A&A 474, 1-13 (2007)
DOI: 10.1051/0004-6361:20077703
M. Moscibrodzka1,2 - D. Proga2 - B. Czerny1 - A. Siemiginowska3
1 - N. Copernicus Astronomical Center,
Bartycka 18, 00-716, Warsaw, Poland
2 -
Department of Physics, University of Nevada, Las Vegas, NV 89154, USA
3 -
3.Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Cambridge, MA 02138, USA
Received 24 April 2007 / Accepted 28 June 2007
Abstract
Context. Numerical simulations of MHD accretion flows in the vicinity of a supermasssive black hole provide useful insights in to the problem of why and how systems - such as the Galactic center - are underluminous and variable. In particular, the simulations indicate that low angular-momentum accretion flow is highly variable both quantitatively and qualitatively. This variability and a relatively low mass-accretion rate are caused by interplay between a rotationally supported torus, its outflow, and a nearly non-rotating inflow.
Aims. To investigate the applicability of such flows to real objects, we examine the dynamical MHD studies with computations of the time-dependent radiation spectra predicted by the simulations.
Methods. We calculated the synthetic broadband spectra of accretion flows using Monte Carlo techniques. Our method computes the plasma electron temperature allowing for pressure work, ion-electron coupling, radiative cooling, and advection. The radiation spectra are calculated by taking thermal synchrotron and bremsstrahlung radiation, self absorption, and Comptonization processes into account. We also explored the effects of non-thermal electrons. We applied this method to calculating spectra predicted by the time-dependent model of an axisymmetic MHD flow accreting onto a black hole presented by Proga and Begelman.
Results. Our calculations show that variability in an accretion flow is not always reflected in the corresponding spectra, at least not in all wavelengths. We find no one-to-one correspondence between the accretion state and the predicted spectrum. For example, we find that two states with different properties - such as the geometry and accretion rate - could have relatively similar spectra. However, we also find two very different states with very different spectra. The existence of nonthermal radiation may be needed to explain X-ray flaring because thermal bremsstrahlung, dominates X-ray emission, is produced at relatively large radii where the flow changes are small and slow.
Key words: magnetohydrodynamics (MHD) - radiation mechanisms: general - radiative transfer
The overall radiative output from the Galactic center (GC) is very low for a system hosting a supermassive black hole (SMBH). This under-performance of the nearest SMBH challenges our understanding of mass accretion processes.
Certain properties of GC are relatively well-established. For example,
the existence of a
SMBH in GC is supported by
stellar dynamics (e.g. Schoedel et al. 2002; Ghez et al. 2003). The nonluminous matter within 0.015 pc of the GC is associated
with Sgr A
,
a bright compact radio source (Balick & Brown
1974). Observations of Sgr A
in X-ray and radio bands reveal a
luminosity substantially below the Eddington limit,
erg s-1. In particular, Chandra
observations show a luminosity in 2-10 keV X-rays of
erg s-1, ten orders of magnitude below
(Baganoff et al. 2003). Chandra observations have also revealed an
X-ray flare rapidly rising to a level about 45 times as high, lasting
for only
104 s, indicating that the flare must originate near
the black hole (Baganoff et al. 2001, 2003). A
strong variability for Sgr A
was also observed in radio and
near infrared (e.g., Eckart et al. 2005).
Despite the relative wealth of the observational data and significant
theoretical developments, the character of the accretion flow in the GC is
poorly understood. In particular, it is unclear whether the emission
comes from inflowing or outflowing material (e.g., Markoff et al. 2001; Yuan et al. 2002). Although the accretion fuel is
most likely captured from the stellar winds,
the mass accretion rate
onto the central black hole is uncertain. For example, the
quiescent X-ray emission measured with Chandra implies
a few
at
(assuming
spherically symmetric Bondi accretion, Baganoff et al. 2003), whereas
measurements of the Faraday rotation in the millimeter band imply
a few
(Bower et al. 2003, 2005). Recent measurements of Faraday rotation
of Marrone et al. (2006) imply even
lower mass accretion rate
,
where the
lower limit is valid for a sub-equipartition disordered or toroidal
magnetic field.
The increasing amount of the available
observational data of the GC has motivated a recent
development in the theory and simulations of accretion flows.
Most theoretical work has been focused on steady-state models,
which assume that the specific angular momentum, l, is high and
the inflow proceeds through a form of the so-called radiatively
inefficient accretion flow (RIAF; e.g., Narayan et al. 1998;
Blandford & Begelman 1999; Stone et al. 1999;
Quataert & Gruzinov 2000; Stone & Pringle 2001).
The high specific angular
momentum considered in these studies is a reasonable assumption, but
one not necessarily required by the data. In fact,
many observational aspects of Sgr A
can be better explained by low-l accretion flows.
For example, Moscibrodzka (2006, hereafter M06) showed
that models of purely spherical accretion flows underpredict the total
observed luminosity if the flow radiative efficiency is not
arbitrarily adopted but calculated self-consistently from the
synchrotron, bremsstrahlung, and Compton emission of the accreting
material. M06's results imply that low-l flows can also
reproduce the observed total luminosity as well as the high-lflows.
