25.03.2013 Views

Contents List of Figures

Contents List of Figures

Contents List of Figures

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

CONTENTS 265<br />

<strong>Contents</strong><br />

6 Macro– and Micro–Jets Driven by Black Holes 267<br />

6.1 DRAG(o)Ns – FR I and FR II . . . . . . . . . . . . . . . . . . . . . . 268<br />

6.2 Jets as Super(magneto)sonic Collimated Plasma Flows . . . . . . . . 275<br />

6.3 Fundamental Parameters for Jets . . . . . . . . . . . . . . . . . . . . 278<br />

6.4 A 3–Phase Model for Kiloparsec–Scale Jets in Clusters . . . . . . . . 283<br />

6.5 Relativistic Jet Propagation . . . . . . . . . . . . . . . . . . . . . . . 288<br />

6.6 Structure and Emission <strong>of</strong> Micro–Jets . . . . . . . . . . . . . . . . . . 291<br />

6.6.1 Core–Jets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 292<br />

6.6.2 Emission Properties . . . . . . . . . . . . . . . . . . . . . . . . 293<br />

6.7 Formation <strong>of</strong> Micro–Jets . . . . . . . . . . . . . . . . . . . . . . . . . 301<br />

6.7.1 The Stationary MHD Model . . . . . . . . . . . . . . . . . . . 304<br />

6.7.2 The Collimation Zone . . . . . . . . . . . . . . . . . . . . . . 307<br />

6.7.3 Nondiffusive Relativistic MHD Approach . . . . . . . . . . . . 309<br />

6.7.4 Knot Ejection Mechanisms . . . . . . . . . . . . . . . . . . . . 312<br />

7 The First Black Holes in the Universe 318<br />

7.1 The Dark Age . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 318<br />

7.2 The First Stars . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 320<br />

7.3 The First Quasars . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 320<br />

<strong>List</strong> <strong>of</strong> <strong>Figures</strong><br />

122 VLA Radio interferometer . . . . . . . . . . . . . . . . . . . . . . . . 267<br />

123 Space VLBI VSOP . . . . . . . . . . . . . . . . . . . . . . . . . . . . 268<br />

124 M 87 – the first Jet . . . . . . . . . . . . . . . . . . . . . . . . . . . . 270<br />

125 Structure <strong>of</strong> the DRAGN Cygnus A . . . . . . . . . . . . . . . . . . . 271<br />

126 Host galaxies <strong>of</strong> FR I and II . . . . . . . . . . . . . . . . . . . . . . . 272<br />

127 DRAGN M 84 in the Virgo cluster . . . . . . . . . . . . . . . . . . . 273<br />

128 FR I versus FR II . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 274<br />

129 Jet Paradigm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 276<br />

130 Fundamental plane for hydro jets . . . . . . . . . . . . . . . . . . . . 278<br />

131 Luminosity <strong>of</strong> radio galaxies . . . . . . . . . . . . . . . . . . . . . . . 280<br />

132 Electron density in the Hydra cluster . . . . . . . . . . . . . . . . . . 283<br />

133 Simulated dark matter and gas pr<strong>of</strong>iles . . . . . . . . . . . . . . . . . 284<br />

134 Cyg A in X–rays and radio . . . . . . . . . . . . . . . . . . . . . . . . 285<br />

135 Bipolar jets in the cluster gas . . . . . . . . . . . . . . . . . . . . . . 286<br />

136 Jet evolution <strong>of</strong> 3C galaxies . . . . . . . . . . . . . . . . . . . . . . . 288<br />

137 The radio halo <strong>of</strong> M 87 . . . . . . . . . . . . . . . . . . . . . . . . . . 289<br />

138 Electron positron jet . . . . . . . . . . . . . . . . . . . . . . . . . . . 291


266 LIST OF FIGURES<br />

139 VLBA image <strong>of</strong> 3C 120 . . . . . . . . . . . . . . . . . . . . . . . . . . 292<br />

140 RXTE light curves <strong>of</strong> 3C 273 . . . . . . . . . . . . . . . . . . . . . . 293<br />

141 Global spectrum for 3C 273 . . . . . . . . . . . . . . . . . . . . . . . 294<br />

142 Dust emission in 3C 273 . . . . . . . . . . . . . . . . . . . . . . . . . 295<br />

143 VLBA image at 8.4 GHz <strong>of</strong> NGC 4261 . . . . . . . . . . . . . . . . . 298<br />

144 Disk geometry in the center <strong>of</strong> NGC 4261 . . . . . . . . . . . . . . . . 299<br />

145 Magnetic jet formation . . . . . . . . . . . . . . . . . . . . . . . . . . 303<br />

146 Black Hole magnetospheres . . . . . . . . . . . . . . . . . . . . . . . 308<br />

147 Particle density along flux tube . . . . . . . . . . . . . . . . . . . . . 310<br />

148 Koide disk–jet . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 311<br />

149 3C 120 sequence <strong>of</strong> images . . . . . . . . . . . . . . . . . . . . . . . . 313<br />

150 RXTE light curves <strong>of</strong> 3C 120 . . . . . . . . . . . . . . . . . . . . . . 315<br />

151 Knot formation in Jets . . . . . . . . . . . . . . . . . . . . . . . . . . 316<br />

152 Dark ages <strong>of</strong> the Universe . . . . . . . . . . . . . . . . . . . . . . . . 319<br />

153 Mass range for Black Hole formation . . . . . . . . . . . . . . . . . . 321<br />

154 Cloud formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 322


6 Macro– and Micro–Jets Driven by Black Holes<br />

267<br />

Active galaxies were first identified in the 1950s when early radio telescopes found<br />

that certain galaxies emit strong radio signals. Something very exotic had to be<br />

going on in those galaxies to produce the energies needed to emit the signals. As<br />

researchers developed new tools that made it possible to see across the entire electromagnetic<br />

spectrum - infrared and ultraviolet emissions, X-ray and gamma ray<br />

sources - astronomers found active galaxies also to be strong emitters <strong>of</strong> other radiation.<br />

This was more evidence that some kind <strong>of</strong> extremely energetic process<br />

had to be at work. Scientists have long thought that matter falling into supermassive<br />

black holes might be the power source <strong>of</strong> active galaxies. That hypothesis<br />

Figure 122: The structure <strong>of</strong> kiloparsec–scale jets has mainly been investigated by<br />

the VLA radio interferometer.<br />

was strengthened by astronomers’ recent understanding <strong>of</strong> microquasars, objects<br />

that show characteristics similar to active galaxies but which are less massive and<br />

also much closer to the earth. Being closer - only thousands <strong>of</strong> light years away -<br />

microquasars have been observed with greater resolution and detail than has been<br />

possible with active galaxies up until now. The active galaxy 3C 120 is 450 million<br />

light years away. The supermassive black hole the researchers expected to be at its


268 6 MACRO– AND MICRO–JETS DRIVEN BY BLACK HOLES<br />

core would have the mass <strong>of</strong> at least 30 million suns, yet would be squeezed into<br />

a region smaller than the distance between the earth and the sun. The tremendous<br />

gravitational force <strong>of</strong> this mass would continuously attract more matter, which<br />

would spiral around the black hole into a thin doughnut-shape called an accretion<br />

disk. A jet was first observed in the radio galaxy M 87 in 1917 by Heber Curtis using<br />

Figure 123: Space VLBI VSOP provides the highest spatial resolution down to 0.1<br />

mas.<br />

an optical telescope. However, it was not until the 1970’s, when observations with<br />

large, high resolution radio telescopes, like the Very Large Array in New Mexico,<br />

revealed the nature <strong>of</strong> the ”curious straight ray” connected to the nucleus <strong>of</strong> M 87.<br />

Jets are present in many, if not all, quasars and can extend millions <strong>of</strong> light years<br />

from the central core <strong>of</strong> the galaxy.<br />

6.1 DRAG(o)Ns – FR I and FR II<br />

The jets <strong>of</strong> quasars and Radio galaxies are certainly the most spectacular collimated<br />

plasma flows in the Universe. Their plasma is probably very exotic and is only


6.1 DRAG(o)Ns – FR I and FR II 269<br />

visible in non–thermal emission. In contrast to these objects, jets <strong>of</strong> normal stellar<br />

objects (young stellar objects, protostars and White Dwarfs) can be detected in<br />

line emission <strong>of</strong> a normal gas. As such one has a much more direct observational<br />

evidence for the flow velocities, they are moderate a few hundred to a few thousand<br />

kilometers per second in stellar objects.<br />

What are DRAGNs ? DRAGNs are large–scale double radio sources<br />

produced by outflows (jets) that are launched by processes in active<br />

galactic nuclei (AGN). They are clouds <strong>of</strong> radio-emitting plasma which have<br />

been shot out <strong>of</strong> active galactic nuclei via narrow jets. The term DRAGN is an<br />

acronym for ”Double Radio Source Associated with a Galactic Nucleus” . It was<br />

coined by Patrick Leahy in 1993 [17].<br />

Most <strong>of</strong> the extragalactic sources detected by early radio surveys were DRAGNs,<br />

but modern deep radio surveys also detect many sources produced by ”starbursts”<br />

in galaxies. Starbursts can produce radio emission throughout an extended region<br />

<strong>of</strong> a galaxy, but do not produce significant radio emission in regions far from the<br />

stars that have formed. DRAGNs fundamentally involve the mass concentration in<br />

the nucleus <strong>of</strong> the galaxy and produce radio emission throughout regions that are<br />

much larger than the host galaxies themselves.<br />

DRAGNs are formed when an active galactic nucleus produces two persistent,<br />

oppositely-directed plasma outflows that contain cosmic ray electrons and magnetic<br />

fields. We do not know exactly what these ”jet” outflows contain, but they clearly<br />

include fast-moving electrons and partially ordered magnetic fields which make them<br />

visible at radio (and higher) frequencies by their synchrotron radiation (emission<br />

that occurs when electrons move across a magnetic field at almost the velocity <strong>of</strong><br />

light). To preserve overall charge neutrality, there must also be either protons or<br />

positrons in the outflows.<br />

DRAGNs are made, so it seems, <strong>of</strong> the stuff that flows down the jets, the socalled<br />

”synchrotron plasma”. We do not know exactly what this is. It radiates in the<br />

radio band (and sometimes at higher frequencies) via the synchrotron process, from<br />

which we can tell that it contains magnetic fields and cosmic-ray electrons. Since<br />

cosmic plasmas must be neutral, there must be also protons or positrons in the jet.<br />

Near the AGN the jets are supersonic, in the sense that the flow speed is faster than<br />

the speed <strong>of</strong> sound in the jet plasma. As a result, shockwaves form easily, giving rise<br />

to small regions with high pressure, which radiate intensely. Further away from the<br />

AGN, the weaker jets become subsonic and turbulent. Both shocks and turbulence<br />

may help to give the cosmic rays their high energies, via the Fermi mechanism.<br />

Structure <strong>of</strong> DRAGNs: The main radio structures in a powerful DRAGN whose<br />

jets stay supersonic to great distances are illustrated above with the VLA image <strong>of</strong><br />

the radio galaxy 3C 405 (Cygnus A, Fig. 125). A compact radio source usually de-


270 6 MACRO– AND MICRO–JETS DRIVEN BY BLACK HOLES<br />

Figure 124: M 87 in the Virgo cluster is historically the first jet detected in optical<br />

emission (Curtis 1918). With HST the structure <strong>of</strong> this jet is nicely visible, the jet<br />

ends beyond the extension <strong>of</strong> the core region <strong>of</strong> the optical galaxy (top). This jet<br />

has been imaged in radio, optical and X–ray emission (bottom). M 87 is however<br />

not a typical DRAGN.


6.1 DRAG(o)Ns – FR I and FR II 271<br />

Figure 125: Structure <strong>of</strong> the archetypical DRAGN Cygnus A.<br />

scribed as the ”core” coincides with the galactic nucleus. Most <strong>of</strong> the radio emission<br />

does not come directly from the well-collimated ”jets”, but from the broader ”lobes”<br />

that are found around the jet paths. The plasma in the lobes is believed to have been<br />

supplied by the jets over millions <strong>of</strong> years. In the more powerful DRAGNs, there<br />

are usually small, bright radio ”hot spots” near the boundaries <strong>of</strong> each lobe. These<br />

hot spots thought to be where strong shocks form near the ends <strong>of</strong> the supersonic<br />

jet outflows in the powerful DRAGNs.<br />

The environs <strong>of</strong> a galaxy are not a perfect vacuum. The jets must travel first<br />

through the atmosphere <strong>of</strong> the galactic nucleus, then through the interstellar medium<br />

<strong>of</strong> the host galaxy, then (if they get that far without being slowed down by interaction)<br />

through the successively lower densities and pressures <strong>of</strong> the outer halo <strong>of</strong><br />

the galaxy, the intra-cluster medium (ICM) <strong>of</strong> any surrounding group or cluster <strong>of</strong><br />

galaxies, and finally into the low-density inter-galactic medium. The jets do not flow<br />

freely away from the AGN, but must push their way through these external media.<br />

Fanar<strong>of</strong>f–Riley Classification: A landmark in the study <strong>of</strong> extragalactic radio<br />

sources was the demonstration by Fanar<strong>of</strong>f and Riley (1974) <strong>of</strong> the existence <strong>of</strong> a<br />

relatively sharp morphological transition at a radio luminosity P178 MHz 2.5 ×


272 6 MACRO– AND MICRO–JETS DRIVEN BY BLACK HOLES<br />

10 26 h −2<br />

50 W Hz −1 . The great majority <strong>of</strong> sources below this luminosity (FR I type)<br />

are characterized by having diffuse radio lobes, with their brightest region within<br />

the inner half <strong>of</strong> the radio source (edge–dimmed). On the contrary, more powerful<br />

sources are usually straighter, exhibit edge–brightened (FR II) morphology, and<br />

typically contain hotspots near the outer edges <strong>of</strong> their radio lobes. Most recently,<br />

Figure 126: Correlation between radio luminosity, absolute optical red magnitude<br />

and Fanar<strong>of</strong>f–Riley class. 1’s are FR I’s, 2’s are FR II’s and F’s are fat doubles.<br />

[Ledlow et al. 2000]<br />

it was realized that the critical radio luminosity separating the FR I and FR II<br />

actually increases with the optical luminosity <strong>of</strong> the host galaxy, P ∗ R ∝ L1.65 opt (Fig.<br />

126). The more luminous the host galaxy is, the more powerful the radio source<br />

must be in order to attain the FR II morphology.<br />

There is good evidence that the outflow speeds in DRAGNs are initially relativistic,<br />

i.e. that the material travels outwards at a large fraction <strong>of</strong> the velocity <strong>of</strong> light,<br />

<strong>of</strong>ten with Lorentz factors <strong>of</strong> order 10. The velocities appear to decrease along many<br />

<strong>of</strong> the the jets, however. In the less powerful sources, it appears that the jets are<br />

slowed down enough on the scale <strong>of</strong> the galaxy to become subsonic and turbulent.


