ArticlePDF Available

An Inverse Compton Scattering Origin of X-ray Flares from Sgr A*

Authors:

Abstract and Figures

The X-ray and near-IR emission from Sgr A* is dominated by flaring, while a quiescent component dominates the emission at radio and sub-mm wavelengths. The spectral energy distribution of the quiescent emission from Sgr A* peaks at sub-mm wavelengths and is modeled as synchrotron radiation from a thermal population of electrons in the accretion flow, with electron temperatures ranging up to $\sim 5-20$\,MeV. Here we investigate the mechanism by which X-ray flare emission is produced through the interaction of the quiescent and flaring components of Sgr A*. The X-ray flare emission has been interpreted as inverse Compton, self-synchrotron-Compton, or synchrotron emission. We present results of simultaneous X-ray and near-IR observations and show evidence that X-ray peak flare emission lags behind near-IR flare emission with a time delay ranging from a few to tens of minutes. Our Inverse Compton scattering modeling places constraints on the electron density and temperature distributions of the accretion flow and on the locations where flares are produced. In the context of this model, the strong X-ray counterparts to near-IR flares arising from the inner disk should show no significant time delay, whereas near-IR flares in the outer disk should show a broadened and delayed X-ray flare.
Content may be subject to copyright.
The Astronomical Journal, 144:1 (10pp), 2012 July doi:10.1088/0004-6256/144/1/1
C
2012. The American Astronomical Society. All rights reserved. Printed in the U.S.A.
AN INVERSE COMPTON SCATTERING ORIGIN OF X-RAY FLARES FROM Sgr A*
F. Yusef-Zadeh1, M. Wardle2, K. Dodds-Eden3,C.O.Heinke
4, S. Gillessen3,
R. Genzel3, H. Bushouse5, N. Grosso6, and D. Porquet6
1Department of Physics and Astronomy, Northwestern University, Evanston, IL 60208, USA
2Department of Physics and Astronomy, Macquarie University, Sydney NSW 2109, Australia
3Max Planck Institut f¨
ur Extraterrestrische Physik, Postfach 1312, D-85741 Garching, Germany
4Department of Physics, University of Alberta, 4-183 CCIS, Edmonton, AB T6G 2E1, Canada
5Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, MD 21218, USA
6Observatoire astronomique de Strasbourg, Universit´
e de Strasbourg, CNRS, INSU, 11 rue de l’Universit´
e, 67000 Strasbourg, France
Received 2011 December 25; accepted 2012 March 4; published 2012 May 21
ABSTRACT
The X-ray and near-IR emission from Sgr A* is dominated by flaring, while a quiescent component dominates
the emission at radio and submillimeter (sub-mm) wavelengths. The spectral energy distribution of the quiescent
emission from Sgr A* peaks at sub-mm wavelengths and is modeled as synchrotron radiation from a thermal
population of electrons in the accretion flow, with electron temperatures ranging up to 5–20 MeV. Here, we
investigate the mechanism by which X-ray flare emission is produced through the interaction of the quiescent and
flaring components of Sgr A*. The X-ray flare emission has been interpreted as inverse Compton, self-synchrotron
Compton, or synchrotron emission. We present results of simultaneous X-ray and near-IR observations and show
evidence that X-ray peak flare emission lags behind near-IR flare emission with a time delay ranging from a
few to tens of minutes. Our inverse Compton scattering modeling places constraints on the electron density and
temperature distributions of the accretion flow and on the locations where flares are produced. In the context of this
model, the strong X-ray counterparts to near-IR flares arising from the inner disk should show no significant time
delay, whereas near-IR flares in the outer disk should show a broadened and delayed X-ray flare.
Key words: black hole physics Galaxy: center infrared: general ISM: clouds ISM: general X-rays: general
Online-only material: color figures
1. INTRODUCTION
Observations of stellar orbits in the proximity of the enigmatic
radio source Sgr A*, located at the dynamical center of our
galaxy, have shown compelling evidence that it is associated
witha4×106Mblack hole (Ghez et al. 2005; Gillessen et al.
2009; Reid & Brunthaler 2004). The extremely high spatial
resolution made possible by its relative proximity provides the
best laboratory for studying the properties of low-luminosity
accreting black holes; 1 corresponds to 0.039 pc at the Galactic
center distance of 8 kpc (Reid 1993). The emission from Sgr A*
is assumed to be produced from radiatively inefficient accretion
flow as well as outflows. The bulk of the continuum flux from
Sgr A* is considered to be generated in an accretion disk, where
identifying the source of variable continuum emission becomes
essential for our understanding of the launching and transport
of energy in the nuclei of galaxies.
The emission from Sgr A* consists of both quiescent and
variable components. The strongest variable component is
detected as flares at near-IR and X-ray wavelengths (Baganoff
et al. 2003; Genzel et al. 2003; Goldwurm et al. 2003; Eckart
et al. 2006; Yusef-Zadeh et al. 2006a; Hornstein et al. 2007;
Porquet et al. 2008; Dodds-Eden et al. 2009; Sabha et al.
2010;Trapetal.2011), whereas only moderate flux variation
is found at radio and submillimeter (sub-mm) wavelengths
(Falcke et al. 1998; Zhao et al. 2001; Herrnstein et al. 2004;
Miyazaki et al. 2004; Yusef-Zadeh et al. 2006b; Marrone et al.
2008). The spectral energy distribution (SED) of the quiescent
component peaks at sub-mm wavelengths and is identified in
radio, millimeter, and sub-mm wavelengths (see Genzel et al.
2010, and references therein). This emission is thought to be
produced by synchrotron radiation from a thermal population of
electrons with kT 10–30 MeV participating in an accretion
flow. A variety of models have been proposed to explain the
quiescent emission from Sgr A* by fitting its SED, including
a thin accretion disk, a disk and jet, an outflow, an advection-
dominated accretion flow, a radiatively inefficient accretion flow,
and advection-dominated inflow/outflow solutions (Blandford
& Begelman 1999; Melia & Falcke 2001; Yuan et al. 2003;
Liu et al. 2004; Genzel et al. 2010, and references cited
therein). Unlike the quiescent component, which originates over
a wide range of physical conditions and length scales of the
accretion flow, flares are localized, allowing emission models
to be directly tested with observations. As a supermassive black
hole candidate, Sgr A* presents an unparalleled opportunity to
closely study the process by which gas is captured, accreted,
or ejected, by characterizing the emission variability over
timescales of minutes to months. Because the timescale for
variability is proportional to the mass of the black hole, this
corresponds to variability on timescales 100 times longer than
that of more massive black holes in the nuclei of other galaxies.
Studying near-IR emission from Sgr A* is crucial to track the
acceleration of energetic particles as well as the accretion flow.
Near-IR flares are produced by synchrotron radiation from a
transient population of accelerated electrons. The near-IR emis-
sion is dominated by flaring activity that occurs a few times
per day, with a small fraction of events showing simultaneous
X-ray flares. The X-ray flare mechanism has been interpreted
as either inverse Compton scattering (ICS), self-synchrotron
Compton (SSC), or synchrotron emission (Markoff et al. 2001;
Liu & Melia 2002; Yuan et al. 2004; Yusef-Zadeh et al. 2006a;
Eckart et al. 2009; Marrone et al. 2008; Dodds-Eden et al.
2009). The X-ray synchrotron mechanism implies that the ac-
celeration mechanism must continuously resupply the 100 GeV
1
The Astronomical Journal, 144:1 (10pp), 2012 July Yusef-Zadeh et al.
electrons for the 30 minute duration of the observed flares, be-
cause the synchrotron loss time of the 100 GeV electrons that
are responsible for the synchrotron emission is 30 s. The syn-
chrotron self-Compton model requires that the local magnetic
field be extremely large or that the number density of electrons
is high. This is necessary to avoid overproducing the near-IR
synchrotron emission from the large number of energetic elec-
trons that are required to upscatter infrared photons into the
X-ray band (Dodds-Eden et al. 2009; Marrone et al. 2008;
Sabha et al. 2010; Trap et al 2011). The typical parame-
ters of the magnetic field—B1–10 G or electron density
ne109cm3—correspond to an energy density in the accel-
erated electrons a thousand times larger than that in the magnetic
field. It is then difficult to understand how these particles are
accelerated and confined.
In the case of X-ray emission produced by ICS, two possibili-
ties have been explored. First, sub-mm photons arising from the
quiescent component of Sgr A* may be upscattered by the tran-
sient electron population that is producing the IR synchrotron
emission during IR flares (Yusef-Zadeh et al. 2006a). Alterna-
tively, near-IR photons emitted during the flare may be upscat-
tered by the mildly relativistic 20 MeV electrons responsible
for the quiescent radio–submm emission (Yusef-Zadeh et al.
2006a,2008,2009). If the sub-mm emission region were opti-
cally thin, this would produce a similar X-ray luminosity as the
upscattering of sub-mm seed photons. However, because the
sub-mm source is optically thick below 1000 GHz, the ob-
served sub-mm flux is produced by a fraction of the underlying
electrons. The exact frequency at which the quiescent emission
becomes optically thick is unknown. However, sub-mm mea-
surements between 230 and 690 GHz (Marrone et al. 2006)
indicate a flattening of the spectral index and thus a deviation
from the rising spectrum observed at lower frequencies (An et al.
