis is debeamed. This leads to the asymmetry of the inner ridge in our computed image, which we have pointed out for the left panel of Figure 1 in Section 3.

Figure 4.

Figure 4. Distribution of jω along the line of sight of (X, Y) = (±2 mas, 25 mas) (dotted lines), (±1 mas, 25 mas) (dashed lines), (±0.5 mas, 25 mas) (solid lines). The red (blue) lines are for negative (positive) X cases. The horizontal axis represents the height z at the point on the line of sight. All the values of jω are normalized by the maximum value for (X, Y) = (−2 mas, 25 mas).

Standard image High-resolution image

The lines of sight passing (X, Y) = (±1 mas, 25 mas) and (±2 mas, 25 mas), for which the valleys and outer ridges are seen, respectively (see Figures 1 and 3), penetrate only the jet edge. Interestingly, jω at the rear part of jet edge z ≲ 1.25 × 104 RS is comparable to that at the front part of jet edge z ≳ 1.4 × 104 RS for each of these sight-lines. That is because the fluid motion at the rear part directs away from the line of sight and then the emission is debeamed, but n' and B' are larger there than those at the front part with higher z. Figure 4 implies that the intensities ${I}_{\omega }=\int {j}_{\omega }{dZ}$ for (X, Y) = (±1 mas, 25 mas) and (±2 mas, 25 mas) should be comparable, which means that the valleys are not deep.

The observed deep valleys may indicate more complex structure of the jet. For example, if the non-thermal electrons distributed separately at the spine and a thin layer of the jet edge, then the brightness for (X, Y) = (±1 mas, 25 mas) would become lower. A thin layer with dense non-thermal electrons would produce bright outer ridges with deep valleys.6 We also consider that a non-axisymmetric jet may lead to deep valleys. For example, if only the rear part of the jet edge at $-1\,\mathrm{mas}\lesssim X\lesssim 1\,\mathrm{mas}$ is intrinsically dim, the intensities at the valleys halve when keeping the outer ridges bright.

4.2. Bulk Lorentz Factor at the Far Zone

We assume that the force-free approximation is valid even at the far zone from the BH. Then the bulk Lorentz factor Γ keeps increasing. Figure 5 represents the Lorentz factor profiles along the magnetic field lines that satisfy Ψ = Ψ (r+, π/2), Ψ = Ψ (r+, π/2)/2, and Ψ = Ψ (r+, π/2)/10 in the model of Figure 1. These obey the analytical relation $1/{{\rm{\Gamma }}}^{2}=1/{{\rm{\Gamma }}}_{1}^{2}+1/{{\rm{\Gamma }}}_{2}^{2}$, where ${{\rm{\Gamma }}}_{1}^{2}={B}^{2}/{B}_{{\rm{p}}}^{2}\propto {R}^{2}$ and ${{\rm{\Gamma }}}_{2}^{2}={B}^{2}/({B}_{\phi }^{2}-{E}^{2})\propto ({R}_{{\rm{c}}}/R)$ at the far region $R{{\rm{\Omega }}}_{F}\gg c$, where Rc is the curvature radius of the poloidal field line. This asymptotic relation is shown in Tchekhovskoy et al. (2008) for the force-free case7 and in Komissarov et al. (2009) and Lyubarsky (2009) for the Poynting-dominated cold ideal MHD case.

Figure 5.

Figure 5. Upper panel: Lorentz factor profiles along the outermost field line of Ψ = Ψ(r+, π/2) (black solid line), Ψ = Ψ(r+, π/2)/2 (red dashed line), and Ψ = Ψ(r+, π/2)/10 (blue dotted line) in the model of Figure 1. Lower panel: the shape of the field lines of the upper panel. The upper panel is plotted in log scale and the lower panel is in linear scale.

Standard image High-resolution image

The Lorentz factor increases up to ∼80 in the computed area along each field line shown in Figure 5. This is much higher than the Lorentz factors deduced from blob motions observed in the VLBI observations (Mertens et al. 2016; Hada 2017, and references there in), Γ ≲ 10, although it is possible that faster blobs are not identified due to the apparent decrease of the number of fast blobs (Komissarov & Falle 1997) and the low time resolution of the current monitoring (Nakamura et al. 2018). If the Lorentz factors of currently identified blob motions are similar to those of the steady flows in jets, we require MHD models with Poynting to kinetic energy flux ratio (i.e., σ parameter) at the jet base as low as ∼10, which gives lower-saturation Lorentz factors.

