Issue |
A&A
Volume 512, March-April 2010
|
|
---|---|---|
Article Number | A2 | |
Number of page(s) | 17 | |
Section | Astrophysical processes | |
DOI | https://doi.org/10.1051/0004-6361/200913186 | |
Published online | 17 March 2010 |
The extreme luminosity states of Sagittarius A*
N. Sabha1 - G. Witzel1 - A. Eckart1,2 - R. M. Buchholz1 - M. Bremer1 - R. Gießübel2,1 - M. García-Marín1 - D. Kunneriath1,2 - K. Muzic1 - R. Schödel3 - C. Straubmeier1 - M. Zamaninasab2,1 - A. Zernickel1
1 - I.Physikalisches Institut, Universität zu Köln,
Zülpicher Str. 77, 50937 Köln, Germany
2 -
Max-Planck-Institut für Radioastronomie,
Auf dem Hügel 69, 53121 Bonn, Germany
3 -
Instituto de Astrofísica de Andalucía (CSIC),
Camino Bajo de Huétor 50, 18008 Granada, Spain
Received 26 August 2009 / Accepted 22 December 2009
Abstract
We discuss mm-wavelength radio, 2.2-11.8 m NIR and
2-10 keV X-ray light curves of the super massive black hole (SMBH)
counterpart of Sagittarius A* (SgrA*) near its lowest and
highest observed luminosity states.
We investigate the structure and brightness of the central
S-star cluster harboring the SMBH to obtain reliable flux density estimates of SgrA*
during its low luminosity phases.
We then discuss the physical processes responsible for
the brightest flare as well as the faintest flare or quiescent
emission in the NIR and X-ray domain.
To investigate the low state of SgrA*
we use three independent methods to remove or strongly suppress the
flux density contributions of stars in the central 2
diameter region
around SgrA*.
The three methods are: a) low-pass filtering the image;
b) iterative identification and removal of individual stars;
c) automatic point spread function (PSF) subtraction.
For the lowest observed flux density state all 3 image reduction methods result in
the detection of faint
extended emission with a diameter of 0.5
-1.0
and
centered on the position of SgrA*.
We analyzed two datasets that cover the lowest luminosity states of SgrA* we observed to date.
In one case we detect a faint K-band (2.2
m)
source of
4 mJy brightness (de-reddened with AK=2.8) which we identify as SgrA* in its low state.
In the other case no source brighter or equal to a de-reddened
K-band flux density of
2 mJy was detected at that position.
As physical emission mechanisms for SgrA* we discuss bremsstrahlung,
thermal emission of a hypothetical optically thick disk,
synchrotron and synchrotron self-Compton (SSC) emission,
and in the case of a bright flare the associated radio response due to
adiabatic expansion of the synchrotron radiation emitting source component.
The luminosity during the low state can be interpreted as
synchrotron emission from a continuous or even spotted accretion disk.
For the high luminosity state SSC emission from THz peaked source components
can fully account for the flux density variations observed in the
NIR and X-ray domain.
We conclude that at near-infrared wavelengths the SSC mechanism
is responsible for all emission from the lowest to
the brightest flare from SgrA*.
For the bright flare event of 4 April 2007 that was covered from the radio to the X-ray
domain, the SSC model combined with adiabatic expansion can explain the related peak
luminosities and different widths of the flare profiles obtained in
the NIR and X-ray regime
as well as the non detection in the radio domain.
Key words: black hole physics - X-rays: general - infrared: general - accretion, accretion disks - Galaxy: center - Galaxy: nucleus
1 Introduction
Table 1: The brightest X-ray flares from SgrA*. For references see text.
Table 2: Details on observing runs during which SgrA* was in a very low NIR luminosity state.
At the center of the Milky Way
stellar motions and variable emission allow us to firmly
associate Sagittarius A*
(SgrA*) with a 6
super-massive black hole
(Eckart & Genzel 1996; Genzel et al. 1997, 2000; Ghez et al. 1998, 2000, 2003, 2005a, 2008;
Eckart et al. 2002; Schödel et al. 2002, 2003, 2009; Eisenhauer 2003, 2005;
Gillessen et al. 2009).
Recent radio, near-infrared and X-ray observations have detected variable
and polarized emission and give detailed insight into the
physical emission mechanisms at work in SgrA*, which may include any or all of synchrotron, SSC,
and bremsstrahlung emission
(e.g. Baganoff et al. 2001, 2002, 2003;
Eckart et al. 2003, 2004, R2006a,b, 2008a,b, 2009,
Porquet et al. 2003, 2008; Goldwurm et al. 2003; Genzel et al. 2003;
Ghez et al. 2004a,b; Eisenhauer et al. 2005;
Belanger et al. R2006; Hornstein et al. 2007;
Yusef-Zadeh et al. R2006a,b, 2007, 2008, 2009; Marrone et al. 2009;
Dodds-Eden et al. 2009).
Multi wavelength detections of the radio point source
at sub-millimeter, X-ray, and infrared wavelengths have
also been made, showing that the luminosity associated
with SgrA* is on the order of
10
-9...-10 times
below the Eddington luminosity
and many orders of magnitudes below
that of SMBHs in active galactic nuclei (AGN) with
comparable masses (of about 4
106
).
The surprisingly low luminosity has motivated many theoretical
and observational efforts to explain the processes that are at work
in Sgr A*.
For a recent summary of accretion models and variable accretion
of stellar winds onto Sgr A* see Yuan (R2006) and Cuadra & Nayakshin (R2006, 2009).
Sgr A* is - in terms of Eddington luminosity - the faintest super-massive black hole known. However, due to its proximity it is bright enough to be studied in great detail. With the possible exception of the closest galaxies, no extragalactic super-massive black hole with a similar feeble Eddington rate would be observable. Motivated by the need to explain the variable luminosity of SgrA* we have analyzed data from it highest and lowest states. In order to separate the weak NIR emission of Sgr A* from the surrounding stars, the use of large telescopes and adaptive optics (AO) is required. Here we present the first attempt to do this with near-infrared data obtained with the VLT. Likewise, high resolution is needed in the X-ray regime to separate Sgr A* from the surrounding diffuse X-ray background.
The temporal correlation between
the rapid variability of the near-infrared (NIR) and X-ray emission
suggests that the emission showing 10
33-35 erg/s flares
arises from a compact source within a few tens of
Schwarzschild radii of the SMBH.
For SgrA* we assume
= 2
= 2GM/c2
1010 mfor a
4
106
black hole.
Here
is one
Schwarzschild radius and
the gravitational radius of the SMBH.
One
then corresponds
to an angular diameter of
8
as at a distance to the Galactic
center of d=8 kpc
(Reid 1993; Eisenhauer et al. 2003; Ghez et al. 2005a,b).
These observations can be explained in a model of
an intrinsically faint accretion disk that is dominated by red-noise
and temporarily harbors
a bright orbiting spot possibly in conjunction with a short jet
(Eckart et al. R2006b; Meyer et al. R2006a,b, 2007).
The phenomenon can best be studied through the brightest flares
that allow us to derive high signal to noise light curves and
multi wavelength observations throughout the electromagnetic spectrum.
In its lowest luminosity states SgrA* is difficult to detect though
due to the presence of a strong diffuse background
at X-ray wavelengths and the confusion with nearby stellar sources
in the near-infrared.
Here we discuss the relative importance of the bremsstrahlung and
SSC process for the low state of SgrA* and investigate if the
brightest flares of SgrA* can be explained via
a SSC model involving up-scattered
sub-millimeter photons from compact source components.
In Sects. 2 and 3 we describe the observations, the
data reduction and the algorithms we used to extract SgrA* in its low
state from the background emission of the dense cluster of high velocity
stars at angular separations of less than about 1
from the position of SgrA*.
Below we will refer to this association of stars as the
``S-star cluster''.
In Sect. 4
we present and discuss results.
We start by summarizing the results of our deep
imaging
in Sects. 4.1 and 4.2.
In the following subsections
we then discuss various emission mechanisms
for the low and high luminosity states of SgrA*.
We show that the SSC process can successfully discribe both the low and
high luminosity states and that - in particular for the high states - no
extra mechanisms need to be involved (discussion in e.g.
Dodds-Eden et al. 2009; Eckart et al. 2009).
A brief summary and conclusions are presented in Sect. 5.
Throughout the paper we quote de-reddened infrared flux densities
using a -band (2.2
m) extinction correction of
(Scoville et al. 2003;
Schödel et al. 2007, 2010, in prep.; Buchholz et al. 2009).
In Sect. 4.1
we use
and
to describe the projected and three
dimensional structure of the central stellar cluster respectively.
In Table 1 and
starting with Sect. 4.4
we use
and
as the bulk boosting factor and the relativistic
electron Lorentz factor.
![]() |
Figure 1: Identification of individual sources that were used for the iterative subtraction of the Ks-band datasets (Table 2). Here we show the image derived from the 23 September 2004 data. The nomenclature was taken from the deconvolved H-band image given in Fig. 1 by Gillessen et al. (R2006). Sources that are not contained in there have labels starting with ``N''. Relative positions and flux densities of the labeled sources are given in Table 3. |
Open with DEXTER |
Table 3: Sources used for iterative source subtraction for the 23 September 2004 data.
2 Observations and data reduction
Near-infrared (NIR) observations of the Galactic center (GC) were
carried out with the NIR camera CONICA and the adaptive optics (AO)
module NAOS (briefly ``NACO'') at the ESO VLT unit telescope 4 (YEPUN)
on Paranal, Chile.
The data were taken on the night of 30 August and 23 September, 2004.
The dataset taken on 23 September 2004
turned out to be the best available set in which SgrA*
is in a quiet state.
During the 30 August observations some weak activity at 2.2 m could be
detected, making this dataset well-suited to determine
the position of SgrA* with respect to neighboring stars.
The combination of both datasets is ideally suited to obtain a
flux density estimate or limit of SgrA* in its quiet state.
For all observations the infrared wavefront sensor of NAOS was used
to lock the AO loop on the NIR bright (-band magnitude
6.5)
supergiant IRS 7, located about
north of Sgr A*.
The pixel scale was 13.27 mas.
The atmospheric conditions (and consequently the AO correction) were
stable during the observations. The seeing at the telescope measured in the optical was 0.6
.
We used an integration time of DIT = 2 s and a number of integrated images of NDIT = 15.
Details on start and stop times are listed in Table 2.
All observations in the -band (2.2
m)
were dithered to cover a larger area of the GC by mosaic imaging.
The sky background for the
-band observations was extracted from the median of
stacks of dithered exposures
of a dark cloud - a region practically empty of stars -
about 400
north and 713
west of the target.
All exposures were sky-subtracted, flat-fielded, and corrected for
dead or bad pixels.
The images were shifted and stacked
in a cube with a mean average to get
a mosaic image of the GC. All of these steps
were performed with the DPUSER software for
astronomical image analysis (Ott, MPE; see also Eckart & Duhoux 1990).
Subsequently, PSFs were extracted from these
images with StarFinder
(Diolaiti et al. 2000) both for deconvolution
with the Lucy-Richardson (LR Lucy 1974) algorithm and for PSF subtraction.
