Anda di halaman 1dari 15

MNRAS 000, 1–15 (2019) Preprint 3 July 2019 Compiled using MNRAS LATEX style file v3.

Discovery of geometry transition in nearby AGN jets

Y. Y. Kovalev1,2,3⋆ , A. B. Pushkarev4,1, E. E. Nokhrina2, A. V. Plavin1,2,


V. S. Beskin1,2, A. Chernoglazov1,2, M. L. Lister5, T. Savolainen6,7,3
1 LebedevPhysical Institute, Leninsky prosp. 53, Moscow, 119991, Russia
2 Moscow Institute of Physics and Technology, Dolgoprudny, Institutsky per., 9, Moscow region, 141700, Russia
3 Max-Planck-Institut für Radioastronomie, Auf dem Hügel 69, 53121 Bonn, Germany
4 Crimean Astrophysical Observatory, Nauchny 298688, Crimea, Russia
arXiv:1907.01485v1 [astro-ph.GA] 2 Jul 2019

5 Department of Physics and Astronomy, Purdue University, 525 Northwestern Avenue, West Lafayette, IN 47907, USA
6 Aalto University Metsähovi Radio Observatory, Metsähovintie 114, 02540 Kylmälä, Finland
7 Aalto University Department of Electronics and Nanoengineering, PL 15500, 00076, Aalto, Finland

Accepted XXX; Received YYY; in original form 23 June 2019

ABSTRACT
Observational studies of collimation in jets in active galactic nuclei (AGN) are a key
to understanding their formation and acceleration processes. We have performed an
automated search for jet shape breaks in a sample of 367 AGN using VLBA data at
2 cm and 22 cm. This search has found ten nearby jets at redshifts z < 0.07 with a
transition from a parabolic to conical shape, while the full sample is dominated by
distant AGN with a typical z ≈ 1. The ten AGN are UGC 00773, NGC 1052, 3C 111,
3C 120, TXS 0815−094, Mrk 180, PKS 1514+00, NGC 6251, 3C 371, and BL Lac. We
conclude that the geometry transition may be a common effect in AGN jets. It can
be observed only when sufficient linear resolution is obtained. Supplementing these
results with previously reported shape breaks in the nearby AGN 1H 0323+342 and
M87, we estimate that the break typically occurs at a distance of 105 –106 gravitational
radii from the nucleus. We suggest that the jet shape transition happens when the
bulk plasma kinetic energy flux becomes equal to the Poynting energy flux, while the
ambient medium pressure is assumed to be governed by Bondi accretion. Our model
predictions on the jet acceleration and properties of the break point are found to be
supported observationally.
Key words: galaxies: jets – galaxies: active – radio continuum: galaxies – quasars:
general – BL Lacertae objects: general

1 INTRODUCTION equipartition between energy densities of the magnetic


field and the radiating particle populations. It has been
Understanding the physical processes that determine the
shown for cold jets that acceleration should not happen in a
formation, acceleration and collimation of relativistic jets
conical jet. This requires something akin to a parabolic jet
in active galactic nuclei (AGN) continues to be among the
shape closer to the jet base to allow differential expansion
most challenging problems of modern astrophysics. There
(Vlahakis & Königl 2004; Komissarov 2012).
are a wide variety of analytical and numerical models for jet
acceleration and its confinement (e.g., Vlahakis & Königl In order to investigate these theories it is important
2003; Beskin & Nokhrina 2006; McKinney 2006; to collect observational data on jet profile shapes for a
Komissarov et al. 2007; Tchekhovskoy et al. 2011; large enough sample of AGN whose properties are well un-
McKinney et al. 2012; Potter & Cotter 2015) that consider derstood. The first observational evidence for a transition
different solutions for jet shapes, such as cylindrical, conical from parabolic to conical jet shape was detected in M87
and parabolic. General relativistic magnetohydrodynamic (Asada & Nakamura 2012) at a distance of about 900 mas
simulations (e.g., McKinney et al. 2012) predict that a jet near the feature HST-1, about 70 pc in projection, corre-
starting from its apex has a parabolic streamline within the sponding to 105 Schwarzschild radii. A few more studies
magnetically dominated acceleration zone and transitions of nearby AGNs to probe their innermost jet regions were
to a conical geometry at outer scales associated with performed recently: Mkn 501 (Giroletti et al. 2008), Cen-
taurus A (Müller et al. 2014), Cygnus A (Boccardi et al.
2016; Nakahara et al. 2019), NGC 6251 (Tseng et al. 2016),
⋆ E-mail: yyk@asc.rssi.ru 0323+342 (Hada et al. 2018), 3C 273 (Akiyama et al. 2018),

c 2019 The Authors



2 Kovalev et al.
NGC 4261 (Nakahara et al. 2018), 3C 84 (Giovannini et al. with Ωm = 0.27, ΩΛ = 0.73 and H0 = 71 km s−1 Mpc−1
2018), 3C 264 (Boccardi et al. 2019). Hovatta et al. (2019) (Komatsu et al. 2009).
have indirectly addressed this question for the 3C 273 jet
close to the central engine on the basis of a model analy-
sis of ALMA rotation measure data. Larger survey studies 2 A DISCOVERY OF SHAPE TRANSITION IN
(Pushkarev et al. 2009) have typically probed regions far- AGN JETS
ther away from the central nucleus, although Algaba et al.
(2017) have used apparent parsec-scale jet base parameters 2.1 Automated search of candidates with a
closer in. change in jet geometry
In a previous work (Pushkarev et al. 2017), we analyzed For the purposes of our study, we made use of data at 15 GHz
parsec-scale radio VLBI images of jets in 362 active galaxies. from the MOJAVE program, the 2-cm VLBA Survey, and
This sample is dominated by compact radio bright blazars the National Radio Astronomy Observatory (NRAO) data
with jet at a small angle to the line of sight and a typi- archive (Lister et al. 2018) for those sources that have at
cal redshift of one. However, some low luminosity nearby least five VLBA observing epochs at 15 GHz between 1994
radio galaxies were also observed. Pushkarev et al. (2017) August 31 and 2016 December 26 inclusive. We used the
show that while the majority of resolved jets indeed have a 2 cm VLBA total intensity MOJAVE stacked epoch im-
shape close to conical, a significant fraction of the sample ages supplemented by single epoch 22 cm VLBA images to
has observed deviations. A systematic change in jet width derive apparent jet widths (d) as a function of projected
profile has been noted by Hervet et al. (2017), who explain distance r from the jet core, and determined jet shapes
it by using a stratified jet model with a fast spine and slow similar to Pushkarev et al. (2017). In that work we fit-
but relatively powerful outer layer. In this paper, we inves- ted the d–r dependence with a single power law d ∝ r k .
tigate if this outcome is partly affected by the typical finite The index is expected to be k ≈ 0.5 for a quasi-parabolic
resolution of VLBI observations. We probe a possible de- shape and 1.0 for a conical jet. We note that even single-
pendence of the jet shape on the distance r to the nucleus. epoch observations at 1.4 GHz adequately reproduce source
Furthermore, we perform a systematic search for a possible morphology, i.e., effectively fill jet cross-section due to a
transition from one jet shape to another on the basis of 2 cm steep spectrum of synchrotron emission of the outflow, with
and 22 cm VLBA images. a typical spectral index −0.7 measured between 2 and
The observation of sources with the change in a jet 8 GHz (Pushkarev & Kovalev 2012) and −1.0 between 8 and
shape from the parabolic to conical may provide an instru- 15 GHz (Hovatta et al. 2014), making the low-frequency ob-
ment to probe the MHD acceleration mechanism models as servations sensitive enough to probe jet morphology at larger
well as the ambient medium conditions. The change in jet scales
shape in M87 (Asada & Nakamura 2012) coincides with the We have carried out a similar analysis allowing for a
stationary bright feature HST-1, which can be associated change in the jet shape. Using all available data (15 GHz
with the change in ambient pressure profile and appearance only or combined data set at 15 and 1.4 GHz) for each
of a recollimation shock due to pressure drop and jet abrupt source, we performed a double power law fit of the jet width
expansion. This interpretation is supported by the measure- as a function of distance, dividing the jet path length in a
ments of external medium pressure by Russell et al. (2015) logarithmic scale by two parts in proportion of 1:1, 1:2 and
almost down to the Bondi radius, with an observed pres- 1:3 to search for cases when the fitted k-index at inner scales
sure profile P ∝ r −1 . The recently observed jet shape in was 0.5 ± 0.2, while at outer scales it was 1.0 ± 0.2. After
0323+342 (Hada et al. 2018) demonstrates a similar behav- such cases were identified automatically, we tuned the fits
ior. On the other hand, there are models predicting a jet by setting the distance of the transition region by eye.
shape transition for a single power law pressure profile. The We ended up dropping 36 AGN jets as having unsatis-
analytical model by Lyubarsky (2009) predicts the transi- factory fits caused by either (i) non-optimal ridge line recon-
tion from parabolic to conical form for certain regimes, as struction for jets with strong bending, (ii) numerous large
well as quasi-oscillations in jet shape in the conical domain. gaps in jet emission, (iii) too short a jet length (iv) low in-
This solution has been applied to the reconstruction of the tensity regions not captured well by our jet width fitting.
recollimation shock properties of M87 by Levinson (2017), This resulted in a sample comprising 331 AGN jets.
with a predicted total jet power on the order of 1043 erg/s. As a result of this analysis, we found a shape transition
The recent semi-analytical results for the warm jet match- in ten jets (Table 1) out of 367 analyzed. We emphasize that
ing the ambient medium with a total electric current closed all the AGNs with detected transition of the jet shape turned
inside a jet by Beskin et al. (2017) predicts a change in a out to have low redshifts z < 0.07, i.e., have a high linear res-
jet shape from parabolic to conical for the Bondi pressure olution of VLBI observations. This is highly unlikely to oc-
profile P ∝ r −2 . In this work we follow the latter model and cur by chance and provides additional strong evidence that
consider the results for a warm outflow in more detail. this result is not an observational artifact but a real effect.
The structure of the paper is the following: section 2 See discussion of the rest of analyzed low redshift AGN in
presents our results of a search for the jet profile change the sample in subsection 2.4. Among the ten sources, there
from parabolic to conical in a large sample of AGN jets, we is one, the radio galaxy 0238−084 (NGC 1052), that shows
interpret our findings in section 3. We summarize our work a two-sided jet morphology. For this object, we analyzed
in section 4. Throughout this paper we will use the term the approaching, brighter outflow propagating to north-east
“core” as the apparent origin of AGN jets that commonly direction, determining the position of a virtual VLBI core
appears as the brightest feature in VLBI images of blazars using a kinematic-based minimization method described in
(e.g., Lobanov 1998; Marscher 2008). We adopt a cosmology Vermeulen et al. (2003).