6.1 DRAG(o)Ns – FR I and FR II 273<br />

These FR Type I sources form large-scale plumes that meander across the sky at<br />

the mercy <strong>of</strong> pressure gradients, shocks and winds in the large–scale intracluster or<br />

intergalactic media. An example is 3C 31 (identified with the galaxy NGC 383),<br />

shown below with the DRAGN in red and the optical field in blue:<br />

In more powerful (FR Type II) sources, it appears that the jets remain at least<br />

mildly relativistic (and supersonic) out to great distances from their host galaxies,<br />

to form the ”classical” double–lobed structures like that shown above for 3C 405<br />

(Cygnus A). In the Type II sources the ends <strong>of</strong> the jets move outwards more slowly<br />

than material flows along the jet. The plasma arriving at the end <strong>of</strong> their jets is<br />

then deflected back around them to form the lobes, large ”bubbles” in the medium<br />

surrounding the galaxy.<br />

Sizes <strong>of</strong> DRAGNs: Individual DRAGNs will tend to grow while their central<br />

engine remains active. The smallest DRAGNs known are only a few tens <strong>of</strong> parsecs<br />

across, contained within in the active nucleus. An example <strong>of</strong> a weak, small DRAGN<br />

on a galactic scale is 3C 272.1, associated with the Virgo Cluster elliptical galaxy<br />

M 84 (NGC 4374), shown in Fig. 127 with the DRAGN in red and the optical field<br />

in blue.<br />

Figure 127: DRAGN <strong>of</strong> FR type I associated with the Virgo cluster elliptical M 84<br />

(elliptical galaxy in blue, radio structure in red).<br />

At the other extreme, giant DRAGNs are the largest known objects, up to several


274 6 MACRO– AND MICRO–JETS DRIVEN BY BLACK HOLES<br />

Megaparsecs from end to end. The lobes <strong>of</strong> the radio source shown below were<br />

discovered by a low-resolution all-sky survey made with the VLA. They were later<br />

shown to be a single DRAGN by observations that revealed the radio core and the<br />

jets connecting it to these lobes (left panel). The core coincides (central panel)<br />

with a galaxy (right panel) whose redshift is z=0.154. This reveals the double radio<br />

source to be over 4 Mpc in extent, and the second largest DRAGN known. The<br />

largest <strong>of</strong> them all, 3C236, is about 6 Mpc across. Typical powerful DRAGNs are<br />

Figure 128: Difference in morphology between FR I jet <strong>of</strong> 3C 31 and a typical quasar<br />

jet (3C 175 at redshift 0.768). Quasars show the most luminous jets with double<br />

sided lobe structure, but only one–sided jets.


6.2 Jets as Super(magneto)sonic Collimated Plasma Flows 275<br />

hundreds <strong>of</strong> kiloparsecs across, several times bigger than their host galaxies. 1<br />

DRAGNs are almost invariably associated with elliptical galaxies, rather than<br />

with spirals. There appears to be a strong connection, still to be explained in detail,<br />

between the processes that determine the bulge-to-disk ratio <strong>of</strong> the host galaxy and<br />

the ability to form (and maintain) the active nuclear engine for the time it takes to<br />

build up a large-scale DRAGN.<br />

Ages <strong>of</strong> DRAGNs: Powerful DRAGNs have estimated lifetimes <strong>of</strong> order 20 million<br />

years (estimated from synchrotron ages); this makes them a brief outburst in<br />

the life <strong>of</strong> a galaxy (compare the 100 million years a star may take to orbit the<br />

galaxy’s hub). On the other hand, weak DRAGNs are so common in the largest<br />

elliptical galaxies that the jets must be ”on” essentially all the time. Most <strong>of</strong> the<br />

radio emission from typical DRAGNs comes not from the jets themselves but from<br />

twin lobes, which are much broader clouds around and near the inferred path <strong>of</strong><br />

the two jets. The synchrotron plasma in the lobes is believed to have been supplied<br />

over a long period through the jets. The jets themselves can sometimes be seen as<br />

narrow features threading the lobes. In powerful DRAGNs, small, bright hotspots<br />

are found near the end <strong>of</strong> each lobe; these are thought to mark the ends <strong>of</strong> the jets.<br />

6.2 Jets as Super(magneto)sonic Collimated Plasma Flows<br />

The standard interpretation <strong>of</strong> DRAGNs was proposed independently by Scheuer<br />

(1974) and Blandford & Rees (1974), and the underlying physical mechanism was<br />

broadly confirmed by numerical fluid dynamics in the 1980’s. To understand DRAGNs,<br />

it is crucial to bear in mind that space is not a perfect vacuum. The jets travel out<br />

first through the atmosphere <strong>of</strong> the galactic nucleus, then through the interstellar<br />

medium <strong>of</strong> the host galaxy, then (if they get that far) through the successively lower<br />

densities and pressures <strong>of</strong> the outer halo <strong>of</strong> the galaxy, the intra-cluster medium <strong>of</strong><br />

any surrounding group or cluster <strong>of</strong> galaxies, and out into the inter-galactic medium.<br />

Although much <strong>of</strong> this gas is pretty tenuous even by astronomical standards, it is<br />

always denser than the jet plasma. This means the jets cannot flow freely away from<br />

the AGN, but must push their way through the external medium (Fig. 129). As<br />

a result, the ends <strong>of</strong> the jets move outwards much more slowly than material flows<br />

up the jet. As envisaged by Scheuer and Blandford & Rees, the plasma arriving at<br />

the end <strong>of</strong> the jet is deflected back to form the lobe, which can be thought <strong>of</strong> as<br />

a large bubble surrounding the jet. The ”end <strong>of</strong> the jet” is just the point where<br />

1 Patrick Leahy and Richard Strom have collected an Atlas <strong>of</strong> DRAGNs that includes images,<br />

vital statistics and brief descriptions <strong>of</strong> each object. The atlas is served from Patrick’s web site at<br />

Jodrell Bank http://www.jb.man.ac.uk/atlas/. The Atlas is a complete sample <strong>of</strong> all the DRAGNs<br />

from the Third Cambridge Catalogue <strong>of</strong> Radio Sources (Second Revision, known as 3CRR), out<br />

to a redshift <strong>of</strong> z=0.5.


276 6 MACRO– AND MICRO–JETS DRIVEN BY BLACK HOLES<br />

the jet collides with the surface <strong>of</strong> the lobe. If the jet is still supersonic at this<br />

point, this collision will take place through a system <strong>of</strong> strong shockwaves, and the<br />

resulting high–pressure region will be seen as the hotspot. Over the past 20 years,<br />

Figure 129: Structural elements <strong>of</strong> propagating jets: beam, cocoon, hotspot, contact<br />

discontinuity (which is unstable against Kelvin–Helmholtz instabilities) and bow<br />

shock. Parameters: M = 6, η = 0.01, vB = 0.28c, ΓB = 1.04, resolution = 6 ppb.<br />

[Hughes 1996].<br />

numerical simulations <strong>of</strong> time–dependent jet flows have progressed enormously from<br />

the earliest two–dimensional axisymmetric gasdynamical flows (Smith et al. 1985)<br />

to three–dimensional flows (e.g., Cox, Gull & Scheuer 1991) until now fully three<br />

dimensional flows incorporating self-consistent MHD are relatively straightforward,<br />

if not yet easy, to model with modest resolution (Krause 2002 [11]; Krause 2003<br />

[12]). These simulation methods have also been extended to include flows in either<br />

2D or 3D with relativistic bulk motions (see review by Marti & Müller 1999 [15]).<br />

Simultaneously, physical models <strong>of</strong> particle acceleration physics, especially as it relates<br />

to the formation <strong>of</strong> collisionless shocks, have become much better developed<br />

(e.g., Jones 2001), even if that cannot yet be called a solved problem.<br />

Most <strong>of</strong> our information about RGs currently derives from radio synchrotron<br />

emissions reflecting the spatial and energy distributions <strong>of</strong> relativistic electrons convolved<br />

with the spatial distribution <strong>of</strong> magnetic fields. X-ray observations, especially<br />

<strong>of</strong> non–thermal Compton emissions, depending on the electron and ambient photon<br />

distributions, are now beginning to add crucial information, as well. Using these<br />

connections, much effort has been devoted to interpreting observed brightness, spectral<br />

and polarization properties <strong>of</strong> the non–thermal emissions for estimates <strong>of</strong> the<br />

key physical source properties, such as the energy and pressure distributions and


6.2 Jets as Super(magneto)sonic Collimated Plasma Flows 277<br />

kinetic power, as well as to find self-consistent models for the particle acceleration<br />

and flow patterns. As telescopes and analysis techniques have improved the level <strong>of</strong><br />

detail obtained, it has become apparent, however, that the observed properties are<br />

not very simple and much harder to interpret than most simple models predict.<br />

Radiative MHD with NIRVANA C: In the following we work in the one–<br />

component approximation for the dynamical part, but include cooling for the evolution<br />

<strong>of</strong> different atomic species [11]. The evolution <strong>of</strong> the jet plasma is given by the<br />

continuity equation for the density ρ, the Euler equations for the momenta ρ V which<br />

have pressure gradient and Lorentz forces as source terms. The magnetic fields B<br />

evolve according to the induction equation, and the time evolution <strong>of</strong> the internal<br />

energy e is determined by advection, compression and cooling (the K–term)<br />

∂ρ<br />

∂t + ∇ · (ρ V ) = 0 (673)<br />

∂ρ V<br />

∂t + ∇ · (ρ V ⊗ V ) = −∇P − 1<br />

8π ∇ B 2 + 1<br />

4π ( B · ∇) B (674)<br />

∂e<br />

∂t + ∇ · (e V ) = −P (∇ · V ) − K (675)<br />

∂ B<br />

∂t = ∇ × ( V × B) (676)<br />

P = (Γ − 1)e . (677)<br />

These equations are implemented in codes such as ZEUS3D and NIRVANA C [22],<br />

or more modern versions based on shock capturing methods [23].<br />

Cooling Functions for Low–Density Plasmas: Sutherland & Dopita [21] have<br />

presented the equilibrium cooling functions for optically thin plasmas. These can<br />

easily built into the code NIRVANA C. Some functions are terminated before reaching<br />

10 4 K when the internal photoionization halts the cooling. The functions represent<br />

a self-consistent set <strong>of</strong> curves covering a wide grid <strong>of</strong> temperature and metallicities<br />

using recently published atomic data and processes. The results have implications<br />

for phenomena such as cooling flows and for hydrodynamic modelling which<br />

include gas components. NIRVANA C can however also handle non–equilibrium<br />

cooling for an entire set <strong>of</strong> atomic species. Then the above equations must be supplemented<br />

with a network <strong>of</strong> time evolution for various species. In cases, where<br />

cooling is very rapid compared to dynamical time–scales, this network has to be<br />

solved by means <strong>of</strong> time–implicit methods.


278 6 MACRO– AND MICRO–JETS DRIVEN BY BLACK HOLES<br />

6.3 Fundamental Parameters for Jets<br />

Already in early simulations, it turned out that one <strong>of</strong> the fundamental parameters<br />

for jet propagation is the density contrast η = ρB/ρ0, where ρB denotes the beam<br />

density and ρ0 the external density <strong>of</strong> the surrounding medium. For relativistic<br />

beams propagating with a Lorentz bulk factor ΓB, the density contrast is given by<br />

η = ρBhBΓ 2 B<br />

ρ0<br />

, (678)<br />

with hB as the specific enthalpy <strong>of</strong> the beam plasma (= 1 for cold plasma) and ΓB<br />

as the Lorentz factor <strong>of</strong> the beam propagation. For Quasar jets we typically find<br />

ΓB 5 − 10 from superluminal motion on the parsec–scale. The third parameter<br />

which determines the morphology <strong>of</strong> a jet is the internal Mach number M = VB/cS <br />

5−10 for the beam plasma. The Mach number is not an independent quantity, since<br />

the beam is steadily heated up by means <strong>of</strong> internal shocks. In addition, the ratio<br />

Figure 130: Fundamental plane for hydro jets.<br />

between external pressure and internal pressure is also important for the initial<br />

launch <strong>of</strong> the jet. Overpressured jets will initially expand and then recollimate.<br />

The density contrast determines the hot spot advance speed over momentum<br />

balance at the working surface<br />

vHS = vB<br />

√<br />

ηɛ<br />

1 + √ , (679)<br />

ηɛ


6.3 Fundamental Parameters for Jets 279<br />

provided M >> 1. ɛ < 1 is the ratio <strong>of</strong> beam to head cross–section. For heavy<br />

jets, η > 1, we find therefore vHS vB, i.e. the hot spot advances with the beam<br />

speed. For light jets, η


280 6 MACRO– AND MICRO–JETS DRIVEN BY BLACK HOLES<br />

Cyg A<br />

L_R = 10^45 erg/s<br />

L_J = 10^46 erg/s<br />

_________________________________________<br />

3C295<br />

___________ FR II / FR I Transition ___________<br />

Figure 131: Luminosity <strong>of</strong> radio galaxies as a function <strong>of</strong> redshift. Open circles:<br />

galaxies in the 3C catalog. The radio power varies over many orders <strong>of</strong> magnitude.<br />

At high redshifts, only the most luminous sources with power comparable to Cyg A<br />

can be detected.<br />

On the FR I – FR II Transition: Many explanations have been given for the<br />

explanation <strong>of</strong> the FR I / FR II transition. The most widely known explanation is<br />

that, while the jets in both cases start out moving at very high (relativistic speeds),<br />

those in FR II sources remain that way out to multi–kpc distances, while those in<br />

FR I’s decelerate to much slower speeds within a few kpc <strong>of</strong> the galaxy core. Bicknell<br />

(1995) developed detailed models for decelerating relativistic jets.<br />

An alternative approach assumes that the FR I and FR II sources differ primarily<br />

in the importance <strong>of</strong> the beam thrust relative to the basic parameters <strong>of</strong> the ambient<br />

medium (Gopal–Krishna & Wiita 1988; 2001). In this version, the emphasis is on<br />

the slowing down <strong>of</strong> the advance <strong>of</strong> the hotspot at the end <strong>of</strong> the jet, rather than<br />

the slowing down <strong>of</strong> the bulk flow <strong>of</strong> the beam. In this scenario, the hotspots <strong>of</strong> FR<br />

II’s have supersonic advance speed with respect to the ambient medium. When this<br />

advance speed becomes transonic relative to the ambient medium, its Mach disk<br />

weakens due to the fall in ram pressure, and the jet becomes decollimated.


