1: %\documentclass[12pt,preprint]{aastex}
2: \documentclass{emulateapj}
3:
4: \usepackage{amsmath}
5: \usepackage{xspace}
6: \usepackage{graphicx}
7: %\usepackage{txfonts}
8: \usepackage{natbib}
9:
10: \newcommand{\myemail}{mfalanga@cea.fr}
11:
12: \shorttitle{The Stress Edge in Sgr A*}
13:
14: \begin{document}
15:
16: \title{Modulated X-ray Emissivity near the Stress Edge in Sgr A*}
17:
18: \author{Maurizio Falanga\altaffilmark{1,2}, Fulvio
19: Melia\altaffilmark{3,4},
20: Martin Prescher\altaffilmark{3}, Guillaume
21: B\'elanger\altaffilmark{5},
22: Andrea Goldwurm\altaffilmark{1,6}}
23:
24: \altaffiltext{1}{CEA Saclay, DSM/IRFU/Service d'Astrophysique,
25: 91191 Gif-sur-Yvette, France; mfalanga@cea.fr}
26: \altaffiltext{2}{AIM - Unit\'e Mixte de Recherche CEA - CNRS -
27: Universit\'e Paris Diderot}
28: \altaffiltext{3}{Physics Department and Steward Observatory, The
29: University of Arizona, Tucson, AZ 85721}
30: \altaffiltext{4}{Sir Thomas Lyle Fellow and Miegunyah Fellow}
31: \altaffiltext{5}{ESA/ESAC, Apartado 50727, 28080 Madrid, Spain}
32: \altaffiltext{6}{UMR Astroparticule et Cosmologie,
33: 10, rue Alice Domont et L\'eonie Duquet, 75005 Paris Cedex 13, France}
34: \begin{abstract}
35:
36: Sgr A* is thought to be the radiative manifestation of a $\sim 3.6\times
37: 10^6$ $M_\odot$ supermassive black hole at the Galactic center. Its
38: mm/sub-mm spectrum and its flare emission at IR and X-ray wavelengths may
39: be produced within the inner ten Schwarzschild radii of a hot, magnetized
40: Keplerian flow. The lightcurve produced in this region may exhibit
41: quasi-periodic variability. We present ray-tracing simulations to
42: determine the general-relativistically modulated X-ray luminosity
43: expected from plasma coupled magnetically to the rest of the disk
44: as it spirals inwards below the innermost stable circular orbit
45: towards the ``stress edge" in the case of a Schwarzschild metric.
46: The resulting lightcurve exhibits a modulation similar to that
47: observed during a recent X-ray flare from Sgr A*.
48: \end{abstract}
49:
50: \keywords{accretion---black hole physics---Galaxy:
51: center---magnetohydro\-dynamics---plasmas---Instabilities}
52:
53: \section{Introduction}
54: \label{intro}
55: Sgr A*'s time-averaged spectrum is roughly a power law below 100\ GHz,
56: with a flux density $S_\nu\propto\nu^\alpha$, where $\alpha\sim$\ 0.19--0.34.
57: In the mm/sub-mm region, however, Sgr A*'s spectrum is dominated by a
58: ``bump'' \citep{Zylka92}, indicative of two different emission components
59: \citep{Melia00,Agol00}. Higher frequencies correspond to smaller spatial
60: scales \citep{Melia92,Narayan95}, so the mm/sub-mm radiation is likely
61: produced near the black hole (BH). X-ray flares detected from Sgr A*
62: \citep{Baganoff01,Goldwurm03,Porquet03,Belanger05} may also have been
63: produced within this compact region, either from a sudden increase in
64: accretion accompanied by a reduction in the anomalous viscosity, or from
65: the quick acceleration of electrons near the BH \citep{LiuMelia02,LiuPetrosian04}.
66: The energized electrons may also manifest themselves via enhanced emission
67: in a hypothesized jet \citep{markoff01}.
68:
69: Near-IR flares detected from Sgr A* appear to be modulated
70: with a variable period $\approx 17$ minutes
71: \citep{Genzel03,Eckart06,Mayer06,Eckart07}. The X-ray
72: and near-IR flares may be coupled via the same electron population,
73: so one may expect similarities in their lightcurves. A long
74: X-ray flare detected with XMM-{\it Newton} in 2004 also appears to have a
75: modulated lightcurve, though not characterized by a fixed period
76: \citep{Belanger08}. If real, the modulation in both the near-IR and
77: X-ray events is almost certainly quasi-periodic rather than periodic,
78: with a decreasing cycle from start to end.