MNRAS 000, 1–15 (2019)


Transition of AGN jet geometry 3

Figure 1. Jet profile with a sign of transition from parabolic to conical shape in ten well resolved nearby active galaxies. Dependence
of the jet width on projected distance to apparent jet base is shown. The cyan and orange dots show measurements at 15 and 1.4 GHz,
respectively. The red and black stripes represent Monte Carlo fits for jet regions before and beyond the jet shape transition region,
respectively. The projected distance is shown in mas for targets with unknown redshift and in pc for known redshift cases. General
properties of these AGN are presented in Table 1, parameters of the fits — in Table 2, parameters of the shape transition region — in
Table 3.

MNRAS 000, 1–15 (2019)


4 Kovalev et al.

Table 1. Source properties for 12 sources with the detected jet shape break. Ten of them are from this study (Figure 1), they are supple-
mented by 0321+340 (Hada et al. 2018) and M87 (Asada & Nakamura 2012). The full online only table adds properties of MOJAVE-1
sources for which redshift and Doppler factor estimates are available, the full table also contains references. Columns are as follows:
(1) B1950 name; (2) alias; (3) optical class, where Q = quasar, B = BL Lac, G = radio galaxy, N = Narrow Line Seyfert 1 (NLSy1);
(4) redshift; (5) maximum apparent radial speed from Lister et al. (2019), (6) variability Doppler factor from Hovatta et al. (2009); (7)
viewing angle; (8) black hole mass estimated basing on assumption of virialized broad lines region (BLR) movement and correlation
between the size of BLR and UV/optical luminosity; (9) black hole mass estimated by a stellar velocity dispersion method and associated
fundamental plane method (for 2200+420).

Source Alias Opt. z βapp δ θ MBH MBH


ID (c) (deg) (log(M⊙ )) (log(M⊙ ))
(1) (2) (3) (4) (5) (6) (7) (8) (9)

0111+021 UGC 00773 B 0.047 0.17 ... 5.0b ... ...


0238−084 NGC 1052 G 0.005 0.42 0.3a 49.0 5.51 8.19
0321+340 1H 0323+342 N 0.061 9.02 ... 6.3 7.30 ...
0415+379 3C 111 G 0.049 8.15 2.0a 13.4 8.21 ...
0430+052 3C 120 G 0.033 5.27 2.1a 18.7 7.52 8.13
0815−094 TXS 0815−094 B ... ... ... 5.0b ... ...
1133+704 Mrk 180 B 0.045 ... ... 5.0b ... 8.21
1228+126 M87 G 0.004 ... ... 14.0 ... 9.82
1514+004 PKS 1514+00 G 0.052 0.45 ... 15.0c ... ...
1637+826 NGC 6251 G 0.024 ... ... 18.0 ... 8.78
1807+698 3C 371 B 0.051 0.01 1.1 7.3 7.14 8.51
2200+420 BL Lac B 0.069 9.96 7.2 7.6 ... 8.23

Note. León Tavares et al. (2014) argue that using the single-epoch spectra scaling relations might lead to the significant black hole
mass underestimation for 0321+340, they provide an independent estimate of 108.6 M⊙ .
a Doppler factor is from Liodakis et al. (2017).
b Assumed θ value as typical for BL Lacs.
c Assumed θ value as typical for radio galaxies in the list which do not show a strong counter-jet.

Following our discovery of the shape transition pref- in PYMC3 (Salvatier et al. 2016), which automatically ac-
erentially occurring in nearby AGN, we supplemented our counts for uncertainties of all the parameters in further infer-
initial AGN sample of 362 targets from Pushkarev et al. ences. The best fit parameters are listed in Table 2, showing
(2017) with stacked images of five more low-z AGN which that initially the jet is quasi-parabolic with k1 close to 0.5,
had five or more 15 GHz VLBA observing epochs after while beyond the break point region the outflow manifests
the Pushkarev et al. analysis was finished. These were: a streamline close to conical, with k2 ≈ 1. The location of
0615−172, 1133+704, 1200+608, 1216+061, 1741+196. the jet shape break given in Table 3 is estimated as the in-
All the stacked images are available from the MOJAVE tersection point of these two d − r dependencies. Note that
database1 . Table 3 includes results on the jet shape transition region
for two more sources, 0321+340 and 1228+126 (M87) taken
from Hada et al. (2018) and Nokhrina et al. (2019), respec-
2.2 Rigorous fitting of the jet shape tively.
For each of the found 10 sources with a sign of jet geometry For other sources without a detected shape break, we
transition, we fit the data with the following dependencies: fit a single power-law d = a(r + r0 )k for consistency. Ob-
d = a1 (r + r0 )k1 and d = a2 (r + r1 )k2 , describing a jet shape jects with unreliable ridge line detection or patchy structure
before and after the break. Here r0 is understood as the in images (15 sources) and those with nonphysical d − r
separation of the 15 GHz apparent core from the true jet dependence (24 sources) were excluded after visual inspec-
origin due to the synchrotron opacity (e.g., Lobanov 1998; tion. They constitute only about one tenth of the dataset
Pushkarev et al. 2012a), while r1 shows how much one un- and thus the exclusion should not bias our estimates. To ac-
derestimates the jet length if it is derived from the data only count for increased uncertainties of jet width measurements
beyond the geometrical transition of the jet. We note that further from the core, the power law model is complemented
this approach is more accurate but more computationally as following:
(
intensive than that used by Pushkarev et al. (2017) and ap- a(r + r0 )k + N (0, σ12 ), if r < R
plied in the original selection of jet break candidates. It is d=
b + N (0, σ22 ), if r > R
needed in order to better fit for jet shape close to the apex.
We fit these dependencies with Bayesian modeling us- Here all of a, r0 , k, R, σ1 , b, σ2 are treated as unknown param-
ing the NUTS Markov Chain Monte Carlo sampler based on eters and inferred simultaneously using a Nested Sampling
the gradient of the log posterior density. It was implemented algorithm as implemented in PolyChord (Handley et al.
2015). As expected, σ2 is typically significantly larger than
σ1 . We find that this model generally captures the d − r de-
1 http://www.physics.purdue.edu/MOJAVE/ pendence and its uncertainty well. Fitting results are given

MNRAS 000, 1–15 (2019)


Transition of AGN jet geometry 5

Table 2. Derived best-fit parameters of the two fitted dependencies d = a1 (r + r0 )k1 and d = a2 (r + r1 )k2 before and after the jet
break, respectively. The online only table presents properties of a single fit for 319 AGN without detected shape break, their k values
are presented in Figure 2.

Source a1 r0 r0 k1 Band rmin rmax


(pc) (mas) (mas) (mas)

0111+021 0.179 ± 0.010 0.143 ± 0.085 0.157 ± 0.093 0.497 ± 0.077 U 0.2 1.5
0238−084 0.078 ± 0.007 0.074 ± 0.038 0.740 ± 0.380 0.391 ± 0.048 U 0.3 2.5
0415+379 0.305 ± 0.011 0.042 ± 0.020 0.044 ± 0.021 0.468 ± 0.026 U 0.2 6.0
0430+052 0.202 ± 0.015 0.122 ± 0.071 0.188 ± 0.109 0.556 ± 0.070 U 0.2 2.0
0815−094a 0.294 ± 0.015 ... 0.163 ± 0.048 0.527 ± 0.044 U 0.2 1.0
1133+704 0.437 ± 0.013 0.061 ± 0.046 0.069 ± 0.052 0.528 ± 0.040 U 0.3 1.5
1514+004 0.171 ± 0.011 0.189 ± 0.088 0.189 ± 0.088 0.564 ± 0.048 U 0.2 3.5
1637+826 0.155 ± 0.005 0.098 ± 0.044 0.204 ± 0.092 0.506 ± 0.041 U 0.2 3.0
1807+698 0.207 ± 0.016 0.130 ± 0.089 0.133 ± 0.091 0.388 ± 0.087 U 0.2 1.4
2200+420 0.505 ± 0.029 0.087 ± 0.096 0.067 ± 0.074 0.537 ± 0.057 U 0.9 2.0

a2 r1 r1 k2 Band rmin rmax


(pc) (mas) (mas) (mas)

0111+021 0.252 ± 0.031 −1.126 ± 0.181 −1.237 ± 0.199 0.934 ± 0.054 U 1.5 14.0
0238−084 0.252 ± 0.021 −0.069 ± 0.067 −0.690 ± 0.670 1.052 ± 0.081 U 2.5 14.1
0415+379 0.123 ± 0.019 −2.043 ± 0.100 −2.151 ± 0.105 1.175 ± 0.046 UL 6.0 60.8
0430+052 0.229 ± 0.022 −0.500 ± 0.188 −0.769 ± 0.289 1.131 ± 0.027 UL 2.0 121.9
0815−094a 0.282 ± 0.033 ... −0.085 ± 0.097 1.032 ± 0.049 U 1.0 14.3
1133+704 0.921 ± 0.057 −0.753 ± 0.083 −0.857 ± 0.094 0.828 ± 0.047 U 1.5 5.1
1514+004 0.185 ± 0.010 −1.167 ± 0.102 −1.167 ± 0.102 0.886 ± 0.022 U 3.5 14.3
1637+826 0.175 ± 0.007 −0.089 ± 0.048 −0.185 ± 0.100 0.730 ± 0.029 U 3.0 8.9
1807+698 0.179 ± 0.018 −0.120 ± 0.188 −0.122 ± 0.192 1.023 ± 0.025 UL 1.4 85.2
2200+420 0.433 ± 0.016 −1.142 ± 0.093 −0.885 ± 0.072 1.124 ± 0.009 UL 2.0 49.0

a Redshift unknown.