Additionally, the intrinsic time variability of low-l magnetized flows
found by Proga & Begelman (2003, PB03 hereafter)
can naturally account for some of Sgr A
variability
(Proga 2005).
In this paper, we present synthetic continuum spectra and Faraday rotation
measurements (RM)
calculated based on time-dependent two-dimensional (2D)
axisymmetric MHD simulations of accretion flows performed PB03.
These simulations are for slowly
rotating flows and are complementary to other simulations that consider
strongly rotating accretion flows initially set to be torii supported
by pressure-rotation (e.g., Stone & Pringle 2001; Hawley
& Balbus 2002 (hereafter HB02); Krolik & Hawley 2002; Igumenshchev & Narayan
2002). PB03 attempt to mimic the outer boundary conditions of classic
Bondi accretion flows modified by the introduction of a small,
latitude-dependent angular momentum at the outer boundary, a
pseudo-Newtonian gravitational potential, and weak poloidal magnetic
fields.
Such outer boundary conditions allow for the density distribution at
infinity to approach spherical symmetry. Recent X-ray images taken by
the Chandra show that the gas
distribution in the vicinity of an SMBH at the centers of nearby
galaxies is close to spherical (e.g., Baganoff et al. 2003; Di Matteo
et al. 2003; Fabbiano et al. 2003). Thus the outer boundary
considered by PB03 capture RIAF in Sgr A
better than those used
in simulations of high-l flows.
The PB03 simulations follow the evolution of the material with
a range of l. The material with
(where
is the Schwarzschild radius)
forms an equatorial torus because
its circularization radius is located outside the last stable
orbit. The torus accretes onto a black hole as a result of
magnetorotational instability (MRI, Balbus & Hawley 1991). As the
torus accretes, it produces a corona and an outflow that can be strong
enough to prevent accretion of a low-l material (i.e.,
with
)
falling through the polar regions. The torus outflow and the corona
can narrow or even totally close the polar funnel for the accretion of
low-l material. One of their conclusions is that even a slow
rotational motion of the MHD flow at large radii can significantly
reduce
compared to the Bondi rate.
Many properties of the torus found by PB03 are typical of the MHD
turbulent torii presented in other global MHD simulations (e.g., Stone
& Pringle 2001; HB02). Radial density profiles,
properties of the corona and outflow, as well as rapid variability are
quite similar in spite of different initial and the outer boundary
conditions. The main difference between simulations
of PB03' and the others is that the former showed
that the typical torus accretion can be quasi-periodically
supplemented or even
replaced by a stream-like accretion of the low-l material occurring
outside the torus (e.g., see the bottom right panel of Fig. 2 in PB03).
When this happens,
sharply increases and then gradually decreases.
The mass-accretion rate due to this "off
torus'' inflow can be one order of magnitude higher than that from
the torus itself (see Fig. 1 in PB03).
The off torus accretion is a consequence of
the outer boundary and initial conditions that introduce the low-lmaterial to the system. This material can reach a black hole because
the torus, the corona, and the outflow are not always strong enough to
push it away.
![]() |
Figure 1:
2D color maps of electron temperature
for four characteristic accretion sates A, B, C, and D.
The length scale for the horizontal axies is logarithmic
in units of
![]() ![]() |
Open with DEXTER |
![]() |
Figure 2:
2D color maps of the emission
at four frequencies:
![]() ![]() ![]() |
Open with DEXTER |
One can expect that in the vicinity of SMBH at the centers of
galaxies, some gas has a very little angular momentum and could be
directly accreted. Such a situation very likely occurs in that part of the GC where a
cluster of young, massive stars losing mass surrounds a SMBH
(e.g., Loeb 2004; Moscibrodzka et al. 2006; Rockefeller et al. 2004).
On the other hand, 3D hydrodynamical numerical models
of wind from the close stars show that the wind can
form a cold disk around Sgr A* with the inner
radius of
103-104 Schwarzschild radii
(Cuadra et al. 2006).
But unless the disk includes the viscosity, it cannot accrete closer.
In the context of such models we can again expect
that only a matter with initial relatively low angular momentum
can inflow at smaller radii at which PB03 started their computations.
Also the winds from inner 0.5
stars (SO-2, SO-16, SO-19)
can supply a lower angular momentum matter for the accretion (Loeb 2004).
To test other properties of the low angular momentum scenario,
we supplement here the dynamical MHD study of PB03 with the computations
of the time-dependent radiation spectra of accreting/outflowing material.
There are only a few papers (HB02; Goldston et al. 2005; Ohsuga et al. 2005)
that estimate radiation spectra, or at least the integrated emission,
predicted by MHD simulations of accretion flows
in GC.
These papers assume different dynamical situations than the one
considered in the PB03 simulations.
HB02 made an
order-of-magnitude estimates of the emission emerging from an MHD torus
in the GC. In particular, they estimated the peak
frequency of the synchrotron emission coming
from the inner parts of the flow at
Hz
(where n0 is the scaling density of the ions).