6.3 Fundamental Parameters for Jets 281<br />

The critical density contrast ηcrit can be estimated from the condition that the<br />

bow shock becomes transonic, vHS = cS. For relativistic beams vB c, we find<br />

ηcrit 10 −5 . For η < ηcrit, no bow shock will be found.<br />

For a more quantitative formulation one needs certain empirical relations between<br />

the elliptical’s blue magnitude MB, its s<strong>of</strong>t X–ray emission LX, the stellar<br />

velocity dispersion σ∗ (the Faber–Jackson relation), and the core radius Rc (Kormendy<br />

relation), assuming H0 = 75 (Bicknell 1995)<br />

log LX = 22.3 − 0.872 MB (685)<br />

log σ∗ = 5.412 − 0.0959 MB (686)<br />

log Rc = 11.7 − 0.436 MB . (687)<br />

We furthermore assume a density scaling for the ISM <strong>of</strong> the elliptical<br />

n(d) =<br />

n0<br />

[1 + (d/Rc) 2 ] δ<br />

(688)<br />

with δ 0.75 typically. Then the hotspot advance speed follows from pressure<br />

balance<br />

vHS(d) = vB<br />

X[1 + (d/Rc) 2 ] δ/2<br />

d/Rc + X[1 + (d/Rc) 2 . (689)<br />

] δ/2<br />

Here X = <br />

4LB/πΘ 2 R 2 cn0µmpc 3 with LB as the beam power, Θ the effective opening<br />

angle for the beam. A reasonable number is Θ 0.1 rad, at least for the inner<br />

jet regions. X corresponds to the above quantity √ χη for cylindrical beams. The<br />

temperature <strong>of</strong> the X–ray gas is not independent, but tied to the central stellar velocity<br />

dispersion, kT = 2.2µmpσ 2 ∗/δ (Falle 1987). There is evidence that the central<br />

gas density n0 is somewhat higher than the X–ray gas density nX, n0 = κnX with<br />

κ 3. The interstellar density declines quite rapidly at distances beyond a few kpc,<br />

observationally Rc 1 kpc. Hence the most likely regime for the jet’s decollimation<br />

due to hotspots having slowed down to subsonic speeds lies within roughly 10 kpc<br />

<strong>of</strong> the core.<br />

Gopal–Krishna & Wiita (2001) now evaluate the critical beam power L ∗ B for<br />

which the hotspot deceleration to subsonic velocities occurs at a distance d ∗ 3−10<br />

kpc from the core. 10 kpc is a typical distance at which jets flare in a sample <strong>of</strong><br />

radio galaxies. Laing et al. (1999) have studied a sample <strong>of</strong> 38 FR I sources giving<br />

a mean projected value <strong>of</strong> 3.5 kpc for the radial distance <strong>of</strong> the point where the<br />

kpc–scale jet first becomes visible, after passing through an emission gap.<br />

Now X can be scaled to<br />

X = C2<br />

√ <br />

LB/ n0Rc<br />

(690)


282 6 MACRO– AND MICRO–JETS DRIVEN BY BLACK HOLES<br />

where C2 = <br />

4/πΘc 3 µmp = 2.14 × 10 −3 for µ = 0.620. LX can be expressed over<br />

the Bremsstrahlung cooling function, sot hat<br />

log LX = log C1 + log n0 + 3 log aX (691)<br />

log nX = log n0 − log κ = 4.188 + 0.218 MB (692)<br />

log X = −0.109 MB + log C2 + 0.25 log C1 + 3.20 + 0.5 log LB . (693)<br />

We next impose the transonic condition<br />

vHS(d ∗ <br />

) = cS = 2.2γ/δ σ∗ 1.5 σ∗ . (694)<br />

For γ = 5/3 and δ = 0.75 this becomes<br />

log vHS(d ∗ ) = 0.345 + log σ∗ = −4.720 − 0.0959 MB . (695)<br />

Equating this to the original expression for vHS gives<br />

dex(−4.720 − 0.0959 MB) =<br />

X<br />

(d/Rc)/[1 + (d/Rc) 2 ] δ/2 . (696)<br />

+ X<br />

This is the relation which must be solved for the critical beam power L ∗ B as a<br />

function <strong>of</strong> MB. The resulting values are given in the following table (GKW 2001).<br />

For MB > −23.5 this corresponds to a scaling L ∗ B ∝ L 1.6<br />

opt, in good agreement with<br />

MB Rc n0 L ∗ B β ∗<br />

[kpc] [cm −3 ] [erg/s]<br />

-19.5 0.052 2.82 4.35(41) 1.74<br />

-20.0 0.085 2.02 8.28(41) 1.40<br />

-20.5 0.141 1.57 1.71(42) 1.57<br />

-21.0 0.233 1.22 3.57(42) 1.60<br />

-21.5 0.384 0.951 7.23(42) 1.56<br />

-22.0 0.635 0.741 1.49(43) 1.58<br />

-22.5 1.05 0.576 3.06(43) 1.55<br />

-23.0 1.73 0.447 6.21(43) 1.55<br />

-23.5 2.86 0.348 1.23(44) 1.47<br />

Table 8: Galaxy and jet parameters for given elliptical galaxy.<br />

the dividing line in Fig. 126 with β ∗ = 1.65. The actual position for the dividing line<br />

follows from the efficiency <strong>of</strong> conversion <strong>of</strong> beam power into synchrotron radiation<br />

<strong>of</strong> about 10%.


6.4 A 3–Phase Model for Kiloparsec–Scale Jets in Clusters 283<br />

Research Project: The ultimate reason for the FR II/FR I transition is still not<br />

yet clear–cut. The density distribution <strong>of</strong> the hot gas around elliptical galaxies embedded<br />

into a cluster background is more complex than discussed here. It consists<br />

<strong>of</strong> two β–laws, one with the core radius <strong>of</strong> the galactic nucleus, and a second one<br />

with the much bigger core radius <strong>of</strong> the cluster medium (see recent Chandra measurements).<br />

This produces a jump in the pressure gradients where the two density<br />

laws cross each other. The flaring <strong>of</strong> the radio jet in FR I sources could just occur<br />

at this pronounced radius.<br />

6.4 A 3–Phase Model for Kiloparsec–Scale Jets in Clusters<br />

Many <strong>of</strong> the bright 3C galaxies are members <strong>of</strong> clusters. The density pr<strong>of</strong>ile is therefore<br />

an important ingredient into the modelling <strong>of</strong> jet propagation. The observed<br />

electron density distribution for the Hydra cluster is shown in Fig. 132.<br />

Figure 132: Density distribution in the Hydra cluster, as measured with Chandra.<br />

Beyond 200 kpc, the density distribution considerably steepens, ne ∝ r −3 , as expected<br />

from cosmological simulations. [Chandra observation]


284 6 MACRO– AND MICRO–JETS DRIVEN BY BLACK HOLES<br />

The Archetype Cyg A: Cyg A is the central galaxy <strong>of</strong> a cluster. The ambient<br />

density is therefore not constant, but follows typically a β–law<br />

ρ0(r) =<br />

ρc<br />

(1 − r2 /r2 . (697)<br />

3β/2<br />

c)<br />

rc 20 − 50 kpc is the core radius for the cluster gas. The Cygnus A radio source<br />

Figure 133: Density pr<strong>of</strong>iles for dark matter and baryonic gas as obtained from<br />

simulations. Inside 100 kpc, the gas density pr<strong>of</strong>ile is quite shallow, beyond 200 kpc<br />

it adjust to the density pr<strong>of</strong>ile <strong>of</strong> dark matter.<br />

is situated at the center <strong>of</strong> a dense cluster atmosphere that extends to a radius <strong>of</strong><br />

at least 0.5 Mpc (Smith et al. 2002). The total X–ray luminosity <strong>of</strong> the cluster is<br />

LX = 1. × 10 45 erg/s at an average temperature <strong>of</strong> 8 keV, and an electron density at<br />

the position <strong>of</strong> the radio hotspots (70 kpc from the center) <strong>of</strong> 0.006 cm −3 . The total<br />

gas mass is 2 × 10 13 M⊙ and the total gravitational mass is 2 × 10 14 M⊙. A number<br />

<strong>of</strong> signatures in X–rays are expected from the interaction between the jet and the


6.4 A 3–Phase Model for Kiloparsec–Scale Jets in Clusters 285<br />

cluster gas, including emission from the unperturbed atmosphere, excess emission<br />

from the shocked ICM surrounding the radio lobes, and a deficit <strong>of</strong> emission from<br />

the evacuated radio lobes themselves. X–ray observations provide direct constraints<br />

Figure 134: A low frequency radio image (VLA at 330 MHz, blue to red) superposed<br />

on the Chandra image (yellow–red) <strong>of</strong> the inner 200 kpc <strong>of</strong> the Cygnus A cluster.<br />

This shows for the first time the interaction <strong>of</strong> the radio source and the cluster gas.<br />

on the physical conditions in the emitting regions. In contrast, observations <strong>of</strong> the<br />

non–thermal radio synchrotron emission provide only information on the properties<br />

<strong>of</strong> the cocoon and beam plasma. The pressure following from X–ray observations<br />

is 1 × 10 10 dyne cm −2 . Fig. 134 also reveals X–ray emission coincident with radio<br />

hot spots. This is non–thermal IC emission from the same population <strong>of</strong> electrons<br />

emitting the radio synchrotron photons (SSC emission). While the radio emissivity<br />

is a function <strong>of</strong> the relativistic electron density and the magnetic field strength, IC<br />

X–ray emissivity constrains the number density alone. From this Wilson derives<br />

magnetic fields <strong>of</strong> 150 µG for the radio hot spots, the minimum energy field would<br />

be 250 µG.<br />

(i) The Sedov–Phase <strong>of</strong> Jet Evolution: For density contrasts η < 0.01, the<br />

initial expansion <strong>of</strong> the bow shock can be described in terms <strong>of</strong> a Sedov wave propagating<br />

into a medium with decreasing density n(r) = n0(Rrc/r) κ with κ 1.4. The<br />

energy in the bubble increases steadily, E = LBt, as long as the jet power does not<br />

fade away. Similarly to the analysis for supernovae, the expansion <strong>of</strong> the blast wave


286 6 MACRO– AND MICRO–JETS DRIVEN BY BLACK HOLES<br />

Figure 135: Bipolar jets propagating in the cluster medium. The light blue color<br />

traces the bow shock and the shocked cluster medium, only visible in X–ray emission;<br />

the dark blue marks the shocked cocoon plasma with the contact discontinuity<br />

(visible in synchrotron emission). The cocoon plasma is extremely hot and thin,<br />

typical temperatures are ≥ 10 11 K. The dark colors mark bow shock plasma mixed<br />

into the cocoon via KH instabilities. The core region stays more or less round, while<br />

the accelerating bow shocks lead to a cigar–shape, once the hot spots left the core<br />

region <strong>of</strong> the cluster atmosphere. [Krause 2002].


6.4 A 3–Phase Model for Kiloparsec–Scale Jets in Clusters 287<br />

is given by [11]<br />

R(t) =<br />

For rc = 10 kpc, this leads to<br />

(3 − κ)(5 − κ)<br />

12πρ0r κ c<br />

1/(5−κ)<br />

t 3/(5−κ) . (698)<br />

R(t) = 3.3 kpc (t/Myr) 1/1.2 . (699)<br />

It is interesting, how accurate this simple formula is when compared to complicated<br />

numerical simulations [11]. For a synchrotron age <strong>of</strong> Cyg A <strong>of</strong> 27 Myr, this exactly<br />

corresponds to the radius <strong>of</strong> the X–ray bubble measured with Chandra (Fig. 134).<br />

Jets must be considered as bipolar outflows working against the background gas <strong>of</strong><br />

the cluster medium. For a density contrast η < 0.01, the initial phase is similar to a<br />

supernova explosion. We call therefore this phase the Sedov–phase. As shown by<br />

Cyg A, the bow shock advances only slightly supersonically, the typical sonic speed<br />

in the cluster medius is about 0.5 kpc/Myr, while the bow shock advance speed is<br />

a few kpc/Myr.<br />

(ii) The Cigar–Phase <strong>of</strong> Jet Evolution: Cygnus A just terminates the Sedov–<br />

phase. The hot spots are on the way to break out <strong>of</strong> the roundish bow shock due<br />

to acceleration in the decaying cluster atmosphere. In a few million years, Cyg A<br />

will appear as 3C 132 (Fig. 136) consisting <strong>of</strong> a central cylindrical cocoon filled<br />

with synchrotron plasma and two jets breaking out from this bubble and forming<br />

a cigar–like feature, such as seen in 3C 341. The long cylindrical jets visible in<br />

synchrotron light trace the cocoon plasma which appears highly collimated. The<br />

surrounding bow shock could only be detected in X–rays, is however fairly weak at<br />

distances <strong>of</strong> a few hundred kpc from the central source.<br />

(iii) The Late Phase Evolution: Since the central activity decays after typically<br />

20 – 50 million years, the radio lobes are no longer fuelled with high pressure<br />

material. The synchrotron plasma gets mixed up with the cluster gas. This leads<br />

to the formation <strong>of</strong> radio haloes in clusters <strong>of</strong> galaxies. In Fig. 137 we show the full<br />

extent <strong>of</strong> the radio source in the Virgo cluster at 327 MHz, about 80 kpc end–to–end.<br />

The inner lobes and jet can just be seen as the red structures in the center (and<br />

the nuclear black hole sits in the center <strong>of</strong> this inner structure). This picture thus<br />

covers an area larger than the optical galaxy; we are studying the inner region <strong>of</strong> the<br />

Virgo cluster. The galaxy, and its radio halo, sit in a large atmosphere <strong>of</strong> hot, X–ray<br />

loud gas (as already detected by the EINSTEIN and ROSAT satellites). From this<br />

image <strong>of</strong> M 87 we can see that the radio halo is a bubble, with a well–defined outer<br />

edge, sitting inside the X-ray emitting atmosphere. The bright plumes emanating<br />

from the inner lobes tell us that the halo is still ”alive”, being supplied with energy<br />

from the black hole. From our knowledge <strong>of</strong> the radio jet, we can tell that the jet is