The flux densities of the sources were measured by aperture photometry with circular apertures of 40 mas radius and corrected for extinction, using AK = 2.8 (Scoville et al. 2003; Schödel et al. 2007; Schödel et al. 2010, in prep.; Buchholz et al. 2009). Possible uncertainties in the extinction by a few tenths of a magnitude alter the infrared flux densities of SgrA* by a few 10%. These variations are easily compensated for by insignificant variations in the spectral indices or the boosting which do not influence the general results obtained in this paper.
The flux density calibration was carried out with the
known Ks-band flux densities of IRS16C, IRS16NE, and IRS21 by
Blum et al. (1996).
Subsequently the relative photometry for Sgr A*
was done using 10 sources within 1
6 of Sgr A* as secondary calibrators
(S67, S92, S35, S8, S76, S1, S2, S87, S65, S30; Gillessen et al. R2006).
This results in a
-band flux density of the high velocity star
S2 of 22
1 mJy, which
compares well with the magnitude and flux for S2 quoted by
Ghez et al. (2005b) and Genzel et al. (2003).
The measurement uncertainties for Sgr A* were obtained on the nearby reference star S2.
The background flux in the immediate vicinity of Sgr A* was
obtained by averaging the measurements at six random locations
in a field located about 0
6 west of Sgr A* that is
free of obvious stellar sources.
A Ks-band image of the central 4.3
2.5
is shown in Fig. 1.
This field is at the center of the stellar cluster of which we show
a 15
15
overview in Fig. 2a.
The source positions and relative astrometry was established
via StarFinder.
The positions in
Table 3 are given relative to the position of
SgrA* as determined from the 23 September 2004 data.
We list the name following the nomenclature used by Gillessen et al. (2009)
and in the second column the name used by Do et al. (2009b).
Sources labeled with N had not been identified before.
Identifications with question marks are tentative and suffer most likely
from the proper motion of faint sources between the present date and the date
used for identification by Gillessen et al. (2009).
Then we list in Table 3 the sky projected relative offsets
,
and radial distance R from SgrA* in arcseconds.
The uncertainties in the offsets are
better than a third of a pixel i.e. 0.004
We then list in Table 3
the observed
magnitude derived in this paper
and the observed K' magnitude listed by Do et al. (2009b)
corrected by an offset value of -0.2 mag.
The scatter between both values after correction for this offset is 0.2 mag, so that the
offset corresponds to 1
in magnitudes (20% in flux)
and is well within the expected calibration uncertainties for faint
stars in the crowded field.
The difference between both mangitude values is listed in the last column.
For sources S0-8, S0-18 and S0-38 no magnitudes are given by Do et al. (2009b).
The large magnitude difference in the case of S0-12 is most
likely due to the blend of S59 and S60.
The two NIR filters
and K
(Do et al. 2009b) are very similar with
Ks(NACO):
m and width
m
and
K'(NIRC):
m and width
m.
No transformation between magnitudes derived with these filters was applied.
The 0.2 mag scatter relates to
stars of different brightnesses at different epochs between which they are
moving at different locations in the cluster with respect to different
local backgrounds.
The relative flux density uncertainty for the star S2 - which we use as an estimate for SgrA* -
is determined within a single epoch at fixed position.
3 Subtracting stars
To investigate the low luminosity state of SgrA*
we use three independent methods to remove or strongly suppress the
flux density contributions of stars in the central 2
diameter region
around SgrA*.
All three methods give comparable results and allow us to clearly
determine the stellar light background at the center of the Milky Way
against which SgrA* has to be detected.
We assume that the PSF as determined from stars in the central few arcseconds of the
image is uniform across the central S-star cluster.
Investigations of larger images
(e.g. Buchholz et al. 2009) show that on scales of a few arcesconds
this is a reasonable assumption and that
PSF variations have to be taken into account
only for fields 10
.
Whether the residual extended emission we find close to SgrA* is due to
residuals from the subtracting algorithms or associated with additional
stellar emission from the S-star cluster is discussed in
Sect. 4.1.
![]() |
Figure 2:
The Ks-band image of the Galactic center region
a)
taken on 23 September 2004. The central
15
|
Open with DEXTER |
3.1 Linear extraction of extended flux density
With the point spread function P and the object distribution Owe can write the formation of an image I as
![]() |
(1) |
Here the symbol * denotes the convolution operator. The variable


![]() |
(2) |
Here we use a narrow Gaussian G normalized to an integral power of unity. This can be rewritten as
![]() |
(3) |
Since the width of the extended emission is expected to be much larger than the full width at half maximum of the PSF, and in particular since the Gaussian G is narrow with respect to P, we can infer that

![]() |
(4) |
Since


![]() |
(5) |
Here the symbol

![]() |
(6) |
or
![]() |
(7) |
The advantage of this new method is that it is independent of input models, i.e. it does not depend on the identification of individual stars or on assumptions on the existence or properties of a diffuse background flux density distribution. Also the method only uses linear operations, i.e. image subtraction and linear convolution and deconvolution (Fourier multiplications and divisions). Therefore it is not dependent on the flux density of sources and numbers of iterations, as is the case for many non-linear algorithms (e.g. the Lucy-Richardson or maximum entropy algorithm). In order to demonstrate the validity of the approach we show the results we obtained on the central 15








![]() |
Figure 3:
The inner 2
|
Open with DEXTER |
3.2 Iterative PSF subtraction
The iterative PSF subtraction was carried out within
the central 2
to 3
around the position of SgrA*.
This was performed by resizing, shifting and scaling
the PSF to the position and the value of each
star and then subtract it.
The sources used for this process are listed in
Table 3 and are shown in Fig. 1.
A total number of 51
stars were subtracted so that the resulting
background was as smooth as possible.
In Fig. 4 we show the resulting subtracted image
of the area of interest around SgrA*.
All stars brighter than about 17.5
(see KLF in Fig. 6)
have been subtracted with a scaled version of the PSF.
The thin lines in Figs. 4a and b are used to
triangulate the position of SgrA*.
Figure 4a shows the image obtained from a section
of the 30 August dataset with SgrA* flaring.
Compared to S2 with
22 mJy SgrA* has a brightness of about 4 mJy.
Fig. 4b shows the image obtained from a section of the
23 September dataset with SgrA* in a quiescent phase.
Here SgrA* is clearly not detected and is fainter than
about 2 mJy (de-reddened).
![]() |
Figure 4: Results of the iterative star subtraction obtained for the 30 August a) 23 September data b) as described in Sect. 3.2. |
Open with DEXTER |
Figure 4 also reveals the presence of faint residual
extended emission - centered on the position of SgrA*.
This residual emission is most likely from the
central S-star cluster of high velocity stars.
We determined the position of SgrA* from the data with
SgrA* in a flare state taken on 30 August, 2004.
Taking an upper limit of the observed two dimensional projected
stellar velocity dispersion of 1000 km s-1 and the fact that
for the Galactic center proper motion of 1 mas/yr corresponds to
about 39 km s-1, the relative positional uncertainty for stars observed
between both epochs is expected to be well below 2 mas.
Thin lines in Figs. 4a and b are used to
triangulate the position of SgrA*.
Figure 4 shows that for the dataset taken
on 23 September 2004 no NIR counterpart of SgrA* was detected.
For 30 August 2004 we find a NIR counterpart of SgrA* with a
dereddened flux density of about 4 mJy with a 3
uncertainty of about 2 mJy.
Estimates of the flux density limits and uncertainties are given in
Sect. 4.2.
3.3 Automatic PSF subtraction
We also analyzed the images with the StarFinder (Diolaiti et al. 2000) program package.
It not only provides accurate
point source photometry but also a reliable estimation of the diffuse
background emission.
With a point spread
function (PSF) extracted from bright stars near
the guide star (in this case IRS 7), we performed
photometry and astrometry on the deconvolved image via PSF fitting with
the StarFinder package.
The resulting diffuse background emission is shown in Fig. 5.
In addition to some residual emission from unresolved stars
this method also reveals extended emission centered on the position of SrgA*
within the dashed 0.5
diameter circle centered on the position of SgrA*.
This finding agrees with the low-pass filtered image in Fig. 3.
![]() |
Figure 5: Result of the automatic PSF subtraction obtained for the 23 September 2004 data. |
Open with DEXTER |
![]() |
Figure 6:
KLF histogram of the stars detected in the central field
derived from the 23 September 2004 data.
The straight full and dashed lines indicate the KLF slope of
0.21 |
Open with DEXTER |
4 Results and discussion
First we analyze the properties of the central cluster of high velocity stars and compare them to the results of published investigations. We can then confidently derive the flux density limits on SgrA* and show that contributions to the lowest luminosity states from bremsstrahlung are unlikely. We then discuss the extreme luminosity states of SgrA* in the framework of a synchrotron self-Compton mechanism.
4.1 Imaging the S-star cluster and the detection of SgrA* in its low luminosity state
The detected number of 51 stars within a radial distance of about 0.5
from SgrA* results in a surface number density of
68
8 arcsec-2, taking the square-root as the uncertainty of that value.
This result agrees with the central number density of
60
10 arcsec-2 given by Do et al. (2009b) in
their Figs. 9 and 10.
Figure 6 shows the KLF histogram (K-band luminosity function)
that can be derived for the stars detected in the central field.
The binning of 0.75 magnitudes allows for a sufficient number of bins
to clearly detect the linear slope in the double logarithmic plot and to
have a sufficiently large number of sources per bin (about 10) at the same time.
With d log(N)/d log(Ks) = 0.3 0.1 the data compare well with the KLF slope of
0.21
0.02 found for the inner field (
)
by Buchholz et al. (2009).
Within the central 0.5 arcsec radius region around SgrA* and in
the magnitude interval ranging from Ks = 16.75 to 17.50 no significant deviation
from a straight powerlaw line can be detected - implying that the
completeness is high and probably comparable to the 70% value derived
for magK=17 by Schödel et al. (2007).
In the interval ranging from Ks = 17.50 to 18.25 the stellar counts drop
quickly to about 20% of the value expected from the straight powerlaw line.
All three methods we used to correct for the flux density contribution of the stars in the
central 2
reveal faint extended emission around SgrA*
(see Figs. 3b, 4b and 5).
For the iterative and automatic PSF subtraction the PSFs were extracted
with a radius of 1
,
which is about twice the FWHM of the S-star cluster.
In case of a PSF misplacement a significant flux density contribution
to the central position can only come from the about five stars that are located within
one FWHM of SgrA*. They have a median brightness of about 2 mJy. To explain all of
the
2 mJy at the center by this effect, each star would have to
provide about 0.4 mJy or 20% of its flux density. This can only be realized by a
systematic positional shift of these stars towards the center by about 1 pixel
13 mas
each. Still, the positional accuracy which is typically reached is on the order of a few tenths of
a pixel. Larger displacements result in a clearly identifiable characteristic plus/minus pattern
in the residual flux distribution along the shift direction.