They also calculated the location of the peak of
a total bremsstrahlung emission
at about 1021 Hz.
They conclude that the dynamical variability in the MHD simulations
is generally consistent with the observed variation in Sgr A*.
However, HB02 focus mainly on the dynamical evolution of a high-l
non-radiative torus and do not consider the detailed
spectral properties of the flow.
Goldston et al. (2005) used the numerical simulation of HB02 and estimated
the synchrotron radiation assuming the scaling of the electron
temperature,
.
They calculated the synchrotron part of
the radiation spectrum and modeled polarization variability
including effects of self-absorption. The emission is variable (by a
factor of 10) at optically thin radio frequencies and originates in the very
inner parts of the flow (timescales on the order of hours).
The authors predict that the variability at different frequencies should
be strongly correlated. The general conclusion is that
a variable synchrotron emission in Sgr A* could be generated by
a turbulent-magnetized accretion flow.
Ohsuga et al. (2005) present the spectral features predicted by the 3D MHD flows simulated by Kato et al. (2004). These simulations follow
the evolution of a turbulent torus.
The heating/cooling balance equation used by Ohsuga et al. (2005) for
calculating the electron temperature includes radiative cooling and
heating via Coulomb collisions. The authors find that MHD flows in
general overestimate the X-ray emission in Sgr A*, because the
bremsstrahlung radiation originated at large radii dominates the X-ray
band. However, the quiescent-state spectrum of Sgr A* cannot be
reconstructed simultaneously in the radio and X-ray bands.
The authors reconstructed the observed flaring state, with assumption of rather
high
,
and restrict the size of the emission region
to only 10
.
A small size of the emission region is
consistent with the short duration of the observed X-ray flares, and it
results in the X-ray variability, so there is no need for
additional radiation processes, e.g. radiation from
nonthermal particles.
On the other hand, the observed quiescent X-ray
emission is extended, which contradicts the assumption made in Ohsuga et al. (2005).
Our goal is to compute the spectral signatures of the highly time-dependent flow found by PB03. In particular, we check whether a reoccurring switching between various accretion modes present in the PB03 simulations is consistent with the observed flare activity in Sgr A*. Our spectral model assumes that, at each moment of the accretion flow, there is a stationary distribution of both hydrodynamical and magnetic quantities. Radiative transfer is calculated based on this "stationary'' solution. The radiation computations are independent of the dynamics (Sect. 2). Our results show which parts of the accretion flow create characteristic features in the radiation spectra. In particular, we map the emission for synchrotron radiation including self-absorption and Comptonization (Sect. 3). We also show the time evolution of the spectra emerging from the flow for a number of moments in the evolution of an inner torus. In Sect. 3 we discuss these results in the context of the GC, and conclude in Sect. 4.
We computed the synthetic spectra for the simulation denoted as "run D''
in PB03. This simulation assumed a slowly rotating
accretion flow (with an angular momentum parameter
at
the equatorial plane) and were performed in the spherical polar
coordinates. The computational domain for run D was defined to occupy
the radial range
,
(where
is the Bondi radius,
a sound speed at infinity),
and the angular range
.
PB03
considered models with
.
The
domain was
discretized into zones with 140 zones in the r direction and 100
zones in the
direction. PB03 fixed the zone size ratios,
,
and
for
.
The MHD simulation were performed using dimensionless
variables. Therefore we first rescale all the quantities. The gas
density
is scaled by a multiplication
factor
,
which is the density at infinity.
The magnetic field
scales with
,
where
,
and
are the radial, latitudinal, and azimuthal components of the magnetic field,
respectively.
The internal energy density e scales with
and
.
To compute the synchrotron emissivity
we calculated the strength of the magnetic field,
,
at
each grid point. The ion temperature of the gas was calculated using politropic
relation
,
where
,
,
and
is the Boltzmann constant, mean particle weight
(
), and proton mass, respectively. We set the adiabatic index
.
We use scaling constants appropriate for the case of Sgr A*; i.e.
was chosen to fit the observed mass accretion
rate (see Sect. 3.1).
PB03 did not include radiative cooling in their calculations;
therefore, we used e and
to compute the distribution of the
temperature of ions
,
using the politropic relation mentioned above.
However, determination of the electron temperature is the key step in spectral
calculations. Ions and electrons undergo different types
of cooling and heating processes, and they
are not coupled well in a low density plasma. Ions and electrons
are likely to have different temperatures.
We calculate the electron temperature distribution by solving the
heating-cooling balance at each grid point at a given time in the
simulation. We present a detailed description of the method in Sect. 2.1
and outline the radiative transfer method in Sect. 2.2.
To calculate the electron temperature in each cell of the grid, we
solve the cooling-heating balance equation for electrons, which is given by
The advection term
consists of the radial part, shown for example
in Narayan et al. (1998), which can be expressed by
![]() |
(2) |
![]() |
(3) |
The compression heating (
in Eq. (1)) by the accretion for a 2D accretion
flow can be written as (M06):
The radiative cooling rate,
,
in Eq. (1) includes
the thermal synchrotron and bremsstrahlung radiation, reduced by
a mean self-absorption. We compute the self-absorption along fifty
directions. Comptonization terms are
computed for both synchrotron and bremsstrahlung radiation.