79: But are the fluctuations due to a single azimuthal perturbation
80: (i.e., a ``hotspot"), or from a global pattern of disturbance
81: with a speed not directly associated with the underlying Keplerian
82: period \citep[][]{TM06,Fal07}? In this
83: {\it Letter} we examine the nature of the observerd quasi-period,
84: and focus on its implications for the flow of matter through the
85: innermost stable circular orbit (ISCO). A principal result of this
86: study is a ray-tracing simulation of the general-relativistically (GR)
87: modulated lightcurve produced as the disrupted plasma spirals
88: inwards towards the disk's ``stress edge" \citep{Krolik02}.
89:
90: \section{Background}
91: \label{back}
92: Magnetohydrodynamic (MHD) simulations of Sgr A*'s disk
93: demonstrate the growth of a Rossby-wave instability, enhancing the
94: accretion rate for several hours, possibly accounting for the observed
95: flares \citep{TM06}. The lightcurve produced by GR effects during a
96: Rossby-wave induced spiral pattern in the disk fit the data relatively
97: well, with a quasi-period associated with the pattern speed rather than
98: the Keplerian motion \citep{Fal07}. However, MHD simulations of
99: black-hole accretion suggest that magnetic reconnection might take
100: place within the plunging region, due to the presence of a
101: non-axisymmetric spiral density structure, initially caused by the
102: magnetorotational instability (MRI) associated with differential
103: rotation of frozen-in plasma \citep[see, e.g.,][]{Hawley01}.
104:
105: In this case, the accreting flow is no longer Keplerian because of a
106: radial velocity component. If Sgr A*'s quasi-period of $\sim$ 17--25 minutes is
107: associated with this kind of process rather than a pattern rotation,
108: it would place the corresponding emission region at $0.73$--$0.94\;
109: r_{\rm ISCO}$ radii, below the ISCO (where $r_{\rm ISCO}= 3r_{s}=
110: 6GM/c^2$) for a Schwarzschild BH. Theoretically, we may therefore
111: distinguish the ISCO from the radius at which the inspiraling material
112: actually detaches from the rest of the magnetized disk---the so-called
113: {\it stress} edge \citep{Krolik02}. The X-ray modulation would then be
114: associated with the ever-shrinking period of the emitting plasma as
115: it spirals inwards from the magnetic flare.
116:
117: Interest in ``hotspots'' began in the early 1980's in connection with
118: quasi-periodic flux modulations observed in BHs accreting from a binary
119: companion. The hotspots are possibly overdense
120: emission regions associated with magnetic instabilities. But even with a
121: hotspot, a Newtonian disk does not produce a modulation since its aspect
122: does not affect the total luminosity observed from it. Other than a dynamical
123: periodicity (such as that due to an azimuthal, radial, or vertical oscillation),
124: only GR effects can produce time-dependent photon trajectories resulting in
125: a modulated lightcurve \citep[see e.g.][]{cb73,abramo91,kb92,Holly95,Falcke00,Brom01}.
126: Even so, the ``standard" disk picture of hotspot modulation has been based on
127: Keplerian motion, for which one then expects a time variability directly related
128: to the Keplerian frequency. Here, the modulation is not associated with
129: such a fixed Keplerian frequency, but from a shrinking orbit and
130: a monotonically decreasing period (see \S\ \ref{model}). The relevance
131: of hotspots has already appeared in \citep[][for review]{Holly95,Mayer06,
132: Eckart07,Melia07}. What is lacking, however, is a non Keplerian treatment
133: of the motion with the intent of probing the stress edge itself.
134:
135: So where exactly is the inner edge of the accretion disk in
136: Sgr A*? This is a question asked in a broader context by
137: \citet{Krolik02}, whose MHD simulations
138: of the plunging region in a pseudo-Newtonian potential
139: identified several characteristic inner radii. Here, we assume a
140: non-spinning BH, so our model pertains solely to the Schwarzschild case.