However, allowing for a maximum positional uncertainty of 1 pixel
and approximating the independent shifts of five stars by five shifts of a single
star that follows a random walk pattern, the expected effective displacement is
1/
pixels for a single shift of each star. It will result in a 5% peak flux contribution of each
of the 5 stars - assuming that the displacements are systematically towards the center, i.e. a total
maximum contribution of 0.4 mJy.
This implies that about one 20%-30% of the flux density detected at the center may be explained by a
positional misplacement of the neighboring stars.
Therefore we assume that the bulk (more than 2/3) of this extended emission toward the center
can be associated with the
0.5
-1
diameter S-star cluster around SgrA* and is
due to faint stars at or beyond the completeness limit reached in the KLF.
The extended residual emission is not distributed azimuthally uniform about the position of SgrA*. This is not expected either. The shape of the residual emission will always be dominated by the randomly distributed few brightest stars that contribute to it. Also the light of the central stellar cluster is not distributed azimuthally uniform on any scale: On scales larger than 0.5 pc it is affected by dust absorption (e.g. extinction map by Buchholz et al. 2009), on scales less than 0.5 pc the bright members of the IRS16 cluster are predominantly located to the east, and during the past decade in the central arcsecond most of the brighter S-stars have been located to the south and east of SgrA*.
The diffuse background emission can be compared to the
projected distribution of stars
,
with R being the projected radius.
Independent of the applied method we find that the azimuthally
averaged residual diffuse background emission centered at
the position of SgrA* decreases very gently as a function of radius
(Fig. 7).
The ratio between the brightness measured within or at the edge of
the central resolution element (
)
and at a projected radius of
e.g.
is
,
corresponding to
.
This is consistent with the distribution of number density counts
of the stellar populations in the central arcseconds
derived from imaging VLT and Keck data.
Buchholz et al. (2009) and Do et al. (2009b)
find a
for the young stars, but a much flatter distribution
for the late-type (old) stars
with
0.17
0.09 or even
-0.12
0.09, respectively.
For pure imaging Do et al. (2009b) quote
inside a radius of
.
![]() |
Figure 7:
Azimuthal average of the diffuse background emission as derived from
the three different methods applied: low pass filtering (red),
manual PSF subtraction (green), and automatic PSF subtraction (blue).
The curves
(mean flux and 1 |
Open with DEXTER |
In case of a relaxed stellar cluster around a supermassive black hole, stellar
dynamics predicts to observe a cusp, i.e. a density increase toward
the black hole with
=
to 0.75 (Bahcall & Wolf 1976;
Murphy 1991; Lightman & Shapiro 1977; Alexander & Hopman 2009).
Here
denotes the exponent of the three dimensional
distribution which does not suffer from projection effects
but is harder to deduce from observed data.
Only in the case of very extreme stellar densities collisions
can lead to a
as low as 0.5. But the required high densities
are not reached in the Galactic center.
The observed values of
are
significantly lower than what is predicted by theory.
The principal reason is probably to be sought in the stellar population
in the central arcseconds, which towards Sgr A* becomes
increasingly dominated by young massive stars that are too young to be
dynamically relaxed.
The small value of the projected diffuse light
exponent
may therefore imply the absence of a pronounced, relaxed cusp of stars.
However, there may in fact exist a deficit of
late-type stars near Sgr A* (see also Genzel et al. 2003, Fig. 7).
It may well be that most of the spectroscopically identified late
type stars at small projected distances are at larger
three dimensional radii r R from SgrA*.
In this case the observed
will be a mixture between the much steeper
for early type stars and the lower value
for
late type stars. A detailed discussion is given in
Do et al. (2009b), Buchholz et al. (2009), Schödel et al. (2007),
and Genzel et al. (2003).
The main result for the present analysis of faint luminosity states of SgrA* is that the
observed flat background light and number density distribution
described by the exponent
as well as the high degree of completeness reached around Ks = 17.5
allows us to clearly distinguish the emission of SgrA* against
the stellar light background at the center of the Milky Way.
This also shows that SgrA* is ideally suited to perform a case study
to investigate this super-massive black hole in its
lowest activity state.
4.2 Observing SgrA* at low luminosities
High angular resolution images in the near-infrared allow us to
unambiguously attribute bright dereddened flux levels to SgrA*
(e.g. Genzel et al. 2003; Eckart et al. 2004; Ghez et al. 2004a,b;
Hornstein et al. R2006; Do et al. 2009a).
The identification of fainter emission from
SgrA* becomes increasingly difficult, since one can expect
confusion from
the 1
diameter S-star cluster that is centered on the
location of Sgr A*.
Due to the high proper motions of the stars within the S-star cluster this
confusion also changes on the time scales of months to years.
From the location of SgrA*
Do et al. (2009a) find
observed
mJy (
)
or dereddened
mJy
(assuming their value of
)
as the faintest observed
flux density
(see also Hornstein et al. 2002 and Eckart et al. R2006).
Using
this results in a flux density of
mJy
For the low flux density state Do et al. (2009a) reported red colors and variability
compared to stellar sources. For the very low luminosity levels
they report with an average power law exponent of
a bluer spectrum than that reported earlier
by Hornstein et al. (2002) with a power law slope of
.
The contamination from nearby (predominantly blue) stars is estimated to
contribute a maximum of about 35%
of the flux to account for this difference in the spectral slope.
Do et al. (2009a) conclude that even when the emission is
faint a large fraction of the flux arising from the location
of SgrA* is likely non-stellar and can be attributed
to physical processes associated with the black hole.
Our data allow us to provide a new independent estimate on the flux density of SgrA* in some of its lowest states based on VLT data.
In a circular aperture of a diameter of
66 mas diameter (5 pixels or about one resolution element) we
measure dereddened flux densities
(using AK = 2.8)
1.5 0.4 mJy,
1.3
0.3 mJy, and
1.3
0.2 mJy for the
low-pass filtering,
iterative, and
automatic PSF subtraction method, respectively.
![]() |
Figure 8: The 30 August 2004 light curve of the flux density measured in a 40 mas radius circular aperture centered on the position of SgrA* (black). We show the light curve of the reference star S2 in green. The bottom panel shows the FWHM seeing in arcseconds measured on the PSF extracted for the field. The grey points show the background counts near SgrA*. The width of the error bars is close to the size of the data points. |
Open with DEXTER |
The uncertainties have been derived
from the flux density measurements in the central 5 5 pixel
aperture and the upper limits of the measurement
uncertainties per pixel as plotted in Fig. 7.
These upper limits are
0.33 mJy,
0.24 mJy
and 0.17 mJy for the three methods, respectively.
Based on these data we derived an upper limit on
the formal uncertainties in any random 5
5 pixel aperture
in the central 0.6
diameter field of
0.066 mJy,
0.048 mJy
and 0.034 mJy for the three methods, respectively.
We adopted these values for the central aperture.
In order to obtain a conservative estimate we added in
each case these values five times,
so that the final uncertainty amounts to six times the value expected for
an aperture of this size - based on the radial averages.
There is no clear detection of a source at that position.
If the observed flux density limit is fully attributed to
SgrA* this corresponds to a clear non-detection of any point
source at that position with a flux density limit of 2.4 mJy for SgrA*
(
16.3; observed reddened magnitude).
If all of it is stellar, the upper limit is defined by the uncertainty of the measurement
and the limit for a flux density from a non-stellar source at the position of SgrA* is 0.9 mJy
(3
,
de-reddened and
as an observed reddened magnitude).
While our data do not allow us to confirm the existance of a constant quiescent state,
our flux density limit is consistent with the flux given by Do et al. (2009a).
This flux density limit is also consistent with the light curves of SgrA* shown
in Figs. 8 and 9.
They were obtained in a 40 mas radius aperture centered at the position of SgrA*.
The light curves were taken without removing any stars
from the sourroundings and the aperture size of only 40 mas
was chosen to minimize the contamination
from neighboring stars.
The bottom parts of Figs. 8 and 9 show the FWHM of the PSF extracted on
a nearby star. These graphs show that the seeing and AO performance were very stable
and similar during the measurements on both dates.
Especially in the second half of the 30 August light curve in Fig. 8 SgrA*
was sufficiently bright to be detected, while it was in a low state during the
entire time during which the 23 September light curve in Fig. 9 was taken.
Our flux density limit for the 23 September data is consistent with the completeness
limit we reached in the central star counts
(see Sect. 4.1).
![]() |
Figure 9: The same as Fig. 8 but for 23 September 2004. |
Open with DEXTER |
4.3 Bremsstrahlung emission from SgrA*
4.3.1 The quiescent state X-ray emission from SgrA*
Baganoff et al. (2001, 2003) reported an extended (R>103 )
thermal bremsstrahlung source co-spatial with the position of the
strongly variable point source at the location of SgrA*
(see also Quataert 2003).
The authors derive
1.4
0.14
for the intrinsic FWHM size
of the source. This corresponds to
for a
black hole.
The scale of the extended structure centered on SgrA*
is consistent with the expected
Bondi accretion radius (1
-2
;
Bondi 1952) for
matter accreting hydrodynamically onto the SMBH.
For the bright flare emission of the X-ray point source
bremsstrahlung is not a likely mechanism to produce
the observed NIR flux density excursions
(Genzel et al. 2003; Yuan et al. 2003; Quataert, 2002).
In part motivated by the blue spectrum reported by Do et al. (2009a)
- assuming that not all of the blue character is due to
stellar contamination -
we investigate here if the lowest NIR flux density values
can be explained by a bremsstrahlung contribution from
a hot disk.
4.3.2 Possible low state bremsstrahlung NIR luminosity?
We now estimate the possible contribution of the X-ray emitting gas that is associated
with the compact disk or jet footpoint of SgrA* to its Ks-band luminosity.
The frequency and temperature dependency of the volume emissivity for the bremsstrahlung
is
d
g(
,T)
exp(-
)d
.
Here the Gaunt factor is
.
For the X-ray luminosity in the 2-10 keV range of a H/He plasma this
results in
with
the source radius R given in pc, the electron density
given in cm-3,
and the temperature T8 given in unity of 108 K.
The Gaunt factor
implies that for a change in the observing frequency from
1018 Hz in the X-rays to
1014 Hz in the NIR
for constant
the Gaunt factor changes significantly
by a factor of ln(104)
10.
So in total the Ks-band luminosity expected
over this frequency range
is about 1000 times lower than the X-ray luminosity.
However, already for a
1 mJy Ks-band flux density contribution and high temperatures above
108 K the predicted X-ray flux would well exceed the measured values.
Substantial flux density contributions due to bremsstrahlung can only occur if a
sufficient amount of ``cooler'' gas (i.e. 108 K) is mixed into the central accretion flow or disk.
Such a two phase medium has already been discussed in the context of flares from SgrA*
(Yuan et al. 2003; see also Page et al. 2004; Nayakshin & Sunyaev 2003).
A possible quiescent phase at low infrared Ks-band flux density could in principle be provided
through bremsstrahlung emission from a
cooler denser phase embedded in the hot accretion flow or in the outer parts of a potential accretion disk.
In luminous AGN a cool optically thick disk
with
K is believed to coexist with a hot optically thin corona with
K.