4.3. Inner-ridge Property

The inner ridge of our computed image arises because there is a region where the direction of the velocity is parallel to the line of sight between the toroidal-dominant region (${v}_{\phi }\gg {v}_{{\rm{p}}}$, RΩF/c ≪ 1) and poloidal-dominant region (vpvϕ, RΩF/c ≫ 1), as explained in Section 3.1. We have demonstrated this property by setting the fluid velocity to be the drift velocity (Equation (5)). The cold ideal MHD velocity also has the same property as the drift velocity, i.e., it is toroidal-dominant at RΩF/c ≪ 1, while it is poloidal-dominant at RΩF/c ≫ 1 (Beskin 2010; Toma & Takahara 2013; Pu et al. 2015, T18).

However, if the jet particles are injected along the magnetic field lines with the large Lorentz factor, Γin, at the jet base, the velocity structure will change, i.e., the velocity at the region R ΩF/c ≪ 1 where Bp ≫ Bϕ will become dominated by the poloidal component (Beskin 1997; Beskin & Malyshkin 2000). We plot the Lorentz factor close to the axis for our fiducial model in Figure 6, which shows that 5 ≲ Γ ≲ 9 around the region that produces the inner ridge (cF ≲ 10RS ≲ R ≲ 20RS; see the bottom right panel in Figure 2). If Γin is smaller than 5, the velocity structure at R ≳ 10RS will not be changed, so that the triple-ridge structure will be still observed. For a smaller viewing angle, the inner-ridge radius is larger (as seen in the upper left panel of Figure 3) and the Lorentz factor at that radius is larger, so that the triple-ridge structure will still appear in cases of even larger Γin. We note that the large Γin will also change the field configuration because of the particle inertia. These effects have not been taken into account either in GRMHD simulations. We need a more sophisticated model to consider them in detail.

Figure 6.

Figure 6. Spatial distribution of Γ. The dotted lines are the contours of 3, 5, 7, 9, 11, 13, 15. The white solid line and the gray dashed line are the same as those in Figure 2.

Standard image High-resolution image

The lower right panel of Figure 2 shows that the inner-ridge radius does not depend on z. In contrast, Asada et al. (2016) indicates that the inner-ridge width of the M87 jet varies as a function of the distance from the BH. They showed that farther from the BH, the inner ridge becomes wider at 5 GHz, while it becomes narrower at 1.6 GHz. This might be caused by the synchrotron cooling, as mentioned in Asada et al. (2016), or the time variation of the jet, neither of which are included in our model.

5. Summary

We have examined a steady axisymmetric force-free model of a jet driven by BH, in which the electromagnetic structure is set to be consistent with GRMHD simulation results, and shown that the triple-ridge structure of a relativistic jet can be produced by a model with a simple Gaussian distribution of emitting electrons at the jet base (z = z1). We have found that the fluid-drift velocity associated with such a field structure produces the inner ridge through the relativistic beaming effect, and the magnetic field strength and electron number density are higher near the jet edge, and produce the outer ridges. Thus, we argue that the observed triple-ridge image does not directly indicate the requirement of the two jet-launching processes working simultaneously, such as the BZ process for the rotating BH plus the BP process for the accretion disk. This argument appears to be consistent with the finding of T18 that the jet from the accretion disk produces highly asymmetric images that are unlike the observed limb-brightened images of the M87 jet, and with the GRMHD simulation result of Nakamura et al. (2018) in which the outermost parabolic streamline of the jet driven by the BZ process overlaps the edge of the observed M87 jet.