141:
142: The monotonic decrease of the period during the flares suggests that
143: we are witnessing the evolution of an event moving inwards across
144: the ISCO. The inflow time scale, $t_{\rm inflow}$, which determines
145: the rate at which plasma can move from one orbit to another, is given
146: by $\tau_v=r_g/v_{\rm inflow} \approx 9.6\,(r/r_g)^{1/2}$ minutes
147: \citep{LiuMelia02} and is approximately 23.5 minutes at $r=3r_s=6r_g$,
148: corresponding to the ISCO for a non-rotating (i.e., $a/r_g=0$)
149: BH. This time scale does not explicitly depend on a viscosity parameter
150: since the viscosity is directly tied to the MRI physical process via
151: the induced Maxwell stress \citep{LiuMelia02}. The inflow time scale
152: defined here characterizes local processes occurring within the innermost
153: portion of the disk during the flares. By comparison, the dynamical
154: time scale, $t_d\approx 1.3\,(r/r_g)^{3/2}$, is roughly $19$ minutes
155: at this radius \citep{LiuMelia02}. Thus, the azimuthal asymmetry
156: giving rise to the modulated flux during the flare may be due to a
157: transient event associated with either a dynamical or viscous process
158: close to the ISCO \citep{Melia01a}.
159:
160: For a BH mass of $3.6\times 10^6$ M$_\odot$, the inflow time scale
161: at $r\approx 2.5\,r_s$ (inferred from the {\it average} period)
162: is just slightly larger than the average period, so the event
163: could be due to the sudden reconfiguration of magnetic field lines
164: frozen into plasma rapidly approaching the ISCO and then
165: flowing across it towards the event horizon. Matter flowing past the
166: ISCO may still remain ``magnetically" coupled to the outer accretion
167: flow, so a dynamically more meaningful
168: radius is the so-called {\it stress edge}, where plunging matter loses
169: dynamical contact with the material farther out \citep{Krolik02}.
170: This may simply be defined as the surface on which the inflow speed
171: first exceeds the magnetosonic speed.
172:
173: \begin{figure}[ht]
174: \begin{center}
175: \epsscale{1.0}
176: \includegraphics[scale=0.3,angle=-90]{fig1.ps}
177: \caption{\footnotesize Upper panel: The stress edge radius, $r_{\rm
178: stress}$, in units of $r_{\rm ISCO}$, as a function of $\kappa$,
179: the exponent in the power-law formulation of $\Omega(r)$. The dotted and
180: dashed curves represent a period of 17 and 25 minutes, respectively,
181: using a black-hole mass of $3.6\times 10^6$ M$_\odot$ \citep{Schodel03}.
182: Lower panel: The corresponding ratio of accreted specific angular momentum,
183: $j_{\rm in}$, to the specific angular momentum at the ISCO.
184: \label{fig:fig1}}
185: \end{center}
186: \end{figure}
187: %\vskip -0.1in
188:
189:
190: In their simulations,
191: \citet{Krolik02} determined that this surface occurs somewhere
192: between $0.77r_{\rm ISCO}$ and $r_{\rm ISCO}$. The specific angular
193: momentum $j=r^{2}\Omega(r)$, in terms of the orbital angular frequency
194: $\Omega(r)$, continues to fall below $r_{\rm ISCO}$, though $\Omega$ may
195: not necessarily trace its Keplerian value, $\Omega_K(r)\equiv
196: (GM/r^3)^{1/2}$. In the absence of any magnetic
197: coupling across $r_{\rm ISCO}$, matter would retain all of its specific
198: angular momentum at the ISCO, so that the accreted value of $j$, which
199: we will call $j_{\rm in}$, would then simply be $j_{\rm in}=
200: r_{\rm ISCO}^2\,\Omega_K(r_{\rm ISCO})$. Instead, the MHD simulations
201: show that $j_{\rm in}\approx 0.95\,j(r_{\rm ISCO})$, for which
202: $r_{\rm stress}$ is then $\sim 0.8\,r_{\rm ISCO}$,
203: within the range of values indicated by the location of the
204: trans-magnetosonic surface.
205:
206: If the period in Sgr A* is decreasing monotonically, $j(r)$ will not
207: follow its Keplerian value below $r_{\rm ISCO}$. Therefore we
208: will adopt the formulation $\Omega(r)=\Omega_0\,r^{-\kappa}$ to fit
209: the data in \S\ 3. Clearly, $\kappa=3/2$ corresponds to
210: Keplerian rotation; $\kappa$ is $2$ in the extreme case of angular
211: momentum conservation. A reasonable fit to the data would therefore
212: be associated with $3/2\le\kappa\le 2$. At the boundary $r_{\rm ISCO}$,
213: we expect $\Omega=\Omega_K$, which
214: then forces the constant $\Omega_0$ to have the value $c\sqrt{r_{g}}\,
215: r_{\rm ISCO}^{\kappa-3/2}$. We calculate $r_{\rm stress}$ using the
216: quasi-periods 17 and 25 minutes emerging from the X-ray lightcurve
217: (see \S\ \ref{model}), and this is plotted as a function of $\kappa$
218: in Fig.~\ref{fig:fig1}.