At the lower SMBH mass and luminosity of SgrA* such material
could have an even lower temperature.
However, a lower limit to the temperature of such cool inclosure is given by the fact that no
strong recombination line emission is observed towards SgrA*.
Only for temperatures above 106 K the contribution to the total luminosity
from thermal bremsstrahlung is greater than that from recombination radiation.
This implies for 106 K a Ks-band luminosity of
.
At a distance of
8 kpc and a wavelength of 2.2
m
1 mJy of dereddened flux density corresponds to a luminostity
of
erg/s = 3.1
.
In order to obtain 1 mJy Ks-band flux density the product
needs to be
on the order of
.
This allows us to estimate the required densities:
The Ks-band diffraction limit of the VLT UT4 corresponds to a projected linear size
of 2.4
10-3 pc
8000
at the distance of SgrA*.
Similarly, 1
corresponds to
pc.
For such an unresolved or even more compact source with sizes
ranging from 1 to 8000
we find
electron densities from
12 cm-3 to
6 cm-3
and Thompson scattering optical depths on the order of 1 to 0.01, respectively.
This implies that for an ``unresolved'' bremsstrahlung emitting Ks-band source the
electron density is several orders of magnitude higher than the estimated
density of 107 cm-3 of hotter gas
in the accretion flow (Yuan et al. 2003).
Combined with the high optical depths this shows that a substantial
bremsstrahlung contribution
of a compact source to the Ks-band flux density
even in the quiescent phase
can therefore be regarded to be very unlikely
(see also Yuan et al. 2003).
4.4 Synchrotron self-Compton emission from SgrA*
Rapid variability on time scales of a few minutes and the strong indication of synchronous (polarized) NIR and X-ray flux density variations linked with time delayed sub-mm/mm flux excursions clearly suggests that a nonthermal emission mechanism is at work. Synchrotron emission from compact source components that unavoidably scatter electromagnetic power into the X-ray domain appear to be a prime choice for such a mechanism. Below we investigate the highest and lowest luminosity states to validate the applicability of the SSC mechanism.
4.4.1 Essentials of current SgrA* SED modeling
Generally the electron energy distribution N(E) responsible for the
overall SED of SgrA* is described by a power-law
with a Lorentz factor
and an electron power-law index p.
This power-law exhibits several breaks at
![]() |
(8) |
where











A further usually higher frequency break occurs due to a
modification of the underlying relativistic electron distribution
as a consequence of synchrotron cooling losses
(Özel et al. 2000; Kardashev 1962).
For an electron power-law index p
and for Lorentz factors larger than
the resulting optically thin electromagnetic spectrum
will follow a power-law distribution with a spectral index
.
In the case of no further injection of fresh electrons with
the powerlaw index p the spectrum will steepen at the frequency
towards a spectral index of (2p-1)/3.
If fresh electron injection occurs the spectral index will
only steepen towards p/2 and is therefore closer to the
original index of
.
At this frequency
the synchrotron cooling time equals
the time within which the relativistic electrons can escape the source.
In the case of SgrA* and at about 3.5
(approximate radius of the last
stable orbit) it can be written as
![]() |
(9) |
(Kardashev 1962; Özel et al. 2000; Yuan et al. 2003; Dodds-Eden et al. 2009). The frequency

![]() |
(10) |
(e.g. Marscher 1983). For B>30 G and





In most cases modeling of the overall SED of SgrA* assumes that the synchrotron quiescent emission (see Fig. 1 in Yuan et al. 2003) is dominated by the relativistic tail of the otherwise thermal Boltzmann distribution that describes the quiescent (bulk) of the radio to sub-mm/FIR spectrum. Flares are then attributed to a variation in the distribution of this relativistic part of the accretion flow. In addition the bulk of the quiescent X-ray spectrum is explained by thermal bremsstrahlung due to the extended X-ray bright Bondi sphere surrounding SgrA* (Baganoff et al. 2001, 2003; Liu et al. R2006; Yuan et al. 2003; Quataert 2002; Mahadevan & Quataert 1997).
Here we consider the following additional points: the flare spectrum of SgrA* can be explained independently of the steady, quiescent emission from the accretion flow. We consider ``fresh'' synchrotron components that become optically thick at THz frequencies and inverse Compton scatter into the X-ray domain. We also investigate synchrotron, bremsstrahlung, and thermal emission from the foot point of the accretion flow or a temporary disk as possible significant contributors for the unresolved source components seen in the NIR during the low luminosity states of SgrA*.
4.4.2 Description and properties of the SSC model
We have employed a simple SSC model to describe the observed radio to X-ray properties of SgrA* with the nomenclature given by Gould (1979) and Marscher (1983). Inverse Compton scattering models provide an explanation for both the compact NIR and X-ray emission by up-scattering sub-mm-wavelength photons into these spectral domains. These models are considered as a possibility in most of the recent modeling approaches and may provide important insights into some fundamental physical properties of the source. The models do not explain the entire low frequency radio spectrum and the bremsstrahlung X-ray emission that dominates the luminosity between X-ray flares. They do successfully account for the flare events observed in recent radio to X-ray campaigns though (Eckart et al. 2003, 2004, R2006a,b, 2008a,b, 2009; Yusef-Zadeh et al. R2006a,b, 2007, 2008; Marrone et al. 2009).
We assume a synchrotron source of an angular extent .
The source size can be as big as a few Schwarzschild radii.
The emitting source is assumed to become optically thick at a frequency
with a flux density
and has an optically thin spectral
index
following the law
.
This allows us to calculate the magnetic field strength B and the
inverse Compton scattered flux density
as a function of the
X-ray photon energy
.
The synchrotron self-Compton spectrum
has the same spectral index as the synchrotron spectrum that is
up-scattered
i.e.,
,
and is valid within the
limits
and
corresponding to the wavelengths
and
(see Marscher et al. 1983, for
further details).
The emitting electrons follow a power-law distribution
with energies
.
Here
and
are the electron
Lorenz factors.
Maximum Lorentz factors
for the emitting electrons of the order of
typically 103 are required to produce a sufficient SSC flux in the
observed X-ray domain.
A possible relativistic bulk motion of the emitting source results
in a Doppler
boosting factor
-1(1-
cos
)-1.
Here
is the angle of the velocity vector to the line of sight,
the velocity v in units of the speed of light c, and
= (
2)-1/2 Lorentz factor for the bulk motion.
Relativistic bulk motion
is not needed to produce sufficient SSC flux density, but
we have used modest values for
-2 and
ranging between 1.3 and 2.0 (i.e. angles
between
and
)
since they will occur
in cases of relativistically orbiting gas as well as relativistic
outflows - both of which are likely to be relevant to SgrA*.
The electron spectral number density in units of cm
is given by
![]() |
(11) |
Here




![]() |
(12) |
This quantity can be used to validate the emission mechanism in comparison to other number density estimates obtained for different source regions.
![]() |
Figure 10:
Sketch of the SgrA* surroundings within a 105 Schwarzschild radius.
The SMBH encloses the central 1 |
Open with DEXTER |
The radio polarization data of SgrA* (Bower et al. 2004; Marrone et al. 2007)
indicate that of the total available stellar mass loss of
10-3
yr-1only a small fraction of
10-7
yr-1is close to the central black hole and available for accretion.
This applies a mean daily accretion of
.
Assuming a flare lasts for 100 min this corresponds to
an accretion mass load of typically
per flare.
The SSC models in Table 5
result in estimates that
lie one to five orders of magnitude below this upper mass limit per flare.
Here models A
to A
,
in which the NIR fluxes are produced via optically
thin synchrotron radiation, are favored since the required masses lie well below
the available accretion mass load (as it is for the quiescent models).
4.4.3 SSC model of the bright flares of SgrA*
For X-ray flares of up to several times the quiescent emission the SSC models with source components that become optically thick in the few THz domain provide a successful description of the compact flare emission that originates from the immediate vicinity of the central black hole (e.g. Eckart et al. 2009, and references therein; Marrone et al. 2009, see also Fig. 10). This description also allows the explanation of the correlation between NIR/X-ray flares and time delayed radio flares (e.g. Eckart et al. 2008a,b, 2009; Yusef-Zadeh et al. 2008; Marrone et al. 2009). In order to validate the applicability of this emission mechanism it is useful to study the bright flare emission associated with SgrA*.
In Table 1 we list important properties
of the three brightest X-ray flares that have been observed to date
(Baganoff et al. 2001; Porquet et al. 2003, 2008).
The NIR flare emission shows observed spectral indices of
0.6 (Ghez et al. 2005a,b; Hornstein et al. R2006) or even
steeper (Eisenhauer et al. 2005; Gillessen et al. R2006).
If the X-ray flare flux density is due to the SSC spectrum of the synchrotron component that
gives rise to emission in the NIR then the spectral indices should be the same.
While a spectral index of
0.6 is consistent with the Chandra value,
the XMM flares require a steeper spectral index.
The spectral index averaged over several flares observed with Chandra results in
(Baganoff et al. 2001).
This is consistent with both the XMM data and the
index of
reported by Ghez et al. (2005a,b) and Hornstein et al. (R2006).
General modeling results:
our modeling results show that synchrotron flare components that become
optically thick in the few THz domain stay optically thin throughout the
MIR/NIR domain, and self-Compton scatter into the X-ray domain can
fully account for the observed properties of the brightest flares from SgrA*.
We use maximum electron Lorenz factors
103 resulting in the
fact that the X-ray emission is always inverse self-Compton scattered.
This delivers synchrotron radiation up to near-infrared wavelengths and results
in the fact that the X-ray emission is always inverse self-Compton scattered.
Source component parameters for SSC models of the brightest
X-ray flares from SgrA*
are given in Tables 4 and 5.
Table 4: Source component parameters for SSC models of the brightest X-ray flares from SgrA*.
Here model labels (Col. 1) ,
and
indicate the
bright flares listed in Table 1.
Model labels A and B indicate models in which the the MIR/NIR flux
densities are accounted for in different ways.
In particular
in models A
-
the observed peak MIR/NIR flux densities are accounted for by the
high frequency end of the synchrotron spectrum.
In models B
-
they are accounted for
by the low frequency end of the inverse self-Compton scattered spectrum.
Among themselves the A models (straight red line in Fig. 11)
and the B models (black long dashed line in Fig. 11)
look very similar for the different flares labeled with
,
and
in Fig. 1.
Within the uncertainties infrared and X-ray flux densities as well as
the X-ray spectral indices are met by the
required model spectral indices. As an example (and since it this model is
discussed in more detail) we show
the synchrotron flare spectra and the
corresponding SSC spectra in Fig. 11
for the models A
and B
.
A global variation of a single parameter
by the value listed in the first row and the corresponding column of
Table 4 results in an increase of
.
Here global variation means:
adding the 1
uncertainty
for a single model parameter but for all source components
in a way that a maximum positive or negative
flux density deviation is reached.
Models with significantly different component fluxes, sizes, and
cut-off frequencies fail to match the observed fluxes by more than 30%.