Table 3. Derived parameters of the jet shape break for 10 AGN with addition of 0321+340 adopted from Hada et al. (2018) and M87
from Nokhrina et al. (2019). Columns are as follows: (1) source name (B1950); (2) jet width at the break in mas; (3) jet width at the
break in pc; (4) projected distance of a break from the apparent core along the jet in mas; (5) projected distance of a break from the
BH along the jet in mas; (6) projected distance of a break from a BH along the jet in pc; (7) deprojected distance of a break from a BH
along the jet in pc, the parameter uses value of a viewing angle; (8) presence of a bright low pattern speed feature (Lister et al. 2019).

proj proj proj deproj


Source dbreak rbreak, app rbreak rbreak rbreak Stationary
(mas) (pc) (mas) (mas) (pc) (pc) jet feature
(1) (2) (3) (4) (5) (6) (7) (8)

0111+021 0.30 ± 0.03 0.28 ± 0.03 2.46 ± 0.27 2.62 ± 0.29 2.38 ± 0.26 27.31 Y
0238−084 0.53 ± 0.05 0.05 ± 0.01 2.93 ± 0.57 3.73 ± 0.65 0.37 ± 0.06 0.49 ...
0321+340 1 1.16 10 10.04 11.64 106.07 Y
0415+379 0.78 ± 0.03 0.74 ± 0.03 7.03 ± 0.50 7.07 ± 0.50 6.72 ± 0.47 29.00 ...
0430+052 0.45 ± 0.06 0.29 ± 0.04 2.67 ± 0.40 2.85 ± 0.41 1.85 ± 0.27 5.77 ...
0815−094 0.37 ± 0.05 ... 1.37 ± 0.30 1.54 ± 0.30 ... ... ...
1133+704 0.57 ± 0.02 0.50 ± 0.02 1.39 ± 0.09 1.46 ± 0.10 1.29 ± 0.09 14.80 Y
1228+126 15.73 ± 0.39 2.42 ± 0.06 ... 273 ± 12 21.0 ± 0.9 87.0 Ya
1514+004 0.34 ± 0.02 0.34 ± 0.02 3.10 ± 0.22 3.39 ± 0.30 3.39 ± 0.30 13.10 Y
1637+826 0.32 ± 0.02 0.16 ± 0.01 1.92 ± 0.26 2.13 ± 0.28 1.02 ± 0.13 3.30 ...
1807+698 0.26 ± 0.04 0.25 ± 0.04 1.53 ± 0.33 1.67 ± 0.34 1.63 ± 0.33 12.83 Y
2200+420 0.74 ± 0.03 0.95 ± 0.04 2.45 ± 0.10 2.52 ± 0.13 3.25 ± 0.16 24.57 ...

a This refers to the well-known HST-1 feature (Chang et al. 2010) which is normally not sampled by MOJAVE images.

in Table 2 and the source distribution of exponents k is in the histogram bin at k = 0.5 which corresponds to the
shown in Figure 2. Even though the estimates for individual parabollic jet shape; the number of objects with k ≈ 0.5
sources have a large spread, the median exponent is very is not high enough in the sample to make it significant
close to 1. This indicates a conical average outflow shape, (Table 2).
and agrees with previous results using slightly different esti-
mation method (Pushkarev et al. 2012b). We note the peak

MNRAS 000, 1–15 (2019)


6 Kovalev et al.

Table 4. Angular size of the VLBA core at 15 GHz, dMC c , and its offset from the true jet origin, r0MC , derived by our Monte Carlo
modeling of the jet width compared with independent MOJAVE core size measurements in the visibility plane, duv c (Lister et al. 2019),
and the core offset, r0cs , estimated from the multi-frequency core shift measurements (Pushkarev et al. 2012a). The shown duv c value is a
median over all epochs from Lister et al. (2019).

Source dMC
c duv
c r0MC r0cs
(mas) (mas) (mas) (mas)
(1) (2) (3) (4) (5)

0111+021 0.075 ± 0.025 0.079 ± 0.030 0.157 ± 0.093 0.159 ± 0.050


0238−084 0.282 ± 0.067 0.284 ± 0.042 0.740 ± 0.380 ...
0415+379 0.073 ± 0.017 0.075 ± 0.008 0.044 ± 0.021 0.275 ± 0.050
0430+052 0.096 ± 0.034 0.182 ± 0.014 0.188 ± 0.109 0.051 ± 0.050
0815−094 0.113 ± 0.020 0.062 ± 0.041 0.163 ± 0.048 ...
1133+704 0.116 ± 0.049 0.089 ± 0.048 0.072 ± 0.056 ...
1514+004 0.067 ± 0.018 0.043 ± 0.027 0.189 ± 0.088 ...
1637+826 0.100 ± 0.025 0.069 ± 0.004 0.204 ± 0.092 0.198 ± 0.050
1807+698 0.096 ± 0.031 0.067 ± 0.007 0.133 ± 0.091 0.240 ± 0.050
2200+420 0.105 ± 0.064 0.044 ± 0.003 0.067 ± 0.074 0.090 ± 0.050

Median k = 1.02 Lister et al. (2019), see Table 4. Two sources, 0111+021 and
35 0415+379, show a good agreement between dMC c and duv
c ,
while for the other seven objects dMC c is somewhat larger
30 than duv
c . This is likely due to a non-ideal determination
of the core position throughout the epochs, which is used
25 to align single-epoch maps to produce stacked images. The
radio galaxy 0430+052 is the only source having dMC < duv
20 for reasons that are unclear.
c c
N

15 A bias related to this effect might affect the results. Sta-


tistically analyzing jet shapes for the whole sample of 331
10 sources with stacked VLBA images by introducing different
ridge line path length limits we have found the following. A
5 near-parabolic streamline for quasars and BL Lacs can be
derived if the innermost jet, only up to ∼1 mas from the ap-
0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 parent core, is considered. This is not a real effect. The bias
k is found to be the most pronounced for curved jets or jets
with features emerging at different position angles over time
Figure 2. A histogram of the best fit exponents k assuming a
single power law d ∼ (r + r0 )k for all spatial scales. Shown here
(Lister et al. 2013). This is confirmed by the found artificial
are 319 sources from Table 2, which exclude those listed in this correlation of median jet width with the number of epochs
table with detected shape break. in a stacked image for such AGN. Uncertainties in the core
position also contribute to this effect due to the non-ideal
alignment of images while performing the stacking. Vari-
2.3 Checking consistency of the fits and analyzing ability of opacity conditions and apparent position of the
for possible biases core (Plavin et al. 2019a) affect this partially even though
We complemented our analysis using 43 GHz data from the alignment of the stacked single epoch images is done on
the Boston University (BU) AGN group2 for two sources, the core position. Together, it causes an additional artifi-
BL Lac and 3C 120, which are present in the MOJAVE cial widening of the jet near the core region up to distances
and BU samples. For each of these sources we (i) produced r ≈ 0.3 mas. The effect quickly vanishes at larger scales.
stacked total intensity maps, aligning single epoch-images Thus, if we exclude jet width measurements at distances
by the position of the VLBA core derived from structure . 0.4 mas, the effect becomes much weaker and disappears
modelfitting of the visibility data, (ii) determined the re- completely if we rule out the measurements within 0.5 mas
constructed jet ridge line, and (iii) fitted the transverse jet from the core. We also note that radio galaxies, being at
width as a function of distance from the core. Both sources low redshift and thus having apparently wider outflows, are
showed a jet shape transition at core separations comparable much less subject to this effect. The same is true for the
to those found from the 15 GHz data. sources with a jet shape break shown in Figure 1, as these
By setting r = 0 we can estimate the apparent core size are low-redshift objects. Only for BL Lac, as the most remote
dMC at 15 GHz from the Monte-Carlo fit of the jet width as source among them and also having a bright quasi-stationary
c
a1 r0k1 and compare it with a median value of the core size duv component near the core (Cohen et al. 2014), we put a con-
c
derived from structure modelfit in visibility plane taken from servative limit of 0.9 mas, while for the other sources we
used the non-cut intervals listed in Table 2, because drop-
ping measurements at r < 0.5 mas did not significantly
2 https://www.bu.edu/blazars/VLBAproject.html change the fit parameters. For the rest of the sources, we

MNRAS 000, 1–15 (2019)