We also have found that the jet image is very sensitive to the height z1 of the base of electron spatial distribution and the width of its distribution Δ, as well as to the geometric parameters of jet Θ and ν. This means that the characteristic jet images as observed with recent sensitive VLBI observations at 43 and 86 GHz can strongly constrain the spatial distribution of injected electrons near the BH, when Θ and ν are estimated by other observational information such as the brightness ratio between the approaching and counter jets and the width profile of the jet. Such studies must be complementary to those directly investigating the physics closely around the BH with the upcoming EHT data (Doeleman et al. 2012; Akiyama et al. 2017).

Our model does not reproduce the sharpness of the observed ridges and the width variation of the inner ridge as a function of z in the M87 jet. They may be caused by distinct production of the non-thermal electrons at the spine and sheath, non-axisymmetry of the jet, temporal variation of the jet, and synchrotron cooling and/or reacceleration of electrons. Including such effects in the jet model is work that would have to be done separately.

We thank the participants in the Mizusawa Project Meetings in 2016, 2017, and 2018 for fruitful discussions on relativistic jets. Numerical calculations were performed on Draco, a computer cluster of the Frontier Research Institute for Interdisciplinary Sciences in Tohoku University. T.O. acknowledges financial support from the Graduate Program on Physics for the Universe of Tohoku University. This work is partly supported by JSPS Grants-in-Aid for Scientific Research 17H06362 (K. Takahashi), 18H01245 (K. Toma), and also by a JST grant Building of Consortia for the Development of Human Resources in Science and Technology (K. Toma).

Appendix: T18 Model at the Far Zone

T18 showed the limb-brightened images at ∼1–4 mas from the core with parameter values MBH = 3.4 × 109 M, a = 0.998, ΩF = 0.5ΩH, Θ = 25°, ν = 1, z1 = 5 RS, Δ = 5 RS, and the beam size 0.43 mas × 0.21 mas (Walker et al. 2008). We show the resultant synchrotron images with these parameter values up to 30 mas, which do not exhibit the triple-ridge structure.

Figure 7 shows the result for the case of Rp = 0. The jet image does not have the triple-ridge structure even at Y > 4 mas, but a narrow bright component near the axis is remarkable, which corresponds to the inner ridge from our viewpoint.

Figure 7.

Figure 7. Synchrotron images with the same parameters as the case of Rp = 0 in T18. Left: image with the computational resolution of 5 RS. Right: image convolved with the beam size (0.43 mas × 0.21 mas). The contours represent the normalized intensity at 2k (k = 1, 2, 3, ..., 30 for the left panel and k = 1, 2, 3, ..., 26 for the right panel).

Standard image High-resolution image

Figure 8 shows the result for the case of Rp = 40 RS. The limb-brightened structure is obtained, and no clear inner ridge was found in this case. The ring-like distribution of electrons does not produce high emissivity around the axis.

Figure 8.

Figure 8. Synchrotron images with the same parameters as the case of Rp = 40 RS in T18. Left: image with the computational resolution of 5 RS. Right: image convolved with the beam size (0.43 mas × 0.21 mas). The contours represent the normalized intensity at 2k (k = 1, 2, 3, ..., 23 for the left panel and k = 1, 2, 3, ..., 21 for the right panel).

Standard image High-resolution image

Footnotes

  • Very recently, the physics of the gaps in BH magnetospheres have been studied with particle-in-cell simulations (Chen et al. 2018; Levinson & Cerutti 2018; Parfrey et al. 2019).

  • Some other papers assume that the non-thermal electron density is related to the magnetic energy density (Broderick & McKinney 2010; Porth et al. 2011; Dexter et al. 2012).

  • From a simple geometric consideration for the jet with a thin layer of the width ΔR viewed at Θ = 90°, one can see that the ratio of the path lengths along the lines of sight across the thin layer for the outer ridge and valley scales as $\propto \sqrt{{\rm{\Delta }}R}$.

  • The asymptotic relation (47) of the Lorentz factor in T18 is not complete but valid only for the second acceleration regime. In the first acceleration regime, Γ is given as ${\rm{\Gamma }}\sim \sqrt{1+{\left(\tfrac{R{\rm{\Omega }}}{{cg}(\nu ,\theta )}\right)}^{2}}$, which is proportional to R when the second term dominates.

Please wait… references are loading.
10.3847/1538-4357/ab1909