2005). The emission region is optically thin to near-IR photons,
so all of these electrons are available to upscatter near-IR seed
photons to X-ray energies (Yusef-Zadeh et al. 2009). The ICS
luminosity produced through this scenario compares favorably
with the observed near-IR and X-ray luminosities (Yusef-Zadeh
et al. 2009). This is the model on which we will focus, as de-
scribed below.
One of the predictions of the ICS model, in which near-IR
photons are upscattered by 10–30 MeV electrons, is a time
delay between the peaks of the near-IR and X-ray flares (Yusef-
Zadeh et al. 2009; Dodds-Eden et al. 2009). Wardle (2011)
provided the theoretical framework for the X-ray echo picture
of the ICS. We present evidence for a time delay between the
peaks of X-ray and near-IR flare emission based on seven new
and archival observations. These measurements provide support
for X-ray production via ICS of IR flare photons by relativistic
electrons of the accretion flow. The cross-correlation profiles of
the peaks are generally skewed toward positive time lags, but
show maximum likelihood values that have low signal-to-noise
(S/N), due to the limited number of detections of simultaneous
X-ray and near-IR flares.
2. OBSERVATIONS
2.1. X-Rays
X-ray observations used in this study come from the
Chandra observatory. Data obtained on 2004 July 6–7 and 2005
July 30 consist of 50.2 and 46 kilosecond (ks) observations,
respectively, (ObsIDs 4683,5953), which were described pre-
viously by Eckart et al. (2006) and Muno et al. (2005). Data
obtained in 2008 (not previously reported) consist of six 28 ks
observations, starting May 5, May 6, May 10, May 11, July 25,
and July 26 (ObsIDs 9169, 9170, 9171, 9172, 9174, 9173, re-
spectively), scheduled to match nighttime IR observations in
Chile (see below). All observations placed Sgr A* at the ACIS-I
aim point and took data in FAINT mode.
We checked for any time intervals of strong background
flaring (none were found) and then reprocessed the data using
CIAO 4.3.7This involved applying corrections to the energy
scales to compensate for time-dependent gain changes and
charge-transfer inefficiency, removing pixel randomization and
improving spatial resolution, as well as creating an updated bad
pixel map. We filtered the data for “bad” grades and status.
We extracted light curves (in spacecraft TT time) from a 1
circular region around the position of Sgr A*, in the 1.5–8 keV
energy range, using Gehrels (1986) errors. We converted the
time stamps to UTC time following the prescription by A. Rots.8
The baseline quiescent X-ray emission from Sgr A* is spatially
extended (Baganoff et al. 2003), but we see no variations in
other local background emission. We tested several choices of
binning the data for comparison to other wavelengths, settling
on 1500 s binning for the 2008 data, 300 s binning for the
major flare on 2004 July 6–7, and 600 s binning for the flare
on 2005 July 30. Using the absorbed thermal plasma model
of Baganoff et al., the ratio of 1.5–8 keV counts to 2–10 keV
unabsorbed flux is 8 ×1011 erg cm2s1per count/s. Recent
measurements indicate a distance of 8.3 kpc to Sgr A* (Gillessen
et al. 2009), but we assume a distance of 8 kpc, which gives
LX(2–10)=6×1035 erg s1per count s1, or a typical quiescent
luminosity of 3×1033 erg s1, in agreement with Baganoff et al.
(2003).
We used Kolmogorov–Smirnov (K-S) tests (using the lcstats
FTOOL9) on the 2008 Chandra light curves (binned to 32.41 s)
to search for evidence of variability. We find evidence for
variability in three of the 2008 observations, while another
three show no evidence of variability. ObsIDs 9169, 9172,
and 9173 give K-S probabilities of a constant light curve of
5×106,2×104, and 1 ×108, respectively, while the
remaining observations give K-S probabilities greater than 5%.
This significantly strengthens the evidence of variability from
Sgr A* at very low levels, as Baganoff et al. (2003) reported a
much larger K-S probability of constancy of 7 ×103during
quiescence. Given that the quiescent X-ray emission arises from
much larger scales, presumably due to Bondi–Hoyle accretion
(Baganoff et al. 2003), we suggest that the X-ray variability
noted here is due to low-level flare emission superimposed on the
steady quiescent emission. Alternatively, the X-ray variability
on hourly timescales could arise from coronally active stars
producing giant flares (Sazonov et al. 2012).
2.2. Near-IR
For the near-IR observations we use archival data taken with
the Very Large Telescope (VLT) and Hubble Space Telescope
(HST). The near-IR data taken in 2004, 2005, and 2008 were
observed with the near-IR adaptive optics (AO) assisted imager
NACO at the VLT (Lenzen et al. 2003; Rousset et al. 2003).
We used Ks-band 13 milliarcsecond (mas) pixel imaging data
from 2004 (nights of July 6–7) and 2005 (night of July 30–31)
first presented by Eckart et al. (2006,2008), and Ks-band
7For example, http://cxc.harvard.edu/ciao/threads/createL2/
8http://cxc.harvard.edu/contrib/arots/time/time_tutorial.html/
9http://heasarc.gsfc.nasa.gov/lheasoft/ftools/xronos.html
2
The Astronomical Journal, 144:1 (10pp), 2012 July Yusef-Zadeh et al.
13 mas pixel polarimetric imaging data from 2008 (nights
of May 4/5, 5/6, 9/10, 10/11 and July 25/26, 26/27) first
presented in Dodds-Eden et al. (2011). We did not apply any data
quality cut for the latter observations, except for the elimination
of 11 images from 2008 July 25, due to a bad AO correction
(quadfoils, or a “waffle” pattern).
We used both aperture photometry and point-spread function
(PSF) photometry methods to produce light curves, both carried
out in the way described in Dodds-Eden et al. (2011). In par-
ticular we note that, for the purposes of that paper, the aperture
method used two small apertures, one centered on the position
of Sgr A* and the other on the star S17 (confused with Sgr A*
in 2006–2008), in order to measure their combined flux. Since
S17 was further from Sgr A* in 2004 and the combined mea-
surement of the flux is not important for the purposes of this
paper, the use of the above method unnecessarily decreases the
S/N. As a result, for the near-IR/X-ray comparison we supple-
mented the data set with higher S/N light curves obtained from
PSF photometry, though the data were sparsely sampled. We
provide the light curve of S17 in the 2008 data set and the light
curve of the comparison star S7 for the 2004 and 2005 data sets.
The stellar background is estimated to be 3.4±0.2 mJy in 2008
(Dodds-Eden et al. 2011).
Near-IR HST observations used in this study are NICMOS
archival data obtained on 2007 April 4 as part of a larger Sgr A*
monitoring campaign. Full observational details have been
presented in Yusef-Zadeh et al. (2009). Briefly, the exposures
used NICMOS camera 1, which has a pixel scale of 0.
043, and
the medium-band filters F170M and F145M, which have central
wavelengths of 1.71 and 1.45 μm, respectively, and FWHMs of
0.2 μm. Individual exposures had a duration of 144 s, with
non-destructive detector readouts occurring every 16 s. We
averaged the readouts to sampling intervals of 64 and 128 s
in the 1.71 and 1.45 μm bands, respectively, to obtain adequate
S/N. Aperture photometry was performed on Sgr A* in each
sampling interval, using an aperture diameter of 3 pixels in order
to limit contamination by nearby stars.
All of the near-IR measurements presented here have been
corrected for reddening using extinction values of Aλ=2.42
(2.2 μm), 4.34 (1.71 μm), and 6.07 (1.45 μm) from Fritz et al.
(2011).
3. RESULTS
The middle two panels of Figure 1show the light curves
from the archival Ks-band and X-ray observations of Sgr A*
that were taken on 2005 July 30. The near-IR light curve of the
comparison star S7 is shown in the top panel. The X-ray and
near-IR light curves of Sgr A* were sampled at intervals of 200
and 600 s, respectively. These measurements, first reported in
Eckart et al. (2008), indicate a flare with a peak X-ray luminosity
of 8×1033 erg s1. The bottom panel shows the cross-correlation
of these light curves. The cross-correlation analysis uses the
Z-transformed discrete correlation function (ZDCF) algorithm
(Alexander 1997). The ZDC function is an improved solution
to the problem of investigating correlation in unevenly sampled
light curves. Maximum likelihood values as well as 1σand 2σ
confidence intervals around those values are estimated using the
start time of each bin. This analysis finds that the X-ray peak
lags the near-IR peak in Figure 1by 8.0 (+10, 10.1) and
8.0 (+20.2, 17.9) minutes for 1σand 2σmaximum likelihood
values, respectively. We varied the sampling interval in the near-
IR and X-ray data between 1.5 and 10 minutes, but the maximum
likelihood lag value remained the same. Eckart et al. (2008)
Figure 1. Light curve of the comparison star S7 is shown in the top panel
with a time sampling of 65 s. Two middle panels show the light curves of
Ks-band (2.2 μm) and X-ray (2–8 keV) data taken simultaneously by the VLT
and Chandra on 2005 July 30 (Eckart et al. 2008). The time sampling for the
X-ray and near-IR data is 600 and 200 s, respectively. The cross-correlation
of the light curves and the corresponding 2σmaximum likelihood values are
shown in the bottom panel. The 1σerror bar is given in Table 2.Abaselevel
of 4.2 mJy have been subtracted from the light curve of Sgr A* (Dodds-Eden
et al. 2011).