Transition of AGN jet geometry 7
have dropped all measurement for r < 0.5 mas while ana-
lyzing the data.
Another possible problem might be related to cases
where the jet width is completely unresolved. Indeed, this
was found for some AGN targets at some epochs from the
visibility model fitting of the core (e.g., Kovalev et al. 2005;
Lister et al. 2019). This issue is taken care of by the require-
ment to drop all measurement for r < 0.5 mas. Interestingly,
the rest of the measured deconvolved jet width values are
always positive. If we assume that this is some sort of a pos-
itive bias overestimating the width, it should not depend on
r for unresolved jets and will result in k values close to zero.
This behavior was not seen in our fitting results.
We have also compared the fitted parameter r0 with the
core offset from the jet base estimated from the core shift
measured between 15 and 8 GHz (Pushkarev et al. 2012a)
assuming an inverse frequency dependence r ∝ ν −1 . These Figure 3. Best fit k-index values against deprojected distance
quantities, also listed in Table 4, agree well within the er- from 15 GHz VLBA core (Table 2) for the sources listed in Table 1
rors in four out of six sources having measured core shifts. with measured redshift and observing angle. Filled dots show fits
at 15 GHz only, while empty circles denote results from analyz-
The large discrepancy for two sources can be explained by
ing measurements at 15 and 1.4 GHz. Horizontal lines denote the
the recently recently established phenomenon of significant
scale at which the k-index was measured for every target. The
core shift variability (Plavin et al. 2019a) or the difference symbols are placed at the median distance of the analyzed jet por-
between the true jet shape derived by us and the assumed tion. Eleven AGN with detected jet shape transition are shown
conical jet shape in (Pushkarev et al. 2012a). We note that in blue: 0111+021, 0238−084, 0321+340, 0415+379, 0430+052,
this result opens a new way to estimate the distance to the 1133+704, 1514+004, 1637+826, 1807+698, 2200+420, and M87.
true jet origin which does not require an assumption regard- The data for 0321+340 and M87 are taken from Hada et al.
ing the jet geometry. (2018) and Nokhrina et al. (2019), respectively.
We warn readers about deriving jet shapes from struc-
ture model fitting of single-epoch data, as the jet may ap-
pear quasi-parabolic (k < 1) up to a certain (typically short) detection of the geometry transition in nearby AGN jets
distance from the core and then change its shape to conical are varied. Some jets, e.g., 0007+106 and 1959+650, have
(k ≈ 1). This effect takes place in the sources that show vari- too compact structure to study the shape. Some others,
ations in their inner jet position angle. Lister et al. (2013) e.g., 0241+622, 0316+413, 1216+061, show purely parabolic
established this as a common, decade-timescale phenomenon streamlines (Table 2), and their transition regions are ex-
for the most heavily monitored AGNs in the MOJAVE sam- pected at larger angular scales than those probed by our ob-
ple. Thus, single-epoch VLBI maps may not reveal the whole servations. MERLIN or low-frequency VLBA observations
jet cross-section, but rather a portion of it, especially in the are needed. For example, the nearby radio galaxy 1216+061
inner jet regions where images are dynamic range limited. (z = 0.0075, scale factor 0.15 pc mas−1 ) not shown in
Therefore, the conclusions on jet geometry based strictly on Figure 3 has a parabolic streamline with k = 0.64 ± 0.05
a modelfit approach should be treated with caution. up to 7 mas at 15 GHz, corresponding to a deprojected dis-
tance of about 1 pc only. We are studying the remaining 12
low-redshift jets that show no sign of a profile break in a
2.4 Jet shape transition: a common effect in AGN
followup program that has been approved by the VLBA.
jets, its consequences and prospects
The other jets in the sample (Table 2), namely 97 %, do
We have found evidence for geometry transition in many not show a clear significant change in jet geometry. We ex-
jets for which sufficient linear resolution was achieved. This plain this by (i) a large scale factor of the order of 8 pc mas−1
means that a change in jet shape is a common phenomenon for a typical source in the sample at a redshift of z ∼ 1 and
which has significant consequences for many high angu- (ii) a small viewing angle typically about several degrees
lar resolution astrophysical and astrometric studies. It is (Pushkarev et al. 2017). Jet power may also play a role,
difficult to conclude if the geometry transition with mea- since the MOJAVE sample is flux-density limited and the
sured properties is specific to only nearby radio galaxies and AGN with z > 0.1 typically have jet luminosities ∼ 2 orders
BL Lacs, or can be extended to the AGN class in general. of magnitude higher than the lower-redshift ones. The jets
The radio luminosities of the nearby (z < 0.07) jets are with a detected shape change have an average scaling factor
much lower than the rest of the sample and this might af- of 0.7 pc mas−1 and, on average, larger viewing angle since
fect the geometry and transition zone. We note that Figure 3 6 out of 12 are radio galaxies. Thus, if a transition region
presents a consistent picture of the power index dependence is located at a distance of a few tens of pc, it corresponds
on the downstream distance for nearby and distant jets. to a projected angular separation of < ∼ 1 mas from the ap-
In total, indications of the transition from parabolic to parent jet base at 15 GHz, which is comparable to the typ-
conical shape are found in 10 out of 29 nearby (z < 0.07) ical interferometric restoring beam size. VLBI observations
sources observed as part of the MOJAVE program or by at higher frequencies may be more effective in registering
other investigators (VLBA archival data from the latter were the jet shape transition, since they provide a better angular
processed by the MOJAVE team). The reasons for non- resolution and are less subject to opacity effects. This would

MNRAS 000, 1–15 (2019)


8 Kovalev et al.
probe scales closer to the jet apex and possible dependencies
between acceleration zone extension and the maximum bulk
Lorentz factor or jet power, as predicted by Potter & Cotter
(2015). On the other hand, the steep spectrum of the opti-
cally thin jet emission hinders the tracing of the jet for long
distances. The small viewing angles of the bright AGN jets
set another limit on any jet shape investigation in the in-
nermost parts. The streamline of an outflow can be studied
down to distances at which the jet half-opening angle is still
smaller than viewing angle. As shown by Pushkarev et al.
(2017), the intrinsic jet opening angle reaches values of a
few degrees at scales of the order of 10 pc. This suggests
that the jet shape transition phenomenon might be more ef-
fectively studied for nearby AGNs that are oriented at larger
angles to the line of sight. After considering all the points
discussed above, we have started a dedicated VLBA pro-
gram in 2019 to search for geometry transition in 61 AGN Figure 4. Same as Figure 3 but for the deprojected distance mea-
jets with z < 0.07 from observations at 15 and 1.4 GHz. sured in gravitational radius. The black hole masses are shown
It is a challenging problem to estimate the consequences Table 1. If available, we use the estimates based on the velocity
of this result on astrometry and astrophysics of AGN. VLBI dispersion method, otherwise — those from reverberation tech-
astrometry delivers the position of the true jet apex only nique. Note, the rightmost source with the detected transition
from parabolic to conical shape is 0321+340. Its mass estimate is
if the opacity driven core shift r is proportional to the
based on reverberation mapping and might be strongly underes-
frequency ν as r ∝ ν −1 (Porcas 2009). However, this
timated as argued by León Tavares et al. (2014) and Hada et al.
is expected only for conical jets and synchrotron opacity (2018).
(Lobanov 1998). A non-conical jet base results in an exten-
sion of the true jet length between the apex and the observed
opaque core. This also produces somewhat larger VLBI-Gaia Pushkarev et al. 2017; Liodakis et al. 2017). For the radio
offsets for AGN positions (Kovalev et al. 2017; Plavin et al. galaxy 1514+004 we assumed a viewing angle of 15◦ .
2019b) than predicted by Kovalev et al. (2008). In Figure 3, we plot the corresponding single power-law
k-index values derived from the 15 and 1.4 GHz VLBA data
(Pushkarev et al. 2017) versus deprojected distance from the
2.5 Deprojected position of the jet break 15 GHz VLBA core for 62 sources. There are eleven sources
among them with known deprojected linear jet distance that
We chose the MOJAVE-1 sample (Lister et al. 2009) to have a jet shape transition (Figure 1, Table 1). They are are
perform a direct comparison with the 12 jets showing the shown by a pair of points each from the double power-law
breaks. Our reasoning is as follows. Most of MOJAVE-1 tar- fits. The BL Lac object 0815−094 is not shown in Fig. 3, as it
gets were observed by VLBA not only at multiple 15 GHz does not have a measured spectroscopic redshift. Our results
epochs but also in a single epoch at 1.4 GHz, which in- on jet shape transition in nine targets (Table 2, Table 3,
creases the jet distance being probed by our analysis. In Figure 3) are supplemented by multi-frequency data for M87
addition, VLBI measurements of the apparent kinematics from Nakamura et al. (2018), with k1 = 0.57, k2 = 0.90,
βapp (Lister et al. 2019) and variability Doppler factor esti- and break point position obtained by Nokhrina et al. (2019).
mates δ (Hovatta et al. 2009; Liodakis et al. 2017) are avail- For M87 we adopt θ = 14◦ (Wang & Zhou 2009), consis-
able for a large fraction of the sample. We need this infor- tent with more recent results by Mertens et al. (2016). For
mation to derive deprojected distance values. These require- NLSy1 0321+430 we use 1.4–2.3 GHz measurements from
ments result in a sample of 65 sources (Table 1) described VLBA observations (Hada et al. 2018), with k1 = 0.6 and
in Pushkarev et al. (2017). k2 = 1.41, for which the jet shape break point position is
We derived viewing angle estimates through the relation estimated.
2βapp Horizontal lines represent the scales at which k-indices
θ = arctan 2 2 −1 were derived, starting from several tens of mas distance from
βapp + δvar
the 15 GHz VLBA core (see subsection 2.3) and up to jet
to convert the jet distance from angular projected to lin- distances limited by the sensitivity of our observations. The
ear deprojected. Note that this assumes the same beaming nearby jets, for which we are probing closer to the central
parameters for the flux density variability and jet kinemat- engine, have low k values and show a transition from quasi-
ics. For βapp we used the fastest non-accelerating appar- parabolic values at small scales to quasi-conical at larger
ent jet speeds from the MOJAVE kinematic analysis. For scales (Figure 3). It is possible that at larger scales of a few
0321+340 we use θ = 6.3◦ , based on the observed super- hundred kiloparsec, where jets become diffuse and disrup-
luminal motion (Lister et al. 2016) assuming θ = γ −1 = tive, their geometry further changes from conical to hyper-
(1 + βapp )−0.5 . The other possible viewing angle value for bolic, characterized by more rapid expansion (Owen et al.
this target θ = 4◦ based on variability time scale is discussed 2000).
by Hada et al. (2018). For the BL Lac objects 0111+021 and In order to plot the observed k-index values as
1133+704 we assumed a viewing angle of 5◦ , typical for this a function of the deprojected distance along jets in
class of AGN (Hovatta et al. 2009; Savolainen et al. 2010; gravitational radius rg = 2GM/c2 units, we use the