219: The radius $r_{\rm stress}$ falls within the range $0.73$--$0.96\;
220: r_{\rm ISCO}$ for all permitted values of $\kappa$. The corresponding
221: accreted specific angular momentum, for the same parameters as used
222: before (see Fig.~\ref{fig:fig1}), is $0.85\,j(r_{\rm ISCO})\le j_{\rm in}
223: \le j(r_{\rm ISCO})$ as a function of $\kappa$. The ratio
224: $j_{\rm in}/j(r_{\rm ISCO})=0.95$ from the MHD simulations
225: would require $\kappa\sim 1.8$, for which $r_{\rm stress}
226: \sim 0.77\, r_{\rm ISCO}$. These results are consistent with the
227: MHD simulations, indicating that the infalling plasma below the ISCO
228: remains magnetically coupled to the outer disk, though the dissipation
229: of angular momentum is not quite strong enough in this region to force
230: the gas into Keplerian rotation.
231:
232: \section{The Inspiraling Plasma Model}
233: \label{model}
234: With $\Omega(r)$ known, we now incorporate strong gravitational
235: effects in a geometrically and optically thin disk, describing
236: the inspiraling disturbance using coordinates ($r,\theta,\varphi$)
237: in the co-rotating frame. In Fig.~\ref{fig:fig2}, we show the inspiraling
238: trajectory and duration of the emitting plasma. The
239: observer is located at infinity with a viewing angle {\it i}
240: relative to the $z'$-axis in the non-rotating frame, at (observer)
241: polar coordinates ($r',\theta',\varphi'$). The deflection angle of
242: a photon emitted by plasma in the inspiraling region is $\psi$,
243: varying periodically with $\cos\, \psi = \cos\,i\, \cos \,\varphi$,
244: for a disk in the plane $\theta=\pi/2$. Also, for $G=c=1$,
245: the BH's horizon occurs at $r_{s} = 2M$, and the last stable
246: orbit is located at $r_{\rm ISCO} = 3r_{\rm s}$.
247:
248: \begin{figure}[ht]
249: \begin{center}
250: \epsscale{1.0}
251: \includegraphics[scale=0.45,angle=-90]{fig2.ps}
252: \caption{\footnotesize Upper panel: The inspiraling trajectory
253: of the hotspot, beginning at $r_{\rm ISCO}$ and terminating at
254: $0.74r_{\rm ISCO}$. The dotted circle represents the location of
255: the event horizon. Lower panel: The period as a function of
256: the stress radius for the two extreme values of $\kappa$ adopted
257: here, assuming a black-hole mass of $3.6\times10^6\;M_\odot$.
258: \label{fig:fig2}}
259: \end{center}
260: \end{figure}
261: \vskip -0.1in
262:
263: We calculate the lightcurve using a full ray-tracing algorithm
264: \citep[see][]{Luminet79,Fal07}.
265: The disk from $r_{\rm ISCO}$ to 90$r_{s}$ is
266: an unperturbed, Keplerian flow, with angular velocity $\Omega_{\rm
267: K}$, and with specific angular momentum $j_{\rm K}=
268: r^{2}u^{\varphi}/u^{t}=r^{2}\Omega_{\rm K}$.
269: The corresponding four-velocity of the effective flow is then
270: $(u^{t},u^{r},u^{\theta},u^{\varphi})=u^{t}(1,0,0,\Omega_{K})$,
271: where $u^{t}=(1-3M/r)^{-1/2}$ \citep{MTW73}. The accretion
272: flow is no longer Keplerian below the ISCO.