A more detailed description of the modeling procedure is
given in Eckart et al. (2009).
Models A
and B
give the source component for a single and multiple spot
(multiples of the dominant spot described by model Q
)
quiescent state model.
There are two approaches to model the NIR/X-ray spectra:
1) The observed peak MIR/NIR flux densities can be accounted for either by the
high frequency end and of the synchrotron spectrum (models A-
,
in
Tables 4 and 5)
or 2) by the low frequency end of the inverse self-Compton
scattered spectrum (models B
-
).
Here the upper cutoff frequency of the scatter spectrum
has a special importance.
With
and
we find that this cutoff lies in the few micrometer MIR/NIR domain.
With
it lies at about 1 keV just below the lower
energy cutoff of the X-ray Chandra and XMM observatories.
In general the models A
to A
are constrained by the
synchrotron spectral index
defined by the THz and NIR/MIR flux densities.
Models B
-
are constrained by the
SSC spectral index (that equals
)
defined by the NIR/MIR, the X-ray flux densities and spectral index.
Therefore
cannot be steeper than about unity
in these SSC models.
Possible peculiarities in the relativistic electron spectrum:
a modification in the upper cutoff of the relativistic electron spectrum
(e.g. Lorentz factor )
may also have a strong influence on the X-ray output of the emitting source component.
An increase in
may reflect a more efficient acceleration mechanism.
Synchrotron losses will lead to a
decrease in
or to the involvement of a cooling break.
The width of the relativistic electron distribution for flare models
A
to A
is narrower than that of models B
to B
or even
models which explain additionally the low frequency radio part of the SgrA* SED
(Yuan et al. 2003; Quataert 2002; Mahadevan & Quataert 1997).
Kardashev (1962) points out that the higher frequency synchrotron
cooling break does not become displaced towards
lower frequencies with time if the energy distribution of the injected
relativistic electron spectrum has a very small width.
A narrow relativistic electron spectrum may therefore support a less
variable cooling break frequency during the main part of the flare.
Table 5: Model densities, lower SSC cutoff frequencies and lower limits on the source component masses.
For models A the lower energy cutoff

The April 4, 2007, flare: the flare event of April 4, 2007 (see Table 1), has also been simultaneously observed at the 11.8


![]() |
Figure 11: Two SSC models A (red) and B (black dashed) describing the April 4 2007, flare data (see Table 1, Tables 4 and 5). |
Open with DEXTER |
The SSC model that we present here for the the April 4, 2007, flare is not covered by Dodds-Eden et al. (2009). Here we cover it in more detail and take it as an example for the bright luminosity state of SgrA*. The results may be used to explain the properties of mm/sub-mm flux density variations that have been observed in several other cases (e.g. Eckart et al. 2008a,b, 2009; Yusef-Zadeh et al. 2008; Marrone et al. 2009).
For the April 4, 2007 flare event
our SSC model A
agrees with the simultaneous 3.8
m flux density, the 3
flux limit at 11.8
m and also the
spectral index in the X-ray domain corrected for hydrogen absorption.
A spectral decomposition is given in Fig. 11 and a comparison of the
light curves with model A
and
(see below) is displayed in Fig. 12.
For the favored model A in Fig. 11 the relevant cutoff frequencies are
labeled and marked with a vertical thin solid line.
For model B these frequencies are all close to
13 Hz.
The model data are compared to the SED of SgrA* in its quiet state.
The radio to sub-mm measurements (Markoff et al. 2001; Zhao et al. 2003)
are time averaged measurements, and the error bars include variable emission
of up to 50%. As open black circles we show 30
m, 24.5
m and 8.6
m
upper limits taken from Melia & Falcke (2001).
The X-ray information is taken from Porquet et al. (2008).
The short dashed line shows the quiescent model by Yuan et al. (2003).
With a magnetic field of 10 G the optically thin synchrotron spectrum extends up to
= c (3.4
m)-1 with a spectral index of
.
Above this the spectrum steepens until the upper synchrotron cutoff
(0.85
m)-1.
For a spectral index of unity, i.e.
,
and
with the assumption of a fresh injection of relativistic electroms the spectrum will
steepen to a value of
=
.
This value fully agrees with the
spectral index between the infrared H- and Ks-band of
which was used as an additional fitting criterion by Dodds-Eden et al. (2009).
The model can be further validated by comparing the involved electron spectral number density
with values inferred from other data or different models.
The value of 14 cm
of the electron spectral number density
results in number densities for the relativistic electrons of
cm-3.
Part of the emitting particles may be due to electron/positron pairs.
Integrating over the SSC source component volume of 0.2 R0 diameter we
obtain
as a lower limit to the total mass of this source
component within a compact jet base that feeds an extended
overall jet or outflow with a low surface brightness
(Falcke & Markoff 2000, 2001; Markoff 2005; Markoff et al. 2007).
If a structure of the size and density described above expands
at highly relativistic velocities,
it can reach the flow density of 107 cm-3 (Yuan et al. 2003).
This can happen on an orbital time scale
of a spot which is close to the last stable orbit
and on less than one flare time scale
(see Fig. 10).
![]() |
Figure 12:
A combined SSC and adiabatic expansion model compared to the
SgrA* flare event on April 4, 2007.
The data are described in
Porquet et al. (2003), Dodds-Eden et al. (2009), Yusef-Zadeh et al. (2009).
The 240 GHz and 43 GHz response of the NIR/X-ray flare is plotted
as a solid and dashed red line, respecitvely. The probably unrelated 43 GHz response
of the flare component |
Open with DEXTER |
For the SSC model B
explains the NIR flux densities shortward of about 4
m through
the self-Compton scattered spectrum the
electron spectral number density and source component mass.
This mass is four orders of magnitudes
higher than for model A
.
Reaching the flow density of 107 cm-3 in one flare time scale
and matching the
total estimated accretion rate onto SgrA* is possible as well.
Still, the predicted 11.8
m flux density is well above the upper limit
obtained for the April 4, 2007 flare. Therefore models in which the Ks- and L-band
flux densities are produced via the self-Compton scattered spectrum are
not favored.
Modeling the April 4, 2007 NIR/X-ray light curves:
Dodds-Eden et al. (2009)
report that the width of the X-ray flare is significantly smaller
than the on of the corresponding overall L-band flare.
This implies that the efficiency of producing X-ray emission was lower
before and after the main part of the flare.
In the framework of the evolving spot model presented in Eckart et al. (2008a)
this may be explained by an evolution of the spot characteristics
similar to the models presented by Hawley & Balbus (1991, 1998) and Yuan et al. (2008).
The recent theoretical approach of hot spot evolution due to shearing is
highlighted in Eckart et al. (2008a) and Zamaninasab et al.
(2008, 2009; see also Pechácek et al. 2008).
Yuan et al. (2008) explicitly solved the relativistic hydrodynamics and
found initial expansion velocities of less than 0.01c close to
the accretion disk followed by relativistic expansion towards the end of the flare.
A compression of a single spot or merging of two spots
due to differential rotation within a viscous disk may explain the
different widths of the flares in the X-ray and L-band light curve.
A variation of the source size
or the peak flux density
of the scattered synchrotron component
by only 20% to 30% will result in the
required significant increase or decrease in the SSC scattering efficiency, i.e. the scattered
SSC X-ray flux density
.
Similarly other differences between the NIR and X-ray light curves may be due to different
or even varying flux density contributions and scattering efficiencies
from a source component responsible for an underlying main flare and components
responsible for the sub-structures.
Modeling the April 4, 2007 radio light curves:
Yusef-Zadeh et al. (2009) report radio measurements that were
carried out at 240 GHz (30 m IRAM) during and at 43 GHz (VLA)
following the April four NIR/X-ray flare.
The 240 GHz data cover the entire NIR/X-ray flare event at a
flux density level of about 1.2 Jy with no significant flux density variation.
After a (``transatlantic'') gap in the data a strong increase and decrease
of the 43 GHz flux density with a peak of about 0.4 Jy is
detected. The entire 43 GHz flare has a full bottom width
(at a level of about 1.1 Jy) of about 4 h
and peaks at 43 GHz about 6.5 h past the NIR/X-ray peak.
Performing an adiabatic expansion calculation (van der Laan 1966)
based on models A
and B
using an expansion speed of 0.007c
places the 43 GHz afterglow of the NIR/X-ray flare close to
or even into
the gap between the 100 GHz and 43 GHz measurements.
We use 0.007c as a typical value close to the expansion speeds found for
other SgrA* flare events for which adiabatic expansion models have been
applied (see e.g. Eckart et al. 2008b, 2009; Yusef-Zadeh et al. 2008).
As an example of such a calculation we show in Fig. 11
a combined SSC and adiabatic expansion model
based on the data for A
given in Table 4.
Here we assume that the 240 GHz and 43 GHz flux density level
above which we have to consider the excess flare flux is at
1.08 Jy - which is at the minimum flux density level
given by the data for this event.
The predicted 240 GHz and 43 GHz peak brightnesses
lie well within the noise of the IRAM measurements and the
following 43 GHz data.
A lower expansion velocity like 0.002c gives the time delay needed
to meet the bight 43 GHz flare peaking at about 12:30 UT
but still fails to reproduce its high radio flux.
Since measurements around this time have not given any hint for a possible
NIR or X-ray counterpart for this event, it can only be explained
by a low frequency flare that originated from a synchrotron component
with low turnover frequency and no significant predicted NIR or X-ray flux density
(like component
given in Table 4).
We therefore regard component
as unrelated to the main NIR/X-ray flare
which happened six hour earlier.
The SSC models listed in Table 4 have small source component sizes
to provide a sufficient inverse Compton scattering efficiency to
explain the large observed X-ray flux densities.
This results in comparatively narrow light curves of the individual components
and requires the observed broader flare profile
to be modeled with six source components.
This may explain the presence of substructure in the NIR and X-ray light curves.
We place the components 1 through 6 at 5:24, 5:37, 5:51, 6:00, 6:15, and 6:21 UT
with the flux ratios between them of 0.6:1:1:0.8:0.5:0.4.
We also allowed components 1, 5, and 6 to be larger
by
compared to the central components.
An increase in size or a decrease in flux density lowers the scattering
efficiency (see above) in a way that the weaker components 1, 5, and 6 do not
contribute significantly to the X-ray flare.
With a spectral index of 1.05 and the dependency in scattering efficiency given
above this results in an X-ray flux density ratio
between the components of 0.01:1:1:0.3:0.003:0.001.
This may explain the different widths of the flares in X-ray and L-band light curve
and allows us to model the broader NIR flare profile.
Model B
in Table 4 is not well suited to explain the wings of the
NIR light curve in comparison to the X-ray flare profile since the flux densities
in both wavelength domains are described by the same SSC scattered spectrum.
Involving steeper SSC spectra for the wings violates the 11.8
m flux
density limit.
Therefore the 4 April, 2007, flare event provides additional support
for the assumption that - at least in this case - the NIR luminosity is due to
optically thin synchrotron emission rather than SSC scattered flux.