MNRAS 000, 1–15 (2019)


Transition of AGN jet geometry 9
black hole masses estimated assuming virialized broad netically dominated flow exists (see e.g., Beskin 2009)
lines region (BLR) motion and correlation between r⊥
BLR size and UV/optical luminosity (Torrealba et al. γ(r⊥ ) = , (2)
RL
2012; McLure & Jarvis 2002; Vestergaard & Peterson 2006;
Landt et al. 2017; Palma et al. 2011; Shaw et al. 2012; where RL = c/Ω0 is the light cylinder radius, and r⊥ is the
Liu et al. 2006). We also have masses inferred by stellar distance from the jet axis. For the black hole spin parameter
or gas kinematics methods (e.g., Woo & Urry 2002) for the a∗ = 0.5, RL ≈ 7.4 rg ≈ 2.1 × 1015 (MBH /109 M⊙ ) cm ≈
closest sources. The mass values and references can be found 6.8×10−4 (MBH /109 M⊙ ) pc. For observed pc scale jets, the
in Table 1. Results are presented in Figure 4. It turns out, jet width d reaches 1 pc, so that d/2RL > σM . This means
that the sources with BH masses obtained by stellar velocity that the flow cannot be still magnetically dominated. As was
dispersion method or stellar/gas kinematics measurements, shown by Nokhrina et al. (2015) who have analysed about
are the subset of the sources with the detected jet shape 100 AGN jets, σM ∼ 10 is a reasonable value constrained by
break, being the closest ones. the observations.
Since estimating the black hole mass is a complicated The observed median value of 1.02 for k-index clearly
and strongly model-dependent method, some of the values points to ballistic plasma motion, suggesting the jet is domi-
might be significantly in error. By dropping the highest and nated by the plasma bulk motion kinetic energy rather than
lowest values as possible outliers of the derived jet break po- by the Poynting flux, as expected close to the launching re-
sition rbreak measured in rg we are able to bound its values in gion.
the narrower range rbreak ∈ (105 , 106 )rg . This is an impor- For this reason we aim to explain the break in the d(r)
tant result, especially when taken together with our finding dependence as the consequence of a transition from the mag-
that the jet shape transition may be a common phenomenon netically dominated to the particle dominated regime. Below
in nearby or even most of the AGN. we present the main results of our semi-analytical consider-
We note the following. The black hole mass of 0321+340 ation. Our goal is in evaluating the dependence of the jet
is suspected to be underestimated (León Tavares et al. 2014; width d on an ambient pressure profile Pext (r). The results
Hada et al. 2018). If we use for this source the mass for the cold jet are presented in Beskin et al. (2017), while
M = 108.6 M⊙ , obtained using the relation between black here we consider the semi-analytical results for a warm out-
hole mass and bulge luminosity (León Tavares et al. 2014), flow.
0321+340 yields rbreak = 2.7 × 106 rg , falling within the dis-
cussed above range of rbreak /rg distances. This may provide
an additional argument favoring a higher black hole mass 3.2 Semi-analytical model
for this source. Basic equations describing the internal structure of rela-
tivistic and nonrelativistic jets within the Grad-Shafranov
approach are now well-established (Heyvaerts & Norman
1989; Pelletier & Pudritz 1992; Lery et al. 1998;
3 MODELING RELATIVISTIC JET WITH A Beskin & Malyshkin 2000; Beskin 2009; Lyubarsky 2009).
SHAPE BREAK This approach allows us to formulate the problem of
3.1 Qualitative consideration finding a stationary axisymmetric magnetohydrodynamic
outflow structure, a jet solution, as a set of two differential
Both analytical (see below) and phenomenologi- equations on a magnetic flux function Ψ and an Alfvénic
cal (Potter & Cotter 2013, 2015) considerations as Mach number M. These equations are Bernoulli equation
well as numerical simulations (Komissarov et al. 2009; and Grad–Shafranov equation of a force balance perpen-
Tchekhovskoy et al. 2009; Porth et al. 2011) show that for dicular to magnetic surfaces. The approach allows us to
moderate initial magnetization of a jet σM ∼ 10–102 , where determine the internal structure of axisymmetric stationary
jets knowing in general case five “integrals of motion”, i.e.,
Ω20 Ψ0
σM = (1) energy E(Ψ) and angular momentum L(Ψ) flux, electric
8π 2 µηc2
potential which connects with angular velocity ΩF (Ψ),
is the Michel magnetization parameter, the flow transits entropy s(Ψ), and the particle-to-magnetic flux ratio η(Ψ).
from a magnetically dominated regime at small distances All these values are to be constant along magnetic surfaces
r from the origin to a particle dominated regime at larger Ψ = const. Once the Grad–Shafranov and Bernoulli
distances. Here Ψ0 and Ω0 are the total magnetic flux and equations solved for the given integrals, all the other flow
characteristic angular velocity of the “central engine” respec- properties, such as particle number density, four-velocity,
tively. Accordingly, µ = mp c2 + mp w is the relativistic en- electric current, and Lorentz factor, can be determined from
thalpy, where w is the nonrelativistic enthalpy, and mp is a algebraic equations (see, e.g., Beskin (2009)). In particular,
particle mass. Here we assume a leptonic jet, so mp is the it was shown that a jet with total zero electric current
electron mass. Below for simplicity we consider not so large can exist only in the presence of an external medium with
temperatures, so that w ≪ c2 . Finally, η is the particle-to- finite pressure Pext . Thus, it is the ambient pressure Pext
magnetic flux ratio. that is expected to determine the transverse dimension of
Indeed, the physical meaning of the Michel magneti- astrophysical jets. In general, it is a complicated problem
zation parameter is the maximum Lorentz factor γ of the to solve the set of Bernoulli and Grad–Shafranov equations.
hydrodynamical flow when all the electromagnetic energy Additional complication is connected with the change of
flux is transferred to particles. On the other hand, for quasi- a system type from elliptical to hyperbolic. So, to tackle
cylindrical jets the following asymptotic solution for mag- the problem different simplifications are introduced. In this

MNRAS 000, 1–15 (2019)


10 Kovalev et al.
paper we simplify the problem, assuming the flow is highly determining the functions M2 (r⊥ ) and Ψ(r⊥ ) one can find
collimated and can be described within the cylindrical the jet radius d/2 as well as the profile of the current I(r⊥ ),
geometry, in which case it can be solved numerically the particle energy, and the toroidal component of the four-
(Beskin & Malyshkin 2000). velocity from a solution of a problem under consideration.
On the other hand, careful matching of a solution inside In particular, as
the jet with the external medium has not been produced up 2
I L − ΩF r⊥ E/c2
to now. The difficulty arises with having a very low energy = 2 2
, (8)
2π 1 − ΩF r⊥ /c − M2
2
density of the external medium in comparison with the en-
ergy density inside the relativistic jet. For this reason, in the condition of the closing of the electric current within
most cases an infinitely thin current sheet was introduced. the jet I(Ψ0 ) = 0 can be rewritten as L(Ψ0 ) = 0 and
Moreover, an ambient pressure was often modelled by ho- ΩF (Ψ0 ) = 0, where Ψ0 is the given total magnetic flux in
2
mogeneous magnetic field Bext /8π = Pext . a jet.
Below we present an approach which is free of the dif- For this reason in what follows we use the following
ficulties mentioned above. We consider a relativistic jet sur- expressions for these integrals
rounded by an external medium with finite gas pressure Pext . r
Ω0 Ψ Ψ
Neither an external magnetic field nor infinitely thin current L(Ψ) = 1− , (9)
4π 2 Ψ0
sheet are assumed. The main new aspect involves the bound- r
ary conditions at the jet boundary r⊥ = d/2. In what follows Ψ
ΩF (Ψ) = Ω0 1 − . (10)
we propose to construct a solution so that both magnetic Ψ0
field and flow velocity vanish at the jet boundary r⊥ = d/2. In the vicinity of a rotation axis these integrals correspond
On the other hand, we suppose that the flow remains super- to the well-known analytical force-free solution for a homo-
sonic, so that Alfvénic Mach number M(d/2) > 1. geneous poloidal magnetic field. On the other hand, they
As was shown by Beskin & Malyshkin (2000), for both vanish at the jet boundary, which guarantees the ful-
a cylindrical flow it is convenient to reduce the Grad- filment of the condition I(Ψ0 ) = 0. For a linearly decaying
Shafranov and Bernoulli equations to two first-order or- magnetic field B(δr⊥ ) ∝ (d/2 − δr⊥ ) this root dependence
dinary differential equations for magnetic flux Ψ(r⊥ ) and on Ψ ensures the linear decay of the integrals L(Ψ(r⊥ )) and
poloidal Alfvénic Mach number M(r⊥ ): ΩF (Ψ(r⊥ )) with respect to a transverse distance.
4πµη 2 The analytical analysis of Equation 4 and Equation 7
M2 = . (3) provides that both magnetic field B and flow four-velocity
n
u vanish at the jet boundary r⊥ = d/2. As one can see from
Here n is an electron number density in the co-moving refer- Equation 8, our choice of L(Ψ) and ΩF (Ψ) guarantees that
ence frame. The first of this set of equations is the relativistic toroidal components Bϕ vanish at the jet boundary. One can
Bernoulli equation u2p = γ 2 − u2ϕ − 1 which has the form find that the same holds for the toroidal velocity uϕ . On the
M4