The synchrotron thermal emissivity, as described by Mahadevan
et al. (1996), is a function of
and
three coefficients:
,
,
and
.
For
K, we used
,
,
(there is a misprint in Mahadevan et al. 1996
). For the temperatures lower than
K, we modeled the synchrotron emissivity using the
analytical expression of Petrosian (1981).
Table 1:
Modeled spectral and photons indeces for different energy bands,
accretion states, and
.
We assume that the synchrotron radiation is produced by particles moving in the mean magnetic field B. Comptonization cooling terms for both synchrotron and bremsstrahlung radiation are calculated following the description given by Esin et al. (1996). More details and exact formulas for the radiative cooling are given in M06 (Sect. 2.1).
To fully describe the electron temperature
,
we iteratively solve Eq. (1) including all radial and angular
terms. We consider a very optically thin accretion
where the heating - cooling balance in Eq. (1) is dominated by two
terms, namely the advection and compression energy rates. These two
rates can lead either to heating or cooling, depending on radial and
angular velocity directions, density derivatives, and the
or
sign.
While
always cools electrons,
always heats electrons. If
and
are negative (in that
and
are both cooling terms ), one can find a solution to the balance
equation by decreasing electron temperature.
This means that we are taking more and more energy from ions through
Coulomb collisions, to keep the electrons balanced. Nominally, in the ion
equation of the conservation of energy the Coulomb coupling term
should also be included. This process can indirectly make electrons
dynamically important even if
.
In our calculations,
is small compared to
so that electrons do
not play any role in the dynamics. If, on the other hand,
and
are positive in the sense that
and
are both heating term, radiative processes
could not cool the electrons efficiently to keep
an energy balance. These are two extreme cases that are quite
difficult to treat numerically, especially the second case where
the heating energy cannot be balanced by the radiative cooling
within above described framework.
These issues however can be avoided by including additional cooling processes (for instance creating electron positron pairs) or, as we have found, by smoothing velocity and density radial profiles at each angle. We adopted the latter and used linear interpolations for the velocity and density radial profiles over the whole grid. This operation results in the advection and compression rates being always negative and positive, respectively. The equation of energy can then be solved using the standard formula for cooling mechanisms.
To avoid the same problem in the
direction, we neglect all
angular terms in
and
.
The energy balance
equation reduces to
An initial value of the electron temperature in each cell of the grid
is given by the adiabatic relation
.
In the initial loop we calculate the cooling term
,
and solve Eq. (1) at
each point. After the first
iteration, we correct
in the whole simulation domain and begin the
procedure again.
We stop the iterations when the equilibrium is
reached in each zone. The equilibrium condition
requires that the electron temperature at a given place
does not change in the next iteration very much, so we usually obtain
few percent consistency.
Two most dominant terms in Eq. (5) are
and
,
so that
parameter cannot be very large. We
show this in the case of radial equations. When we approximate
gradients of electron temperature and density with the
algebraic values (Eq.
), we obtain:
![]() |
(6) |
Our calculations show that the ions can reach very high temperatures of 1012 K in some regions, but such temperatures are too high for electrons;
therefore, to have
in order to keep the model self-consistent, we
assume rather low values of the parameter
.
Our analysis is
consistent with Quataert
Gruzinov's (1999) conclusions that, for
weakly magnetized plasma,
should be rather small. We keep
constant in the entire flow as a global parameter. This
approach should be refined in future studies, because
varies across the computation domain; in particular, it is small
very close to the black hole.
With the smoothed vr and
distribution, the values of
higher
would lead to the higher values of
,
and an
additional cooling process would be needed. We suppose that a creation
of electron-positron pairs could be produced and start to dominate the
cooling term.
Currently the maximum value of
in our
calculations is a few
K in the innermost regions.
For higher temperatures, such as 1011 K, and the assumed electron
positron pair equilibrium, the ratio between the pair number density
n+ and the number density of protons
,
z rises steeply
from z=10-4 up to z=16 in the inner parts of the flow.
The effects discussed above could change if we increase the density scaling
factor
.
The bremsstrahlung emissivity increases like
,
so the advection or compression term could be more
easily balanced by the radiation cooling. These results are thus
correct only for very optically thin accretion flows (Thomson
thickness is
), where cooling by radiation is
negligible as far the flow dynamics is concerned.
Our Monte Carlo algorithms were adopted from Pozdnyakov et al. (1983)
and Gorecki
Wilczewski (1984). The distributions of
the photon emissivity and photon energy were calculated
as described in Kurpiewski
Jaroszynski (1999). We considered a 2D
flow and the place of emission is given by the conditional
probability. We took a 2D distribution of the rate of
photon emission into account (for details how do compute
,
see Kurpiewski
Jaroszynski 1999).
When choosing a random location of the photon emission, we calculate
the probability of the emission at a given radius:
![]() |
(7) |
![]() |
(8) |
Finally we note that to calculate the radiation spectra, we followed M06
in using a grid that consists of 100 zones in the radial direction and 100
zones in the
direction.