(A color version of this figure is available in the online journal.)
compared the X-ray and near-IR flare emission and found that
the peaks are coincident within ±7 minutes.
The aperture photometry technique that was used to reanalyze
the near-IR VLT data from 2004 produced a light curve that
is quite similar to that published by Eckart et al. (2006), who
deconvolved their images. The only difference is that the present
analysis uses data extending up to 4 hr UT on July 7, which is
longer than that of Eckart et al. (2006). The bottom three panels
of Figure 2show the cross-correlation of these near-IR and X-ray
data with a maximum likelihood lag of 7.0 (+1.3, 1.1) minutes
and (+7.5, 6.9) minutes with 1σand 2σerrors, respectively.
The lag is larger than zero at the 1σlevel. The near-IR light curve
of the comparison star S7 is shown in the top panel. The light
curve of S7 is constant and supports the variable emission from
Sgr A* between 3 and 4h UT. The sampling of the near-IR data
reduced using PSF photometry is much more sparse than the
aperture photometry data. The 1σcross-correlation peak using
the PSF photometry showed a time lag 7.7 (+2.9, 3.4) minutes
within the error bars of that reduced by aperture photometry.
Eckart et al. (2006) also cross-correlated their data and showed
no time delay within 10 minutes.
Next we examined the X-ray and near-IR light curves ob-
tained on 2007 April 4 using XMM and VLT observations
(flare 2 in Porquet et al. 2008; Dodds-Eden et al. 2009).
These data contain the second brightest X-ray flare (flare 2 in
3
The Astronomical Journal, 144:1 (10pp), 2012 July Yusef-Zadeh et al.
Figure 2. Top panel shows K-band (2.2 μm) VLT data of the comparison star
on S7 with a time sampling of 45 s and Sgr A* on 2004 July 6/7, while the
third panel shows simultaneous Chandra X-ray (2–8 keV) data (Eckart et al.
2006). The middle two panels show the light curves of K-band (2.2 μm) and
X-ray (2–8 keV) data taken simultaneously by the VLT and Chandra on 2004
July 6/7 (Eckart et al. 2006). The time sampling for the X-ray and near-IR data
on Sgr A* are 300 and 140 s, respectively. The cross-correlation of the light
curves is plotted in the bottom panel. For the 2004 light curve determined from
aperture photometry the stellar background estimate is 5.3±0.2 mJy which has
been subtracted. For the light curve determined from PSF photometry (separated
from S17 and S19 and free from any contribution from the seeing halo of S2)
the specific amount of faint stellar contribution is not clear, but small, <1mJy.
A maximum likelihood value with 2σerror bars are shown in the bottom panel.
The 1σerror bar is given in Table 2.
(A color version of this figure is available in the online journal.)
Porquet et al. 2008) coincident with one of the strongest near-IR
flares that has ever been detected. The cross-correlation of the
X-ray and near-IR data for this bright flare shows a 1σmaxi-
mum likelihood time delay of 0.5 (+7.0, 6.5) minutes, which
is consistent with zero time delay (Dodds-Eden et al. 2009;
Yusef-Zadeh et al. 2009). The peak luminosity of the brightest
flare is 24.6×1034 erg s1(Porquet et al. 2008). Two other
moderate X-ray flares (flares 4 and 5) were detected on 2004
April 4 following the bright X-ray flare. X-ray flares 4 and 5,
with peak luminosities of 6×1034 and 8.9×1034 erg s1,
respectively, are covered in the near-IR 1.71 and 1.45 μm
NICMOS data. The cross-correlations of the X-ray and near-
IR light curves for flares 4 and 5 are presented in Figures 3(a)–
(d). The NICMOS observations alternated between the 1.71 and
1.45 μm bands every 6 minutes. In all of the four cases studied,
the maximum likelihood values of flares 4 and 5 show positive
lags ranging between 5 and 10 minutes. Similar to the other
cases analyzed here, the peaks of the cross-correlations are all
skewed toward positive time lags.
Finally, we compared near-IR (VLT) and X-ray data taken
in 2008 May and July. Figure 4shows the light curves from
the two different days of observations. These X-ray flares are
an order of magnitude less luminous than those detected in
earlier observations. We have carried out K-S tests indicating
the reality of these low-level X-ray visibilities (Section 2.1).
The cross-correlations of the light curves from these two days
give maximum likelihood lags of 19 (+6.8, 2.4) and 14.6 (5.6,
2.9) minutes with 2σerror bars. Figure 4also shows the cross-
correlation of X-ray with near-IR light curves derived from PSF
photometry. The resulting time delays of 26.5 (+19, 29) and
16.6 (14.8, 11.8) minutes are well within the error bars of
the aperture photometry data. To provide additional support for
the reality of the variability of Sgr A*, Figure 5compares the
light curves of Sgr A* and S17 using the 2008 May 5 and 2008
July 26 observations, which are based on PSF photometry. In
these data, where Sgr A* and S17 are separated from each other,
each source is detected independently.
Although most of the individual cross-correlation results that
are presented here have low S/N, the 1σmaximum likelihood
peaks in eight different measurements show a tendency for
X-ray emission to lag near-IR emission rather than lead. The
cross-correlation gives maximum likelihood near-IR-to-X-ray
lag values that are systematically higher than zero. The strongest
simultaneous near-IR and X-ray flares (Flare 2 in Porquet et al.
2008) do not show any time delay, whereas the faintest X-ray
flares seem to show the longest time delays.
4. SSC MODELS
Several alternative models for the relationship between near-
IR and X-ray flares have been proposed. Synchrotron emission
from a high-energy tail of the accelerated electron population
responsible for the near-IR flaring may be responsible for the
observed X-ray flaring (Dodds-Eden et al. 2009,2010). SSC
models, in which the same population of electrons produce
the near-IR synchrotron emission and upscatter lower-energy
synchrotron photons, require an unrealistically compact, and
hence over-pressured, source region (Dodds-Eden et al. 2009)
or a very weak magnetic field or a high electron density to
avoid overproducing the IR synchrotron emission (Marrone
et al. 2008; Sabha et al. 2010;Trapetal.2011).
In SSC models the observed ratio of the X-ray and IR
fluxes demands a certain Thomson optical depth in relativistic
electrons. SSC flare models (Marrone et al 2008; Sabha et al.
2010;Trapetal.2011) adopt source region radii R of order Rs.
The requisite electron densities then imply a particular magnetic
field strength, so that the synchrotron emission from the electron
population matches the observed near-IR flaring. These SSC
models have electron energy densities ranging between 103and
2×105times the magnetic energy density. Because electron
acceleration mechanisms invoke magnetic fields, the energy
density in the field should be comparable to or greater than
the energy density of the accelerated particles. Thus, scaling
the SSC models to equipartition fields by reducing the electron
density while increasing the magnetic field to keep the product
neB[(p+1)/2] and hence the near-IR synchrotron flux fixed, one
finds that a reduction in electron density by a factor of 40 or
more is required. Thus, SSC contributes at most 1/40 of the
observed X-ray flux in such a model.
Alternatively, one can attempt to construct SSC models in
which the field is in equipartition by reducing the source
size R, while keeping the products neB[(p+1)/2]R3and neR
fixed to preserve the synchrotron and SSC fluxes, respectively.
Equipartition between the relativistic electrons and magnetic
field is attained when Ris reduced by a factor of a thousand
4
The Astronomical Journal, 144:1 (10pp), 2012 July Yusef-Zadeh et al.
(a) (b)
(c) (d)
Figure 3. (a) The light curves of flare 4 of 2007 April 4 in near-IR (1.70 μm) and X-ray (2–10 keV) are taken by HST/NICMOS, and XMM-Newton/EPIC, with a
time sampling of 64 and 300 s, respectively. (b) Same as (a) except that the 1.45 μm are sampled at 144 s interval to improve the S/N. (c) Same as (a) except that the
light curves of flare 5 are displayed at 1.70 μm. (d) Same as (c) except that the light curves of flare 5 are displayed at 1.45 μm. The cross-correlation and the maximum
likelihood values with 2σerror bars are shown in bottom panels. The 1σerror bars for flares 4 and 5 at 1.70 μmaregiveninTable2.
(A color version of this figure is available in the online journal.)
5
The Astronomical Journal, 144:1 (10pp), 2012 July Yusef-Zadeh et al.