288 6 MACRO– AND MICRO–JETS DRIVEN BY BLACK HOLES<br />

3C295<br />

3C132<br />

3 Phase Evolution<br />

<strong>of</strong> Jets<br />

3C341<br />

Figure 136: Time evolution <strong>of</strong> 3C galaxies as seen by computer simulations. Bright<br />

3C galaxies show various types <strong>of</strong> morphology, from the Cyg A type (3C 295),<br />

through break–out geometry (3C 132) and the typical cigar–shape (3C 341). These<br />

different forms are the result <strong>of</strong> a time–evolution <strong>of</strong> jets in the background <strong>of</strong> the<br />

inhomogeneous cluster medium.<br />

pumping at least at much energy into the local atmosphere, as is being lost by that<br />

atmosphere to X–rays. This and other evidence tells us, indirectly, that the core<br />

<strong>of</strong> the Virgo cluster is a complex, turbulent, magnetized region. It is anything but<br />

well-understood.<br />

6.5 Relativistic Jet Propagation<br />

In the last years, the Newtonian approach has been generalized to include relativistic<br />

propagation <strong>of</strong> the beam fluid. In Minkowski spacetime and Cartesian coordinates<br />

(t, x 1 , x 2 , x 3 ), the conservation equations can be written in conservative form as<br />

(Marti & Müller 1999)<br />

∂ U<br />

∂t + ∂F i ( U)<br />

= 0 . (700)<br />

∂xi


6.5 Relativistic Jet Propagation 289<br />

Figure 137: The kpc–scale jet <strong>of</strong> M 87 is embedded into a much more extended faint<br />

radio halo (VLA image at 327 MHz). Such extended radio halos are now found in<br />

many clusters <strong>of</strong> galaxies.<br />

The state vector is defined by<br />

and the flux vector F i as<br />

U = (D, S 1 , S 2 , S 3 , τ) T<br />

(701)<br />

F i = (Dv i , S 1 v i + P δ 1i , S 2 v i + P δ 2i , S 3 v i + P δ 3i , S i − Dv i ) T . (702)<br />

The five conserved quantities D, S 1 , S 2 , S 3 and τ are the rest–mass density, the three<br />

components <strong>of</strong> the momentum density, and the energy density (measured relative to<br />

the rest mass energy density), respectively. They are all measured in the laboratory<br />

frame, and are related to quantities in the local rest frame <strong>of</strong> the fluid (primitive<br />

variables) through<br />

D = ρΓb (703)


290 6 MACRO– AND MICRO–JETS DRIVEN BY BLACK HOLES<br />

S i = ρhΓ 2 bv i<br />

(704)<br />

τ = ρhΓ 2 b − D − P . (705)<br />

v i = u i /u t is the three velocity and Γb the Lorentz factor. The system <strong>of</strong> equations<br />

is closed by means <strong>of</strong> an equation <strong>of</strong> state (EOS), which we shall assume to be given<br />

in the form P = P (ρ, ɛ).<br />

In the non-relativistic limit D, S i and τ approach their Newtonian counterparts<br />

ρ, ρv i and ρE = ρɛ + ρv 2 /2, and the equations <strong>of</strong> system reduce to the classical<br />

ones. In the relativistic case the equations <strong>of</strong> are strongly coupled via the Lorentz<br />

factor and the specific enthalpy, which gives rise to numerical complications.<br />

An important property <strong>of</strong> system (5) is that it is hyperbolic for causal EOS.<br />

For hyperbolic systems <strong>of</strong> conservation laws, the Jacobians have real eigenvalues<br />

and a complete set <strong>of</strong> eigenvectors. Information about the solution propagates at<br />

finite velocities given by the eigenvalues <strong>of</strong> the Jacobians. Hence, if the solution is<br />

known (in some spatial domain) at some given time, this fact can be used to advance<br />

the solution to some later time (initial value problem). However, in general, it is<br />

not possible to derive the exact solution for this problem. Instead one has to rely<br />

on numerical methods which provide an approximation to the solution. Moreover,<br />

these numerical methods must be able to handle discontinuous solutions, which are<br />

inherent to non-linear hyperbolic systems.<br />

The simplest initial value problem with discontinuous data is called a Riemann<br />

problem, where the one dimensional initial state consists <strong>of</strong> two constant states separated<br />

by a discontinuity. The majority <strong>of</strong> modern numerical methods, the so-called<br />

Godunov–type methods, are based on exact or approximate solutions <strong>of</strong> Riemann<br />

problems [23, 20].<br />

Although MHD and general relativistic effects seem to be crucial for a successful<br />

launch <strong>of</strong> the jet, purely hydrodynamic, special relativistic simulations are adequate<br />

to study the morphology and dynamics <strong>of</strong> relativistic jets at distances sufficiently far<br />

from the central compact object (i.e., at parsec scales and beyond). The development<br />

<strong>of</strong> relativistic hydrodynamic codes based on HRSC techniques has triggered the<br />

numerical simulation <strong>of</strong> relativistic jets at parsec and kilo–parsec scales [18]. In Fig.<br />

138 we show the time evolution <strong>of</strong> a light, relativistic (beam flow velocity equal to<br />

0.99) jet with large internal energy. The logarithm <strong>of</strong> the proper rest-mass density is<br />

plotted in grey scale, the maximum value corresponding to white and the minimum<br />

to black.<br />

Highly supersonic models, in which kinematic relativistic effects due to high<br />

beam Lorentz factors dominate, have extended over-pressured cocoons. These overpressured<br />

cocoons can help to confine the jets during the early stages <strong>of</strong> their evolution<br />

and even cause their deflection when propagating through non-homogeneous<br />

environments. The cocoon overpressure causes the formation <strong>of</strong> a series <strong>of</strong> oblique<br />

shocks within the beam in which the synchrotron emission is enhanced. In long term


6.6 Structure and Emission <strong>of</strong> Micro–Jets 291<br />

Figure 138: Structure <strong>of</strong> a relativistic electron positron jet. [Scheck et al. 2002]<br />

simulations, the evolution is dominated by a strong deceleration phase during which<br />

large lobes <strong>of</strong> jet material (like the ones observed in many FR II’s (e.g., Cyg A)<br />

start to inflate around the jet’s head. These simulations reproduce some properties<br />

observed in powerful extragalactic radio jets (lobe inflation, hot spot advance speeds<br />

and pressures, deceleration <strong>of</strong> the beam flow along the jet) and can help to constrain<br />

the values <strong>of</strong> basic parameters (such as the particle density and the flow speed) in<br />

the jets <strong>of</strong> real sources.<br />

Research Project: Relativistic jet simulations have to be adapted to real cluster<br />

pr<strong>of</strong>iles. In previous simulations, the jets always propagate in a background medium<br />

<strong>of</strong> constant pressure and density.<br />

6.6 Structure and Emission <strong>of</strong> Micro–Jets<br />

There is a wealth <strong>of</strong> information on the structure <strong>of</strong> jets on the parsec–scale in<br />

galaxies. In the following we only a few aspects relevant for modelling jet acceleration<br />

and collimation.


292 6 MACRO– AND MICRO–JETS DRIVEN BY BLACK HOLES<br />

6.6.1 Core–Jets<br />

Parsec–scale jets can only be detected by means <strong>of</strong> VLBI techniques. Since the jets<br />

are obviously relativistic on this scale, they appear as one–sided (due to relativistic<br />

Doppler beaming), including a stationary core and knots moving away from this<br />

core at superluminal speeds.<br />

Figure 139: A VLBA image <strong>of</strong> 3C 120 at a frequency <strong>of</strong> 22 GHz (Walker et al. 1996).<br />

The scale is in milliarcseconds (mas). (At the distance <strong>of</strong> 3C 120, 1 mas corresponds<br />

to a length <strong>of</strong> 0.7 parsecs = 2.3 light-years.)<br />

3C 120: In Fig. 139 we show the structure <strong>of</strong> the parsec–scale jet <strong>of</strong> the radio<br />

galaxy 3C 120 at 22 GHz (1 mas = 0.7 parsecs for a Hubble constant <strong>of</strong> 65<br />

km/s/Mpc). There is an unresolved stationary feature on the left (eastern) end,<br />

customarily called the core. Bright spots (termed blobs by experts) move down the<br />

jet. Of particular interest is a bright blob that emerges from the core after a major<br />

outburst in brightness (flux density) in early 1998 (Fig. 149). The speeds <strong>of</strong> the<br />

superluminally moving components range from 3.5 to 8.3 times the speed <strong>of</strong> light.<br />

The magnetic field behavior is complex, changing with both position and time. The


6.6 Structure and Emission <strong>of</strong> Micro–Jets 293<br />

core, however, is usually unpolarized. 2<br />

6.6.2 Emission Properties<br />

As the brightest and nearest quasar (z = 0.158), 3C 273 is an ideal laboratory to<br />

study emission processes for jets in quasars. This source displays significant flux<br />

variations, has a well measured wide band spectral energy distribution and has<br />

a relativistic jet emanating from the central part in a galaxy. This collimated jet<br />

structure extends up to 150 kpc from the core, depending on the unknown inclination<br />

towards the observer. 3C 273 is classified as a blazar – though I find this notion<br />

somewhat misleading – and is also a prominent gamma–ray source.<br />

3C 273 and the hard X–ray emission: Recently, the results <strong>of</strong> a long–term<br />

monitoring <strong>of</strong> 3C 273 with RXTE have been published (Kataoka et al. 2002). This<br />

covers the longest observation period in the hard X–rays (about 835 ksec, Fig. 140).<br />

Most <strong>of</strong> the 3C 273 photon spectra observed between 1996 and 2000 can be fitted<br />

Figure 140: RXTE light curves <strong>of</strong> 3C 273 in the hard X–rays (Katakoa et al. 2002).<br />

The data are binned in 10 days intervals.<br />

by a power–law spectrum with spectral index αX = 1.6 ± 0.1. This is consistent<br />

with previously published data. Multifrequency spectra provide information for<br />

physical quantities relevant to the jet physics, e.g. the magnetic field, the size <strong>of</strong><br />

the emission region, the maximum energy and the density <strong>of</strong> relativistic electrons.<br />

2 For a movie see http://www.bu.edu/blazars/3C120.html


294 6 MACRO– AND MICRO–JETS DRIVEN BY BLACK HOLES<br />

von Montigny et al. (1997) examined 3 models to reproduce the multifrequency<br />

spectrum <strong>of</strong> 3C 273: (i) the synchrotron–self Compton model (SSC model); (ii) the<br />

external radiation Compton model (ERC model; Sikora, Begelman and Rees 1994),<br />

and, (iii) the proton induced cascade model (PIC model; Mannheim & Biermann<br />

1992). At that time, the authors concluded that the data were still insufficient to<br />

discriminate between the models. In Fig. 141, a simple one–zone model is shown to<br />

explain the spectrum. In this simple model, the radiation is due to a homogeneous<br />

jet component moving with a bulk Lorentz factor ΓBLK at an angle <strong>of</strong> line <strong>of</strong> sight<br />

Θ 1/ΓBLK. It is assumed that the peak emission <strong>of</strong> the low–energy synchrotrn<br />

component (LE) and the high energy inverse Compton components (HE) arise from<br />

the same electron population with Lorentz factor γp. The peak frequencies <strong>of</strong> LE<br />

and HE are related by<br />

νHE,p = 4<br />

(706)<br />

3 γ2 pνLE,p<br />

with νLE,p = 10 13.5 Hz, and νHE,p = 10 20 Hz. This requires a γp = 2 × 10 3 . The<br />

Figure 141: Global spectrum for 3C 273 with X–rays as explained by SSC–emission<br />

(Katakoa et al. 2002).<br />

synchrotron peak frequency νLE,p is given by<br />

νLE,p = 3.7 × 10 6 B γ 2 p<br />

D<br />

Hz (707)<br />

1 + z


6.6 Structure and Emission <strong>of</strong> Micro–Jets 295<br />

where B is the magnetic field strength in Gauss and D = Γ −1<br />

B (1 − β cos Θ) −1 ΓB<br />

is the relativistic Doppler factor. From the above numbers we get then a magnetic<br />

field strength<br />

and a corresponding energy density<br />

B 0.4 × 10<br />

ΓB<br />

UB 6.7 × 10 −3<br />

<br />

10<br />

ΓB<br />

2<br />

G (708)<br />

erg/cm 3 . (709)<br />

The ratio <strong>of</strong> the synchrotron luminosity and the inverse Compton luminosity<br />

Figure 142: Global IR spectra for 3C 273 and similar quasars with possible dust<br />

emission in the mid–infrared (Meisenheimer et al. 2001).<br />

LSSC<br />

LS<br />

= US<br />

UB<br />

(710)


296 6 MACRO– AND MICRO–JETS DRIVEN BY BLACK HOLES<br />

is independent <strong>of</strong> the bulk motion. US is the energy density in the synchrotron<br />

photons measured in the comoving frame <strong>of</strong> the jet. With LHE = 10 47.1 erg/s and<br />

LS = 10 46.8 erg/s we obtain<br />

US 0.02<br />

2 10<br />

erg/cm<br />

D<br />

3<br />

Assuming a spherical geometry for the emission region <strong>of</strong> radius R<br />

providing a radius<br />

(711)<br />

LS = 4πR 2 cD 4 US erg/s , (712)<br />

R 2 × 10 16 cm 10<br />

. (713)<br />

D<br />

The SSC spectrum is then calculated self–consistently from these numbers for an<br />

electron distribution <strong>of</strong> the form N(γ) ∝ γ−2 exp(−γ/γp). This would indicate that<br />

the inverse Compton emission is radiated in an extremely compact region near the<br />

collimation region. The spectrum in the radio band can usually not be fitted by<br />

simple one–zone models, since this low–frequency emission occurs on much larger<br />

scales in the jet.<br />

An ERC model can however also fit the observed spectra. Part <strong>of</strong> the mid–IR<br />

and far–IR radiation is in fact emitted by warm dust in a dusty torus located on the<br />

parsec–scale in the center <strong>of</strong> the bulge <strong>of</strong> the host galaxy <strong>of</strong> 3C 273 (Meisenheimer et<br />

al. 2001, Fig. 142). This is a common feature for strong radio galaxies (Cyg A e.g.)<br />

and radio–loud quasars. This dust emission zone can roughly be approximated by a<br />

shell <strong>of</strong> radius RD 10 pc, so that the energy density inside the shell as measured<br />

in the comoving frame <strong>of</strong> the jet is<br />

UD = LD<br />

4πR 2 Dc Γ2 BLK 10 −3 Γ 2 B erg/cm 3 ξDLUV X<br />

10 47 erg/s<br />

10 pc<br />

RD<br />

2<br />

. (714)<br />

The factor ξD 0.5 represents the fraction <strong>of</strong> the UV–bump luminosity reprocessed<br />

by dust emission. This factor is quite high in Quasars. In particular, all photons<br />

directed inwards towards the central source are blue–shifted, and enhanced by a<br />

factor ΓB with respect to the comoving frame <strong>of</strong> the jet. On the parsec–scale,<br />

this external photon density is the dominant source for inverse Compton<br />

scattering. With such an ERC model the frequencies are now interrelated as<br />

νHE,e = 4<br />

3 γ2 eΓ 2 B νIR . (715)<br />

The dust spectrum also peaks at νIR 10 13.5 Hz (warm dust). For ΓJ 10, this<br />

only requires a γe 100. This is the typical Lorentz factor for the electrons in an