2
K other hand, using relation nup = ηBp , one can conclude
2
= 2 2 − µ2 η 2 . (4) that the conditions up = 0 and Bp = 0 can be realised for
64π 4 r⊥ dr⊥ r⊥ A
finite n and η. For simplicity we consider here the case
Here
η(Ψ) = η = const. (11)
A = 1 − Ω2F r⊥
2
/c2 − M2 (5)
Below we assume that the flow remains supersonic up to the
is the Alfvénic factor, very boundary: M(d/2) > 1. This assumption allows us to
2 simplify our consideration. Indeed, in this case Equation 7
K = r⊥ (e′ )2 (A − M2 ) + M4 r⊥
2
E 2 − M4 L2 c2 , (6)
has no additional singularity at the Alfvénic surface A = 0
and e′ = E−ΩF L. The second equation on the Mach number in the vicinity of the jet boundary. We check this assumption
is as follows: numerically, and we find that it holds for our solution.
 ′ 2
(e ) Ω2F r⊥
2
c2s dM2

M6 L2 c2 Further, as is shown in Beskin et al. (2017), energy in-
2 2
− 1 + 2
− A 2
= tegral E(Ψ) can be written down as
µ η c c dr⊥ Ar 3 µ2 η 2
Ω2 r⊥ M2 (e′ )2 e′ dΨ de′ E(Ψ) = ΩF L + µjet ηγ(Ψ). (12)
 
+ F 2 2− + M2 2 2
c Aµ2 η 2 µ η dr⊥ dΨ Here γ(0) = γin is the injection Lorentz-factor along the jet
M2 r ⊥
2
dΨ dΩ2F Ω2F r⊥2
c2s axis, and γ(Ψ0 ) = 1. The last relation just implies that the
 
2 dΨ 1 dη
+ − M 1 − + 2A flow velocity vanishes at the jet boundary. As to relativistic
2c2 dr⊥ dΨ c2 c2 dr⊥ η dΨ
enthalpy µjet = mp c2 + mp wjet , we consider it here as a
Ω2 r 2 M2 dΨ ds
     
A ∂P
− + 1 − F2 T . (7) constant with wjet = c2jet /(Γ − 1), where the constant cjet
n ∂s n c µ dr⊥ dΨ
corresponds to sonic velocity of a flow at the very boundary.
Here T is the temperature, cs is the sound velocity defined For simplicity in this paper we use polytropic equation of
as c2s = (dP/dn)|s /mp , mp is a particle mass, and P is a gas states with polytropic index Γ. The relativistic enthalpy µ
pressure. As a result, Bernoulli Equation 4 and Equation 7 in Equation 4 and Equation 7 is determined via local value
form the system of two ordinary differential equations for w = c2s /(Γ − 1) (see below). Finally, in what follows we
the Mach number M2 (r⊥ ) and the magnetic flux Ψ(r⊥ ) de- assume
scribing cylindrical relativistic jets.
s(Ψ) = s = const. (13)
As it was already stressed, the solution results from our
choice of five integrals of motion. It is important that by Here the entropy s enters the polytropic equation of state

MNRAS 000, 1–15 (2019)


Transition of AGN jet geometry 11
as P (n, s) = k(s)nΓ with the function k(s) which particu-
lar form is not used explicitly in the following equations. It 10−4 σM = 10
P∝ σM = 100
is this nonzero value that allows us to match magnetically d −3
.7 σM = 1000
dominated flow to the external media with finite pressure. 10−6 P∝
In addition to five integrals of motion, the system of d −3

Pext (dyn/cm 2)
Equation 4 and Equation 7 requires two boundary condi- .8
tions. The first one is the obvious condition at the symmetry 10−8

P∝
axis

d
−3
10−10

.9
Ψ(0) = 0. (14)
P∝d−
2.3
As to the second one, it can be found from the pressure bal-
ance Pext = Γ−1 c2jet mp njet , where njet is the number density
10−12 P∝
d −2.4
of a flow at the jet boundary. Thus, we can determine the jet P∝
d −2.
radius d/2 as a function of Pext self-consistently as the condi- 10−14 9
tions v(d/2) = 0 and B(d/2) = 0 are automatically fulfilled
103 104 105
d (rg )
at Ψ = Ψ0 . Using now the definition (Equation 3) one can
write down the second boundary condition M2jet = M2 (d)
as
Figure 5. Ambient pressure Pext as a function of a jet width
4πΓ−1 m2p c2 c2jet η 2
M2jet = . (15) d shown in units of the gravitational radius rg for for different
Pext magnetization parameters σM . We take the fiducial a∗ = 0.1 and
The local non-relativistic enthalpy w can be written as Ψ0 = 1033 G · cm2 . The start of each curve corresponds to the
Γ−1 start of a super Alfvénic flow down from the axis. The powers
c2jet

n showed in the figure. The position of a break in a power law
w= , (16)
(Γ − 1) njet depends on the flow initial magnetization in correspondence with
our interpretation.
where the local particle number density n is obtained from
the equation
"
2  Γ−1 # which one may neglect by the derivatives over r in com-
2 2 2 1 cjet n
M n = 4πη mp c 1 + . (17) parison with the derivatives over r⊥ in the two-dimensional
Γ − 1 c2 njet Grad–Shafranov and Bernoulli equations. This can be done
Thanks to these definitions the system of Equation 4, for the highly collimated, at least as a parabola, outflows
Equation 7 becomes now fully determined. (Nokhrina et al. 2015) and the flows with small opening an-
We solve the system of Equation 4 and Equation 7 for gles (Tchekhovskoy et al. 2009).
the boundary conditions Equation 14 and We find that for the chosen sound velocity at the bound-
ary c20 = 0.001c2 the thermal effects may be neglected in the
P |r⊥ =d/2−0 = Pext . (18) outflow volume, playing an important role only at the out-
We should note that due to the root dependence of the in- flow boundary. It turns out that the resultant dependence of
tegrals L(Ψ) and ΩF (Ψ) the thickness of the final current pressure at the jet boundary as a function of jet radius for
closure domain tends to zero and is not resolved numeri- the cold (Beskin et al. 2017) and the warm outflow starts to
cally. However, as it has been shown in (Beskin et al. 2017), differ considerably for large M20 (this value is of an order
the total pressure in this region is strictly conserved: of 10, but depends on the initial magnetization), affecting
the flow boundary shape downstream of the equipartition
B2
 
d transition.
P+ = 0. (19)
dr⊥ 8π The proposed jet model with an electric current en-
This means that the solution we obtain up to the bound- closed inside the jet has a natural sheath structure, observed,
ary does contain the residual current and, thus, the toroidal for example, in the M87 jet (Mertens et al. 2016). Due to
magnetic field Bϕ . choice of integrals, the outer parts of a jet have slower ve-
To obtain the solution, we employ the following itera- locities, tending to non-relativistic with γ(d/2) = 1. Such
tive procedure. For each fixed fast magnetosonic Mach num- a sheath may be produced by different mechanisms: it may
ber at the axis M20 we start to set Pext at the jet bound- be a slower disk wind or an outer jet disturbed and slowed
ary. It defines the particle number density at the boundary down by the pinch instability (Chatterjee et al. 2019). In our
njet , and together with M20 — the particle number density model it appears naturally as a consequence of a jet tran-
at the axis n0 . Having set the latter, we solve Equation 4 siting into the ambient medium with the hydrodynamical
and Equation 7 from the axis outwards and calculate the jet discontinuity only (Beskin et al. 2017).
pressure at the boundary provided by the solution P (solution) .
By iterations we find self-consistently such Pext that is equal
3.3 Transition from magnetically dominated to
to one, provided by the solution: P (solution) = Pext .
particle dominated flow
This procedure fully determines the solution of the
problem (Equation 4), (Equation 7), (Equation 14), and It is necessary to stress that this system of equations can
(Equation 18). For each magnitude of the external pressure describe both magnetically and particle dominated flow, the
the obtained solution is a crosscut at r = const. Piling physical answer (including the jet boundary radius d/2) de-
of these different crosscuts is a solution for an outflow in pending on one external parameter only, namely, on the am-

MNRAS 000, 1–15 (2019)


12 Kovalev et al.
bient pressure Pext . In Figure 5 we show the dependence of 8
4.0
0.
the ambient pressure Pext on a jet transverse dimension d/2 ∝
r
d 3.5
obtained by numerical solving the system of Equation 4 and
Equation 7 for central mass M = 109 M⊙ (gravitational ra- 1.0
3.0

dius rg ∼ 3 · 1014 cm) and black hole spin a∗ = 0.1. We also

d (pc)
2.5
set the total magnetic flux in an outflow Ψ0 = 1032 G cm2 ,

σ
which gives the value B(rg ) = 350 G, and BL = 3.5 G. We 2.0

observe (see Figure 5) that the pressure has different power .5

law dependence on the jet radius for small and large d/2. For r0 1.5
0.1 d∝
each magnetization σM , this behavior holds, with the change 1.0