(a) (b)
(c) (d)
Figure 4. (a) Using the aperture photometry technique, the top two panels show the light curves of Ks-band (2.2 μm) and X-ray (2–8 keV) data taken simultaneously
with VLT and Chandra on 2008 May 5. The sampling interval for X-ray and near-IR data are 25 and 2 minutes, respectively. The cross-correlation of the light
curves is plotted in the bottom panel. (b) Similar to (a) except that the data were taken on 2008, July 26+27. (c) Similar to (a) except the light curve of Sgr A* is
calibrated using PSF photometry technique. (d) Similar to (b) except the light curve of Sgr A* is calibrated using PSF photometry technique. The cross-correlation
and the maximum likelihood values with 2σerror bars are shown in bottom panels. The 1σerror bars for aperture photometric data are given in Table 2.
(A color version of this figure is available in the online journal.)
6
The Astronomical Journal, 144:1 (10pp), 2012 July Yusef-Zadeh et al.
(a) (b)
Figure 5. (a) The panels show the light curves of PSF photometrically reduced Sgr A* and S17 at 2.2 μm on 2008 May 5. (b) Similar to (a) except that the data were
taken on 2008, July 26+27. A base level of 3.6 mJy has been subtracted.
(A color version of this figure is available in the online journal.)
or more, with corresponding field strength in the 104–105G
range, which is orders of magnitude more than what is reason-
able.
5. X-RAY ECHO DUE TO ICS
Here we focus on an inverse Compton scenario for the
X-ray flares, suggested by Yusef-Zadeh et al. (2009) and
outlined in more detail by Wardle (2011). In this model, near-IR
flare photons are upscattered to X-ray energies by the thermal
electrons (kTe10 MeV) in the accretion flow. This process
dominates the alternative inverse Compton pathway, in which
the nonthermal energetic electrons responsible for the near-IR
synchrotron emission upscatter sub-millimeter photons emitted
by the thermal electrons in the accretion flow into the X-ray
band (Rybicki & Lightman 1986, chapter 7.5). This alternative
ICS pathway is less effective because the accretion flow is
optically thick in the submillimeter, so that the ratio between
sub-mm photons and the thermal electrons producing them is
reduced by a factor of the optical depth. Then the upscattering of
near-IR photons proportionately produces more emission than
would be inferred by implicitly assuming that the submillimeter
synchrotron flux is optically thin (e.g., Dodds-Eden et al.
2009). In this process, the second-order scattering echo can
also produce MeV γ-ray emission with a luminosity Lγthat is
lower than that of X-rays LXby a factor of a few (i.e., Lγ/LX
LX/LNIR).
One significant difference of the ICS picture from the syn-
chrotron and SSC pictures is that the longer path from the
near-IR source to the observer taken by an upscattered photon
detected in the X-ray compared to the straight-line path taken by
a photon received in the near-IR introduces a time delay between
flaring in the near-IR and X-rays. In addition, because scattering
occurs from a range of locations within the accretion flow, with a
corresponding range of time delays, the reflection signal tends to
be broadened compared to the near-IR seed photon light curve.
While there is some evidence of systematic delays between the
near-IR and X-ray flaring, the X-ray flares appear to generally
have a narrow FWHM compared to their corresponding near-IR
flares.
5.1. Modeling
To explore whether this model can plausibly explain the X-ray
flaring, we compute the X-ray “echoes” of the observed near-IR
flares to compare with our simultaneous X-ray observations. We
make a number of simplifying assumptions, none of which are
severe. We assume that the observed near-IR flare comes from
a point located in the accretion flow with a power-law spectrum
and Gaussian light curve, Sν(t)ν0.5exp((tt0)2/2σ2).
Because the X-ray flares are narrower than the near-IR we have
assumed that the FWHM of the near-IR flaring narrows as λ0.5
below 2.2 μm. The physical justification is that the synchrotron
loss timescale scales as λ0.5and becomes comparable with the
observed FWHM at about 2 μm. The energy of an upscattered
photon with initial energy IR isassumedtobe(4/3)γ2IR
where γis the electron Lorentz factor. Because the upscattered
photon energies are much lower than the electron rest energy,
the scattering occurs in the Thompson regime. Assuming
isotropic upscattering, the total production rate of upscattered
photons per unit volume is nIRneσTc, where nIR and neare the
7
The Astronomical Journal, 144:1 (10pp), 2012 July Yusef-Zadeh et al.
Tab l e 1
Model Parameters for Flare Events
Flare kT a
0nb
0rc
(MeV) (cm3)(cm)
2004 Jul 7 9 2.9×1081.0×1013
2005 Jul 30 20 2.3×1071.0×1013
2007 Apr 4 flare 2 7 2.1×1084.0×1012
2007 Apr 4 flare 4 5 7.9×1081.0×1013
Notes.
aAccretion flow temperature profile Te(r)=T0(r/r0)1where r0=2×
1012 cm.
bDensity profile ne(r, z)=n0(r/r0)0.75 exp(2z2/r2).
cRadial location of near-IR flare in equatorial plane (Rs1.2×1012 cm for
MBH =4×106M).
number densities of infrared photons and relativistic electrons,
respectively. We ignore relativistic effects such as the Doppler
boosting associated with the bulk motion of the accretion flow,
because the corresponding Lorentz factor is small compared to
that of the individual electrons. We also ignore the time delay,
gravitational redshift and lensing effects of the Kerr metric,
which only become important close to the event horizon in
highly inclined systems.
The electron density and temperature profiles in the accretion
flow are assumed to be steady, axisymmetric power laws in
cylindrical radius r, with ner0.75 and Ter1and
the density truncated within 2 Rsand beyond 20 Rs.The
accretion flow is assumed to be confined to a thick disk
with scale height/radius (h/r )=0.5. The electron population
is characterized by an approximate relativistic Maxwellian
f(x)=1/2x2exp(x), where x=E/(kTe), which is a
good approximation for kTe>
2 MeV. The adopted profiles
are within the typical ranges considered in analytic estimates
(e.g., Loeb & Waxman 2007), semi-analytic models for the
accretion flow (e.g., Yuan et al. 2003), and MHD simulations
(e.g., Mo´
scibrodzka et al. 2009).
The remaining parameters specify the flare location relative to
the line of sight and relative to the accretion flow: the inclination
iof the disk to the line of sight, and the flare location (r, φ, z)in
the natural cylindrical coordinate system. The low optical depth
of the accretion flow to near-IR photons means that the results
are insensitive to the inclination iand the azimuthal angle φ
between the flare location and the poloidal plane containing the
line of sight and the z-axis. We therefore fix these at typical
values i=45and φ=90, respectively. Similarly, the results
are insensitive to the height zof the near-IR flare for z<
r,so
we simply assume that the flare occurs in the disk midplane, i.e.,
that z=0.
The free parameters are the electron density n0and tempera-
ture T0at the fiducial radius 2 ×1012 cm, and the radial location
of the flare, r. The noisiness of the observed light curves preclude
formal fitting, so for each near-IR/X-ray flare combination we
adjust these parameters by hand to approximately match the
X-ray light curve. Reasonable matches to the observed X-ray
profile are obtained with flares occurring at r10Rs, and elec-
tron densities 107.5–108.5cm3and temperatures 5–20 MeV,
as listed in Table 1.
In the context of the ICS picture, we fit a sample of light curves
that have good time coverage in near-IR and X-ray wavelengths
in order to illustrate the point that this model can potentially be
a powerful tool to quantify the physical characteristics of the
accretion flow. The light curves presented in Figures 1and 2
are modeled following the second brightest X-ray flare that has
ever been recorded on 2007 April 4 (Porquet et al. 2008). The
light curves of the moderate flare that followed this bright flare
(flare 2) is presented in Figure 3. The time delays of the peak
emission shown in Figures 1,2, and 3are 8, 7, and 8 minutes,
respectively, whereas the bright flare on 2007 April 4 showed a
time delay consistent with zero. Given the limited simultaneous
time coverage of the flares shown in Figures 3and 4, we focus
only on modeling these four flares. Figure 6shows the observed
and modeled light curves for the simultaneous near-IR and X-ray
flares that occurred on 2005 July 30, 2004 July 6/7, and 2007
April 4 (the main flare and flare 4). Parameters of the fit for
each of the four examples are shown in Table 1. Substructures
corresponding to two weak flares in Figure 6(a) are also modeled
with the same parameters as the main flare, as listed in Table 1.A
baseline level has been subtracted from the near-IR light curves
before constructing the theoretical light curves. If we restrict the
evaluation of χ2to just the flare part of the X-ray light curve,
χ2/df is between 3 and 4. The reason is that there is often point-
to-point variability in the light curve that throws points well
away from the smooth “prediction”. In addition, the near-IR light
curve in Figure 6(c) shows substructures that could be arising
from different flares, which would have different time delays in
the context of ICS. Thus we have not formally fit the light curves
as our models are illustrative only. Given these limitations, we
obtain reasonable parameters from the theoretical X-ray light
curves, which are superimposed on the observed light curves in
Figure 6.