273:
274: Triggering a perturbation induces an azimuthal asymmetry in the region
275: $r_{\rm stress} \approx 0.73 <r< 0.9 r_{\rm ISCO}$. Below $r_{\rm ISCO}$,
276: we use a simple representation of the bulk velocity field, in which
277: $\Omega(r)=u^{\varphi}_{\rm sw}/u_{\rm sw}^{t}$, as described e.g.,
278: in \citet{Fukumura04}:
279: \vskip -0.1in
280: \begin{equation}
281: v^{r}_{\rm sw} = - A_{r} e^{-(r-r_{\rm
282: stress})/\Delta_{\rm sw}}\sin^{\gamma_{0}}[k_{r}(r-r_{\rm
283: stress})+m\varphi/2-\varphi_{\rm sw}/2]\;.
284: \end{equation}
285: \vskip 0.01in\noindent
286: In this case, the specific angular momentum is $j_{\rm in}=r^{2}u^{\varphi}_{\rm
287: sw}/u_{\rm sw}^{t}=\Omega_{0} r^{2-\kappa}$. The subscript ``sw" denotes the
288: spiral wave, and the number $m$ is the azimuthal wavenumber, fixed to
289: be $m=1$ for a single-armed spiral wave. The constant $\gamma_{0}=2$ is
290: the width of the spiral wave, $A_{r}=0.1$ and
291: $A_{\varphi}=0.1$ are the amplitudes chosen to be relatively small,
292: $k_{r}$ characterizes a tightness (i.e., the
293: number of windings) of the spiral, and the effective radial range of
294: the spiral motion is controlled by $\Delta_{\rm sw}=30$, and
295: $\varphi_{\rm sw}=0$ denotes the phase of the spiral. Since
296: ($u^{r}_{\rm sw}$, $u^{\varphi}_{\rm sw}$) is not axisymmetric, the
297: net velocity field is also non-axisymmetric. For the effective flow then,
298: $(u_{\rm sw}^{t},u^{r}_{\rm sw},u_{\rm sw}^{\theta},u^{\varphi}_{\rm sw}) =
299: u_{\rm sw}^{t}(1,v^{r}_{\rm sw},0,\Omega_0\,r^{-k})$, where $u_{\rm sw}^{t}=[
300: (1-2/m)-(1-2m/r)^{-1}(v_{\rm sw}^{r})^{2}-r^{2}\Omega]^{1/2}$,
301: corresponding to the four-vector normalization condition
302: $g_{\alpha,\beta}u^{\alpha}u^{\beta}=-1$.
303:
304: We consider four GR effects: (i) light-bending, (ii) gravitational
305: Doppler effect defined as (1+z), taking into account the non-axisymmetric
306: radial and azimuthal components below $r_{\rm ISCO}$, (iii) gravitational
307: lensing, $d\Omega_{\rm obs}=b\,db\,d\varphi/D^2$ (with $D$ the distance
308: to the source), expressed through the impact parameter, and (iv) the travel
309: time delay. The relative time delay between photons arriving at the observer
310: from different parts of the disk are calculated from the geodesic equation.
311: The first photon arrives from phase $\varphi=0$ and $r=r_{\rm ISCO}$, and
312: defines the reference time, $T_{0}$, which is set to zero. The observed
313: time is then the orbital time plus the light-bending travel time delays,
314: i.e., $T_{\rm obs}(\varphi_{\rm sw},r,i) = \Omega^{-1}(r)\varphi_{\rm sw}
315: + \Delta T_{\rm GR}$.
316:
317: The observed flux at energy $E'$ is $F_{\rm obs}(E') =
318: I_{\rm obs}(E')d\Omega_{\rm obs}$, where $I_{\rm obs}(E')$
319: is the radiation intensity observed at infinity and
320: $d\Omega_{\rm obs}$ is the solid angle on the observer's sky
321: including relativistic effects. Using the relation
322: $I_{\rm obs}(E',\alpha')=(1+z)^{-3} I_{\rm em}(E,\alpha)$,
323: a Lorentz invariant quantity that is constant along null geodesics in
324: vacuum, the intensity of a light source integrated over its effective
325: energy range is proportional to the fourth power of the redshift
326: factor, $I_{\rm obs}(\alpha')= (1+z)^{-4}I_{\rm em}(r,\varphi)$,
327: $I_{\rm em}(r,\varphi)$ being the intensity measured in the rest
328: frame of the inspiraling disturbance \citep{MTW73}. The
329: disk radiates an inverse Compton spectrum, $I_{\rm em}$,
330: calculated using the parameter scalings, rather than their absolute
331: values. The spectrum parameters are
332: \citep{Melia01} the disk temperature, $T(r)$, the electron
333: number density, $n_e(r)$, the magnetic field, $B(r)$, and the disk
334: height $H(r)$. This procedure gives correct amplitudes in the
335: lightcurve, though not the absolute value of the flux per se.