6.6 Structure and Emission <strong>of</strong> Micro–Jets 297<br />

ion jet and corresponds to the minimal electron Lorentz factor required by global<br />

spectra (Ghisellini et al. 2001).<br />

Relativistic electrons are a natural consequence in relativistic jets. The plasma<br />

in such jets is automatically heated to high temperatures by internal shocks (see<br />

previous sections). The ion component essentially stays hot and can only cool over<br />

adiabatic expansion. The electrons stay probably extremely hot since the jet material<br />

passes through various shocks. In quasars, inverse Compton cooling is probably<br />

the fastest mechanism on the parsec–scale with a cooling time given by<br />

tIC =<br />

mec 2<br />

(4/3) σT cUphγe<br />

10 yrs (716)<br />

for γe = 1000 and Uph 10 −3 erg/cm 3 . Without reheating, the electrons would cool<br />

down to a mean Lorentz factor γmin 50−100, when traversing the dust shell. The<br />

heat carried by the electrons is therefore a rough measure for the inverse Compton<br />

luminosity<br />

LIC Lheat,e = ˙ MBc 2 < γe > (me/mp) < LB . (717)<br />

The magnetic field structure can essentially be probed by means <strong>of</strong> Faraday<br />

rotation<br />

8.1 × 105<br />

RM =<br />

< γe > 2<br />

<br />

neB dl[pc] rad m −2 , (718)<br />

where the electron density ne is given in units <strong>of</strong> particles per cc and < γe ><br />

is the mean energy <strong>of</strong> the relativistic electrons. The classical expression <strong>of</strong> the<br />

rotation measure has to be corrected by the square <strong>of</strong> the Lorentz factor for the<br />

electrons, since the rotation measure is proportional to the plasma frequency and the<br />

cyclotron frequency. The parsec–scale jet <strong>of</strong> 3C 273 has a typical rotation measure<br />

RM 300 rad m−2 . This indicates that the internal Faraday rotation is negligible.<br />

It is probably produced by the electrons in the ISM <strong>of</strong> the galactic nucleus.<br />

NGC 4261 and the state <strong>of</strong> the disk at 0.1 pc: The structure <strong>of</strong> the inner<br />

disk in nearby radio galaxies can be probed by synchrotron emission from the jets<br />

over free–free absorption. The nearby low–luminosity FRI–galaxy NGC 4261 (3C<br />

270, Fig. 143) is a good candidate for detecting free–free absorption by ionized gas<br />

in the inner disk [9, 24]. The galaxy is known to contain a Black Hole with a mass <strong>of</strong><br />

MH = 7×10 8 M⊙ (Ferrarese, Ford and Jaffe 1996), a nearly edge–on disk <strong>of</strong> gas and<br />

dust with a diameter <strong>of</strong> about 100 pc. The large–scale symmetric radio structure<br />

implies that the radio axis is close to the plane <strong>of</strong> the sky. First VLBA observations<br />

revealed a parsec-scale radio jet and counter–jet aligned with the kiloparsec–scale<br />

jet [9]. The opening angle is less than 20 degrees within 0.2 pc from the core and<br />

less than 5 degrees within 0.8 pc. The free–free optical depth is given by<br />

τff = 9.8 × 10 −3 Ln 2 e T −3/2 ν −2 (17.7 + ln(T 3/2 /ν)) , (719)


298 6 MACRO– AND MICRO–JETS DRIVEN BY BLACK HOLES<br />

Figure 143: The radio galaxy NGC 4261 and its nuclear disk (HST, top panel) with<br />

a VLBA image at 8.4 GHz (bottom panel, Jones et al. 2001). At a distance <strong>of</strong> 40<br />

Mpc, 1 mas corresponds to 0.2 pc. A 700 Mio solar mass Black Hole in the center<br />

generates the jets in this radio galaxy.


6.6 Structure and Emission <strong>of</strong> Micro–Jets 299<br />

where the path length L is in cm, and the electron density in cm−3 . From this we<br />

can estimate the electron density<br />

<br />

ne 3.4 τ/L νT 3/4<br />

(720)<br />

For a path–length L 0.3 pc, τff 3 and a temperature T 10 4 K, this corresponds<br />

to an electron density ne 3 × 10 4 cm −3 . Since the temperature decays in a<br />

Figure 144: Disk geometry in the center <strong>of</strong> NGC 4261 [Jones et al. 2001; Wehrle et<br />

al. 2002].<br />

SAD very rapidly with increasing radius, accretion disks around supermassive Black<br />

Holes are expected to have a partial ionization zone, as in Galactic binaries, and<br />

therefore to be subject <strong>of</strong> a similar thermal instability. This zone forms at a distance<br />

<strong>of</strong> a few hundred Schwarzschild radii from the center (Lin and Shields 1986; Clarke


300 6 MACRO– AND MICRO–JETS DRIVEN BY BLACK HOLES<br />

1989). Depending on the viscosity, the instability can develop in a very narrow<br />

unstable zone and propagate over the entire disk resulting in this way in a large–<br />

amplitude outburst on time–scales <strong>of</strong> the order <strong>of</strong> a few 10 5 years (Siemiginowska et<br />

al. 1996).<br />

In standard disk models one finds a local equilibrium relationship between a<br />

steady state accretion rate ˙ M and the surface density Σ for given mass, viscosity<br />

parameter α and radius R. This relationship has a characteristic S–shape for an<br />

optically thick, geometrically thin disk. Three characteristic regions on the S–curve<br />

describe different physical conditions. The lower branch is thermally stable, cool,<br />

and neutral hydrogen dominates the chemical composition, molecules also contribute<br />

to the opacity. A disk on the upper branch is also thermally stable, but is hot, and<br />

hydrogen is fully ionized. Bound–free transitions in heavy metals, free–free transitions<br />

and electron scattering determine the opacity. The middle branch corresponds<br />

to a partially ionized disk that is thermally unstable due to the rapid increase <strong>of</strong> the<br />

opacity with temperature. The instability strip is located at an effective temperature<br />

<strong>of</strong> TA 4000 K, and its radius RA can be derived from the local values <strong>of</strong> ΣA<br />

and TA (Siemiginowska et al. 1996)<br />

RA 10 3 Rg M −0.6<br />

H,8<br />

˙M 0.4 α −0.05<br />

−1 . (721)<br />

This corresponds to a radius <strong>of</strong> about 100 Schwarzschild radii in the core <strong>of</strong> NGC<br />

4261. Local accretion occurs then on the viscous time–scale, tvisc R/v r , or<br />

tvisc 2 × 10 5 yr α −0.8<br />

−1 M 1/4<br />

8<br />

˙M −0.3 R 1.25<br />

16 . (722)<br />

The disk beyond RA is optically thin (Clarke 1988). All the energy produced<br />

dissipatively is radiated away by free–free emission. Under these conditions, the<br />

ionisation instability requires the disk to be either in an optically thin hot state<br />

or an optically thick cool state. These values are consistent with the particle<br />

densities obtained for a standard accretion disk at a distance <strong>of</strong> 10 4 gravitational<br />

radii (Camenzind 1997),<br />

n 24 cm −3 α −1 ˙m −2 M −1<br />

H,9<br />

R<br />

10 4 Rg<br />

−3/2<br />

. (723)<br />

For a turbulence parameter α 0.1 and a relative mass accretion rate ˙m =<br />

˙M/ ˙ MEd 0.01 this gives a particle density n 10 6 cm −3 with an ionisation fraction<br />

<strong>of</strong> about 10%. The temperature <strong>of</strong> the disk at this position is in fact about 10000 K<br />

TD(R) 10 4 K α −1/4 M −1/4<br />

H,9<br />

R<br />

10 4 Rg<br />

−3/8<br />

. (724)<br />

This absorption through the disk produces a gap in the low–frequency emission<br />

detected from the counter–jet (Fig. 143, Jones et al. 2001).


6.7 Formation <strong>of</strong> Micro–Jets 301<br />

6.7 Formation <strong>of</strong> Micro–Jets<br />

In accretion and jet production theory the principal parameters determining the<br />

appearance and behavior <strong>of</strong> the system are the black hole mass MH, the mass accretion<br />

rate ˙ M, and the black hole angular momentum J, expressed in dimensionless<br />

form as m9 = MH/10 9 M⊙ (where M⊙ represents one solar mass), ˙m = ˙ M/ ˙ MEdd<br />

(where ˙ MEdd = 4GMH/ɛHκesc = 22 M⊙ yr −1 m9 is the accretion rate that produces<br />

one Eddington luminosity for an efficiency ɛH = 0.1 and electron scattering opacity<br />

κes), and j = J/Jmax (where Jmax = GM 2 H/c is the angular momentum <strong>of</strong> a maximal<br />

Kerr black hole). For most AGN and quasar models typical ranges <strong>of</strong> the parameters<br />

are 10 −3 < m9 < 10, 10 −5 < ˙m < 1, and 0 < j < 1. While all parameters will affect<br />

the properties <strong>of</strong> an AGN to a certain extent, the purpose here is to identify the<br />

principal observable effects <strong>of</strong> each.<br />

The accretion paradigm states that most, and perhaps all, AGN are powered<br />

by accretion onto a supermassive black hole. Within this model ˙m plays the most<br />

important role, determining the emission properties, and therefore the appearance,<br />

<strong>of</strong> the central source. Objects with high accretion rate ( ˙m > 0.1) appear as an “optical”<br />

quasar (<strong>of</strong> course, equally bright, if not brighter, in X-rays as well), while low<br />

sub-Eddington accretion ( ˙m ≤ 10 −2 ) produces a weak “radio” core with substantially<br />

less optical emission. A zero accretion rate produces a “dead” quasar - a black<br />

hole detectable only through its gravitational influence on the galactic nucleus. For<br />

a given level, the black hole mass determines mainly the luminosity scaling.<br />

The spin paradigm states that, to first order, it is the normalized black hole<br />

angular momentum j that determines whether or not a strong radio jet is produced<br />

(Wilson & Colbert 1995, Blandford 1999). If correct, this hypothesis has significant<br />

implications for how we should view the jets and lobes in radio sources: the jet radio<br />

and kinetic energy comes directly from the rotational energy <strong>of</strong> a (perhaps formerly)<br />

spinning black hole. Radio sources are not powered (directly) by accretion.<br />

There is significant theoretical basis for this paradigm as well. Several models <strong>of</strong><br />

relativistic jet formation indicate that the jet power should increase as the square<br />

<strong>of</strong> the black hole angular momentum<br />

Ljet 10 44 erg s −1<br />

2<br />

BH<br />

m9j<br />

1000 G<br />

2 . (725)<br />

where BH is the strength <strong>of</strong> the poloidal (vertical/radial) magnetic field threading<br />

the ergospheric and horizon region <strong>of</strong> the rotating hole. In this model rotational<br />

energy is extracted via a Penrose-like process: the frame-dragged accretion disk is<br />

coupled to plasma above and outside the ergosphere via the poloidal magnetic field;<br />

some plasma is pinched and accelerated upward while some disk material is diverted<br />

into negative energy (retrograde) orbits inside the ergosphere, removing some <strong>of</strong><br />

the hole’s rotational energy. The key parameter determining the efficiency <strong>of</strong> this


302 6 MACRO– AND MICRO–JETS DRIVEN BY BLACK HOLES<br />

process is the strength <strong>of</strong> the poloidal magnetic field. The standard approach (e.g.,<br />

Moderski & Sikora 1996) to estimating BH is to set it equal to Bφ, the dominant<br />

azimuthal magnetic field component given by the disk structure equations, yielding<br />

and<br />

Ljet,B = 2 × 10 45 erg s −1 m9 ˙m−2 j 2<br />

(726)<br />

Ljet,A = 3 × 10 49 erg s −1 m 0.8<br />

9 ˙m−1 j 2 . (727)<br />

for Class B (radio galaxy/ADAF) and Class A (quasar/standard disk) objects, respectively.<br />

Note that, while the jet is not accretion-powered in this model, the<br />

efficiency <strong>of</strong> extraction is still essentially linear in ˙m.<br />

Basic physics <strong>of</strong> MHD acceleration and collimation: Though many models<br />

have been proposed to generate jets, the magnetohydrodynamic (MHD) model is still<br />

the leading one (Fig. 145). In this picture, the plasma is simplified in terms <strong>of</strong> a one–<br />

component approximation and the conductivity is assumed to be very high so that<br />

electric fields in the plasma are shorted out. In a first approximation, plasma is not<br />

allowed to cross field lines, it can only flow parallel to them. In regions where the field<br />

is weak or the plasma is dense, the rotating field will be bent backward. To accelerate<br />

and collimate a jet with magnetic fields, all that is needed is a gravitating body to<br />

collect the material which should be ejected, a poloidal magnetic field threading that<br />

material, and some differential rotation, which produces a helical structure in the<br />

field. This rotating field structure drives the confined plasma upwards and outwards<br />

along the field lines. As this twist propagates outwards, the toroidal field pinches<br />

the plasma towards the rotational axis. Depending on the relative importance <strong>of</strong><br />

the magnetic field, the plasma density and rotation, a variety <strong>of</strong> configurations are<br />

possible from spherical winds, slowly collimated and highly collimated outflows.<br />

Poloidal magnetic field strengths are estimated from equation disk pressure, but<br />

(H/R) is <strong>of</strong> order unity for the low cases, and also for the high Kerr case due to<br />

Lens-Thirring bloating <strong>of</strong> the inner disk. Otherwise (H/R) is calculated from the<br />

electron scattering/gas pressure disk model <strong>of</strong> Shakura & Sunyaev (1973), and disk<br />

field strengths are computed from that paper or from Narayan et al. (1998), as<br />

appropriate. The logarithms <strong>of</strong> the resulting poloidal field strengths, and corresponding<br />

jet powers, are represented as field line and jet arrow widths. In the Kerr<br />

cases, the inner disk magnetic field is significantly enhanced over the Schwarzschild<br />

cases, due in part to the smaller last stable orbit (flux conservation) and in part to<br />

the large (H/R) <strong>of</strong> the bloated disks. The high accretion rate, Schwarzschild case has<br />

the smallest field - and the weakest jet - because the disk is thin, the last stable orbit<br />

is relatively large, and the Keplerian rotation rate <strong>of</strong> the field there is much smaller<br />

than it would be in a Kerr hole ergosphere. Enhancement <strong>of</strong> the poloidal field due<br />

to the buoyancy process suggested by Krolik (1999) is ignored here because we find


6.7 Formation <strong>of</strong> Micro–Jets 303<br />

Figure 145: Essential elements for magnetic jet formation. A torus <strong>of</strong> hot gas is<br />

initially magnetized by vertical magnetic fields. The differential rotation <strong>of</strong> the<br />

torus drags the field lines in the azimuthal direction, angular momentum is thereby<br />

carried away from the disk material which allows the material to accrete inwards<br />

towards the horizon. This exerts a torque on the external field with a kind <strong>of</strong><br />

torsional Alfvèn wave. This carries away angular momentum and energy from the<br />

system. [Meier et al., Science 291 (2001)].