between two profiles occurring at different jet widths. For


0.5
fiducial σM = 10 the pressure changes its dependence on d 100 101 102
from P ∝ d−3.7 to P ∝ d−2.3 . r (pc)
In order to model a jet shape break position along the
jet we need to introduce the pressure dependence on r, which
Figure 6. An example of a jet boundary shape (blue solid line)
we choose in the power law form for σM = 50. The jet magnetization at a given distance from
 −b its base is plotted by a red solid line. The transition from one
r
Pext = P0 . (20) power law to the other (green dashed lines) for the jet boundary
r0 roughly coincides with the point where the outflow transits from
Such a pressure profile is consistent with Bondi the magnetically dominated to particle dominated (equipartition)
flow (Quataert & Narayan 2000; Shcherbakov 2008; regime, designated by a star.
Narayan & Fabian 2011) having b ∈ [1; 2.5] for different
models. This power law with b ≈ 2.0 allows to reproduce
well both the parabolic jet form upstream the break holds always, since the Poynting flux behaves at the jet axis
and conical downstream. We obtain for small distances r as
(magnetically dominated regime) 2
|S| ∝ I = πjr⊥ 2
+ o(r⊥ ) (27)
0.54
d∝r . (21)
if the current density j has no singular behavior at r⊥ = 0.
Accordingly, for large distances (saturation regime) Thus, σ → 0 at the axis. The same holds for the boundary
in a case of a full current closure. Due to specific choice of
d ∝ r 0.87 . (22) integrals E(Ψ), L(Ψ), and ΩF (Ψ) (Equations 9, 10, 12), the
As we see, qualitatively, the power indices are in good agree- Poynting flux together with the magnetization reach their
ment the with observational data. maximum values at Ψ = Ψ0 /2. It is at this magnetic field
line the flow attains its highest Lorentz factor across the
jet for the given distance from the central source. Thus, we
3.4 Magnetization choose the maximal magnetization reaching approximately
the unity as a criteria of a flow attaining the ideal MHD
In this subsection we check whether the break in a jet shape
equipartition regime.
corresponds to the transition from magnetically-dominated
In Figure 6 we present the maximal magnetization and
into equipartition regime. The jet magnetization is defined
the break in a jet form. We see that the break in a jet shape
as the ratio of Poynting flux
occurs roughly at the distance from the BH r, where the flow
c magnetization becomes equal to unity. For the higher initial
S= E×B (23)
4π magnetization it takes the larger transverse jet dimension in
to particle kinetic energy flux RL to accelerate the flow up to equipartition, according to
Equation 2.
K = γmc2 nup , (24)
The first change in jet shape has been reported by
where n is particle number density in the jet proper frame. Asada & Nakamura (2012), and this break is associated with
Using the standard expressions for ideal MHD velocities and the standing bright feature HST-1 in M87. The change in jet
electric and magnetic fields, one can obtain the following form may be explained by the change in ambient pressure
expression for the magnetization: at the break radius (Asada & Nakamura 2012). Another ex-
planation has been proposed by Levinson (2017), who con-
|S| ΩF I
σ= = . (25) sidered a disk wind. The shock between the jet and a disk
|K| 2πcγµη wind followed by ambient pressure decrease accounts for
Using the definitions of bulk Lorentz factor γ and total cur- both jet boundary shape change and the presence of HST-
rent I, we rewrite it as 1. This point is supported by the presence of a standing
2 bright feature, approximately coincident with a jet shape
L − ΩF r⊥ E/c2 break location, in 6 of 12 sources (see details in Table 3).
σ = ΩF . (26)
E − ΩF L − M 2 E Another model for a jet boundary shape is an analytical
In order to check σ along the jet, we calculate the max- asymptotic solution by Lyubarsky (2009), which predicts
imal magnetization across the jet for each given distance r both the change in a jet form without the change in an am-
along the jet. The magnetization is always much less than bient pressure profile and the oscillations of a jet form under
the unity at the jet axis and at the jet boundary. The first the conditions of a jet being not in equilibrium state with

MNRAS 000, 1–15 (2019)


Transition of AGN jet geometry 13

18 motion the following kinematics pattern: efficient Lorentz


factor growth before the break point, and cessation of it in
16 the conical region.
This prediction is consistent with observations by the
Lorentz factor

14 MOJAVE program that acceleration is a common property


of jet features (e.g., Homan et al. 2015; Lister et al. 2019),
12 reflecting a tendency for increasing Lorentz factors near the
10 base of the jet with decreasing or constant speeds being
more common at distances & 10 − 20 parsecs projected
8 (Homan et al. 2015). While decreasing speeds are not a pre-
diction of this model for a change in jet shape, they could
6 naturally occur if the reduction in positive acceleration is
also accompanied by entrainment of external material into
0.0 0.5 1.0 1.5 2.0 2.5 3.0 the jet.
Deprojected distance, (106 rg)
Figure 7. A model example of a Lorentz factor (blue solid line)
growth against a deprojected distance along a jet. Model parame- 4 SUMMARY
ters are set so as to position the break at the deprojected distance
Pushkarev et al. (2017) studied the jet shapes by measur-
from a black hole r = 5 × 105 rg — green dashed vertical line. We
predict that there must be an efficient plasma acceleration before
ing the power low index k assuming a d ∝ r k dependence
the break, that becomes very slow after it. of the observed deconvolved jet width on the apparent dis-
tance from its core r. Most of the jets exhibited k values
in the range from 0.5 to 1.5. As it was clearly shown by
the ambient pressure. It may account for the observed os- Pushkarev et al. (2017), high-quality, high-dynamic-range
cillations reported by Pushkarev et al. (2017). Here we pro- stacked images are needed for an analysis of this kind in
pose that the jet form change, observed in M87, as well as order to trace the full jet channel. In view of a few recent
in other sources, is due to an internal flow transition from exciting reports on jet shape transitions from parabolic to
magnetically dominated to particle dominated regime with conical (e.g., Asada & Nakamura 2012; Giroletti et al. 2008;
the smooth external pressure profile. We may speculate in Boccardi et al. 2016; Tseng et al. 2016; Hada et al. 2018;
this framework, that the observed bright feature, associated Akiyama et al. 2018; Nakahara et al. 2018), we have per-
with a standing shock, may be positioned preferably at the formed a systematic search of such transition using MO-
beginning of the particle-dominated part of a jet. JAVE 15 GHz stacked images supplementing some of them
with available single epoch 1.4 GHz VLBA images to trace
larger scales.
3.5 Additional observational evidence of the
Using an automated analysis approach, we have found
break point and predicted evolution of plasma
10 jets with such transition out of 331 analyzed: 0111+021,
acceleration
0238−084, 0415+379, 0430+052, 0815−094, 1133+704,
We find in six out of twelve sources (Table 3) that the jet 1514+004, 1637+826, 1807+698, 2200+420. Their redshift
shape transition region is associated with a low pattern lies in the range z < 0.07 except for 0815−094, whose red-
speed feature from the MOJAVE analysis by Lister et al. shift is unknown. For the full analyzed sample the typical
(2019). They define such features as having angular speed redshift is about 1. This low-z coincidence is unlikely to have
smaller than 20 µas yr−1 and being at least 10 times slower occurred by chance. Taken together with an analysis of pos-
than the fastest component in the same jet. This is a factor sible biases, we conclude that a genuine effect is found. We
of 2.5 larger compared to overall statistics of jet kinematics could predict that the BL Lac object 0815−094 should also
analysis performed at 15 and 43 GHz that reveals a fraction be a nearby AGN.
of quasi-stationary jet features to be about 20% (Lister et al. This finding leads to the following important conclu-
2019; Jorstad et al. 2017). We note that the MOJAVE sion. The transition of jets from the parabolic to the conical
team uses conservative criteria in cross-identifying compo- shape might be a general property of AGN. At the same
nents between epochs and selecting robust ones (Lister et al. time, we note that AGN observed at higher redshifts typi-
2019). This means that the 50 % fraction of sources which cally have higher luminosities and kinetic power which can
show a standing feature in the break point region should be affect the collimation properties. This conclusion has impor-
considered as a lower limit. tant implications for jet models, astrophysics and astrome-
We plot in Figure 7 the maximum Lorentz factor of a try of AGN. Measuring this phenomenon requires a search
bulk plasma motion along a jet, which we obtain within within nearby AGN (the subject of our current followup
our semi-analytical model. The predicted pattern of a bulk study), or increasing the resolution by using Space VLBI
Lorentz factor acceleration in magnetically dominated do- (e.g., Giovannini et al. 2018) or high dynamic range high
main is γ ∝ r⊥ , which provides for a parabolic jet γ ∝ r 0.5 . frequency VLBI imaging.
After the flow reaches equipartition, the acceleration contin- The deprojected distance rbreak from the nucleus to the
ues slower than any power-law (logarithmically slow) (e.g., break zone is found to be typically 10 pc. Even more inter-
Beskin & Nokhrina 2006). There is also a transitional zone esting due to its relation to jet formation and acceleration
between the two regimes. Thus, we would expect for the models is this value measured in Schwarzschild radius units.
sources with the detected jet shape break and superluminal We find typical range of rbreak ∈ (105 , 106 )rg .

MNRAS 000, 1–15 (2019)