4.4.4 SSC modeling of a weak X-ray flare
In Tables 4 and 5 we also list modeling
results for the so far weakest statistically significant
X-ray flare that was modeled (Eckart et al. 2002). This flare was also the first that
was detected simultaneously in both the NIR and X-ray domain.
The SSC model
which explains the NIR flux densities through the high frequency end of
the associated synchrotron spectrum describes the flare successfully.
SSC models (B in Tables 4 and 5)
that explain the NIR fluxes through the inverse Compton scattered part of the spectrum
are strongly constrained by the NIR/X-ray spectral index determined by
.
While they do not result in a satisfactory fit of the observed
data for the very bright X-ray flares
(see Sect. 4.4.3),
the situation for weak X-ray flares is different.
Here
and the corresponding
electron spectral number density and source component masses are of the same order as for the
SSC models A
.
Therefore both model approaches, A
and B
,
result in acceptable explanations
for the NIR and X-ray fluxes and the spectral indices during weak flares.
4.4.5 SSC modeling of the lowest states of SgrA*
Eckart et al. (R2006) have modeled the
emission from SgrA* in its low flux density state.
They find that the upper limits of the compact X-ray emission in the ``interim-quiescent'' (IQ),
low-level luminosity states of SgrA* are consistent with
a SSC model that allows for substantial contributions from both the
SSC and the synchrotron part of the modeled spectrum.
In these models the X-ray emission of the point source is well below
20-30 nJy and contributes considerably less than half of the X-ray flux density
during the weak flare event reported by Eckart et al. (2004).
The flux densities at a wavelength of 2.2 m are of the order
of 1 to 3 mJy
(see Table 2 and Figs. 8 and 9).
Eckart et al. (R2006) list representative SSC models for the low flux state.
For their models IQ1-IQ3
the source component has a size on the order of 0.2 to 2 Schwarzschild radii
with an optically thin radio/sub-mm spectral index ranging from
1.0 to 1.3, a value similar to the
observed spectral index between the NIR and X-ray domain.
In Tables 4 and 5 we give the detailed
model data for a single spot in which the
MIR/NIR flux densities are accounted for by the
high frequency end of the synchrotron spectrum (A
)
or
by the low frequency end of the inverse self-Compton scattered spectrum (B
).
These models account for the quiescent source spectrum given by Yuan et al. (2003),
as shown in Fig. 13.
For model B
the electron spectral number density value
and total masses integrated over the SSC source component volume
are large for a quiescent state compared to the flaring state
(
3
1011 cm-3 and
)
and to the assumed accretion flow density of 107 cm-3.
Therefore model A
is favored.
![]() |
Figure 13: A single spot SSC quiescent spectrum (black dashed) and a corresponding multi-spot SSC quiescent spectrum. |
Open with DEXTER |
As an alternative to a larger continuous disk component one may assume a
continuous disk or a disk consisting of several spots with dominant source
components that contribute only about 0.1 mJy (model Q
)
to the Ks-band flux density each and have
electron spectral number densities on the order of
107 cm-3
and total masses integrated over small SSC source component sizes
of 0.1
with a diameter of
a few times 10-19
.
Here the magnetic field strengths are comparable to those of the larger flare
components, and the densities are comparable to the
assumed accretion flow density of 107 cm-3.
In Fig. 13 - as an example -
20 spots as described by model Q
in
Tables 4 and 5 produce a significant part
of the quiescent spectrum of SgrA* in the Ks-band, as one would
expect for a low level quiescent state of SgrA*.
The 20 spots we used in Fig. 13 have different relativistic
boosting factors as they would occur on an inclined circular orbit of a few
Schwarzschild radii around the SgrA* SMBH
(see Tables 4 and 5,
and data in Eckart et al. 2004).
In Fig. 13 relevant cutoff frequencies are
labeled and marked with a vertical thin solid line.
The quiescent X-ray information is taken from Baganoff et al. (2001, 2003).
The short dashed line shows the quiescent model by Yuan et al. (2003).
The double-dot-dashed line shows the contribution of the relativistic electrons to this
quiescent model.
Therefore both a larger continuous synchrotron emitting disk or a faint disk spotted with SSC components are attractive models that may explain a possible low state of SgrA*. Any quiescent state flux density (e.g. measured in the NIR) would then be expected to be variable and polarized. High angular resolution and sensitivity - as it will be provided by the future large telescopes (ELTs) - are required to measure the quiescent state flux density contribution with sufficient precision.
4.5 Thin disk NIR luminosity
As another way to explain the Ks-band flux density especially in an assumed
quiescent state of SgrA*, we explore the possible presence of a small
permanent disk that radiates as a black body.
Eckart et al. (R2006) show that at least for the flaring state of SgrA* a
disk may exist. For an optically thick disk seen at an inclination angle i
and radiating like a black body one finds a luminosity of
![]() |
(13) |
(Peterson 1997; see also Ishibashi & Courvoisier 2009; Kishimoto et al. 2008). At infrared wavelengths and shorter the spectrum is dominated by the emission of the inner disk radii. In the NIR Ks-band 1 mJy corresponds to 3









5 Summary and conclusion
We have discussed the variable infrared and X-ray emission of SgrA*
in its extreme luminosity states.
We reported on Ks-band imaging of the
SMBH counter part of SgrA* in its low state
and investigated the structure and brightness of the central
S-star cluster that surrounds the SMBH at the position of SgrA*.
We have used three independent methods to remove or strongly suppress the
flux density contributions of stars in the central 2
diameter region
around SgrA*.
All methods revealed faint extended emission around the SgrA* position.
De-reddened with AK=2.8 the peak emission of the extended S-star cluster is
about 1.3 mJy within one resolution element,
corresponding to a surface flux density of about 0.5 Jy (dereddened) per square arcsecond.
For the dataset taken
on 23 September 2004 no NIR counterpart of SgrA* was detected with a
flux density limit of about 2 mJy (dereddened).
For 30 August 2004 we find a NIR counterpart of SgrA* with a
dereddened flux density of about 4 mJy.
We find that the luminosity during the low state can most likely be accounted
for by synchrotron emission from a continuous or spotted accretion disk.
In this case we expect the possible quiescent source associated with SgrA* to be
significantly polarized.
Steep spectral indices observed in the NIR wavelength domain may
reflect the presence of typical cutoff frequencies.
In addition to the cooling break
as discussed in Sect. 4.4.1
there may be a cutoff in the relativistic electron spectrum at work
(e.g. Eckart et al. R2006a; Liu et al. R2006).
This will result in a modulation of the intrinsically flat spectra
with an exponential cutoff proportional to
(see e.g. Bregman 1985; and Bogdan & Schlickeiser 1985)
and a cutoff wavelength
in the infrared.
If
lies in the 4-8
m wavelength range (Eckart et al. R2006a),
the variation in the spectral
index is on the order of
1.0.
In a number of extragalactic jets
these cutoffs have been observed to be relevant in the NIR wavelength domain
(3C 293: Floyd et al. R2006,
M 87: Perlman et al. 2001; Perlman & Wilson 2005;
3C 273: Jester et al. 2001, 2005).
In these sources the cutoff frequency at which synchrotron losses become dominant
is
Hz (or 7.5
m wavelength),
which is remarkably similar to what may be required in the case of
the Galactic center (Eckart et al. R2006a).
For SgrA* such a cutoff can quite naturally explain the steep NIR spectral indices reported by
(Eisenhauer et al. 2005; Gillessen et al. R2006; Krabbe et al. R2006) despite the
low upper flux density limits in the MIR.
For the three brightest X-ray flares the SSC emission from THz peaked source components
can fully account for the observed flux density variations observed in the
NIR and X-ray domain.
Here models in which the MIR/NIR flux density contributions are due
to the high frequency tail of the associated synchrotron spectrum are favored
(models A in Tables 4 and 5).
Combined with the source component size of 0.2-0.3 R0 and relativistic expansion
within or towards the end of the flare, the spectral energy densities of
relativistic electrons are compatible with the electron number density derived for the
accretion flow at 10
distance
and the estimated upper limits of the mass accretion rate onto SgrA*.
For the weak X-ray flare first discussed by Eckart et al. (2004) these boundary conditions
are fulfilled for both model approaches (A and B).
For model A
we expect significant polarization of the NIR flux density,
whereas for model B
the scattering process should lead to a significantly
lower degree of polarization.
For the 4 April 2007 flare our favored model A
of the peak fluxes
meets the 11.8
m 3
limit and is only by 1
off the expected
flux density at 3.8
m wavelength. In the X-ray domain it matches the mean flux exactly and the
spectral slope by 0.8
.
Giving these four quantities the same weight we obtain
a reduced
value of
/d.o.f = 3.7.
The X-ray data cover a comparatively large and rather densely sampled spectral range.
We get
if the X-ray information is weighted higher by a factor of two or more
(see Porquet et al. 2008; Dodds-Eden et al. 2009).
A
-based comparison of different single component emission models
over a large wavelength range (radio or infrared till X-ray) is, however,
only of limited value.
More adequate physical models of the compact emission from SgrA*
contain on some level contributions from several source components, like
the hot-spot/disk (Eckart et al. R2006b; Broderick & Loeb R2006)
or multi-spot models (Eckart et al. 2008a; see also Pechácek et al. 2008),
the spiral arm models (Karas et al. 2007; Falanga et al. 2007),
the more elaborate evolving hot-spot model by (Yuan et al. 2008),
or the jet/jet-base models
(Falcke & Markoff 2000, 2001; Markoff 2005; Markoff et al. 2007).
This implies that in addition to a possible dominant emission component,
secondary source or spectral components will contribute to the flux
and spectral index information in
the different wavelength bands across the electromagnetic spectrum.
This complicates the comparison of different models.
Distinguishing between emission mechanisms
requires more work on detailed flare modeling, using time dependent
dense flux density sampling across a broad frequency band, preferentially including
polarization and spatial information.
Here we presented a new SSC model for the flare event on April 4, 2007 that accounts for the available simultaneous peak flux density and spectral index information. Variations in the SSC scattering explain that the flare profile in the X-ray domain is narrower than that at NIR wavelengths. This suggests at least for this flare event that models in which the NIR flares are due to optically thin synchrotron radiation rather than inverse Compton scattered light are preferred. In general models in which the MIR/NIR flux densities are explained through the low frequency tail of the inverse Compton scattered spectrum are disfavored. They result in higher masses for the emitting sources.
We conclude that both weak and bright flares from SgrA* can be explained by the synchrotron self-Compton mechanism. This emission mechanism may explain the flare activity over up to two orders in magnitude in the NIR or X-ray flux density and over six orders of magnitude in frequency. No other dominant emission mechanisms are required. Densities of relativistic electrons of a few times 107 cm-3(comparable to what is derived from X-ray data on a 1000 Schwarzschild radii scale) are also required close to the SgrA* black hole to explain the flare emission.
The variable synchrotron emission
may be limited at short wavelengths by a NIR cutoff (cooling break) due to
synchrotron losses.
The self-Compton scattered spectrum may be limited at the longest wavelengths by the
low energy cutoff in the relativistic electron distribution.