304 6 MACRO– AND MICRO–JETS DRIVEN BY BLACK HOLES<br />

it not to be a factor in the simulations discussed below. If it were important, then<br />

the grand scheme proposed here would have to be re-evaluated, as the effect could<br />

produce strong jets (up to the accretion luminosity in power) even in the plunging<br />

region <strong>of</strong> Schwarzschild holes. Then even the latter would be expected to be radio<br />

loud as well (Ljet 10 43−46 erg s −1 ).<br />

6.7.1 The Stationary MHD Model<br />

For BHs the above processes must be modelled with relativistic MHD. The time–<br />

dependent analysis <strong>of</strong> this theory is however not yet fully developed. One can get<br />

some insight from considerations <strong>of</strong> the plasma confinement analogue.<br />

Plasma flow near the horizon is presently not observable due to the limited spatial<br />

resolution <strong>of</strong> VLBI. So we cannot test directly the jet formation scenarios. The only<br />

viable model for the generation <strong>of</strong> relativistic jets is however MHD models, where<br />

collimation occurs on scales <strong>of</strong> a few hundred Schwarzschild radii. Such collimated<br />

plasma beams will have therefore a kind <strong>of</strong> conical structure on scales <strong>of</strong> a few<br />

thousand Schwarzschild radii (fraction <strong>of</strong> parsecs) to scales <strong>of</strong> a few hundred parsecs.<br />

We can describe these plasma beams as a stationary equilibrium between various<br />

forces (Appl & Camenzind [1])<br />

κ B2 p<br />

4π (1 − M 2 − x 2 ) = (1 − x 2 B<br />

)∇⊥<br />

2 p B<br />

+ ∇⊥<br />

8π 2 φ<br />

+ ∇⊥P<br />

8π<br />

− B2 pΩF 4πc2 ∇⊥(Ω F R 2 <br />

2 µnj I2<br />

) + −<br />

R3 4πR3 <br />

(−∇⊥R) . (728)<br />

Pressure gradients acting perpendicularly to the magnetic surfaces are in equilibrium<br />

with electric forces, centrifugal forces, pinch forces generated by the poloidal currents<br />

(I), as well as with curvature forces expressed on the left hand side. This is the<br />

correct generalisation <strong>of</strong> well known pressure forces in Newtonian plasma pinches.<br />

This equilibrium assumes a particualr simple form for cylindrical configurations,<br />

κ = 0<br />

(1 − x 2 ) d B<br />

dR<br />

2 p<br />

8π<br />

d B<br />

+<br />

dR<br />

2 φ<br />

8π<br />

− B2 pR<br />

2πR2 −<br />

L<br />

µnj 2<br />

I2<br />

−<br />

R3 4πR3 <br />

= 0 . (729)<br />

This can be written for the dimensionless variables, x = R/RL, y = B 2 p/B 2 0 and<br />

I = RBφ → I/RLB0 as<br />

(1 − x 2 ) dy 1<br />

− 4xy +<br />

dx x2 dI2 8π<br />

+<br />

dx B2 0<br />

dP<br />

dx<br />

− 8πµn<br />

B 2 0<br />

j2 (cRL) 2 = 0 . (730)<br />

x3


6.7 Formation <strong>of</strong> Micro–Jets 305<br />

This is an extension <strong>of</strong> the force–free equation discussed by [1]. Pressure and tension<br />

due to the toroidal field are combined into one expression for the variation <strong>of</strong> the<br />

current, or the pinch force. The electric field also contributes a positive pinch<br />

force, −x 2 dy/dx, which is opposite to the usual pressure force. These two forces<br />

can therefore compensate for all pressure forces and lead to a pinch equilibrium.<br />

Outside the light cylinder, R > RL the current assumes the simple expression<br />

I = −Ω F (R 2 Bp) 1<br />

βJ<br />

(731)<br />

and the specific angular momentum j <strong>of</strong> the plasma stays constant, i.e. the rotational<br />

velocity <strong>of</strong> the plasma decreases as 1/R. This shows that the centrifugal force has a<br />

negligible influence on the plasma equilibrium outside the light cylinder. Under this<br />

condition we get a core–halo structure for the distribution <strong>of</strong> the poloidal magnetic<br />

field<br />

Bp <br />

with the core radius Rc = ΓJβJRL and<br />

Bφ − B0<br />

βJ<br />

B0<br />

1 + R 2 /R 2 c<br />

R/RL<br />

1 + R 2 /R 2 L<br />

(732)<br />

. (733)<br />

The beam radius RJ is probably larger than the core radius and given by pressure<br />

equilibrium with external pressure. The plasma in the beam is rotating with nearly<br />

the speed <strong>of</strong> light at the light cylinder.<br />

The total electron density in the beam is<br />

ne =<br />

˙MJ<br />

πR 2 Jmpc<br />

100 cm−3<br />

˙MJ<br />

0.1 M⊙ yr −1<br />

3 × 10 9 M⊙<br />

MH<br />

2 1000 2<br />

Rg<br />

The mass–flow in the beam is also related to the total jet kinetic energy<br />

LJ = ˙ MJc 2 (ΓJ − 1) 5 × 10 46 erg s −1<br />

˙MJ<br />

0.1 M⊙ yr −1<br />

RJ<br />

. (734)<br />

ΓJ<br />

. (735)<br />

10<br />

This is a reasonable number for the outflow rate <strong>of</strong> bright quasars, since about a<br />

few percent <strong>of</strong> the inflow is probably converted to outflows. The outflow rate in M<br />

87 is estimated to be about a factor 100 lower than this number (Camenzind 1999).<br />

These numbers tell us that the Alfven speed in the beam is about the speed <strong>of</strong> light<br />

for<br />

VA c B0/0.3 G<br />

. (736)<br />

n/10 cm−3 So one has to use the correct expression for the Alfven speed.


306 6 MACRO– AND MICRO–JETS DRIVEN BY BLACK HOLES<br />

Physical Parameter Symbol 3C 273 M 87<br />

Black Hole mass MH [M⊙] 3 × 10 9 3 × 10 9<br />

Gravitational radius Rg [AU] 30 30<br />

Black Hole spin a 0.9 MH 0.8<br />

Light cylinder RL/Rg 20 20<br />

Bulk Lorentz factor ΓJ 10 6<br />

Inclination ΘJ [deg] 8 20<br />

Jet core radius Rc = ΓJRL 0.1 lyr 0.01 lyr<br />

Beam Radius RJ 0.1 pc 0.03 pc<br />

Ion number density nJ 6 cm −3<br />

Ion temperature Ti 100 MeV 100 MeV<br />

Beam sonic Mach number M = ΓJβJ/cS 30<br />

Mass flux ˙ MJ 0.1 M⊙ yr −1<br />

Magnetic field B0 0.1 Gauss 0.03 G<br />

Magnetic flux ΨJ B0R 2 c 10 33 G cm 2<br />

Total current IJ cRcB0 10 18 Amp<br />

Alfvèn Mach number MA = ΓJβJ/VA 10 10<br />

Fast magnetos. M number MF M 5 5<br />

Mean e − Lorentz factor γe 100 50<br />

Cut<strong>of</strong>f e − Lorentz factor γp 2000 2000<br />

Table 9: Parameters <strong>of</strong> the MHD parsec–scale jet in 3C 273 and M 87. The Black<br />

Hole mass in 3C 273 is estimated from the similarity <strong>of</strong> its host galaxy with M 87,<br />

the mass in M 87 has been derived from the HST disk. The spin parameters are<br />

essentially unknown.


6.7 Formation <strong>of</strong> Micro–Jets 307<br />

VLBI jets are extreme plasma pipes, where all communication occurs<br />

over the speed <strong>of</strong> light – the Alfven speed and the fast magnetosonic speed<br />

are relativistic. The plasma in these pipes is also exotic: it is permanently heated<br />

to high temperatures <strong>of</strong> the order <strong>of</strong> 10 12 K by means <strong>of</strong> internal shocks. The ions<br />

are probably still non–relativistic, while the electrons always assume relativistic<br />

temperatures with < γe > 100, which are then subject to acceleration to even<br />

higher energies by various mechanisms.<br />

6.7.2 The Collimation Zone<br />

I am still convinced that jets in quasars are launched by complete MHD processes<br />

and that this will be modelled some day within General Relativistic MHD. The<br />

following elements should be included<br />

• a dipolar magnetosphere that closes radially towards the horizon;<br />

• hot plasma is launched from the innermost part (probably within the ergosphere<br />

region) in radial direction;<br />

• Along the axis, higly relativsitic plasma is injected from a polar gap. This<br />

plasma is energetically not important, but it carries the closure currents.<br />

• The gap is feeded by a pair plasma created from photoproduction in the ion<br />

torus near the horizon.<br />

• The toroidal field amplified by the frame–dragging effect leads to additional<br />

J × B–forces near the ergosphere.<br />

Structure <strong>of</strong> the magnetosphere: The calculation <strong>of</strong> the magnetosphere in the<br />

stationary approach is very cumbersome (Beskin 1996; Camenzind [6]; Fendt &<br />

Greiner [7]), since the current distribution has to be found self–consistently (this<br />

exercise is in fact better done within a time–dependent approach). The transition<br />

from the highly diffusive disk towards the ideal outflow conditions has however<br />

hampered down all modelling so far. What is urgently needed for this process is a<br />

realisation <strong>of</strong> the GR MHD equations on the background <strong>of</strong> compact objects. The<br />

form <strong>of</strong> the magnetosphere shown in Fig. 146 is a first guess.<br />

The Plasma Outflow: It has been shown that the stationary axisymmetric and<br />

polytropic plasma flow along an axisymmetric magnetic flux tube Ψ(r, θ) can be<br />

described by means <strong>of</strong> an algebraic wind equation (Camenzind [6]; Fendt & Greiner<br />

[7]). The essential quantity which determines the asymptotic plasma flow is the


308 6 MACRO– AND MICRO–JETS DRIVEN BY BLACK HOLES<br />

Figure 146: The Black Hole magnetosphere consist <strong>of</strong> a family <strong>of</strong> nested magnetic<br />

surfaces, and plasma can only flow along the flux surfaces. Currents can however<br />

cross the flux surfaces. Only in the force–free approximation, currents are also force<br />

to flow along the flux surfaces. This is in general violated near the Alfven points.<br />

[Fendt 2001]<br />

magnetissation parameter σ∗ at the injection point<br />

σ∗ =<br />

Φ∗<br />

4πmpIp;∗<br />

(737)<br />

measuring the Poynting–flux in terms <strong>of</strong> the particle flux Ip = √ −gnup, where up<br />

is the poloidal velocity along the flux–tube. The form <strong>of</strong> the flux–tube is expressed<br />

in terms <strong>of</strong> the dimensionless flux–tube function Φ, Φ ≡ 1 is a pure monopole<br />

flux surface. The other wind parameters are the total energy E and total angular<br />

momentum L carried by the plasma. At the magnetosonic points the wind equation<br />

becomes singular, and this fixes two parameters (for more details, see Lecture by<br />

Beskin). In terms <strong>of</strong> astrophysical quantities, the magnetization parameter is given


6.7 Formation <strong>of</strong> Micro–Jets 309<br />

as (Camenzind [6])<br />

σ∗ <br />

Ψ 2 ∗<br />

2c ˙ MJ(Ψ)R 2 L<br />

= B2 p,∗R 2 L<br />

c ˙ MJ(Ψ)<br />

4<br />

R∗<br />

RL<br />

. (738)<br />

˙MJ(Ψ) denotes the jet mass flux carried by the flux surface Ψ, and RL = c/Ω F is the<br />

light cylinder radius <strong>of</strong> the flux surface. With an idea on the maximal flux provided<br />

by the disk near the horizon,<br />

Ψ∗ ˙ MaccR 2 ∗<br />

αT<br />

<br />

Rg<br />

R∗<br />

we can estimate the value <strong>of</strong> the magnetization<br />

−1<br />

H<br />

R<br />

(739)<br />

σ∗ (740)<br />

The outflow starts hot, cS,∗ 0.3c and is then continuously accelerated, until collimation<br />

is reached (Fig. 147). thereby, the particle number decreases rapidly and<br />

also the temperature decreases adiabatically.<br />

6.7.3 Nondiffusive Relativistic MHD Approach<br />

Recent work by Koide, Shibata and Kudoh on general relativistic simulations <strong>of</strong> the<br />

accretion <strong>of</strong> magnetized material by black holes is beginning to provide a clearer<br />

picture <strong>of</strong> the development and evolution <strong>of</strong> a black hole magnetosphere and the<br />

resulting jet. Initial indications are that rapid rotation <strong>of</strong> the black hole and rapid<br />

infall <strong>of</strong> the magnetized plasma into this rotating spacetime both contribute to<br />

powerful, collimated, relativistic jet outflows. We find no evidence for buoyant<br />

poloidal field enhancement in the plunging region and, therefore, no reason to expect<br />

Schwarzschild holes to have a jet any more powerful than that estimated in Figure<br />

148. Figure 148 shows the initial and final state <strong>of</strong> a Kerr (j = 0.95) black hole<br />

simulation with a weak (VA = Bp/ √ 4πρdisk = 0.01c) magnetic field and an inner<br />

edge at Rin = 4.5GMH/c 2 . For more details see Koide et al. (1999a). In this first<br />

simulation the disk was non-rotating, and so began to free-fall into the ergosphere<br />

(at Rergo = 2GMH/c 2 ) as the calculation proceeded. As with other MHD disk<br />

simulations, the field lines were wound up by differential rotation and a hollow jet<br />

<strong>of</strong> material was ejected along the poloidal field lines by the J × B forces; here the<br />

jet velocity was rather relativistic (vJ 0.93c or 2.7). However, as the disk was<br />

initially non-rotating, all the action was due to the differential dragging <strong>of</strong> frames by<br />

the rotating black hole. Little or no jet energy was derived from the binding energy<br />

<strong>of</strong> the accreting material or its Keplerian rotation.<br />

As a control experiment, the same simulation was also run with a Schwarzschild<br />

(j = 0) hole. The collapse developed a splash outflow due to tidal focusing and<br />

shocking <strong>of</strong> the inflowing disk, but no collimated MHD jet occurred.