336:
337: The synchrotron emissivity is therefore $j_s \propto B\,n_{\rm nt}
338: \propto B\,T\,n_e$, where the nonthermal particle energy is roughly
339: in equipartition with the thermal. The X-rays are produced via inverse
340: Compton scattering from the seed
341: photon number flux. Thus, with $L_{\rm seed}\propto r^3\, j_s$, where $j_s$
342: is the synchrotron emissivity in units of energy per unit volume per unit
343: time, the soft photon flux scales as the emitted power divided by the
344: characteristic area. That is, $F_{\rm seed}\propto r^3\, j_s / r^2 = r j_s$,
345: which is going to be roughly the same scaling as the seed photon density, so
346: $n_{\rm seed} \propto r j_s \propto r\,B\,T\,n_e$. The inverse
347: Compton scattering emissivity is therefore $j_{ic} \propto n_{\rm nt}\,
348: n_{\rm seed} \propto (T\,n_e)^2\,r\,B$. Thus, $j_{x}\sim j_{ic}$,
349: and the surface intensity is $I_{\rm em} \propto \int j_{x} ds \propto
350: j_{x} H$, which gives finally $I_{\rm em} \propto
351: (T\,n_e)^2\,r\,B\,H$.
352:
353: The flux at a given azimuthal angle $\varphi$ and radius $r$ is calculated
354: from a numerical computation of $\psi(\alpha)$, followed by a calculation
355: of the Doppler shift, lensing effects, and the flux $F_{\rm obs}$ as a
356: function of the arrival time. For the persistent emission we use
357: the best fit spectral parameters to the
358: {\it Chandra} data \citep{Melia01,Baganoff01}, described above as
359: a surface emissivity $I_{\rm em}$. The observed flare
360: normalized flux is modeled with two polynomials, one between 0--100
361: minutes and the second from 100-160 minutes \citep[see also][]{Mayer06}.
362: The value $k_{r}$ is fixed at 11 to have the six observed cycles
363: (see Fig. \ref{fig:fig3}, solid line). The free parameters to
364: fit the data are the inclination angle $i$ and the $\kappa$
365: value. The integrated flux is calculated for an extended spiral
366: wave 90$^{\circ}$ long in the azimuthal direction and $\Delta r =
367: 0.28r_{g}$ in the radial direction, plus the persistent emission.
368: The MHD simulations show that in the innermost part of the
369: disk a spiral-arm often expands out to $\sim90^{\circ}$ (see, e.g.,
370: Hawley 2001). The radial extent of the inspiraling region is set
371: by the observed condition that six cycles
372: should fit within the overall migration of the plasma from the ISCO
373: to the stress edge. In Fig.~\ref{fig:fig3} (solid line), we show the
374: best fit model for $72\pm3^{\circ}$ and $\kappa=1.7\pm0.05$.
375:
376: \begin{figure}[ht]
377: \begin{center}
378: \epsscale{1.0}
379: \includegraphics[scale=0.4,angle=-90]{fig3.ps}
380: \caption{\footnotesize Lightcurve of the August 31, 2004
381: flare in the 2--10 keV energy band \citep{Belanger08},
382: normalized with the observed mean count rate of 0.231 cts s$^-1$ for the
383: flare duration. The best fit model for an inspiraling disturbance
384: is shown by the solid line using an inclination angle 72$^{\circ}$ and
385: $\kappa=1.7$. The dotted curve represents a constant Keplerian period
386: at the last stable orbit, i.e., $r_{\rm ISCO}$, $i= 72^{\circ}$, and
387: $\kappa=1.5$. Panels (a) and (b) show the residuals (in units of sigma)
388: of the inspiraling and constant-period model, respectively, compared
389: to the data.
390: \label{fig:fig3}}
391: \end{center}
392: \end{figure}
393:
394:
395: \section{Conclusion}
396: \label{conclusion}
397:
398: If we adopt the simple view that the last period
399: corresponds to the ISCO, then Sgr A* with a mass of $3.6\times 10^6\;
400: M_\odot$ must be spinning at a rate $a/r_g\gtrsim 0.2-0.4$. With a more
401: realistic analysis of the magnetic coupling between matter in the
402: plunging region and that beyond the ISCO, we conclude that the peak
403: of the instability probably occurs at $\sim 0.97r_{\rm ISCO}$, where
404: the period is $\sim 25$ minutes, and the flaring activity continues
405: as the plasma spirals inwards, ending several orbits later when the
406: matter crosses the stress edge at $\sim 0.8 r_{\rm ISCO}$.