If future sub-mm and X-ray obervations allow for a simultaneous detection of
the THz turnover frequency (
)
and a low energy cutoff of the SSC spectrum
(
),
information on the low energy
cutoff in the relativistic electron distribution can be obtained via
1
.
In a few years further polarimetric NIR observations
at high angular resolution and sensitivity
will help to distinguish between the importance of the different
emission mechanisms in the low state of SgrA*.
Interferometric observations in the near-infrared using GRAVITY
at the VLTI (Eisenhauer et al. 2008; Straubmeier et al. 2008; Zamaninasab et al. 2009)
and in the mm-wavelength domain
(e.g. Fish et al. 2009; Doeleman et al. 2009)
will be important to understand the emission from SgrA*.
N. Sabha is member of the Bonn Cologne Graduate School for Physics and Astronomy. This work was supported in part by the Deutsche Forschungsgemeinschaft (DFG) via grant SFB 494, the Max Planck Society through the International Max Planck Research School, as well as special funds through the University of Cologne. We are grateful to all members of the NAOS/CONICA and the ESO PARANAL team. Macarena García-Marín is supported by the German federal department for education and research (BMBF) under the project numbers 50OS0502 & 50OS0801. M. Zamaninasab and D. Kunneriath are members of the International Max Planck Research School (IMPRS) for Astronomy and Astrophysics at the MPIfR and the Universities of Bonn and Cologne. R. Schödel acknowledges support by the Ramón y Cajal programme by the Ministerio de Ciencia e Innovación of the government of Spain.
References
- Alexander, T., & Hopman, C. 2009, ApJ, 697, 1861 [NASA ADS] [CrossRef] [Google Scholar]
- Bondi, H. 1952, MNRAS, 112, 195 [NASA ADS] [CrossRef] [MathSciNet] [Google Scholar]
- Baganoff, F. K., Bautz, M. W., Brandt, W. N., et al. 2001, Nature, 413, 45 [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
- Baganoff, F. K., Bautz, M. W., Ricker, G. R., et al. 2002, 201st AAS Meeting, #31.08, BAAS, 34, 1153 [Google Scholar]
- Baganoff, F. K., Maeda, Y., Morris, M., et al. 2003, ApJ, 591, 891 [NASA ADS] [CrossRef] [Google Scholar]
- Bahcall, J. N., & Wolf, R. A. 1976, ApJ, 209, 214 [NASA ADS] [CrossRef] [Google Scholar]
- Bélanger, G., Terrier, R., de Jager, O. C., Goldwurm, A., & Melia, F. 2006, JPhCS, 54, 420 [Google Scholar]
- Blum, R. D., Sellgren, K., Depoy, D. L. 1996, ApJ, 470, 864 [NASA ADS] [CrossRef] [Google Scholar]
- Bogdan, T. J., & Schlickeiser, R. 1985, A&A, 143, 23 [NASA ADS] [Google Scholar]
- Bower, G. C., Falcke, H., Herrnstein, R. M., et al. 2004, Science, 304, 704 [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
- Bregman, J. N. 1985, AJ, 288, 32 [NASA ADS] [CrossRef] [Google Scholar]
- Broderick, A. E., & Loeb, A. 2006, MNRAS, 367, 905 [NASA ADS] [CrossRef] [Google Scholar]
- Buchholz, R. M., Schödel, R., & Eckart, A. 2009, A&A, 499, 483 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Cuadra, J., & Nayakshin, S. 2006, JPhCS, 54, 436 [NASA ADS] [Google Scholar]
- Cuadra, J., Nayakshin, S., & Martins, F. 2008, MNRAS, 383, 458 [NASA ADS] [Google Scholar]
- Diolaiti, E., Bendinelli, O., Bonaccini, D., et al. 2000, A&AS, 147, 335 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Do, T., Ghez, A. M., Morris, M. R., Yelda, S., et al. 2009a, ApJ, 691, 1021 [NASA ADS] [CrossRef] [Google Scholar]
- Do, T., Ghez, A. M., Morris, M. R., et al. 2009b, ApJ, 703, 1323 [NASA ADS] [CrossRef] [Google Scholar]
- Doeleman, S. S., Fish, V. L., Broderick, A. E., Loeb, A., & Rogers, A. E. E. 2009, ApJ, 695, 59D [NASA ADS] [CrossRef] [Google Scholar]
- Dodds-Eden, K., Porquet, D., Trap, G., et al. 2009, ApJ, 698, 676 [NASA ADS] [CrossRef] [Google Scholar]
- Eckart, A., & Duhoux, P. R. M. 1990, ASPC, 14, 336 [Google Scholar]
- Eckart, A., & Genzel, R. 1996, Nature, 383, 415 [NASA ADS] [CrossRef] [Google Scholar]
- Eckart, A., Genzel, R., Ott, T., & Schödel, R. 2002, MNRAS, 331, 917 [NASA ADS] [CrossRef] [Google Scholar]
- Eckart, A., Moultaka, J., Viehmann, T., et al. 2003, ANS, 324, 557 [Google Scholar]
- Eckart, A., Baganoff, F. K., Morris, M., et al. 2004, A&A, 427, 1 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Eckart, A., Baganoff, F. K., Schödel, R., et al. 2006a, A&A, 450, 535 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Eckart, A., Schödel, R., Meyer, L., et al. 2006b, A&A, 455, 1 [Google Scholar]
- Eckart, A., Baganoff, F. K., Zamaninasab, M., et al. 2008a, A&A, 479, 625 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Eckart, A., Schödel, R., Garcia-Marin, M., et al. 2008b, A&A, 492, 337 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Eckart, A., Baganoff, F. K., Morris, M. R., et al. 2009, A&A, 500, 935 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Eisenhauer, F., Schödel, R., Genzel, R., et al. 2003, ApJ, 597, L121 [NASA ADS] [CrossRef] [Google Scholar]
- Eisenhauer, F., Genzel, R., Alexander, T., et al. 2005, ApJ, 628, 246 [NASA ADS] [CrossRef] [Google Scholar]
- Eisenhauer, F., Perrin, G., Brandner, W., et al. 2008, SPIE, 7013, 69 [NASA ADS] [Google Scholar]
- Fish, V. L., Doeleman, S. S., Broderick, A. E., Loeb, A., & Rogers, A. E. E. 2009, ApJ, 706, 1353 [NASA ADS] [CrossRef] [Google Scholar]
- Falanga, M., Melia, F., Tagger, M., Goldwurm, A., & Belanger, G. 2007, ApJ, 662, L15 [NASA ADS] [CrossRef] [Google Scholar]
- Falcke, H., & Markoff, S. 2000, A&A, 362, 113 [NASA ADS] [Google Scholar]
- Falcke, H., Markoff, S., & Fender, R. 2001, AGM, 18, 107 [Google Scholar]
- Floyd, D. J. E., Eric Perlman, E., Leahy, J. P., et al. 2006, ApJ, 639, 23 [NASA ADS] [CrossRef] [Google Scholar]
- Genzel, T., Eckart, A., Ott, T., & Eisenhauer, F. 1997, MNRAS, 291, 219 [NASA ADS] [CrossRef] [Google Scholar]
- Genzel, R., Pichon, C., Eckart, A., Gerhard, O. E., & Ott, T. 2000, MNRAS, 317, 348 [NASA ADS] [CrossRef] [Google Scholar]
- Genzel, R., Schödel, R., Ott, T., et al. 2003, Nature, 425, 934 [CrossRef] [PubMed] [Google Scholar]
- Gezari, S., Ghez, A. M., Becklin, E. E., et al. 2002, ApJ, 576, 790 [NASA ADS] [CrossRef] [Google Scholar]
- Ghez, A., Klein, B. L., Morris, M., & Becklin, E. E. 1998, ApJ, 509, 678 [NASA ADS] [CrossRef] [Google Scholar]
- Ghez, A., Morris, M., Becklin, E. E., Tanner, A., & Kremenek, T. 2000, Nature, 407, 349 [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
- Ghez, A. M., Duchéne, G., Matthews, K., et al. 2003, ApJ, 586, L127 [NASA ADS] [CrossRef] [Google Scholar]
- Ghez, A. M., Wright, S. A., Matthews, K., et al. 2004a, ApJ, 601, 159 [Google Scholar]
- Ghez, A. M., Hornstein, S. D., Bouchez, A., et al. 2004b, A&AS 205, 2406 [Google Scholar]
- Ghez, A. M., Salim, S., Hornstein, S. D., et al. 2005a, ApJ, 620, 744 [NASA ADS] [CrossRef] [Google Scholar]
- Ghez, A. M., Hornstein, S. D., Lu, J. R., et al. 2005b, ApJ, 635, 1087 [NASA ADS] [CrossRef] [Google Scholar]
- Gillessen, S., Eisenhauer, F., Quataert, E., et al. 2006, ApJ, 640, L163 [NASA ADS] [CrossRef] [Google Scholar]
- Gillessen, S., Eisenhauer, F., Trippe, S., et al. 2009, ApJ, 692, 1075 [NASA ADS] [CrossRef] [Google Scholar]
- Goldwurm, A., Brion, E., Goldoni, P., et al. 2003, ApJ, 584, 751 [NASA ADS] [CrossRef] [Google Scholar]
- Gould, R. J. 1979, A&A, 76, 306 [NASA ADS] [Google Scholar]
- Hawley, J. F., & Balbus, S. A. 1991, ApJ, 376, 223 [NASA ADS] [CrossRef] [Google Scholar]
- Hawley, J. F., & Stone, J. M. 1998, ApJ, 501, 758, [NASA ADS] [CrossRef] [Google Scholar]
- Hornstein, S. D., Ghez, A. M., Tanner, A., et al. 2002, ApJ, 577, L9 [NASA ADS] [CrossRef] [Google Scholar]
- Hornstein, S. D., Matthews, K., Ghez, A. M., et al. 2006, JPhCS, 54, 399 [NASA ADS] [Google Scholar]
- Hornstein, S. D., Matthews, K., Ghez, A. M., et al. 2007, ApJ, 667, 900 [NASA ADS] [CrossRef] [Google Scholar]
- Igumenshchev, I. V. 2002, ApJ, 577, 31 [NASA ADS] [CrossRef] [Google Scholar]
- Ishibashi, W., & Courvoisier, T. J.-L. 2009, A&A, 495, 113 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Jester, S., Röser, H.-J., Meisenheimer, K., Perley, R., & Conway, R. 2001, A&A, 373, 447 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Jester, S., Röser, H.-J., Meisenheimer, K., & Perley, R. 2005, A&A, 431, 477 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Karas, V., Dovciak, M., Eckart, A., & Meyer, L. 2007, Proceedings of the Workshop on the Black Holes and Neutron Stars, ed. S. Hledik, & Z. Stuchlik, 19-21 September 2007 (Opava: Silesian University), 99 [arXiv:0709.3836] [Google Scholar]
- Kardashev, N. S. 1962, SvA, 6, 317 [Google Scholar]
- Kishimoto, M., Antonucci, R., Blaes, O., et al. 2008, Nature, 454, 492 [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
- Krabbe, A., Iserlohe, C., Larkin, J. E., et al. 2006, ApJ, 642, L145 [NASA ADS] [CrossRef] [Google Scholar]
- Lightman, A. P., & Shapiro, S. L. 1977, ApJ, 211, 244 [NASA ADS] [CrossRef] [Google Scholar]
- Liu, S., Petrosian, V., Melia, F., & Fryer, C. 2006, ApJ, 648, 1020 [NASA ADS] [CrossRef] [Google Scholar]