14 Kovalev et al.
We suggest the following model to explain the observed Beskin V. S., 2009, MHD Flows in Compact Astrophysical Ob-
jet shape break. The accurate matching of a jet outflow with jects: Accretion, Winds and Jets
an ambient medium (Beskin et al. 2017) predicts a change Beskin V. S., Malyshkin L. M., 2000, Astronomy Letters, 26, 208
in jet shape from parabolic to conical if the ambient medium Beskin V. S., Nokhrina E. E., 2006, MNRAS, 367, 375
pressure is assumed to be governed by Bondi accretion. Beskin V., Chernoglazov A., Kiselev A., Nokhrina E., 2017,
MNRAS, 472, 3971
Within this model, a smaller external pressure is needed
Blandford R. D., Königl A., 1979, ApJ, 232, 34
to support a jet. The transition of predicted jet shape from Boccardi B., Krichbaum T. P., Bach U., Mertens F., Ros E., Alef
parabolic to conical occurs in the domain where the bulk W., Zensus J. A., 2016, A&A, 585, A33
plasma kinetic energy flux becomes equal to the Poynting Boccardi B., Migliori G., Grandi P., Torresi E., Mertens F., Kara-
energy flux, i.e., where the bulk flow acceleration reaches sat- manavis V., Angioni R., Vignali C., 2019, arXiv e-prints,
uration (Beskin & Nokhrina 2006). From studying the break p. arXiv:1905.06634
properties it allows us to estimate black hole spin and/or Chang C. S., Ros E., Kovalev Y. Y., Lister M. L., 2010, A&A,
mass, jet total magnetic flux, and ambient medium proper- 515, A38
ties as discussed by Nokhrina et al. (in prep.). Chatterjee K., Liska M., Tchekhovskoy A., Markoff S. B., 2019,
The following two model predictions are supported ob- arXiv:1904.03243,
Cohen M. H., et al., 2014, ApJ, 787, 151
servationally. The break point, where jets start to be plasma
Giovannini G., et al., 2018, Nature Astronomy, 2, 472
dominated energetically, might be a preferable domain for
Giroletti M., Giovannini G., Cotton W. D., Taylor G. B., Pérez-
shocks. We detect standing jet features in this region from Torres M. A., Chiaberge M., Edwards P. G., 2008, A&A,
the MOJAVE analysis (Lister et al. 2019) in at least a half 488, 905
of the AGN targets. The plasma acceleration is predicted to Hada K., et al., 2018, ApJ, 860, 141
decrease significantly at the transition region which is consis- Handley W. J., Hobson M. P., Lasenby A. N., 2015, MNRAS,
tent with MOJAVE acceleration results (Homan et al. 2015; 453, 4384
Lister et al. 2019). Hervet O., Meliani Z., A. Z., C. B., V. C., C. S., H. S., 2017,
Our finding also implies the following. The well-known A&A, 606, A103
effect of the apparent shift of the core position with fre- Heyvaerts J., Norman C., 1989, ApJ, 347, 1055
quency due to synchrotron self-absorption does not follow Homan D. C., Lister M. L., Kovalev Y. Y., Pushkarev A. B.,
Savolainen T., Kellermann K. I., Richards J. L., Ros E., 2015,
the rcore ∝ ν −1 law all the way up to the true jet base,
ApJ, 798, 134
since a −1 power low index is expected only for a conical Hovatta T., Valtaoja E., Tornikoski M., Lähteenmäki A., 2009,
jet (Blandford & Königl 1979; Lobanov 1998). Geometrical A&A, 494, 527
and physical estimates made on the basis of core shift mea- Hovatta T., et al., 2014, AJ, 147, 143
surements will need to take this into account while VLBI Hovatta T., O’Sullivan S., Martı́-Vidal I., Savolainen T.,
and VLBI-Gaia astrometry applications will need to correct Tchekhovskoy A., 2019, A&A, 623, A111
for it (Porcas 2009) in cases where very high accuracy is Jorstad S. G., et al., 2017, ApJ, 846, 98
required. Komatsu E., et al., 2009, ApJS, 180, 330
Komissarov S., 2012, Central Engines: Acceleration,
Collimation and Confinement of Jets. pp 81–114,
doi:10.1002/9783527641741.ch4
ACKNOWLEDGMENTS Komissarov S. S., Barkov M. V., Vlahakis N., Königl A., 2007,
MNRAS, 380, 51
We thank Dan Homan and Eduardo Ros for valuable com-
Komissarov S. S., Vlahakis N., Königl A., Barkov M. V., 2009,
ments and suggestions. This study has been supported in
MNRAS, 394, 1182
part by the Russian Science Foundation: grant 16-12-10481. Kovalev Y. Y., et al., 2005, AJ, 130, 2473
TS was supported by the Academy of Finland projects Kovalev Y. Y., Lobanov A. P., Pushkarev A. B., Zensus J. A.,
274477, 284495, and 31249. This research has made use of 2008, A&A, 483, 759
data from the MOJAVE database that is maintained by the Kovalev Y. Y., Petrov L., Plavin A. V., 2017, A&A, 598, L1
MOJAVE team (Lister et al. 2018). The MOJAVE program Landt H., et al., 2017, MNRAS, 464, 2565
is supported under NASA-Fermi grant NNX15AU76G. This León Tavares J., et al., 2014, ApJ, 795, 58
study makes use of 43 GHz VLBA data from the VLBA-BU Lery T., Heyvaerts J., Appl S., Norman C. A., 1998, A&A,
Blazar Monitoring Program, funded by NASA through the 337, 603
Fermi Guest Investigator Program. The VLBA is an instru- Levinson A.and Globus N., 2017, MNRAS, 465, 1608
Liodakis I., et al., 2017, MNRAS, 466, 4625
ment of the National Radio Astronomy Observatory. The
Lister M. L., et al., 2009, AJ, 137, 3718
National Radio Astronomy Observatory is a facility of the Lister M. L., et al., 2013, AJ, 146, 120
National Science Foundation operated by Associated Uni- Lister M. L., et al., 2016, AJ, 152, 12
versities, Inc. This research has made use of NASA’s Astro- Lister M. L., Aller M. F., Aller H. D., Hodge M. A., Homan D. C.,
physics Data System. Kovalev Y. Y., Pushkarev A. B., Savolainen T., 2018, ApJS,
234, 12
Lister M. L., et al., 2019, ApJ, 874, 43
Liu Y., Jiang D. R., Gu M. F., 2006, ApJ, 637, 669
REFERENCES
Lobanov A. P., 1998, A&A, 330, 79
Akiyama K., Asada K., Fish V., Nakamura M., Hada K., Nagai Lyubarsky Y., 2009, ApJ, 698, 1570
H., Lonsdale C., 2018, Galaxies, 6, 15 Marscher A. P., 2008, in Rector T. A., De Young D. S., eds,
Algaba J. C., Nakamura M., Asada K., Lee S. S., 2017, ApJ, Astronomical Society of the Pacific Conference Series Vol. 386,
834, 65 Extragalactic Jets: Theory and Observation from Radio to
Asada K., Nakamura M., 2012, ApJ, 745, L28 Gamma Ray. pp 437–443

MNRAS 000, 1–15 (2019)


Transition of AGN jet geometry 15
McKinney J. C., 2006, MNRAS, 368, 1561
McKinney J. C., Tchekhovskoy A., Blandford R. D., 2012,
MNRAS, 423, 3083
McLure R. J., Jarvis M. J., 2002, MNRAS, 337, 109
Mertens F., Lobanov A. P., Walker R. C., Hardee P. E., 2016,
A&A, 595, A54
Müller C., et al., 2014, A&A, 569, A115
Nakahara S., Doi A., Murata Y., Hada K., Nakamura M., Asada
K., 2018, ApJ, 854, 148
Nakahara S., Doi A., Murata Y., Nakamura M., Hada K., Asada
K., 2019, ApJ, 878, 61
Nakamura M., Asada K., Hada K., et al. 2018, ApJ, 868, 146
Narayan R., Fabian A. C., 2011, MNRAS, 415, 3721
Nokhrina E. E., Beskin V. S., Kovalev Y. Y., Zheltoukhov A. A.,
2015, MNRAS, 447, 2726
Nokhrina E. E., Gurvits L. I., Beskin V. S., Nakamura M., Asada
K., Hada K., 2019, arXiv e-prints,
Owen F. N., Eilek J. A., Kassim N. E., 2000, ApJ, 543, 611
Palma N. I., et al., 2011, ApJ, 735, 60
Pelletier G., Pudritz R. E., 1992, ApJ, 394, 117
Plavin A. V., Kovalev Y. Y., Pushkarev A. B., Lobanov A. P.,
2019a, MNRAS, 485, 1822
Plavin A. V., Kovalev Y. Y., Petrov L. Y., 2019b, ApJ, 871, 143
Porcas R. W., 2009, A&A, 505, L1
Porth O., Fendt C., Meliani Z., Vaidya B., 2011, ApJ, 737, 42
Potter W. J., Cotter G., 2013, MNRAS, 431, 1840
Potter W. J., Cotter G., 2015, MNRAS, 453, 4070
Pushkarev A. B., Kovalev Y. Y., 2012, A&A, 544, A34
Pushkarev A. B., Kovalev Y. Y., Lister M. L., Savolainen T.,
2009, A&A, 507, L33
Pushkarev A. B., Hovatta T., Kovalev Y. Y., Lister M. L.,
Lobanov A. P., Savolainen T., Zensus J. A., 2012a, A&A,
545, A113
Pushkarev A. B., Hovatta T., Kovalev Y. Y., Lister M. L.,
Lobanov A. P., Savolainen T., Zensus J. A., 2012b, A&A,
545, A113
Pushkarev A. B., Kovalev Y. Y., Lister M. L., Savolainen T.,
2017, MNRAS, 468, 4992
Quataert E., Narayan R., 2000, ApJ, 528, 236
Russell H. R., Fabian A. C., McNamara B. R., Broderick A. E.,
2015, MNRAS, 451, 588
Salvatier J., Wiecki T. V., Fonnesbeck C., 2016, PyMC3: Python
probabilistic programming framework, Astrophysics Source
Code Library (ascl:1610.016)
Savolainen T., Homan D. C., Hovatta T., Kadler M., Kovalev
Y. Y., Lister M. L., Ros E., Zensus J. A., 2010, A&A, 512, A24
Shaw M. S., et al., 2012, ApJ, 748, 49
Shcherbakov R. V., 2008, ApJS, 177, 493
Tchekhovskoy A., McKinney J. C., Narayan R., 2009, ApJ,
699, 1789
Tchekhovskoy A., Narayan R., McKinney J. C., 2011, MNRAS,
418, L79
Torrealba J., Chavushyan V., Cruz-González I., Arshakyan T. G.,
Berton E., Rosa-González D., 2012, Revista Mexicana de As-
tronomı́a y Astrofı́sica, 48, 9
Tseng C.-Y., Asada K., Nakamura M., Pu H.-Y., Algaba J.-C.,
Lo W.-P., 2016, ApJ, 833, 288
Vermeulen R. C., Ros E., Kellermann K. I., Cohen M. H., Zensus
J. A., van Langevelde H. J., 2003, A&A, 401, 113
Vestergaard M., Peterson B., 2006, ApJ, 641, 689
Vlahakis N., Königl A., 2003, ApJ, 596, 1080
Vlahakis N., Königl A., 2004, ApJ, 605, 656
Wang C.-C., Zhou H.-Y., 2009, MNRAS, 395, 301
Woo J. H., Urry C. M., 2002, ApJ, 579, 530

This paper has been typeset from a TEX/LATEX file prepared by


the author.

MNRAS 000, 1–15 (2019)

Anda mungkin juga menyukai