Therefore, we transformed all the input data (ion
temperature, magnetic field induction and density taken
from the PB03 simulation, and calculated electron temperature)
into a new grid as defined in M06.
The input parameter is
.
We chose a value for this parameter that is
suitable for the conditions in low-luminosity active
Galactic Nuclei (LLAGN). In particular, our basic model was calculated
for the case of the GC. We assume a black hole mass of
and the asymptotic value of the density
of
.
To determine which characteristic flow patterns are responsible for specific spectral properties, we calculate broadband radiation spectra for 32 snapshots from the MHD simulation, including the four characteristic accretion states, A, B, C, and D, discussed by PB03.
State A is an accretion state at the early phase of the simulations,
while states B and C correspond to accreting and non-accreting tori,
respectively. Finally, state D corresponds to
a stream-like accretion of the very low-l material
that is approaching the SMBH, not through the torus, but from outside of
the torus (see the bottom-right panel
of Fig. 8 in PB03b where the stream
is below the equatorial plane).
The four characteristic states differ not only in the flow and magnetic
structure but also in their gross properties, e.g.,
.
In particular,
,
0.009, 0.001,
and 0.048, for states A, B, C, and D, respectively (
is the Bondi accretion rate).
To study the time variation of the radiative properties, in particular
to construct the light curves for various wavelengths, we compute radiation
spectra not only for these four states but also
for another 28 snapshots, for the time tfrom 2.35 to 2.40
(as in PB03, all times here are in units of the Keplerian orbital
time at
). Here, we include only thermal
particles and assume the inclination
angle
(edge-on view).
We begin presentation of our results by showing maps of
for the
four characteristic states (Fig. 1).
The figure shows only the innermost 20
where the temperature ranges between
K
and
K. The temperature distribution traces
the density distribution somewhat, but there is no one-to-one correspondence
with the density (see Fig. 8 in PB03b),
because
is a non-linear function of a few quantities,
including the complex velocity field.
In addition, the temperature maps compared to the density maps
lack fine details due to our smoothing procedure as described in Sect. 2.1.
The maximum electron temperature in our models is higher
than the maximum temperature in ADAF models. For example,
the Quataert & Narayn (1999) calculations
show that, to obtain
as high as in our calculations, one should
assume
and a relatively strong outflow
(e.g., model 5b in their Table 2).
Table 2: The modeled and observed values of the rms function.
Figure 2 shows the maps of synchrotron emission for the four
characteristic states and for four frequencies;
.
At these frequencies, emission is dominated by synchrotron emission (
and 12),
Comptonization (
), and bremsstrahlung (
).
The figure shows only the innermost 20
where most of the synchrotron emission is produced.
First, we describe our results for synchrotron emission (two left columns in Fig. 2).
In state A, the photon emission is distributed over a
wide range of .
In state B, the majority of the photons
come from the very inner cusp of the accretion torus. In state C,
there is no cusp, because the torus is truncated by the magnetic field,
and the synchrotron emissivity is reduced in the very inner region. In
state D, the emissivity is relatively high, and most of
the synchrotron photons are created in the stream of matter
approaching the black hole from below the equatorial plane.
As expected, the synchrotron photons undergo Comptonization.
For states B, C, and D, relatively few scatterings
occur near the black hole (see the third column from left).
Generally, the scattering regions are similar in all four states, except
for the enhanced Comptonization in a narrow elongated region below
the equator for
in state D. This narrow region corresponds to
the stream of low-l matter mentioned above.
For state A, scattering is more uniform compared to the above three states
because the density distribution is also more uniform.
In all four states, the bremsstrahlung photons are created relatively far
from the center (i.e.,
of the photons are created beyond
,
see the right most column).
In the central part of the accretion flow, the distribution of
the bremsstrahlung emission has a cusp for all states but state C.
Because the bremsstrahlung emissivity traces the density distribution,
the farther it is from the center, the more uniform
the bremsstrahlung emissivity.
Only the highest energy bremsstrahlung photons are created close to
the black hole.
![]() |
Figure 3:
Radiation spectra
corresponding to accretion states A, B, C, and D, respectively
(only thermal electrons were included). The mass accretion
rate at the outer boundary of the simulation is
![]() ![]() ![]() ![]() ![]() |
Open with DEXTER |
Figure 3 shows the radiation spectra for the four accretion states.
All spectra have three distinct components:
the synchrotron, Compton, and bremsstrahlung bump at low,
medium, and high frequencies, respectively. The spectra differ
from each other mainly in the location and strength of the synchrotron
and Compton bumps. The changes in the synchrotron radiation
reflect that this radiation
is produced in the central part of the flow where the flow changes
the most from state to state. In particular,
the low (
108-1011 Hz)
and high (
1011-1013 Hz) frequency synchrotron radiation changes
because this radiation is produced respectively in the thick torus
and in the plunging region.
Comptonization of synchrotron photons occurs predominantly in the
outer regions of the torus. However, the Comptonization bump
(
Hz) changes from state to state mainly because
of the changes of the energy in the input synchrotron photons.