310 6 MACRO– AND MICRO–JETS DRIVEN BY BLACK HOLES<br />

Figure 147: Poloidal flow velocity along a collimated flux tube (top panel) and<br />

particle density (bottom panel) (Fendt & Greiner 2001). x = R/Rg


6.7 Formation <strong>of</strong> Micro–Jets 311<br />

Figure 148: General relativistic simulation <strong>of</strong> a magnetized flow accreting onto a<br />

Kerr black hole (after Koide et al. 1999a). Left panel shows one quadrant <strong>of</strong><br />

the initial model with the j = 0.95 hole at lower left, freely-falling corona, and<br />

non-rotating disk. The hole rotation axis (Z) is along the left edge. The initial<br />

magnetic field (vertical lines) is weak compared to the matter rest energy density<br />

(VA = 0.01c). Right panel shows final model at t = 130GMH/c 3 . Some <strong>of</strong> the disk<br />

and corona have been accreted into the hole, threading the ergosphere and horizon<br />

with magnetic field lines that develop a significant radial component, and hence an<br />

azimuthal component as well, due to differential frame dragging. The resulting jet<br />

outflow is accelerated by J × B forces to a Lorentz factor <strong>of</strong> 2.7. [Koide et al. 2001]<br />

The authors also studied counter-rotating and co-rotating Keplerian disks with<br />

otherwise similar initial parameters (Koide et al. 1999b). The counter-rotating<br />

case behaved nearly identically to the non-rotating case: because the last stable<br />

retrograde orbit is at Rms = 9GMH/c 2 , the disk began to spiral rapidly into the<br />

ergosphere, ejecting a strong, black-hole-spin-driven MHD jet. On the other hand,<br />

for prograde orbits Rms = GMH/c 2 , so the co-rotating disk was stable, accreting<br />

on a slow secular time scale. At the end <strong>of</strong> the calculation (tmax = 94GMH/c 3 ),<br />

when the simulation had to be stopped because <strong>of</strong> numerical problems, the disk<br />

had not yet accreted into the ergosphere and had produced only a weak MHD jet.<br />

Future evolution <strong>of</strong> this case is still uncertain. The Keplerian Schwarzschild case was<br />

previously reported by Koide et al. (1998) and developed a moderate sub-relativistic<br />

MHD jet similar to the aforementioned simulations <strong>of</strong> magnetized Keplerian flows<br />

around normal stars, plus a pressure-driven splash outflow internal to the MHD jet.<br />

In the Kerr cases, significant magnetic field enhancement over the Schwarzschild


312 6 MACRO– AND MICRO–JETS DRIVEN BY BLACK HOLES<br />

cases occurs due to compression and differential frame dragging. It is this increase<br />

that is responsible for the powerful jets ejected from near the horizon. In the<br />

Schwarzschild cases, on the other hand, - particularly in the Keplerian disk case<br />

in Koide et al. (1998) - we do not see any MHD jet ejected from well inside Rms.<br />

That is, we do not see any increase in jet power that could be attributed to buoyant<br />

enhancement <strong>of</strong> the poloidal field in the plunging region. While there is a jet<br />

ejected from inside the last stable orbit, it is pressure-driven by a focusing shock<br />

that develops in the accretion flow. The magnetically-driven jet that does develop<br />

in the Schwarzschild case emanates from near the last stable orbit as expected, not<br />

well inside it. It is therefore concluded that a powerful MHD jet will be produced<br />

near the horizon if and only if the hole is rotating rapidly.<br />

The spin paradigm even <strong>of</strong>fers an explanation for why present-day giant radio<br />

sources occur only in elliptical galaxies. The e-folding spindown time (Erot/Ljet) is<br />

<strong>of</strong> the order <strong>of</strong> the Hubble–time (Camenzind 1997)<br />

tspindown = JH<br />

dJH/dt<br />

4 MHc 2<br />

B 2 Hr 3 H<br />

rH<br />

c 1011 yr<br />

−2 9<br />

BH 10 M⊙<br />

. (741)<br />

kG MH<br />

As a result, all AGN should be radio quiet at the present epoch, their black holes<br />

having spun down when the universe was very young. In order to continually produce<br />

radio sources up to the present epoch, there must be periodic input <strong>of</strong> significant<br />

amounts <strong>of</strong> angular momentum from accreting stars and gas, or from a merger with<br />

another supermassive black hole (see also Wilson & Colbert 1995). Such activity<br />

is triggered most easily by violent events such as galaxy mergers. Since only elliptical<br />

galaxies undergo significant mergers (the merging process is believed to be<br />

responsible for their elliptical shape), only ellipticals are expected to be giant radio<br />

sources in the present epoch. With little merger activity, spiral galaxies are<br />

expected to be relatively radio quiet. Since merging and non-merging galaxies will<br />

fuel and re-kindle their black hole’s spin in very different ways, bi-modality in the<br />

radio luminosity distribution is to be expected.<br />

6.7.4 Knot Ejection Mechanisms<br />

According to theoretical models (see, for example, computer simulations by Meier,<br />

Koide, et al.), this drives jets <strong>of</strong> ultra-high-energy particles (including electrons) at<br />

speeds <strong>of</strong> about 98% the speed <strong>of</strong> light. (The spiral swirling <strong>of</strong> the jet corresponds to<br />

the expected spiral pattern <strong>of</strong> the magnetic field.) About once every 10 months, an<br />

instability in the accretion disk causes a chunk <strong>of</strong> the inner portion to break <strong>of</strong>f and<br />

fall into the black hole. Although most <strong>of</strong> the matter falls past the event horizon<br />

and disappears, a small amount <strong>of</strong> the mass and energy is injected into the jet. This<br />

causes a bright spot to appear about 0.41 parsecs (1.3 light-years) downstream <strong>of</strong><br />

the black hole. The bright spot appears to be moving at almost 5 times the speed


6.7 Formation <strong>of</strong> Micro–Jets 313<br />

Figure 149: A sequence <strong>of</strong> VLBA images (year <strong>of</strong> observation in decimal form marked<br />

on the left) at a frequency <strong>of</strong> 43 GHz. The scale is in milliarcseconds (mas). The<br />

apparent diameter <strong>of</strong> the moon in the sky is about 1.8 million mas, which means<br />

that the resolution <strong>of</strong> the VLBA is extraordinarily fine. There is a stationary core<br />

at the left. One can see bright spots coming out at speeds <strong>of</strong> about 1.8 mas per<br />

year, or between 4 and 5 times the speed <strong>of</strong> light. (At the distance <strong>of</strong> 3C 120, 1 mas<br />

corresponds to a length <strong>of</strong> 0.7 parsecs = 2.3 light-years.) The diagonal lines pass<br />

through the centers <strong>of</strong> the bright spots.


314 6 MACRO– AND MICRO–JETS DRIVEN BY BLACK HOLES<br />

<strong>of</strong> light. This is an illusion that occurs when a jet moving at slightly less than the<br />

speed <strong>of</strong> light is pointing nearly at us. Why the jet is invisible between the black<br />

hole and 0.41 parsecs is not yet understood. (Marscher thinks that the jet expands<br />

too fast until its pressure falls below that <strong>of</strong> the interstellar medium, which causes<br />

a stationary shock wave to form. The shock wave then energizes electrons and<br />

amplifies the magnetic field, which makes the jet shine from synchrotron radiation<br />

and inverse Compton scattering. But, he admits that the evidence in favor <strong>of</strong> this is<br />

pretty thin.) The top panel in Fig. 150 shows the X-ray flux as a function <strong>of</strong> time<br />

measured with RXTE. The red points correspond to what we call the X-ray dips,<br />

when the flux was low for at least 4 consecutive observations. The blue arrows show<br />

the times when a bright spot first showed up on the VLBA radio images, about 0.1<br />

years after each X-ray dip. The middle panel is the slope <strong>of</strong> the X-ray spectrum;<br />

a low value means that the higher frequency X-rays are more prominent relative to<br />

the lower frequency X-rays. The bottom panel shows the flux at 3 radio frequencies<br />

(green is 14.5 GHz, red is 22 GHz, and blue is 43 GHz) obtained at the University<br />

<strong>of</strong> Michigan Radio Astronomy Observatory and the Metsahovi Radio Observatory<br />

in Finland. The black points on the bottom indicate the flux <strong>of</strong> the radio core as<br />

seen on our 43 GHz VLBA images.<br />

It is apparent from this Figure that each X–ray dip is followed by the appearance<br />

<strong>of</strong> a superluminal knot at the side <strong>of</strong> the radio core. The mean time–delay between<br />

the minimum in the X–ray flux and the appearance <strong>of</strong> the knot ejection is 0.1 ± 0.03<br />

yr. At a typical superluminal apparent speed <strong>of</strong> 4.7 c, a knot moves a distance <strong>of</strong><br />

0.14 pc in 0.1 year, projected on the plane <strong>of</strong> the sky (the real distance is about 0.5<br />

pc for an inclination <strong>of</strong> 20 degrees). The jet probably starts as a broad wind until it<br />

gets collimated at the radio core on the scale <strong>of</strong> a few 100 Schwarzschild radii (see<br />

M 87, Junor, Biretta and Livio, Nature 1999). this would correspond to a minimu<br />

distance <strong>of</strong> 0.3 pc from the X-ray source to the radio core. The core <strong>of</strong> the radio<br />

jet is therefore not coincident with the position <strong>of</strong> the Black Hole. The jet might be<br />

invisible between the site <strong>of</strong> the origin and the radio core far downstream where the<br />

electrons are reaccelerated (e.g. by a recollimation shock).<br />

This correlated radio and X–ray behaviour is similar to the microquasar GRS<br />

1915+105: here, X–ray dips, where the spectrum hardens, are followed by the ejection<br />

<strong>of</strong> a bright superluminal radio knot. This is seen as resulting from an instability<br />

where a piece <strong>of</strong> the inner disk breaks <strong>of</strong>f. Most <strong>of</strong> this matter is accreted towards<br />

the horizon, but a smaller amount is ejected from the inner region and accelerated<br />

by magnetic forces to high velocities. The X–ray flux is due to Compton scattering<br />

in the inner accretion torus. The bulk velocity <strong>of</strong> the jets in 3C 120 and GRS<br />

1915+105 are also similar, about 0.98 c in each case (i.e. a Lorentz factor <strong>of</strong> about<br />

5). The jet in the microquasar is however seen at a much higher inclination <strong>of</strong> 66<br />

degrees. Therefore, the counterjet is also visible.<br />

Nakamura et al. (2001) investigated the wiggled structure <strong>of</strong> AGN radio


6.7 Formation <strong>of</strong> Micro–Jets 315<br />

Figure 150: RXTE light curves <strong>of</strong> 3C 120 in the hard X–rays (top panel, Marscher<br />

et al. 2002). The middle panel is the slope <strong>of</strong> the X-ray spectrum. The bottom<br />

panel shows the flux at 3 radio frequencies (green is 14.5 GHz, red is 22 GHz, and<br />

blue is 43 GHz). The blue arrows show the times when a bright spot first showed<br />

up on the VLBA radio images. [Marscher et al. 2002]


316 6 MACRO– AND MICRO–JETS DRIVEN BY BLACK HOLES<br />

Figure 151: Knot formation in rotating magnetized jets. [Nakamura et al. 2001].<br />

jets by performing 3-dimensional magnetohydrodynamic (MHD) simulations based<br />

on the ‘Sweeping Magnetic-Twist’ model. The correlation between the wiggled<br />

structures <strong>of</strong> AGN radio jets and tails and the distribution <strong>of</strong> magnetic field in them<br />

suggests that the magnetic field plays an essential role, not only in the emission <strong>of</strong><br />

synchrotron radiation, but also in the dynamics <strong>of</strong> the production <strong>of</strong> the AGN–jets–<br />

lobes systems (core, jets, tails, lobes, and hotspots) themselves. In order to produce<br />

such a systematic magnetic configuration, a supply <strong>of</strong> a huge amount <strong>of</strong> energy, in an<br />

organized form, is necessary. The supply <strong>of</strong> this energy must come from AGN core.<br />

The most natural means <strong>of</strong> carrying this energy is in the form <strong>of</strong> the Poynting flux<br />

<strong>of</strong> torsional Alfven wave train (TAWT) produced in the interaction <strong>of</strong> the rotating<br />

accretion disk and the large scale magnetic field brought into it by the gravitational<br />

contraction. The propagation <strong>of</strong> the TAWT can produce a slender jet shape by the<br />

sweeping pinch effect from the initial large scale magnetic field. Further, wiggles <strong>of</strong><br />

the jet can be produced by the MHD processes due to the TAWT. Our numerical<br />

results reveal that the structure <strong>of</strong> the magnetic jet can be distorted due to the<br />

helical kink instability. This results in the formation <strong>of</strong> wiggled structures in the<br />

jets as the TAWT encounters a domain <strong>of</strong> reduced Alfven velocity (either towards


6.7 Formation <strong>of</strong> Micro–Jets 317<br />

the boundary <strong>of</strong> the ‘cavity’ from which the mass contracted to the central core, or<br />

encounting smaller and denser clouds in the domain). The toroidal component <strong>of</strong> the<br />

field accumulates due to the lowered Alfven velocity, and produces a strongly pinched<br />

region, as well as the deformation <strong>of</strong> the jet into a writhed structure. The condition<br />

for this to occur corresponds to the Kruskal–Shafranov criterion <strong>of</strong> the linear case.<br />

If the intrinsic relationship between the wiggled structure <strong>of</strong> the jet shape and the<br />

magnetic field in the jets and tails as proposed in this paper is confirmed, it will<br />

influence the understandings <strong>of</strong> the AGN jets in a vital way.