In each case a Gaussian form of the near-IR light curve has
been adopted to represent the observed (extinction-corrected)
near-IR flare at 3.8 μm, 2.2 μm, and 1.7 μm. However, in our
models, the X-ray flare arises from scattered optical photons.
We assume the peak flux of the synchrotron flare (emitting at
near-IR to optical) scales as ν0.5with frequency. Because
the X-ray flare is narrower than the near-IR we have assumed
that the FWHM is constant below 2.2 μm and narrows as λ0.5
shortward of this. Reasonable matches to the observed X-ray
profile are obtained with the flare occurring at the inner edge
of the density profile and electron densities, as their estimated
values are given in Table 1. The FWHM assumption requires
both the rise and the decay timescale of the flare to be faster at
optical frequencies than at near-IR. For the decay timescale, this
is a natural result of synchrotron cooling (the synchrotron loss
timescale scales as ν0.5). The rise timescale depends on the
acceleration mechanism. While the acceleration mechanism of
flare production is not understood, it is plausible that optically
emitting electrons are produced later than IR emitting electrons
(Kusunose & Takahara 2011).
Another issue involves the prediction that the spectral index
between X-rays and near-IR/optical emission be identical in the
context of ICS with the assumption that the electron distribution
has a single power-law spectrum. It is, however, possible that
the energy spectrum of electrons has a broken power law, thus
producing a different spectral index in near-IR and X-rays. The
mismatch in spectral index is in fact noted for the bright X-ray
flare coincident with a strong near-IR flare of 2007 April 4
(Dodds-Eden et al. 2009). Although these authors discuss the
difference in the spectral index using a synchrotron mechanism,
the broken power law of NIR emitting electrons with a steeper
spectral index shortward of 3.8 μm was also argued in the ICS
scenario (Yusef-Zadeh et al. 2009).
Although the spectral index measurements in X-ray and
near-IR cannot distinguish between the synchrotron and ICS
8
The Astronomical Journal, 144:1 (10pp), 2012 July Yusef-Zadeh et al.
0
2
4
6
8
Sν 2.2µm (mJy)
02468
0.0
0.5
1.0
1.5
2.0
UT (hours)
LX (2–8 keV) (1034 erg/s)
2005 Jul 30
0
2
4
6
8
Sν 2.2µm (mJy)
02468
0
2
4
6
UT (hours)
2004 Jul 07
LX (2–8 keV) (1034 erg/s)
0
10
20
30
Sν 3.8µm (mJy)
45678
0
1
2
3
4
UT (hours)
LX (2–10 keV) (1035 erg/s)
2007 Apr 04
0
2
4
Sν 1.7µm (mJy)
12 13 14 15 16
0.5
1
UT (hours)
LX (2–10 keV) (1035 erg/s)
2007 Apr 04 4A
(a) (b)
(c) (d)
Figure 6. Adopted near-IR light curves and the corresponding ICS produced X-ray light curves (solid lines) superimposed on the near-IR and X-ray flares observedon
2005 July 30 (a), 2004 July 7 (b), and two flares on 2007, April 4 ((c) and (d)), respectively. Near-IR flare data are an input to the ICS model. The X-ray and near-IR
flare of 2005 July 30 and 2004 July 7 are taken from Eckart et al. (2006,2008) whereas the 2007, April 4 data are taken from Porquet et al. (2008) and Dodds-Eden
et al. (2009). The ICS model parameters are listed in Table 1. Two weak flares before and after the main flare in the 2005 data in (a) have also been modeled. A possible
second flare of the 2004 data near 4h UT, as shown in Figure 2, has not been modeled in (b).
(A color version of this figure is available in the online journal.)
models for the production of X-rays, it is predicted that the ratio
of near-IR to X-ray flare emission can increase with increasing
time delay in the ICS scenario. This is because bright X-ray
flares are generated in the inner disk where the time delay is
expected to be small. Although the available data are limited
to test this aspect of the proposed model, we note a trend that
is consistent with this expectation. Table 2shows the ratio of
2.2 μm peak flux (mJy) to peak X-ray luminosity (1035 erg s1)
for seven different measurements. The 1σerror bars of the
maximum likelihood values are given in Column 7. The smallest
to largest flux ratios, as shown in Column 6 of Table 2, support
the trend that near-IR/X-ray flux ratios increase with the time
9
The Astronomical Journal, 144:1 (10pp), 2012 July Yusef-Zadeh et al.
Tab l e 2
Flux Ratios versus Time Delay
Flare IR X-Ray IR Peak X-Ray Peak Peak Ratio Time Delay
Backg. Backg. (mJy) (1 ×1035 erg s1)IR/X-Ray (minutes) 1σ
2007 Apr 4 flare 2 5.0 0.2 16.5 4.8 3.4 0.5 (+7, 6.5)
2007 Apr 4 flare 5 0.2 0.2 2.63 1.23 2.1 5.0 (+1.9, 1.5)
2007 Apr 4 flare 4 0.2 0.2 4.67 1.22 3.8 5.0 (+1, 1.4)
2004 Jul 7 1.5 0.03 6 0.39 15.5 7 (+1.3 1.2)
2005 Jul 30 0.0 0.002 5.9 0.13 45.4 8 (+10, 10.1)
2008 Jul 26+27 5.5 0.003 3.7 0.05 68.5 14.6 (+5.6, 7.4)
2008 May 5 6 0.003 2.74 0.021 130.0 19 (+6.8, 2.4)
delay, as expected in the context of the ICS model. For the near-
IR observations on 2007 April 4 obtained at 3.8 and 1.7 μm,
we convert the peak flux to 2.2 μm using the spectral index
of 0.7, where Sνν0.7, before we estimate the flux ratio.
Future simultaneous measurements of X-ray and near-IR flares
should examine the correlation of the peak near-IR to X-ray flux
as a function of increasing observed time delay.
In summary, we have presented cross-correlations of simul-
taneous X-ray and near-IR flare light curves from Sgr A* and
found a time lag of the X-ray peak flare emission with respect
to the near-IR. Such an X-ray echo provides support for ICS of
near-IR flare photons by 5–20 MeV electrons. A fraction of
near-IR flare photons must upscatter from the accretion flow into
the X-ray band. This can be significant for plausible accretion
models, and therefore may explain the observed X-ray flares, or
at least place significant constraints on the accretion flow. Future
cross-correlations based on more continuous near-IR and X-ray
observations should give us better S/N in the maximum likeli-
hood values of the time lag. In the context of the ICS model,
future measurements will place better constraints on the density
and temperature profiles of the accretion flow and the location
of near-IR flares.
This work is partially supported by grants AST-0807400
from the National Science Foundation and DP0986386 from the
Australian Research Council. C.O.H. is supported by NSERC
and an Ingenuity New Faculty Award.
REFERENCES
Alexander, T. 1997, MNRAS, 285, 891
An, T., Goss, W. M., Zhao, J.-H., et al. 2005, ApJ,634, L49
Baganoff, F. K., Maeda, Y., Morris, M., et al. 2003, ApJ,591, 891
Blandford, R. D., & Begelman, M. C. 1999, MNRAS,303, L1
Dodds-Eden, K., Gillessen, S., Fritz, T. K., et al. 2011, ApJ,728, 37
Dodds-Eden, K., Porquet, D., Trap, G., et al. 2009, ApJ,698, 676
Dodds-Eden, K., Sharma, P., Quataert, E., et al. 2010, ApJ,725, 450
Eckart, A., Baganoff, F. K., Morris, M. R., et al. 2009, A&A,500, 935
Eckart, A., Baganoff, F. K., Sch ¨
odel, R., et al. 2006, A&A,450, 535
Eckart, A., Sch¨
odel, R., Garc´
ıa-Marn, M., et al. 2008, A&A,492, 337
Falcke, H., Goss, W. M., Matsuo, H., et al. 1998, ApJ,499, 731
Fritz, T. K., Gillessen, S., Dodds-Eden, K., et al. 2011, ApJ,737, 73
Gehrels, N. 1986, ApJ,303, 336
Genzel, R., Eisenhauer, F., & Gillessen, S. 2010, Rev. Mod. Phys.,82, 3121
Genzel, R., Sch¨
odel, R., Ott, T., et al. 2003, Nature,425, 934
Ghez, A. M., Salim, S., Hornstein, S. D., et al. 2005, ApJ,620, 744
Gillessen, S., Eisenhauer, F., Trippe, S., et al. 2009, ApJ,692, 1075
Goldwurm, A., Brion, E., Goldoni, P., et al. 2003, ApJ,584, 751
Herrnstein, R. M., Zhao, J.-H., Bower, G. C., & Goss, W. M. 2004, AJ,127,
3399
Hornstein, S. D., Matthews, K., Ghez, A. M., et al. 2007, ApJ,667, 900
Kusunose, M., & Takahara, F. 2011, ApJ,726, 54
Lenzen, R., Hartung, M., Brandner, W., et al. 2003, Proc. SPIE, 4841, 944
Liu, S., & Melia, F. 2002, ApJ,566, L77
Liu, S., Petrosian, V., & Melia, F. 2004, ApJ,611, L101
Loeb, A., & Waxman, E. 2007, J. Cosmol. Astropart. Phys.,JCAP03(2007)011
Markoff, S., Falcke, H., Yuan, F., & Biermann, P. L. 2001, A&A,379, L13
Marrone, D. P., Baganoff, F. K., Morris, M. R., et al. 2008, ApJ,682, 373
Marrone, D. P., Moran, J. M., Zhao, J.-H., & Rao, R. 2006, in J. Phys. Conf.
Ser. 54, 354
Melia, F., & Falcke, H. 2001, ARA&A,39, 309
Miyazaki, A., Tsutsumi, T., & Tsuboi, M. 2004, ApJ,611, L97
Mo´
scibrodzka, M., Gammie, C. F., Dolence, J. C., Shiokawa, H., & Leung,
P. K. 2009, ApJ,706, 497
Muno, M. P., Lu, J. R., Baganoff, F. K., et al. 2005, ApJ,633, 228
Porquet, D., Grosso, N., Predehl, P., et al. 2008, A&A,488, 549
Reid, M. J. 1993, ARA&A,31, 345
Reid, M. J., & Brunthaler, A. 2004, ApJ,616, 872
Rousset, G., Lacombe, F., Puget, P., et al. 2003, Proc. SPIE, 4839, 140
Rybicki, G. B., & Lightman, A. P. (ed.) 1986, Radiative Processes in Astro-
physics (New York: Wiley), 382, chap. 7.5
Sabha, N., Witzel, G., Eckart, A., et al. 2010, A&A,512, A2
Sazonov, S., Sunyaev, R., & Revnivtsev, M. 2012, MNRAS,420, 388
Trap, G., Goldwurm, A., Dodds-Eden, K., et al. 2011, A&A,528, A140
Wardle, M. 2011, in ASP Conf. Ser. 439, The Galactic Center: A Window to the
Nuclear Environment of Disk Galaxies ed. M. R. Morris, Q. Daniel Wang,
& F. Yuan (San Francisco, CA: ASP), 450
Yuan, F., Quataert, E., & Narayan, R. 2003, ApJ,598, 301
Yuan, F., Quataert, E., & Narayan, R. 2004, ApJ,606, 894
Yusef-Zadeh, F., Bushouse, H., Dowell, C. D., et al. 2006a, ApJ,644, 198
Yusef-Zadeh, F., Bushouse, H., Wardle, M., et al. 2009, ApJ,706, 706
Yusef-Zadeh, F., Roberts, D., Wardle, M., Heinke, C. O., & Bower, G. C.
2006b, ApJ,650, 189
Yusef-Zadeh, F., Wardle, M., Heinke, C., et al. 2008, ApJ,682, 361
Zhao, J.-H., Bower, G. C., & Goss, W. M. 2001, ApJ,547, L29
10
... The high linear polarization degree (∼ 20% − 40%) of the flares at NIR wavelengths suggests a synchrotron origin, but the emission mechanism for the X-ray flares is not fully conclusive. Many models have been proposed, the possible radiative processes invoked in the models include inverse Compton (IC) scattering of submillimeter photons, synchrotron-self-Compton (SSC) scattering of IR/ NIR photons, or pure synchrotron radiation from non-thermal electron (Markoff et al. 2001;Yuan et al. 2004;Yusef-Zadeh et al. 2006;Broderick & Loeb 2006;Eckart et al. 2008;Dodds-Eden et al. 2009;Yuan et al. 2009;Dodds-Eden et al. 2010;Yusef-Zadeh et al. 2012;Ponti et al. 2017;Subroweit et al. 2020). It is argued in some works that, although IC and SSC models can self-consistently explain the simultaneous IR/X-ray multiwavelength observations and the time delay between flares at high and low frequencies, synchrotron radiation model has the advantage that it can be achieved with physically plausible magnetic field strength and electron density compared with IC and SCC models (Dodds-Eden et al. 2009. ...
Preprint
High-resolution near infrared observations with GRAVITY instrument have revealed rapid orbital motions of a hot spot around Sgr A*, the supermassive black hole in our Galactic center, during its three bright flares. The projected distances of the spot to the black hole are measured and seems to increase with time. The values of distance, combined with the measured orbiting time, imply that the spot is rotating with a super-Keplerian velocity. These results are hard to understand if the spot stay within the accretion flow thus provide strong constraints on theoretical models for flares. Previously we have proposed a "ME" model for the flares by analogy with the coronal mass ejection model in solar physics. In that model, magnetic reconnection occurred at the surface of the accretion flow results in the formation of flux ropes, which are then ejected out. Energetic electrons accelerated in the current sheet flow into the flux rope region and their radiation is responsible for the flares. In this paper, we apply the model to the interpretation of the GRAVITY results by calculating the dynamics of the ejected flux rope, the evolution of the magnetic field and the energy distribution of accelerated electrons, and the radiation of the system. We find that the model can well explain the observed light curve of the flares, the time-dependent distance and the super-Keplerian motion of the hot spot. It also explains why the light curve of some flares have double peaks.
... Some report simultaneity between the X-ray and IR peaks but do not report a time frame within which that claim can be considered valid (Yusef-Zadeh et al. 2006bTrap et al. 2011). Those that constrain timing between X-ray and IR activity (Eckart et al. 2004(Eckart et al. , 2006bHornstein et al. 2007;Eckart et al. 2008a;Dodds-Eden et al. 2009;Eckart et al. 2012;Ponti et al. 2017;Hornstein et al. 2007;Yusef-Zadeh et al. 2012) are plotted in Figure B1 along with the Spitzer-Chandra results of this campaign. Table B2 reports all updated time lags measured from the Spitzer-Chandra campaign. ...
Article
Full-text available
We report a timing analysis of near-infrared (NIR), X-ray, and submillimeter data during a 3 day coordinated campaign observing Sagittarius A*. Data were collected at 4.5 μ m with the Spitzer Space Telescope, 2–8 keV with the Chandra X-ray Observatory, 3–70 keV with NuSTAR, 340 GHz with ALMA, and 2.2 μ m with the GRAVITY instrument on the Very Large Telescope Interferometer. Two dates show moderate variability with no significant lags between the submillimeter and the infrared at 99% confidence. A moderately bright NIR flare ( F K ∼ 15 mJy) was captured on July 18 simultaneous with an X-ray flare ( F 2−10 keV ∼ 0.1 counts s ⁻¹ ) that most likely preceded bright submillimeter flux ( F 340 GHz ∼ 5.5 Jy) by about + 34 − 33 + 14 minutes at 99% confidence. The uncertainty in this lag is dominated by the fact that we did not observe the peak of the submillimeter emission. A synchrotron source cooled through adiabatic expansion can describe a rise in the submillimeter once the synchrotron self-Compton NIR and X-ray peaks have faded. This model predicts high GHz and THz fluxes at the time of the NIR/X-ray peak and electron densities well above those implied from average accretion rates for Sgr A*. However, the higher electron density postulated in this scenario would be in agreement with the idea that 2019 was an extraordinary epoch with a heightened accretion rate. Since the NIR and X-ray peaks can also be fit by a nonthermal synchrotron source with lower electron densities, we cannot rule out an unrelated chance coincidence of this bright submillimeter flare with the NIR/X-ray emission.
... Among the physical mechanisms proposed as possible origins of the NIR flares such as the population of non-thermal electrons, magnetic reconnection, and shocks (Yuan et al. 2003;Dodds-Eden et al. 2010;Yusef-Zadeh et al. 2012;Ponti et al. 2017), the one most independent from the accretion rate is magnetic reconnection (Ripperda et al. 2020(Ripperda et al. , 2021. We therefore suggest that the brightest NIR flares of Sgr A * are likely caused by magnetic reconnection. ...
Article
In 2019, Sgr A*—the supermassive black hole in the Galactic Center—underwent unprecedented flaring activity in the near-infrared (NIR), brightening by up to a factor of 100 compared to quiescent values. Here we report Atacama Large Millimeter/submillimeter Array (ALMA) observations of Sgr A*'s continuum variability at 1.3 mm (230 GHz)—a tracer of the accretion rate–conducted one month after the brightest detected NIR flare and in the middle of the flaring activity of 2019. We develop an innovative light-curve extraction technique which (together with ALMA’s excellent sensitivity) allows us to obtain light curves that are simultaneously of high time resolution (2 s) and high signal-to-noise ratio (∼500). We construct an accurate intrinsic structure function of the Sgr A* submm variability, improving on previous studies by about two orders of magnitude in timescale and one order of magnitude in sensitivity. We compare the 2019 June variability behavior with that of 2001–2017 and suggest that the most likely cause of the bright NIR flares is magnetic reconnection.
... To explain the mysterious X-ray flares, astronomers have suggested that there exists a gas cloud around Sgr A* containing hundred-trillions of asteroids, comets, and planets that are stripped from their parent stars by the tidal forces of the massive black hole. When these objects rain down or are accreted onto the massive black hole, X-ray flares take place via physical processes such as the non-thermal synchrotron emission [47], the inverse-Compton scattering [48], and stochastic electron acceleration [49]. To emit the high-energy X-rays detected, an object that was striped from its parent star had to be torn apart into gases during its falling and the gases when arriving nearly at the massive black hole had to spike to hundreds of million degrees Celsius, which is ten or more times hotter than the center of the Sun. ...
Article
Recently the author has fully addressed the first two days of Genesis according to his well-developed black hole universe model (see Paper-I and Paper-II). In accordance with this new interpretation of Genesis, God first created the infinite entire universe called “earth” with matter named “water”, and light, space and time, fundamental forces and motion. Then, he hierarchically structured the entire universe by separating the matter and space with infinite layers bounded by event horizons (called “vaults”) and further formed our finite black hole universe. The efforts bridged the gap between Genesis and observations of the universe and brought us a scientific understanding of the Genesis. In this sequence study as Paper-III, we describe how God constructed the interiors of our finite black hole universe. It includes the formation of celestial objects by gathering the water or gravitationally collapsing the initial super fluidal matter under the sky or inside the even horizon of our black hole universe. These formed celestial objects could be stars and planets called dry grounds or lands, in which matter is not in the water state any more, and galaxies and clusters called, respectively, seas of stars and seas of galaxies.
... The upper limit from Hornstein et al. (2007) indicates an X-ray flare whose peak occurred 36 minutes before IR observations began. Yusef-Zadeh et al. (2012) is the only work to report any time lag between the X-ray and IR with error bars. We reanalyze the seven flares presented in their work and plot the results of our reanalysis here. ...
... There is no consensus about the precise mechanism of the flares and about their physical origin. They may be caused by a population of nonthermal electrons (Dodds-Eden et al. 2010;Ponti et al. 2017), magnetic reconnection events (Yusef-Zadeh et al. 2012;Yuan et al. 2003), or a short term increase of the accretion flow rate into the black hole at the event horizon. ...
Article
In 2019, the Galactic center black hole Sgr A* produced an unusually high number of bright near-infrared flares, including the brightest-ever detected flare. We propose that this activity was triggered by the near simultaneous infall of material shed by G1 and G2 objects due to their interaction with the background accretion flow. We discuss mechanisms by which S-stars and G-objects shed material, and estimate both the quantity of material and the infall time to reach the black hole.
... In light of the interferometric results, the variability data obtained over the last two decades are valuable as a complementary source of information about the physical processes at event-horizon scales. Many multiwavelength campaigns in the submm, NIR, and the Xrays have been organized in the hope of determining-or at least constraining-the underlying radiative processes (Baganoff et al. 2001;Eckart et al. 2004Eckart et al. , 2006Gillessen et al. 2006;Yusef-Zadeh et al. 2006a,b;Eckart et al. 2008a,b;Marrone et al. 2008;Yusef-Zadeh et al. 2008;Dodds-Eden et al. 2009;Yusef-Zadeh et al. 2009;Trap et al. 2011;Eckart et al. 2012;Yusef-Zadeh et al. 2012;Haubois et al. 2012;Mossoux et al. 2016;Rauch et al. 2016;Ponti et al. 2017). ...
Preprint
Full-text available
Sagittarius A* (Sgr A*) is the variable radio, near-infrared (NIR), and X-ray source associated with accretion onto the Galactic center black hole. We have analyzed a comprehensive submillimeter (including new observations simultaneous with NIR monitoring), NIR, and 2-8 keV dataset. Submillimeter variations tend to lag those in the NIR by $\sim$30 minutes. An approximate Bayesian computation (ABC) fit to the X-ray first-order structure function shows significantly less power at short timescales in the X-rays than in the NIR. Less X-ray variability at short timescales combined with the observed NIR-X-ray correlations means the variability can be described as the result of two strictly correlated stochastic processes, the X-ray process being the low-pass-filtered version of the NIR process. The NIR--X-ray linkage suggests a simple radiative model: a compact, self-absorbed synchrotron sphere with high-frequency cutoff close to NIR frequencies plus a synchrotron self-Compton scattering component at higher frequencies. This model, with parameters fit to the submillimeter, NIR, and X-ray structure functions, reproduces the observed flux densities at all wavelengths, the statistical properties of all light curves, and the time lags between bands. The fit also gives reasonable values for physical parameters such as magnetic flux density $B\approx13$ G, source size $L \approx2.2R_{S}$, and high-energy electron density $n_{e}\approx4\times10^{7}$ cm$^{-3}$. An animation illustrates typical light curves, and we make public the parameter chain of our Bayesian analysis, the model implementation, and the visualization code.
Article
High-resolution near infrared observations with GRAVITY instrument have revealed rapid orbital motions of a hot spot around Sgr A*, the supermassive black hole in our Galactic center, during its three bright flares. The projected distances of the spot to the black hole are measured and seems to increase with time. The values of distance, combined with the measured orbiting time, imply that the spot is rotating with a super-Keplerian velocity. These results are hard to understand if the spot stay within the accretion flow thus provide strong constraints on theoretical models for flares. Previously we have proposed a ‘CME’ model for the flares by analogy with the coronal mass ejection model in solar physics. In that model, magnetic reconnection occurred at the surface of the accretion flow results in the formation of flux ropes, which are then ejected out. Energetic electrons accelerated in the current sheet flow into the flux rope region and their radiation is responsible for the flares. In this paper, we apply the model to the interpretation of the GRAVITY results by calculating the dynamics of the ejected flux rope, the evolution of the magnetic field and the energy distribution of accelerated electrons, and the radiation of the system. We find that the model can well explain the observed light curve of the flares, the time-dependent distance and the super-Keplerian motion of the hot spot. It also explains why the light curve of some flares have double peaks.
Preprint
Full-text available
We report timing analysis of near-infrared (NIR), X-ray, and sub-millimeter (submm) data during a three-day coordinated campaign observing Sagittarius A*. Data were collected at 4.5 micron with the Spitzer Space Telescope, 2-8 keV with the Chandra X-ray Observatory, 3-70 keV with NuSTAR, 340 GHz with ALMA, and at 2.2 micron with the GRAVITY instrument on the Very Large Telescope Interferometer. Two dates show moderate variability with no significant lags between the submm and the infrared at 99% confidence. July 18 captured a moderately bright NIR flare (F_K ~ 15 mJy) simultaneous with an X-ray flare (F ~ 0.1 cts/s) that most likely preceded bright submm flux (F ~ 5.5 Jy) by about +34 (+14 -33) minutes at 99% confidence. The uncertainty in this lag is dominated by the fact that we did not observe the peak of the submm emission. A synchrotron source cooled through adiabatic expansion can describe a rise in the submm once the synchrotron-self-Compton NIR and X-ray peaks have faded. This model predicts high GHz and THz fluxes at the time of the NIR/X-ray peak and electron densities well above those implied from average accretion rates for Sgr A*. However, the higher electron density postulated in this scenario would be in agreement with the idea that 2019 was an extraordinary epoch with a heightened accretion rate. Since the NIR and X-ray peaks can also be fit by a non-thermal synchrotron source with lower electron densities, we cannot rule out an unrelated chance coincidence of this bright submm flare with the NIR/X-ray emission.
Article
Sagittarius A* (Sgr A*) is the variable radio, near-infrared (NIR), and X-ray source associated with accretion onto the Galactic center black hole. We have analyzed a comprehensive submillimeter (including new observations simultaneous with NIR monitoring), NIR, and 2–8 keV data set. Submillimeter variations tend to lag those in the NIR by ∼30 minutes. An approximate Bayesian computation fit to the X-ray first-order structure function shows significantly less power at short timescales in the X-rays than in the NIR. Less X-ray variability at short timescales, combined with the observed NIR–X-ray correlations, means the variability can be described as the result of two strictly correlated stochastic processes, the X-ray process being the low-pass-filtered version of the NIR process. The NIR–X-ray linkage suggests a simple radiative model: a compact, self-absorbed synchrotron sphere with high-frequency cutoff close to NIR frequencies plus a synchrotron self-Compton scattering component at higher frequencies. This model, with parameters fit to the submillimeter, NIR, and X-ray structure functions, reproduces the observed flux densities at all wavelengths, the statistical properties of all light curves, and the time lags between bands. The fit also gives reasonable values for physical parameters such as magnetic flux density B ≈ 13 G, source size L ≈ 2.2 R S , and high-energy electron density n e ≈ 4 × 10 ⁷ cm ⁻³ . An animation illustrates typical light curves, and we make public the parameter chain of our Bayesian analysis, the model implementation, and the visualization code.
Article
Full-text available
Context: .We report new simultaneous near-infrared/sub-millimeter/X-ray observations of the Sgr A* counterpart associated with the massive 3{-}4×106 M⊙ black hole at the Galactic Center. Aims: . We investigate the physical processes responsible for the variable emission from Sgr A*. Methods: . The observations have been carried out using the NACO adaptive optics (AO) instrument at the European Southern Observatory's Very Large Telescope and the ACIS-I instrument aboard the Chandra X-ray Observatory as well as the Submillimeter Array SMA on Mauna Kea, Hawaii, and the Very Large Array in New Mexico. Results: . We detected one moderately bright flare event in the X-ray domain and 5 events at infrared wavelengths. The X-ray flare had an excess 2-8 keV luminosity of about 33×1033 erg/s. The duration of this flare was completely covered in the infrared and it was detected as a simultaneous NIR event with a time lag of ≤10 min. Simultaneous infrared/X-ray observations are available for 4 flares. All simultaneously covered flares, combined with the flare covered in 2003, indicate that the time-lag between the NIR and X-ray flare emission is very small and in agreement with a synchronous evolution. There are no simultaneous flare detections between the NIR/X-ray data and the VLA and SMA data. The excess flux densities detected in the radio and sub-millimeter domain may be linked with the flare activity observed at shorter wavelengths. Conclusions: . We find that the flaring state can be explained with a synchrotron self-Compton (SSC) model involving up-scattered sub-millimeter photons from a compact source component. This model allows for NIR flux density contributions from both the synchrotron and SSC mechanisms. Indications for an exponential cutoff of the NIR/MIR synchrotron spectrum allow for a straightforward explanation of the variable and red spectral indices of NIR flares.
Article
Full-text available
We have carried out Very Large Array (VLA) continuum observations to study the variability of Sgr A* at 43 GHz (λ = 7 mm) and 22 GHz (λ = 13 mm). A low level of flare activity has been detected with a duration of ~2 hr at these frequencies, showing the peak flare emission at 43 GHz leading the 22 GHz peak flare by ~20-40 minutes. The overall characteristics of the flare emission are interpreted in terms of the plasmon model of van der Laan by considering the ejection and adiabatic expansion of a uniform, spherical plasma blob due to flare activity. The observed peak of the flare emission with a spectral index, ν-α, of α = 1.6 is consistent with the prediction that the peak emission shifts toward lower frequencies in an adiabatically expanding self-absorbed source. We present the expected synchrotron light curves for an expanding blob, as well as the peak frequency emission, as a function of the energy spectral index constrained by the available flaring measurements in near-IR, submillimeter, millimeter, and radio wavelengths. We note that the blob model is consistent with the available measurements; however, we cannot rule out the jet of Sgr A*. If expanding material leaves the gravitational potential of Sgr A*, the total mass-loss rate of nonthermal and thermal particles is estimated to be ≤2 × 10-8M☉ yr-1. We discuss the implication of the mass-loss rate, since this value matches closely the estimated accretion rate based on polarization measurements.
Article
Full-text available
Recent X-ray and radio observations by Muno et al. and Bower et al. have identified a transient low-mass X-ray binary (LMXB) located only 0.1 pc in projection from the Galactic center, CXOGC J174540.0-290031. In this paper, we report the detailed analysis of X-ray and infrared observations of the transient and its surroundings. Chandra observations detect the source at a flux of FX=2×10-12 ergs cm-2 s-1 (2-8 keV). After accounting for absorption both in the interstellar medium (ISM) and in material local to the source, the implied luminosity of the source is only LX=4×1034 ergs s-1 (2-8 keV; D=8 kpc). However, the diffuse X-ray emission near the source also brightened by a factor of 2. The enhanced diffuse X-ray emission lies on top of a known ridge of dust and ionized gas that is visible in infrared images. We interpret the X-ray emission as scattered flux from the outburst and determine that the peak luminosity of CXOGC J174540.0-290031 was LX>~2×1036 ergs s-1. We suggest that the relatively small observed flux results from the fact that the system is observed nearly edge-on, so that the accretion disk intercepts most of the flux emitted along our line of sight. We compare the inferred peak X-ray luminosity to that of the radio jet. The ratio of the X-ray to radio luminosities, LX/LR~105). This is probably because the jets are radiating with unusually high efficiency at the point where they impact the surrounding ISM. This hypothesis is supported by a comparison with mid-infrared images of the surrounding dust. Finally, we find that the minimum power required to produce the jet, Ljet~1037 ergs s-1, is comparable to the inferred peak X-ray luminosity. This is the most direct evidence yet obtained that LMXBs accreting at low rates release about half of their energy as jets.
Article
Full-text available
We observed Sgr A* using the Very Large Array (VLA) and the Giant Metrewave Radio Telescope (GMRT) at multiple centimeter and millimeter wavelengths on 2003 June 17. The measured flux densities of Sgr A*, together with those obtained from the Submillimeter Array (SMA) and the Keck II 10 m telescope on the same date, are used to construct a simultaneous spectrum of Sgr A* from 90 cm to 3.8 μm. The simultaneous spectrum shows a spectral break at about 3.6 cm, a possible signature of synchrotron self-absorption of the strong radio outburst that occurred near epoch 2003 July 17. At 90 cm, the flux density of Sgr A* is 0.22 ± 0.06 Jy, suggesting a sharp decrease in flux density at wavelengths longer than 47 cm. The spectrum at long cm wavelengths appears to be consistent with free-free absorption by a screen of ionized gas with a cutoff ~100 cm. This cutoff wavelength appears to be three times longer than that of ~30 cm suggested by Davies, Walsh, & Booth based on observations in 1974 and 1975. Our analysis suggests that the flux densities of Sgr A* at wavelengths longer than 30 cm could be attenuated and modulated by stellar winds from massive stars close to Sgr A*.
Book
This clear, straightforward, fundamental introduction to radiative processes in astrophysics is designed to present - from a physicist's viewpoint - radiation processes and their applications to astrophysical phenomena and space science. The book covers such topics as radiative transfer theory, relativistic covariance and kinematics, bremsstrahlung radiation, Compton scattering, some plasma effects, and radiative transitions in atoms. The discussion begins with first principles, physically motivating and deriving all results rather than merely presenting finished formulas. Much of the prerequisite material is provided by brief reviews, making the book a self-contained reference tool. Also included are about 75 problems with solutions, illustrating applications of the material and methods for calculating results
Conference Paper
NAOS is the first adaptive optics system installed at the VLT 8m telescopes. It was designed, manufactured and tested by a french Consortium under an ESO contract, to provide compensated images to the high angular resolution IR spectro-imaging camera (CONICA) in the 1 to 5 mum spectral range. It is equipped with a 185 actuator deformable mirror, a tip/tilt mirror and two wavefront sensors, one in the visible and one in the near IR spectral range. It has been installed in November at the Nasmyth focus B of the VLT UT4. During the first light run in December 2001, NAOS has delivered a Strehl ratio of 50 under average seeing conditions for bright guide stars. The diffraction limit of the telescope has been achieved at 2.2 mum. The closed loop operation has been very robust under bad seeing conditions. It was also possible to obtain a substantial correction with mV=17.6 and mK=13.1 reference stars. The on-sky acceptance tests of NAOS-CONICA were completed in May 2002 and the instrument will be made available to the European astronomical community in October by ESO. This paper describes the system and present the on-sky performance in terms of Strehl ratio, seeing conditions and guide star magnitude.
Article
The Adaptive Optics NIR Instrument NAOS-CONICA has been commissioned at the VLT (UT4) between November 2001 and March 2002. After summarizing the observational capabilities of this multimode instrument in combination with the powerful AO-system, we will present first on sky results of the instrumental performance for several non-direct imaging modes: High spatial resolution slit-spectroscopy in the optical and thermal NIR region has been tested. For compact sources below 2 arcsec extension, Wollaston prism polarimetry is used. For larger objects the linear polarization pattern can be analyzed by wire grids down to the diffraction limit. Coronographic masks are applied to optimize imaging and polarimetric capabilities. The cryogenic Fabry-Perot Interferometer in combination with an 8m-telescope AO-system is shown to be a powerful tool for imaging spectroscopy (3D-scans).© (2003) COPYRIGHT SPIE--The International Society for Optical Engineering. Downloading of the abstract is permitted for personal use only.
Article
NAOS is the first adaptive optics system installed at the VLT 8m telescopes. It was designed, manufactured and tested by a french Consortium under an ESO contract, to provide compensated images to the high angular resolution IR spectro-imaging camera (CONICA) in the 1 to 5 μm spectral range. It is equipped with a 185 actuator deformable mirror, a tip/tilt mirror and two wavefront sensors, one in the visible and one in the near IR spectral range. It has been installed in November at the Nasmyth focus B of the VLT UT4. During the first light run in December 2001, NAOS has delivered a Strehl ratio of 50 under average seeing conditions for bright guide stars. The diffraction limit of the telescope has been achieved at 2.2 μm . The closed loop operation has been very robust under bad seeing conditions. It was also possible to obtain a substantial correction with m V=17.6 and m K=13.1 reference stars. The on-sky acceptance tests of NAOS-CONICA were completed in May 2002 and the instrument will be made available to the European astronomical community in October by ESO. This paper describes the system and present the on-sky performance in terms of Strehl ratio, seeing conditions and guide star magnitude.© (2003) COPYRIGHT SPIE--The International Society for Optical Engineering. Downloading of the abstract is permitted for personal use only.