- Lucy, L. B. 1974, AJ, 79, L745 [NASA ADS] [CrossRef] [Google Scholar]
- Mahadevan, R., & Quataert, E. 1997, ApJ, 490, 605 [NASA ADS] [CrossRef] [Google Scholar]
- Markoff, S. 2005, ApJ, 618, L103 [NASA ADS] [CrossRef] [Google Scholar]
- Markoff, S., Falcke, H., Yuan, F., & Biermann, P. L. 2001, A&A, 379, L13 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Markoff, S., Bower, G. C., & Falcke, H. 2007, MNRAS, 379, 1519 [NASA ADS] [CrossRef] [Google Scholar]
- Marrone, D. P., Moran, J. M., Zhao, J.-H., & Rao, R. 2007, ApJ, 654, L57 [NASA ADS] [CrossRef] [Google Scholar]
- Marrone, D. P., Baganoff, F. K., Morris, M., et al. 2008, ApJ, 682, 373 [NASA ADS] [CrossRef] [Google Scholar]
- Marscher, A. P. 1983, ApJ, 264, 296 [NASA ADS] [CrossRef] [Google Scholar]
- Melia, F., & Falcke, H. 2001, ARA&A, 39, 309 [NASA ADS] [CrossRef] [Google Scholar]
- Meyer, L., Eckart, A., Schödel, R., et al. 2006a, A&A, 460, 15 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Meyer, L., Schödel, R., Eckart, A., et al. 2006b, A&A, 458, L25 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Meyer, L., Eckart, A., Schödel, R., et al. 2007, A&A, 473, 707 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Murphy, B. W., Cohn, H. N., & Durisen, R. H. 1991, ApJ, 370, 60 [NASA ADS] [CrossRef] [Google Scholar]
- Nayakshin, S., & Sunyaev, R. 2003, MNRAS, 343, L15 [NASA ADS] [CrossRef] [Google Scholar]
- Özel, F., Psaltis, D., & Narayan, R. 2000, ApJ, 541, 234 [NASA ADS] [CrossRef] [Google Scholar]
- Ott, T., Eckart, A., & Genzel, R. 1999, ApJ, 523, 248 [NASA ADS] [CrossRef] [Google Scholar]
- Page, K. L., Reeves, J. N., O'Brien, P. T., Turner, M. J. L., & Worrall, D. M. 2004, MNRAS, 353, 133 [NASA ADS] [CrossRef] [Google Scholar]
- Pechácek, T., Karas, V., & Czerny, B. 2008, A&A, 487, 815 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Perlman, E. S., & Wilson, A. S. 2005, ApJ, 627, 140 [NASA ADS] [CrossRef] [Google Scholar]
- Perlman, E. S., Biretta, J. A., Sparks, W. B., Macchetto, F. D., & Leahy, J. P. 2001, ApJ, 551, 206 [NASA ADS] [CrossRef] [Google Scholar]
- Peterson, B. M. 1997, An introduction to active galactic nuclei, 44 (Cambridge University Press) [Google Scholar]
- Porquet, D., Predehl, P., Aschenbach, B., et al. 2003, A&A, 407, L17 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Porquet, D., Grosso, N., Predehl, P., et al. 2008, A&A, 488, 549 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Quataert, E. 2002, ApJ, 575, 855 [NASA ADS] [CrossRef] [Google Scholar]
- Quataert, E. 2003, Astron. Nachr., 324, No. S1, Special Supplement, The central 300 parsecs of the Milky Way, ed. A. Cotera, H. Falcke, T. R. Geballe, & S. Markoff, 435 [arXiv:astro-ph/0304099] [Google Scholar]
- Reid, M. J. 1993, ARA&A, 31, 345 [NASA ADS] [CrossRef] [Google Scholar]
- Straubmeier, C., Eisenhauer, F., Perrin, G., Brandner, W., & Eckart, A. 2008, SPIE, 7013, 93 [NASA ADS] [Google Scholar]
- Schödel, R., Ott, T., Genzel, R., et al. 2002, Nature, 419, 694 [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
- Schödel, R., Genzel, R., Ott, T., et al. 2003, ApJ, 596, 1015 [NASA ADS] [CrossRef] [Google Scholar]
- Schödel, R., Eckart, A., Alexander, T., et al. 2007, A&A, 469, 125 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Schödel, R., Merritt, D., & Eckart, A. 2009, A&A, 502 91 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Scoville, N. Z., Stolovy, S. R., Rieke, M., Christopher, M., & Yusef-Zadeh, F. 2003, ApJ, 594, 294 [Google Scholar]
- Tanner, A., Ghez, A. M., Morris, M., et al. 2002, ApJ, 575, 860 [NASA ADS] [CrossRef] [Google Scholar]
- Tanner, A., Ghez, A. M., Morris, M. R., & Christou, J. C. 2005, ApJ, 624, 742 [NASA ADS] [CrossRef] [Google Scholar]
- van der Laan, H. 1966, Nature, 211, 1131 [NASA ADS] [CrossRef] [Google Scholar]
- Yuan, F. 2006, JPhCS, 54, 427 [Google Scholar]
- Yuan, F., Quataert, E., & Narayan, R. 2003, ApJ, 598, 301 [NASA ADS] [CrossRef] [Google Scholar]
- Yuan, F., Lin, J., Wu, K., & Ho, L. C. 2008, MNRAS, 395, 2183 [Google Scholar]
- Yusef-Zadeh, F., Roberts, D., Wardle, M., Heinke, C. O., & Bower, G. C. 2006a, ApJ, 650, 189 [NASA ADS] [CrossRef] [Google Scholar]
- Yusef-Zadeh, F., Bushouse, H., Dowell, C. D., et al. 2006b, ApJ, 644, 198 [NASA ADS] [CrossRef] [Google Scholar]
- Yusef-Zadeh, F., Wardle, M., Cotton, W. D., Heinke, C. O., & Roberts, D. A. 2007, ApJ, 668, L47 [NASA ADS] [CrossRef] [Google Scholar]
- Yusef-Zadeh, F., Wardle, M., Heinke, C., et al. 2008, ApJ, 682, 361 [NASA ADS] [CrossRef] [Google Scholar]
- Yusef-Zadeh, F., Bushouse, H., Wardle, M., et al. 2009, ApJ, 706, 348 [NASA ADS] [CrossRef] [Google Scholar]
- Zamaninasab, M., Eckart, A., Meyer, L., et al. 2008, Proc. of a conference on Astrophysics at High Angular Resolution (AHAR 08) held 21-25 April 2008 in Bad Honnef, Germany, JPhCS, 131, 12008 [Google Scholar]
- Zamaninasab, M., Eckart, A., Witzel, G., et al. 2009, A&A, 510, A3 [Google Scholar]
- Zhao, J.-H., Young, K. H., Herrnstein, R. M., et al. 2003, ApJ, 586, L29 [NASA ADS] [CrossRef] [Google Scholar]
All Tables
Table 1: The brightest X-ray flares from SgrA*. For references see text.
Table 2: Details on observing runs during which SgrA* was in a very low NIR luminosity state.
Table 3: Sources used for iterative source subtraction for the 23 September 2004 data.
Table 4: Source component parameters for SSC models of the brightest X-ray flares from SgrA*.
Table 5: Model densities, lower SSC cutoff frequencies and lower limits on the source component masses.
All Figures
![]() |
Figure 1: Identification of individual sources that were used for the iterative subtraction of the Ks-band datasets (Table 2). Here we show the image derived from the 23 September 2004 data. The nomenclature was taken from the deconvolved H-band image given in Fig. 1 by Gillessen et al. (R2006). Sources that are not contained in there have labels starting with ``N''. Relative positions and flux densities of the labeled sources are given in Table 3. |
Open with DEXTER | |
In the text |
![]() |
Figure 2:
The Ks-band image of the Galactic center region
a)
taken on 23 September 2004. The central
15
|
Open with DEXTER | |
In the text |
![]() |
Figure 3:
The inner 2
|
Open with DEXTER | |
In the text |
![]() |
Figure 4: Results of the iterative star subtraction obtained for the 30 August a) 23 September data b) as described in Sect. 3.2. |
Open with DEXTER | |
In the text |
![]() |
Figure 5: Result of the automatic PSF subtraction obtained for the 23 September 2004 data. |
Open with DEXTER | |
In the text |
![]() |
Figure 6:
KLF histogram of the stars detected in the central field
derived from the 23 September 2004 data.
The straight full and dashed lines indicate the KLF slope of
0.21 |
Open with DEXTER | |
In the text |
![]() |
Figure 7:
Azimuthal average of the diffuse background emission as derived from
the three different methods applied: low pass filtering (red),
manual PSF subtraction (green), and automatic PSF subtraction (blue).
The curves
(mean flux and 1 |
Open with DEXTER | |
In the text |
![]() |
Figure 8: The 30 August 2004 light curve of the flux density measured in a 40 mas radius circular aperture centered on the position of SgrA* (black). We show the light curve of the reference star S2 in green. The bottom panel shows the FWHM seeing in arcseconds measured on the PSF extracted for the field. The grey points show the background counts near SgrA*. The width of the error bars is close to the size of the data points. |
Open with DEXTER | |
In the text |
![]() |
Figure 9: The same as Fig. 8 but for 23 September 2004. |
Open with DEXTER | |
In the text |
![]() |
Figure 10:
Sketch of the SgrA* surroundings within a 105 Schwarzschild radius.
The SMBH encloses the central 1 |
Open with DEXTER | |
In the text |
![]() |
Figure 11: Two SSC models A (red) and B (black dashed) describing the April 4 2007, flare data (see Table 1, Tables 4 and 5). |
Open with DEXTER | |
In the text |
![]() |
Figure 12:
A combined SSC and adiabatic expansion model compared to the
SgrA* flare event on April 4, 2007.
The data are described in
Porquet et al. (2003), Dodds-Eden et al. (2009), Yusef-Zadeh et al. (2009).
The 240 GHz and 43 GHz response of the NIR/X-ray flare is plotted
as a solid and dashed red line, respecitvely. The probably unrelated 43 GHz response
of the flare component |
Open with DEXTER | |
In the text |
![]() |
Figure 13: A single spot SSC quiescent spectrum (black dashed) and a corresponding multi-spot SSC quiescent spectrum. |
Open with DEXTER | |
In the text |
Copyright ESO 2010
Current usage metrics show cumulative count of Article Views (full-text article views including HTML views, PDF and ePub downloads, according to the available data) and Abstracts Views on Vision4Press platform.
Data correspond to usage on the plateform after 2015. The current usage metrics is available 48-96 hours after online publication and is updated daily on week days.
Initial download of the metrics may take a while.