On the other hand, most bremsstrahlung emission
(
Hz) is produced in the outer regions where
all four states are quite similar.
Therefore this part of the spectrum does not change, except for the high energy
cutoff which is sensitive to the conditions in the inner flow
where the highest energy bremsstrahlung photons are produced there.
Cross comparison between Figs. 2 and 3 shows that there is no
one-to-one correspondence between the accretion state
and the spectrum. For example, states A and D are quite different
(including their corresponding
), yet their predicted
spectra are relatively similar.
However, this is not
to say that radiation spectra do not depend on the the accretion state
because two very different states: C and D have very different spectra.
In particular,
the flux of the higher energy
synchrotron radiation and Compton radiation is orders of magnitude
higher for the stream-accreting state D than for the suppressed accretion state C.
Similarly, the emissivity is also significantly
higher for the torus accreting state B than for state C.
We find that the radiation spectrum responds to changes in the accretion state and mass accretion rate in a complex non-linear way. Because the flow is asymmetric the spectrum can also depend on the inclination angle, i.
To check the spectrum dependence on i, we computed several spectra for various inclinations. The emerging radiation spectrum is relatively insensitive to i, which is consistent with our optically thin approximation (i.e., low gas and column densities). However, even for these conditions, self-absorption can be appreciable for the synchrotron radiation because the opacity at low frequencies ( 108-1011 Hz) is sensitive to the electron temperature, the density, and the magnetic field distribution function (which depends on emissivity function). At higher frequencies (higher than the synchrotron peak frequency), the spectra are independent of i, because self-absorption does not play a role in this part of the spectrum.
![]() |
Figure 4: Synchrotron spectra emitted at accretion state D in the radio band, seen by observers at the different locations. |
Open with DEXTER |
Figure 4 shows our results for four inclination
angles:
,
and
.
The spectra were computed for state D, for which
effects of i are the strongest, as this state
represents the episode of the highest flow asymmetry due to the
presence of the stream of the low-l material below the equator.
We present results for
Hz because,
for high
,
the effects of "i'' are very small as discussed above.
The largest spectral differences are
for
and
,
bacause the flux can be an order of magnitude
lower for the former compared to the latter.
For
,
an observer can see low-energy photons
produced in the outer parts of the flow whereas for
,
low energy photons are self-absorbed
in the dense accreting material.
For
,
an observer sees the center through the accreting
material, the self-absorption is somewhat stronger, and the emission
decreases at low synchrotron energies.
We conclude that the spectra depend weakly on i
and it will be very difficult
to infer the inclination angle based on spectral analysis.
![]() |
Figure 5:
Upper panel: time evolution of 31 spectra in time intervals from 2.33 to 2.4.
Time is given in Keplerian orbital time (see PB03);
![]() ![]() |
Open with DEXTER |
Figure 5 shows light curves at four frequencies from radio to X-rays
(the top panel) and the corresponding time evolution of
(the bottom panel).
We analyze the time behavior for
.
This time interval is representative of the late-time evolution of
the flow as it captures flow switching between three of
the four characteristic states: B, C, and D (state A is only characteristic
of the early-time evolution).
For example, at t=2.368 and 2.394, the flow
accretes onto the SMBH through a stream of low-l material (state D)
and a torus (state B), respectively. These accretion episodes were
described and discussed in detail by PB03. Additionally,
our analysis includes times (i.e., at t=2.341 and 2.348) when
the accretion is suppressed.
During these two episodes the flow is in a state similar to
state C using PB03's terms.
Moreover, the flow at t=2.388 is in
a transition
between state D and short-lived state C.
![]() |
Figure 6:
Spectra emitted at accretion states
B, C, and D. The model was calculated for the
parameters of Sgr A*. Spectra for thermal electrons are marked by B, C, D,
and hybrid electron distribution including non-thermal electrons is marked
by "nt B'', "nt C'', and "nt D''. The parameters for non-thermal electrons are:
![]() ![]() |
Open with DEXTER |
As one might expect based on the spectra presented in Fig. 3,
the strongest variability, up to 3 orders of magnitude, is at
and 1014 Hz.
There is a smaller variability in the
synchrotron radiation at
Hz
(yet up to 2 orders of magnitude),
whereas there is practically no
variability in the X-ray flux (
Hz).
The X-ray flux ratio for different states varies by a factor of 1.06.
The light curves in the radio and IR frequencies show that the response of
the radiation properties to the changes in the flow dynamics
is the strongest when there is a significant temporary decrease in
(by nearly two orders of magnitude)
corresponding to a suppression of SMBH accretion
(i.e., at t=2.341, 2.348, and 2.388).
However, even during these changes, the X-ray light remains constant
because, as described above, bremsstrahlung photons are
produced far away from the center where the flow changes are small.
In summary, the accretion states of the inner flow cannot be identified in
the X-ray energy band, but they can be identified in the radio and IR bands.
To model a nonthermal population of electrons
with modified synchrotron emissivities, we follow
Yuan et al. (2003). For simplicity, we neglect the cooling
break in nonthermal distribution of electron velocities.