407:
408: The significance of the fit for an inspiraling disturbance
409: is $\chi^2/d.o.f. = 92.4/39$, compared to $\chi^2/d.o.f. = 285.2/46$
410: for a fixed Keplerian period (see dotted curve in
411: Figure~\ref{fig:fig3}). An inspiraling disturbance
412: is preferred over a fixed orbit by a factor 2.6 in the reduced $\chi^2$.
413: The residuals in the lower panels of Figure ~\ref{fig:fig3} show
414: that the model using a fixed period produces modulations that are
415: progressively shifted in phase with respect to the data, by as much as
416: $\sim 16.5$ minutes by the end of the flare. The inspiraling
417: model, on the other hand, follows the evolution of the flare
418: and therefore fits the data much better. Plasma on such an orbit
419: also produces a constant pulsed fraction$=(I_{\rm max}-I_{\rm mean})/
420: (I_{\rm max}+I_{\rm min})$ of $\sim 9\%$, compared with a linear
421: increase from $\sim9\%$ to $\sim 11\%$ for the inspiralling wave;
422: this effect is due to a radially-dependent gravitational
423: lensing effect. Together, these two effects render the inspiraling
424: scenario a better explanation for the data than the fixed orbit
425: disturbance.
426:
427: \acknowledgments
428: \noindent MF is grateful to Keigo Fukumura for helpful discussions. This
429: research was supported by NSF grant AST-0402502 in Arizona,
430: and by the French Space Agency (CNES).
431:
432: \begin{thebibliography}{}
433:
434: \bibitem[\protect\citeauthoryear{Abramowicz et al.}{1991}]{abramo91}
435: Abramowicz, M. A., Bao, G., Lanza, A., Zhang, X.-H., A\&A, 245, 454
436:
437: \bibitem[\protect\citeauthoryear{Agol}{2000}]{Agol00}
438: Agol, E. 2000, ApJ, 538, L121
439:
440: \bibitem[\protect\citeauthoryear{Baganoff et al.}{2001}]{Baganoff01}
441: Baganoff, F. et al. 2001, Nature, 413, 45
442:
443: \bibitem[\protect\citeauthoryear{B\'elanger et al.}{2005}]{Belanger05}
444: B\'elanger, G. et al. 2005, ApJ, 635, 1095
445:
446: \bibitem[\protect\citeauthoryear{B\'elanger et al.}{2008}]{Belanger08}
447: B\'elanger, G. et al. 2008, ApJ, submitted (astro-ph/0604337)
448:
449: \bibitem[\protect\citeauthoryear{Bromley et al.}{2001}]{Brom01}
450: Bromley, B. C., Melia, F., Liu, S. 2001, ApJ Letters, 555, L83
451:
452: \bibitem[\protect\citeauthoryear{Cunningham \& Bardeen}{1973}]{cb73}
453: Cunningham, C. T. \& Bardeen, J. 1973, ApJ, 183, 237
454:
455: \bibitem[\protect\citeauthoryear{Eckart et al.}{2006}]{Eckart06}
456: Eckart, A., Sch\"odel, R., Meyer, L. et al. 2006, A\&A, 450, 535
457:
458: \bibitem[\protect\citeauthoryear{Eckart et al.}{2007}]{Eckart07}
459: Eckart, A, et al. 2007, A\&A, in press, [arXiv:0712.3165]
460:
461: \bibitem[\protect\citeauthoryear{Falanga et al.}{2007}]{Fal07}
462: Falanga, M. et al. 2007, ApJ, 662, L15
463:
464: \bibitem[\protect\citeauthoryear{Falcke et al.}{2000}]{Falcke00}
465: Falcke, H., Melia, F., Agol, E. 2000, ApJ Letters, 528, L13
466:
467: \bibitem[\protect\citeauthoryear{Fukumura \& Tsuruta}{2004}]{Fukumura04}
468: Fukumura, K. \& Tsuruta, S. 2004, ApJ, 613, 700