318 7 THE FIRST BLACK HOLES IN THE UNIVERSE<br />

7 The First Black Holes in the Universe<br />

Recent ground- and space-based observations over the course <strong>of</strong> the last decade have<br />

directly measured the kinematic signature <strong>of</strong> supermassive black holes at the centers<br />

<strong>of</strong> approximately forty nearby galaxies. Even more interesting than this confirmation<br />

<strong>of</strong> the existence <strong>of</strong> these black holes, however, is the fact that black holes have been<br />

found in all <strong>of</strong> the galaxies which have a substantial spheroid component (i.e. they<br />

are ellipticals or are spirals with a substantial bulge) and these black holes are<br />

millions to billions <strong>of</strong> times as massive as our sun. This 100% success rate strongly<br />

suggests that all galaxies have supermassive black holes at their centers. Present<br />

research focuses on how these black holes have grown to their present size.<br />

Black holes are known to form during the death throes <strong>of</strong> certain massive stars.<br />

However, the black holes that form during the end stages <strong>of</strong> stellar evolution are<br />

typically only a few times as massive as the sun, rather than millions to billions <strong>of</strong><br />

times greater. The main focus <strong>of</strong> this section on black hole evolution is how the<br />

supermassive black holes at the centers <strong>of</strong> galaxies grew to their present size.<br />

Most <strong>of</strong> this research is therefore aimed at identifying and understanding different<br />

mechanisms that can ”feed” these black holes, determining which mechanisms are<br />

more or less important, and how long these phases <strong>of</strong> black hole evolution last. If the<br />

black hole at the center <strong>of</strong> a galaxy is growing relatively ”quickly,” which means it<br />

may double in mass in a few tens <strong>of</strong> millions <strong>of</strong> years. In this case, the host galaxy is<br />

said to have a high-luminosity active galactic nucleus (AGN), or quasar, due to the<br />

great amount <strong>of</strong> radiation that is emitted as material falls onto the black hole. At<br />

a more moderate growth rate, the emitted radiation is not as intense, and the host<br />

galaxy would more likely appear to be a low-luminosity AGN, such as a Seyfert or<br />

LINER galaxy. For very low growth rates, no emission from the host galaxy nucleus<br />

may be visible at all, and in this case the galaxy may not be classified as an AGN.<br />

7.1 The Dark Age<br />

After its creation 10 – 15 billion years ago in the mother <strong>of</strong> all explosions, called the<br />

Big Bang, the universe was a homogenous sea <strong>of</strong> cooling gas. Slight fluctuations in<br />

gravity rippled this smoothness, however, and gravity’s force pulled the ripples into<br />

waves <strong>of</strong> matter that condensed into the first objects with separate identities.<br />

The Universe literally entered a dark age about 300,000 years after the big bang,<br />

when the primordial radiation cooled below 3000K and shifted into the infrared.<br />

Unless there were some photon input from (for instance) decaying particles, or string<br />

loops, darkness would have persisted until the first non-linearities developed into<br />

gravitationally-bound systems, whose internal evolution gave rise to stars, or perhaps<br />

to more massive bright objects.<br />

Spectroscopy from the new generation <strong>of</strong> 8 – 10 metre telescopes now comple-


7.1 The Dark Age 319<br />

Figure 152: Dark age and reionization <strong>of</strong> the Universe.


320 7 THE FIRST BLACK HOLES IN THE UNIVERSE<br />

ments the sharp imaging <strong>of</strong> the Hubble Space Telescope (HST); these instruments<br />

are together elucidating the history <strong>of</strong> star formation, galaxies and clustering back,<br />

at least, to redshifts z = 5. Our knowledge <strong>of</strong> these eras is no longer restricted to<br />

‘pathological’ objects such as extreme AGNs - this is one <strong>of</strong> the outstanding astronomical<br />

advances <strong>of</strong> recent years. In addition, quasar spectra (the Lyman forest,<br />

etc) are now observable with much improved resolution and signal-to-noise; they <strong>of</strong>fer<br />

probes <strong>of</strong> the clumping, temperature, and composition <strong>of</strong> diffuse gas on galactic<br />

(and smaller) scales over an equally large redshift range, rather as ice cores enable<br />

geophysicists to probe climatic history.<br />

7.2 The First Stars<br />

Currently, we do not have direct observational constraints on how the first stars<br />

(Population III stars) formed at the end <strong>of</strong> the cosmic dark age. What we know<br />

is how Population I stars form out <strong>of</strong> cold, dense molecular gas. The molecular<br />

clouds are supported against gravity by turbulent velocity fields and are pervaded<br />

by large–scale magnetic fields. Stars tend to form in clusters, ranging from a few<br />

hundred to one million stars. The IMF <strong>of</strong> Pop I stars is observed to have about the<br />

Salpeter form<br />

dN<br />

d ln M ∝ M x , (742)<br />

where x −1.35 for M ≥ 0.5 M⊙. For lower masses it turns over. The lower limit<br />

corresponds roughly to the opacity limit for fragmentation. Therefore, 1 M⊙ is<br />

the typical mass–scale <strong>of</strong> Pop I star formation, in the sense that most <strong>of</strong> the mass<br />

goes into stars with masses close to this value.<br />

How did the first stars form ? The simplest thing we can do is estimate the corresponding<br />

characteristic mass–scale. To investigate the collapse and fragmentation<br />

<strong>of</strong> primordial gas, one has to carry out complicated simulations (Bromm et al. 2002<br />

[4]). In primordial gas clumps, molecular cooling, H2, in a metal–free environment<br />

provides characteristic temperatures <strong>of</strong> 200 K and densities 10 4 cm −3 . Evaluating<br />

the Jeans mass for these characteristic values results in MJ 10 3 M⊙. These<br />

masses can still grow by accretion processes.<br />

Current simulations indicate that the first stars were predominantly<br />

very massive, and therefore rather different from star formation in the<br />

local Universe. These very massive stars would rapidly explode and form seed<br />

Black Holes which can grow by accretion.<br />

7.3 The First Quasars<br />

The brightest quasars at z ≥ 6 are most likely hosted by rare galaxies, more massive<br />

than 10 12 M⊙. They would end up today as the most massive elliptical galaxies.


7.3 The First Quasars 321<br />

Figure 153: Formation <strong>of</strong> compact objects in stellar evolution (adapted from Heger<br />

et al. (2001)). In the mass range from 140 - 250 M⊙, no Black Holes are left over<br />

(pair creation instability).


322 7 THE FIRST BLACK HOLES IN THE UNIVERSE<br />

There are essentially two routes towards the formation <strong>of</strong> supermassive BHs in the<br />

center <strong>of</strong> galaxies:<br />

1. Black Holes with masses <strong>of</strong> a few hundred solar masses, left over from the<br />

collapse <strong>of</strong> very massive stars, could grow by accretion.<br />

2. Black Holes with masses in the range <strong>of</strong> 10 5 − 10 6 M⊙ are formed in a direct<br />

collapse <strong>of</strong> primordial gas clouds at high redshifts. It has been shown that<br />

without a pre–existing central point mass, this direct collapse is difficult by<br />

the negative feedback resulting from star formation in the collapsing cloud<br />

(Loeb & Rasio 1994). The input <strong>of</strong> kinetic energy from supernovae explosions<br />

prevents the gas from assembling in the center <strong>of</strong> the dark matter potential.<br />

If however star formation can be suppressed in a cloud that undergoes overall<br />

collapse, such a negative feedback would not occur.<br />

Figure 154: Formation <strong>of</strong> a cloud at redshift 10.3. The box size is 200 pc. One<br />

compact object has formed in the center with a mass <strong>of</strong> 2.7 Mio solar masses.<br />

[Bromm 2002]


7.3 The First Quasars 323<br />

Bromm & Loeb (2003) [4] have carried out SPH simulations <strong>of</strong> isolated peaks in<br />

the dark matter fluctuations with total masses <strong>of</strong> 10 8 M⊙ that collapse at z 10 3 .<br />

The virial temperature <strong>of</strong> these dwarf galaxies exceeds 10 4 K, which allows collapse<br />

<strong>of</strong> their gas through cooling by atomic hydrogen H2. Such a scenario would include<br />

lower–mass halos that would have collapsed earlier on. Those lower mass systems<br />

would have virial temperatures below 10 4 K, and consequently rely on the presence<br />

<strong>of</strong> molecular hydrogen for cooling. This H2 is however fragile and readily destroyed<br />

by photons in the energy bands <strong>of</strong> 11 – 13.6 ev (Werner bands), just below the<br />

Lyman limit.<br />

What is the future fate <strong>of</strong> the central object ? Once the gas has collapsed to<br />

densities above 10 7 cm −3 and radii les than 0.01 light year, Thomson scattering traps<br />

the photons, and the cooling time becomes now much larger than both the free–fall<br />

time and the viscous time–scales. The gas cloud settles into a radiation–pressure<br />

supported configuration resembling a rotating supermassive star. Fully relativistic<br />

calculations <strong>of</strong> the collapse <strong>of</strong> such stars have recently been done by Baumgarte et al.<br />

(1999) [3]. These objects are expected to collapse towards a supermassive spinning<br />

Black Hole. A substantial fraction <strong>of</strong> the mass is expected to end up in the Black<br />

Hole.<br />

The main question is whether a complete destruction <strong>of</strong> molecular hydrogen<br />

will occcur. The effective suppresion <strong>of</strong> H2 formation will crucially depend on the<br />

presence <strong>of</strong> a stellar–like radiation background. It is therefore likely that stars<br />

precede the first quasars. The above discussion shows that there is only a small<br />

window in redshift where supermassive stars could be formed at redshifts around<br />

10. At later times, the strong UV radiation field will certainly destroy molecular<br />

hydrogen. Another question is whether there is some initial spin in the collapsing<br />

cloud. This would lead to a binary system with a typical separation <strong>of</strong> one parsec.<br />

3 see homepage http://www.ukaff.ac.uk/starcluster/


324 REFERENCES<br />

References<br />

[1] Appl, S., Camenzind, M. 1993, The structure <strong>of</strong> relativistic MHD jets: a<br />

solution to the nonlinear Grad-Shafranov equation, A&A 274, 699.<br />

[2] Barkana, R., Loeb, A. 2001, The First Sources <strong>of</strong> Light and the Rionisation<br />

<strong>of</strong> the Universe, Phys. Reports 349, 125<br />

[3] Baumgarte, T.W., Shapiro, S.L. 1999, , ApJ 526, 941.<br />

[4] Bromm, V., Loeb, A. 2002, Formation <strong>of</strong> the First Supermassive Black<br />

Holes, ApJ<br />

[5] Bromm, V., Loeb, A. 2003, The first Sources <strong>of</strong> Light, astro–ph/0301406<br />

[6] Camenzind, M. 1998, Magnetohydrodynamics <strong>of</strong> Rtoating Black Holes, in<br />

Relativistic Astrophysics, eds. H. Riffert et al., Vieweg, pp 82–117.<br />

[7] Fendt, C., Greiner, J. 2001, Magnetically driven superluminal motion<br />

from rotating black holes. Solution <strong>of</strong> the magnetic wind equation in Kerr<br />

metric, A&A 369, 308.<br />

[8] Gopal–Krishna, Wiita, P. 2001, The Fanar<strong>of</strong>f–Riley Transition and the<br />

Optical Luminosity <strong>of</strong> the Host Elliptical Galaxy, A&A 373, 100.<br />

[9] Jones, et al. 2001, In the Shadow <strong>of</strong> the Accretion Disk: Higher Resolution<br />

Imaging <strong>of</strong> the Central Parsec in NGC 4261, ApJ 553, 968.<br />

[10] Kataoka, Jun et al. 2002, RXTE observations <strong>of</strong> 3C 273 between 1996<br />

and 2000: variability time-scale and jet power, MNRAS 336, 932.<br />

[11] Krause, M. 2002, On the interaction <strong>of</strong> jets with the dense medium <strong>of</strong> the<br />

early universe, PhD thesis, Univ. Heidelberg.<br />

[12] Krause, M. 2003, Very light jets. I. Axisymmetric parameter study and<br />

analytic approximation, A&A 398, 113.<br />

[13] Laing, R.A. et al. 1999, MNRAS 306, 513.<br />

[14] Ledlow, M.J., Owen, F.N., Eilek, J.A. 2001, Rich-cluster and non–cluster<br />

radio galaxies and the (P, D) diagram for a large number <strong>of</strong> FR I and<br />

FR II sources, New Astron. Rev. 46, 343.<br />

[15] Marti, J.M., Müller, E. 1999, Living Reviews in Relativity.<br />

[16] Marscher, A.P. et al. 2002, Observational evidence for the accretion-disk<br />

origin for a radio jet in an active galaxy, Nature 417, 625.


REFERENCES 325<br />

[17] Leahy, P. 1993, in Jets in Extragalactic Radio Sources, eds. Röser, H.-J.,<br />

Meisenheimer, K.; LNP, Springer-Verlag, Heidelberg.<br />

[18] Scheck, L. et al. 2002, Does the plasma composition affect the long term<br />

evolution <strong>of</strong> relativistic jets ?, MNRAS 331, 615.<br />

[19] Smith, D.A. et al. 2002, A Chandra X-Ray Study <strong>of</strong> Cygnus A. III. The<br />

Cluster <strong>of</strong> Galaxies, ApJ 565, 195.<br />

[20] Spindeldreher, S. 2002, The discontinous Galerkin method applied on the<br />

equations <strong>of</strong> ideal relativistic hydrodynamics, PhD Thesis, Univ. Heidelberg<br />

[21] Sutherland, R.S., Dopita, M.A. 1993, Cooling functions for low–density<br />

astrophysical plasmas, ApJSup 88, 253.<br />

[22] Thiele, M. 2000, Simulation protostellarer Jets, PhD thesis, Univ. Heidelberg.<br />

[23] Toro, E.F. 2001, Godunov Methods: Theory And Applications, Kluwer<br />

Academics.<br />

[24] Wehrle, A. et al. 2002, The radio jets and accretion disk in NGC 4261,<br />

New Astron. Rev. 46, 235.<br />

[25] Wilson, A. S., Yang, Y. 2002, Chandra X-Ray Imaging and Spectroscopy<br />

<strong>of</strong> the M87 Jet and Nucleus, ApJ 568, 133.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!