The nonthermal power-law electron distribution,
as a function of Lorenz factor ,
is given by:
![]() |
(9) |
We adopt the same electron temperature distribution
as in the thermal case ; i.e., we do not include nonthermal emission in the
cooling processes while computing .
We assume that
,
which is the nonthermal-to-thermal energy ratio,
is constant. Parameter
is one of the three free
parameters in calculations of nonthermal radiation.
The two other free parameters are
and p.
The maximum of a nonthermal emission is at the critical frequency of
,
where
is the cyclotron
frequency (Mahadevan
Quataert 1997).
To compare nonthermal spectra we assumed p=2.5.
This choice of p is motivated by simulations of particle acceleration in
nonrelativistic and relativistic shocks, which showed that p should be
slightly larger than 2, (e.g., Kirk
Schneider 1987).
In a nonthermal case, we assumed
to be
very low,
,
which was motivated by the work of
Ghisellini et al. (1998), who showed that self-absorption plays the main role in
thermalization of nonthermal electrons in optically thin hot sources.
We computed the emerging radiation spectra
for
and 106. These rather high
allow for a strong nonthermal emission
at X-rays.
![]() |
Figure 7:
As in Fig. 6 for different parameters of non-thermal electron distribution.
Note the change in the non-thermal spectra.
The parameters for non-thermal electrons are:
![]() ![]() |
Open with DEXTER |
Figures 6 and 7 compare spectra computed for B, C,
and D accretion states, with and without nonthermal electrons.
Nonthermal electrons produce a power-law spectrum extending from
thermal synchrotron peak up to the bremsstrahlung bump, or even up to
the -ray energy range for some parameters.
The IR emission due to nonthermal electrons is much larger in the case with nonthermal
emission, than in the case of pure thermal radiation.
The synchrotron
thermal Comptonization bump cannot be observed in any of the examples,
because it is "hidden'' under nonthermal photons.
For fixed p, ,
and
parameters, changes between
accretion states B, C, and D naturally produce variability in the whole
range of frequencies (see below for a discussion of this in
reference to GC flaring behavior). We note, however, that
a significant variability in the high-energy part of the spectrum
and spectral slopes of the power-law emission
can also be caused by changes in these three free parameters for
a given accretion state. Changes in p affect the spectral
slope of the nonthermal spectrum, whereas changes in
affect the nonthermal emission in the whole spectral
range. Generally the flux increases with increasing
at all frequencies. If
is high enough, no thermal synchrotron peak can be seen in the
radiation spectra, because the whole radiation spectrum is then dominated
by nonthermal emission. The position of the high-energy cutoff of
the nonthermal emission depends on
.
If the
and
parameters are high
enough, the thermal bremsstrahlung
peak in the high-energy part of the spectrum
will be covered by nonthermal emission.
We also note that timescales for the
variability of nonthermal emission in various accretion states can be
very short because nonthermal electrons are created much closer to
the black holes horizon. This effect could lead to a timescale
variability in high energy bands when the flow changes between different accretion states.
We have computed synthetic spectra based on MHD simulations of an accretion flow performed by PB03, using Monte Carlo techniques. We have studied time-dependent radiative properties of MHD flow by investigating light curves and broad band spectra, and the effects of inclination. We have also finished studing the effects of a possible non-thermal contribution to the thermalized electrons,so now we move to discussign the results in the context of Sgr A* observation.
As described in Sect. 1, Sgr A* is an underluminous and variable source.
For example, in the X-ray band, multiple flares (e.g. Eckart et al. 2004;
Belanger et al. 2005) are superimposed on a steady, extended emission at the
level of
erg s-1, with
occasional, extremely bright eruptions (Baganoff et al. 2001; Porquet
et al. 2003; maximum flux of
and
erg/s, respectively). The duration of the
flares ranges from half an hour to several hours, while the rise/decay
time is found to be a few hundred seconds (Baganoff 2003).
A variable emission is also seen in the NIR band (Genzel et al. 2003, 2004). A 17 min periodicity was
found in two of the events (Genzel et al. 2003)
and recently confirmed by Eckart et al. (2006a). The X-ray and NIR outbursts are directly related, as shown by
the detection of simultaneous NIR/X-ray events (Eckart et al. 2004,
2006b). The duration of events is around tens of
minutes. Quiescence emission is at the level of 1.9 mJy (Eckart
et al. 2004).
For GHz, and
GHz, the observed spectral index in Sgr A*
is
(
)
and
respectively (Narayan et al. 1998; Yuan et al. 2003, and references there in). In the X-ray
band, the quiescent emission is fitted well by the power-law photon index
(
), or it
can be explained by an absorbed, optically-thin thermal plasma with
kT=1.9 keV. In the intermediate and strongest flare state,
;
(Baganoff et al. 2001, 2003; Porquet et al. 2003).
Figure 7 shows the available observational data
for Sgr A*.
For pure thermal particles, the predicted emission is definitely too
faint in most of the energy bands.
Our models reproduce the level and slope of the radio emission for low .
However, the model cannot reproduce the observed position of the peak of the synchroton emission.
IR-optical synchrotron thermal emission does not match observational data. The
Comptonization bump is too weak due to weak input emission. X-ray
emission is well-fitted to the quasi-stationary spectrum, but shows
no variability when the accretion rate changes in the inner flow.
![]() |
Figure 8: Faraday rotation RM for four accretion states. The x axis shows the angle of the observer, and the y axis shows the RM values as a function of inclination as predicted by the models. Different lines show modeled RM for different simulation time (A, B, C, D). A state is marked as a solid line, B as a long dash line, C as a short dash line, and D as a dotted line, Two straight lines show the lower and upper limits from Marrone et al. (2006). |
Open with DEXTER |
The modeled X-ray photon index in the energy band 2-10 keV for thermal
emission is constant for the B, C, and D states. Its value is too low to
fit the measured power-law photon index in quiescence emission.
In our thermal model this part of the spectrum is due to bremsstrahlung
emission (or power law with
),
which does not agree with the interpretation of observational
data that the bremsstrahlung emission is produced
by the 2 keV plasma (or power law with
).
Our model could agree with the thermal bremsstrahlung from the 2 keV plasma,
if we were to assume a much lower mass accretion rate and used
a large computational domain. We note
that the outer radius of our
grid is
,
as in PB03, which is 100 times
smaller than the field of view of spectroscopic observations of Sgr A*
(
). To make our model more
self-consistent, we would have to perform our calculations over
a wide range of radii and then
compare them with the observations. This would increase
X-ray emissivity because the integration would be dominated by the
outer parts of the flow. Thus, to compensate for this increase in emissivity,
we would have to decrease
,
which would affect the synchrotron
part of the radiation spectrum. To reduce the discrepancy, one
would also have to compute MHD models in full 3D. Non axi-asymmetric
effects very likely contribute to the time-variability of Sgr A* and also
to the overall strength and topology of the magnetic field. We plan to
carry out 3D MHD in the near future.
In the case of a nonthermal emission, the general behavior of the radio spectral index is similar to the thermal radio emission. The difference is that nonthermal electrons in general produce a flatter spectral index in comparison to the thermal model,because a photon distribution function is not as centered as in the thermal case. Photons in partially nonthermal plasma are created at slightly greater distances than in thermal plasma. The best agreement with the data is obtained for the accretion state C.
Our nonthermal models in the X-ray energy band agree with the data for both a
weaker flare with
,
which can be explained by models with
,
and for the strongest flare with
,
which can
be explained well by models with
.
These results
would suggest that
changes are not necessarily
correlated with the flare strength. The flare is not
extended (Quataert 2003), in contrast to the quiescent emission, which also supports
our suggestion of flares being created by photons emitted by
nonthermal particles.
It seems that, apart from the changes of accretion state (which can be
responsible for the IR variability and some kind of periodicity in
this frequency range), changes in
are needed to
describe the X-ray flaring behavior. To fit the observation
well, a range of
should be on the order of
104-106.
Summarizing, the model with a contribution of nonthermal electrons
offers a much better representation of the spectral variability of Sgr
A*, although it cannot reconstruct a few
observational points at the synchrotron peak. In addition, we would like to stress
that, although low energy radio photons are created
in the outer parts of a torus, our results are consistent with recent
measurements of Sgr A* size, which depends on
(Shen 2006).
In particular for
Hz),
the intrinsic size of Sgr A* is 1 AU (
) cm. For
Hz, its size
increases by 25
.
Our calculations of the size of Sgr A* are
cm and
cm for
and
Hz,
respectively.
Our models can also be constrained by RM observations.
For example, Marrone et al. (2006) found that over a period of two months,
RM varied
within a range between
and
.
Figure 8 shows our computed RM as a function of the inclination angle
and of the lower and upper limits from observations taken by Marrone et al. (2006).
The figure shows that RM depend strongly on an orientation angle of a torus
with respect to the observer. In state A, RM are very high compared to
the state B (an accreting torus), and the measurements at various
inclinations can differ by 2 orders of magnitude.
Generally, the RM increasing with decreasing inclination angle.
This is simply because the magnetic field is highly ordered close to the
poles where there is an outflow, whereas the
magnetic field can be strongly tangled close to the equatorial plane. The
model does not overpredict the observational limits either at the
equatorial plane, or around the inclinations of 30-50,
where the inflow mixes with outflow and the region is particularly
turbulent.
Acknowledgements
We would like to thank the referee for his/her useful comments. Thanks go to Ryuichi Kurosawa and Agnieszka Siemiginowska for reading the manuscript. This work was supported in part by grants 1P03D 008 29 of the Polish State Committee for Scientific Research (KBN). D.P. acknowledges support from NASA under grant HST-AR-10305 from the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS5-26555. Partial support for this work was provided by the National Aeronautics and Space Administration through the Chandra award TM7-8008X issued by the Chandra X-Ray Observatory Center, which is operated by the Smithsonian Astrophysical Observatory for and on behalf of NASA under contract NAS8-39073.