469:
470: \bibitem[\protect\citeauthoryear{Genzel et al.}{2003}]{Genzel03}
471: Genzel, R. et al. 2003, Nature, 425, 934
472:
473: \bibitem[\protect\citeauthoryear{Goldwurm et al.}{2003}]{Goldwurm03}
474: Goldwurm, A. et al. 2003, ApJ, 584, 751
475:
476: \bibitem[\protect\citeauthoryear{Hawley}{2001}]{Hawley01}
477: Hawley, J., F., 2001, ApJ, 554, 534
478:
479: \bibitem[\protect\citeauthoryear{Hollywood et al.}{1995}]{Holly95}
480: Hollywood, J. M., Melia, F., Close, L. M. et al., 1995, ApJ, 448, L21
481:
482: \bibitem[\protect\citeauthoryear{Karas \& Bao}{1992}]{kb92}
483: Karas, V. \& Bao, G., 1992, A\&A, 257, 531
484:
485: \bibitem[\protect\citeauthoryear{Krolik \& Hawley}{2002}]{Krolik02}
486: Krolik, J. H. \& Hawley, J. F. 2002, ApJ, 573, 754
487:
488: \bibitem[\protect\citeauthoryear{Liu \& Melia}{2002}]{LiuMelia02}
489: Liu, S. \& Melia, F., 2002, ApJ, 566, L77
490:
491: \bibitem[\protect\citeauthoryear{Liu et al.}{2004}]{LiuPetrosian04}
492: Liu, S., Petrosian, V. \& Melia, F. 2004, ApJ, 611, L101
493:
494: \bibitem[\protect\citeauthoryear{Luminet}{1979}]{Luminet79}
495: Luminet, J., -P. 1979, A\&A, 75, 228
496:
497: \bibitem[\protect\citeauthoryear{Markoff et al.}{2001, and references
498: cited therein}]{markoff01}
499: Markoff, S., Falcke, H., Yuan, F., Biermann, P. L. 2001, A\&A, 379, L13
500:
501: \bibitem[\protect\citeauthoryear{Meyer et al.}{2006}]{Mayer06}
502: Meyer, L., Eckart, A., Sch\"odel, R., et al. 2006, A\&A, 460, 15
503:
504: \bibitem[\protect\citeauthoryear{Melia}{1992}]{Melia92}
505: Melia, F. 1992, ApJ, 387, L25
506:
507: \bibitem[\protect\citeauthoryear{Melia}{2007}]{Melia07}
508: Melia, F. 2007, The Galactic Supermassive Black Hole, PUP (New York)
509:
510: \bibitem[\protect\citeauthoryear{Melia et al.}{2000}]{Melia00}
511: Melia, F. Liu, S. \& Coker, R. 2000, ApJ, 545, L117
512:
513: \bibitem[\protect\citeauthoryear{Melia et al.}{2001}]{Melia01}
514: Melia, F., Liu, S. \& Coker, R., 2001, ApJ, 553, 146
515:
516: \bibitem[\protect\citeauthoryear{Melia et al.}{2001}]{Melia01a}
517: Melia, F., Bromley, B., C., Liu, S., Walker, C., K., 2001, ApJ, 554, L37
518:
519: \bibitem[\protect\citeauthoryear{Misner et al.}{1973}]{MTW73}
520: Misner, C. W., Thorne, K., S., \& Wheeler, J., A. 1973, {\it Gravitation}
521: (San Francisco: Freeman)
522:
523: \bibitem[\protect\citeauthoryear{Narayan et al.}{1995}]{Narayan95}
524: Narayan, R., Yi, I. \& Mahadevan, R., 1995, Nature, 374, 623
525:
526: \bibitem[\protect\citeauthoryear{Porquet et al.}{2003}]{Porquet03}
527: Porquet, D. et al. 2003, A\&A, 407, L17
528:
529: \bibitem[\protect\citeauthoryear{Sch\"odel et al.}{2003}]{Schodel03}
530: Sch\"odel, R., Ott, R., Genzel, R. et al. 2003, ApJ, 596, 1015
531:
532: \bibitem[\protect\citeauthoryear{Tagger \& Melia}{2006}]{TM06}
533: Tagger, M. \& Melia, F. 2006, ApJ, 636, L33
534:
535: \bibitem[\protect\citeauthoryear{Zylka et al.}{1992}]{Zylka92}
536: Zylka, R., Mezger, P. \& Lesch, H., 1992, A\&A, 261, 119
537:
538: \end{thebibliography}
539:
540: \end{document}
541:
542: