1: %%
2: %% Beginning of file 'sample.tex'
3: %%
4: %% Modified 2005 December 5
5: %%
6: %% This is a sample manuscript marked up using the
7: %% AASTeX v5.x LaTeX 2e macros.
8:
9: %% The first piece of markup in an AASTeX v5.x document
10: %% is the \documentclass command. LaTeX will ignore
11: %% any data that comes before this command.
12:
13: %% The command below calls the preprint style
14: %% which will produce a one-column, single-spaced document.
15: %% Examples of commands for other substyles follow. Use
16: %% whichever is most appropriate for your purposes.
17: %%
18: %%\documentclass[12pt,preprint]{aastex}
19:
20: %% manuscript produces a one-column, double-spaced document:
21:
22: %\documentclass[manuscript]{aastex}
23: \documentclass[preprint2]{aastex}
24: %\usepackage{psfig}
25: \usepackage{graphicx}
26: \usepackage{epsfig}
27:
28:
29: %% preprint2 produces a double-column, single-spaced document:
30: %% \documentclass[preprint2]{aastex}
31:
32: %% Sometimes a paper's abstract is too long to fit on the
33: %% title page in preprint2 mode. When that is the case,
34: %% use the longabstract style option.
35:
36: %% \documentclass[preprint2,longabstract]{aastex}
37:
38: %% If you want to create your own macros, you can do so
39: %% using \newcommand. Your macros should appear before
40: %% the \begin{document} command.
41: %%
42: %% If you are submitting to a journal that translates manuscripts
43: %% into SGML, you need to follow certain guidelines when preparing
44: %% your macros. See the AASTeX v5.x Author Guide
45: %% for information.
46:
47: \newcommand{\vdag}{(v)^\dagger}
48: \newcommand{\myemail}{skywalker@galaxy.far.far.away}
49:
50: %% You can insert a short comment on the title page using the command below.
51:
52: %\slugcomment{Not to appear in Nonlearned J., 45.}
53:
54: %% If you wish, you may supply running head information, although
55: %% this information may be modified by the editorial offices.
56: %% The left head contains a list of authors,
57: %% usually a maximum of three (otherwise use et al.). The right
58: %% head is a modified title of up to roughly 44 characters.
59: %% Running heads will not print in the manuscript style.
60:
61: \shorttitle{Sgr A$^*$: a laboratory to measure the central black
62: hole and cluster parameters} \shortauthors{De Paolis et al.}
63:
64: %% This is the end of the preamble. Indicate the beginning of the
65: %% paper itself with \begin{document}.
66:
67: \begin{document}
68:
69: %% LaTeX will automatically break titles if they run longer than
70: %% one line. However, you may use \\ to force a line break if
71: %% you desire.
72:
73: \title{Sgr A$^*$: a laboratory to measure the central black hole and
74: stellar cluster parameters}
75:
76: %% Use \author, \affil, and the \and command to format
77: %% author and affiliation information.
78: %% Note that \email has replaced the old \authoremail command
79: %% from AASTeX v4.0. You can use \email to mark an email address
80: %% anywhere in the paper, not just in the front matter.
81: %% As in the title, use \\ to force line breaks.
82:
83: \author{A.A. Nucita, F. De Paolis and G. Ingrosso}
84: \affil{Dipartimento di Fisica, Universit\`a di Lecce, and {\it
85: INFN}, Sezione di Lecce, CP 193, I-73100 Lecce, Italy}
86:
87: \author{A. Qadir}
88: \affil{Center for Advanced Mathematics and Physics, National
89: University of Science and Technology, Campus of College of
90: Electrical and Mechanical Engineering, Peshawar Road, Rawalpindi,
91: Pakistan}
92:
93: \and
94:
95: \author{A.F. Zakharov\altaffilmark{1,2}}
96: \affil{Institute of Theoretical and Experimental Physics, 25,
97: B.Cheremushkinskaya St., Moscow, 117259, Russia}
98:
99: %% Notice that each of these authors has alternate affiliations, which
100: %% are identified by the \altaffilmark after each name. Specify alternate
101: %% affiliation information with \altaffiltext, with one command per each
102: %% affiliation.
103:
104: \altaffiltext{1}{Astro Space Centre of Lebedev Physics Institute,
105: Moscow, Russia} \altaffiltext{2}{Bogoliubov Laboratory for
106: Theoretical Physics, Joint Institute for Nuclear Research, Dubna,
107: Russia}
108:
109:
110: %% Mark off your abstract in the ``abstract'' environment. In the manuscript
111: %% style, abstract will output a Received/Accepted line after the
112: %% title and affiliation information. No date will appear since the author
113: %% does not have this information. The dates will be filled in by the
114: %% editorial office after submission.
115:
116: \begin{abstract}
117: Several stars orbit around a black hole candidate of mass
118: $3.7\times 10^6$ M$_{\odot}$, in the region of the Galactic Center
119: (GC). Looking for General Relativistic (GR) periastron shifts is
120: limited by the existence of a stellar cluster around the black
121: hole that would modify the orbits due to classical effects that
122: might mask the GR effect. Only if one knows the cluster parameters
123: (its mass and core radius) it is possible to unequivocally deduce
124: the GR effects expected and then test them. In this paper it is
125: shown that the observation of the proper motion of Sgr~A$^*$,
126: $v_{Sgr~A^*} = (0.4\pm 0.9)$ km s$^{-1}$ (\citealt{reid2004}),
127: could help us to constrain the cluster parameters significantly
128: and that future measurements of the periastron shifts for at least
129: three stars may adequately determine the cluster parameters and
130: the mass of the black hole.
131: \end{abstract}
132:
133: %% Keywords should appear after the \end{abstract} command. The uncommented
134: %% example has been keyed in ApJ style. See the instructions to authors
135: %% for the journal to which you are submitting your paper to determine
136: %% what keyword punctuation is appropriate.
137:
138: \keywords{Gravitation --- Galaxy: center --- Physics of black
139: holes}
140:
141: %% From the front matter, we move on to the body of the paper.
142: %% In the first two sections, notice the use of the natbib \citep
143: %% and \citet commands to identify citations. The citations are
144: %% tied to the reference list via symbolic KEYs. The KEY corresponds
145: %% to the KEY in the \bibitem in the reference list below. We have
146: %% chosen the first three characters of the first author's name plus
147: %% the last two numeral of the year of publication as our KEY for
148: %% each reference.
149:
150:
151: %% Authors who wish to have the most important objects in their paper
152: %% linked in the electronic edition to a data center may do so by tagging
153: %% their objects with \objectname{} or \object{}. Each macro takes the
154: %% object name as its required argument. The optional, square-bracket
155: %% argument should be used in cases where the data center identification
156: %% differs from what is to be printed in the paper. The text appearing
157: %% in curly braces is what will appear in print in the published paper.
158: %% If the object name is recognized by the data centers, it will be linked
159: %% in the electronic edition to the object data available at the data centers
160: %%
161: %% Note that for sources with brackets in their names, e.g. [WEG2004] 14h-090,
162: %% the brackets must be escaped with backslashes when used in the first
163: %% square-bracket argument, for instance, \object[\[WEG2004\] 14h-090]{90}).
164: %% Otherwise, LaTeX will issue an error.
165:
166: \section{Introduction}
167:
168: GR predicts that orbits about a massive central body suffer
169: periastron shifts yielding {\textit {rosette}} shapes. However,
170: the classical perturbing effects of other objects on inner orbits
171: give an opposite shift. Since the periastron advance depends
172: strongly on the compactness of the central body, the detection of
173: such an effect may give information about the nature of the
174: central body itself. This would apply for stars orbiting close to
175: the GC, where there is a \lq\lq dark object", the black hole
176: hypothesis being the most natural explanation of the observational
177: data. A cluster of stars in the vicinity of the GC (at a distance
178: $< 1\arcsec $) has been monitored by ESO and Keck teams for
179: several years (\cite{Genzel03,Schoedel03,Ghez03,Ghez04,Ghez05}).
180: In particular, Ghez et al. (\citeyear{Ghez03}) have reported on
181: observations of several stars orbiting close to the GC massive
182: black hole. Among those, the S2 star, with mass $M_{\rm S2}\simeq
183: 15$ M$_{\odot}$, appears to be a main sequence star orbiting the
184: black hole with a Keplerian period of $\simeq 15$ yrs. This yields
185: \citet{Ghez05} a mass estimate of $M_{\rm Sgr~A^*}\simeq
186: 3.67\times 10^6 ~M_{\odot}$ within $4.87\times 10^{-3}$ pc, that
187: is the S2 semi-major axis.
188:
189: Several authors have discussed the possibility of measuring the
190: GR corrections to Newtonian orbits for Sgr~A$^*$ (see e.g.
191: \citealt{Jaroszynski98, Jaroszynski99, Jaroszynski00, Fragile00,
192: Rubilar01, Weinberg05}), usually assuming that the central body is
193: a Schwarzschild black hole. However, since black holes generally
194: rotate, and there is no reason why they should not be rotating
195: fast, the Kerr metric should be used instead. Not only stellar
196: mass black holes but also supermassive black holes are believed to
197: be spinning. Indeed, X-ray observations of Seyfert galaxies,
198: microquasars and binary systems (\cite{fabian1, tanaka1, Fabi00,
199: Fabian04} and references therein) show that the data could be
200: explained by a rotating black hole model (see e.g.
201: \cite{zak_rep03_aa, Zak_rep03_AR,ZKLR02} and \cite{ ZR_ASR04}).
202: Further, supermassive black holes at the center of QSOs, AGNs and
203: galaxies show beamed jet emission implying that they have non zero
204: angular momentum. Hence, Kerr black holes may be fairly common in
205: nature. The relatively short orbital period of the S2 star
206: encourages a search for genuine GR effects like the orbital
207: periastron shift. Quite possibly, more suitable stars, close to
208: the GC black hole, will be found in the future. Bini et al.
209: (\citeyear{bini2005}) studied the GR periastron shift around
210: Sgr~A$^*$ and estimated it for various solutions belonging to the
211: Weyl class, including the Schwarzschild and Kerr black holes.
212: However, they did not take into account the presence of a stellar
213: cluster, which could in principle be sizable.
214:
215: The purpose of this paper is to try to find limits for the extent
216: and density of the cluster about Sgr~A$^*$ and if those limits
217: yield a measurement of its spin.
218:
219: Clearly, a thorough knowledge of the cluster mass and density
220: distribution is necessary to be able to infer the mass and spin of
221: the black hole at the GC by measuring the periastron shift and
222: subtracting the Newtonian shift. Unfortunately, the star cluster
223: parameters are poorly known. \footnote{We remark that the star
224: cluster we are considering around the central black hole might
225: contain not only normal stars but also white dwarfs, neutron stars
226: and/or stellar mass black holes.} However, the measure of the
227: Brownian motion of the central black hole due to the surrounding
228: matter may be used to constrain the black hole to cluster mass
229: ratio.\footnote{Other methods for estimating the black hole
230: parameters (i.e. mass and angular momentum) based on gravitational
231: retrolensing have been proposed. For more details on this topic we
232: refer to \citet{depaolis1,depaolis2,depaolis3,depaolis4,ZNDI_04}
233: and reference therein.} The latest observations of the $Sgr~A^*$
234: proper motion, $v_{Sgr~A^*}= (0.4\pm 0.9)$ km $\rm s^{-1}$
235: (\citealt{reid2004}), is much tighter than the earlier one of
236: $2\mathrm{-}20$ km $\rm{s}^{-1}$ (see Reid, Readhead, Vermeulen et
237: al. \citeyear{reid1999}).
238:
239: For a test particle orbiting a Schwarzschild black hole of mass
240: $M_{\rm BH}$, the periastron shift is given by (see e.g. Weinberg,
241: \citeyear{Weinberg72})
242: \begin{equation}
243: \Delta \phi_S \simeq \frac{6\pi G
244: M_{BH}}{d(1-e^2)c^2}+\frac{3(18+e^2)\pi
245: G^2M_{BH}^2}{2d^2(1-e^2)^2c^4}~, \label{schshift}
246: \end{equation}
247: $d$ and $e$ being the semi-major axis and eccentricity of the test
248: particle orbit, respectively. For a rotating black hole with spin
249: parameter $a=|{\bf a}|=J/GM_{\rm BH}$, the space-time is described
250: by the Kerr metric and, in the most favorable case of equatorial
251: plane motion ({\bf a.v} = 0), the shift is given by (Boyer and
252: Price \citeyear{boyerprice}, but see also Bini et al.
253: \citeyear{bini2005} for more details)
254: \begin{equation}
255: \begin{array}{l}
256: \displaystyle{\Delta \phi_K \simeq \Delta \phi_S +\frac{8a\pi
257: M_{BH}^{1/2}G^{3/2}}{d^{3/2}(1-e^2)^{3/2}c^3}+\frac{3a^2\pi
258: G^2}{d^{2}(1-e^2)^{2}c^4}~,} \label{kershift}
259: \end{array}
260: \end{equation}
261: which reduces to eq. (\ref{schshift}) for $a\rightarrow 0$. In the
262: more general case, {\bf a.v} $\neq 0$, the expected periastron
263: shift has to be evaluated numerically.
264:
265: The expected periastron shifts (mas/revolution), $\Delta \phi$ (as
266: seen
267: from the center) and $\Delta \phi _E$ (as seen from Earth at
268: the distance $R_0\simeq~8$ kpc from the GC), for the Schwarzschild
269: and the extreme Kerr black holes, for the S2 and S16 stars turn
270: out to be $\Delta\phi^{S2}=6.3329\times 10^5$ and $6.4410\times
271: 10^5$ and $\Delta \phi _E^{S2}=0.661$ and $0.672$ respectively,
272: and $\Delta\phi^{S16}=1.6428\times 10^6$ and $1.6881\times 10^6$
273: and $\Delta \phi _E^{S16}=3.307$ and $3.399$ respectively. Recall
274: that
275: \begin{equation}
276: \Delta \phi _E = \frac{d(1+e)}{R_0} \Delta \phi_{S,K}~.
277: \end{equation}
278: Notice that the differences between the periastron shifts for the
279: Schwarzschild and the maximally rotating Kerr black hole is at
280: most $0.01$ mas for the S2 star and $0.009$ mas for the S16 star.
281: In order to make these measurements with the required accuracy,
282: one needs to know the S2 orbit with a precision of at least $10$
283: $\mu$as.
284:
285: There is a proposal to improve the angular resolution of VLTI with
286: the PRIMA facility (\cite{Rottgering03, Delplancke03,
287: Quirrenbach03} but see also the related
288: web-site\footnote{http://obswww.unige.ch/PRIMA/home/introduction.}),
289: which, by using a phase referenced imaging technique, will get
290: $\sim 10$ $\mu$as angular resolution. Hence, at least in
291: principle, the effect of a maximally rotating Kerr black hole on
292: the periastron shift of the S2 star can be distinguished from that
293: produced by a Schwarzschild black hole with the same mass.
294:
295: The plan of the paper is as follows: In the next section we
296: briefly discuss the effect of a central star cluster on the
297: periastron advance. In Section 3 we use the Sgr~A$^*$ Brownian
298: motion to constrain the black hole to star cluster mass ratio.
299: Then we consider whether the detection of the spin of the black
300: hole from the periastron shift of the S2 star is possible, once
301: the cluster density and size have been adequately constrained. In
302: Section 5 we show how future measurements of the periastron shifts
303: for at least three stars close to the GC black hole may be used to
304: estimate the black hole mass and the star cluster mass density
305: distribution. In the next section we consider what the
306: observational requirements would be for adequate determination of
307: the cluster parameters to be able to resolve the Kerr effect.
308: Finally, in section 7, we present some concluding remarks.
309:
310: \section{Retrograde shift due to a central stellar cluster}
311:
312: The star cluster surrounding the central black hole in the GC
313: could be sizable. At least 17 members have been observed within 15
314: mpc up to now (\citealt{Ghez05}). However, the cluster mass and
315: density distribution, that is to say its mass and core radius, is
316: still unknown. The presence of this cluster affects the periastron
317: shift of stars orbiting the central black hole. The periastron
318: advance depends strongly on the mass density profile and
319: especially on the central density and typical length scale.
320:
321: We model the stellar cluster by a Plummer model density profile
322: (\citealt{binneytremaine})
323: \begin{equation}
324: \rho_{CL}(r) = \rho_0 f(r)~,~~~~~~~~{\rm
325: with}~~~~~~~~~~f(r)={\left[1+\left(\frac{r}{r_c}\right)^2\right]^{-\alpha/2}}~,\label{plummer1}
326: \end{equation}
327: where the cluster central density $\rho _0$ is given by
328: \begin{equation}
329: \rho _0 = \frac{M_{CL}}{\int _0^{R_{CL}} 4\pi r^2 f(r)~dr}~,
330: \label{plummer2}
331: \end{equation}
332: $R_{CL}$ and $M_{CL}$ being the cluster radius and mass,
333: respectively. According to dynamical observations towards the GC,
334: we require that the total mass $M(r)=M_{BH}+M_{CL}(r)$ contained
335: within $r\simeq 5\times 10^{-3}$ pc is $M\simeq 3.67\times 10^
336: 6~M_{\odot}$. Useful information is provided by the cluster mass
337: fraction, $\lambda_{CL}=M_{CL}/M$, and its complement,
338: $\lambda_{BH}=1-\lambda_{CL}$. As one can see, the requirement
339: given in eq. (\ref{plummer2}) implies that $M(r)\rightarrow
340: M_{BH}$ for $r\rightarrow 0$. The total mass density profile
341: $\rho(r)$ is given by
342: \begin{equation}
343: \rho(r) = \lambda_{BH} M \delta ^{(3)}(\overrightarrow{r}) +\rho_0
344: f(r)~ \label{totaldensity}
345: \end{equation}
346: and the mass contained within $r$ is
347: \begin{equation}
348: M(r) = \lambda_{BH}M + \int_0^r 4\pi r'^2\rho_0 f(r')~dr'~.
349: \end{equation}
350: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
351: % figure Density
352: \begin{figure}
353: \begin{center}
354: \includegraphics[scale=0.50]{f1.eps}\qquad
355: \end{center}
356: \caption{The cluster mass density profile is shown for different
357: values of $\lambda_{BH}$. Solid, dotted and dashed lines
358: correspond to $\lambda_{BH}= 0,~0.7,~0.9$, respectively. Thick,
359: red lines have been obtained for $r_c=3$ mpc with the same values
360: $\lambda_{BH}$ as given above.} \label{density}
361: \end{figure}
362: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
363: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
364: % figure Density
365: \begin{figure}
366: \begin{center}
367: \includegraphics[scale=0.50]{f2.eps}\qquad
368: \end{center}
369: \caption{The mass enclosed within the distance $r$ is shown for
370: different fractions $\lambda_{BH}$ of the total mass $M$ which is
371: contained in the central black hole. Solid, dotted and dashed
372: lines correspond to $\lambda_{BH}= 0$, $\lambda_{BH}= 0.7$ and
373: $\lambda_{BH}=0.9$, respectively.
374: %Thick (red lines) have been obtained for $r_c=3$ mpc while thin
375: %(black) lines are for $r_c=5.8$mpc.
376: Note that the case corresponding to $\lambda_{BH} = 0$ is not
377: realistic as shown by some observations (\cite{shen}).}
378: \label{ratiomass}
379: \end{figure}
380: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
381: In Figure \ref{density} we show the cluster mass density profile
382: $\rho_{CL}(r)$ as given by eq. (\ref{plummer1}), for selected
383: values of $\lambda_{BH}$. The total mass $M(r)$ enclosed within
384: the radius $r$ is also shown in Figure \ref{ratiomass}. In both
385: Figures, solid, dotted and dashed lines correspond to
386: $\lambda_{BH}= 0,~0.7,~0.9$, and we have assumed $r_c=3$ mpc
387: (thick lines) and $r_c=5.8$ mpc (thin lines).
388:
389: The Newtonian gravitational potential $\Phi_N$ at a distance $r$
390: due to the mass contained within it can be evaluated as
391: \begin{equation}
392: \Phi_N(r) = -G \int_r^{\infty} \frac{M(r')}{r'^2}~dr'~.
393: \label{gravitationalpotential}
394: \end{equation}
395: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
396: % figure Density
397: \begin{figure}
398: \begin{center}
399: \includegraphics[scale=0.50]{f3.eps}\qquad
400: \end{center}
401: \caption{The gravitational potential at distance $r$ as due to the
402: mass $M(r)$ is shown for different fractions $\lambda_{BH}$ of the
403: total mass $M$. Solid, dotted and dashed lines correspond to
404: $\lambda_{BH}= 0$, $\lambda_{BH}= 0.7$ and $\lambda_{BH}=0.9$,
405: respectively. Thick red lines have been obtained for $r_c=3$ mpc
406: while thin black lines are for $r_c=5.8$ mpc.} \label{potential}
407: \end{figure}
408: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
409: In Figure \ref{potential}, the gravitational potential $\Phi_N(r)$
410: due to the mass density distribution in eq. (\ref{totaldensity})
411: is given for selected values of $\lambda_{BH}$.
412:
413: According to GR, the motion of a test particle can be fully
414: described by solving the geodesic equations. Under the assumption
415: that the matter distribution is static and pressureless, the
416: equation of motion of the test particle becomes (see e.g. Weinberg
417: \citeyear{Weinberg72})
418: \begin{equation}
419: \frac{d\textbf{v}}{dt}\simeq-\nabla(\Phi_N +2\Phi_N
420: ^2)+4\textbf{v}(\textbf{v}\cdot \nabla)\Phi_N -v^2\nabla \Phi_N~.
421: \end{equation}
422: For a spherically symmetric mass distribution \footnote{We would
423: like to mention that the dynamical state of the region around Sgr
424: A$^*$ is known to be complex, with a significant population of
425: young stars of unclear origin making the assumption of an
426: undisturbed spherical cluster likely uncorrect. Considering the
427: effects caused by a non spherically symmetric mass distribution
428: makes the passage to an equation similar to eq. (\ref{setode}) not
429: analytically solvable. This problem will be addressed in a
430: subsequent work.} with a density profile given by eq.
431: (\ref{plummer1}) and for a gravitational potential given by eq.
432: (\ref{gravitationalpotential}), the previous relation becomes (see
433: for details Rubilar et al. \citeyear{Rubilar01})
434: \begin{equation}
435: \frac{d\textbf{v}}{dt}\simeq-\frac{GM(r)}{r^3}\left[\left(1+\frac{4\Phi_N}{c^2}+
436: \frac{v^2}{c^2}\right)\textbf{r}-\frac{4\textbf{v}(\textbf{v}\cdot\textbf{r})}{c^2}\right]~,
437: \label{setode}
438: \end{equation}
439: $\textbf{r}$ and $\textbf{v}$ being the radius vector of the test
440: particle (with respect to the center of the stellar cluster) and
441: the velocity vector, respectively. Once the initial conditions for
442: distance and velocity are given, the orbit of a test particle can
443: be found by solving the set of ordinary differential equations in
444: eq. (\ref{setode}) numerically.
445:
446: Now consider the S2 star, which is moving around the central
447: distribution of matter on an elliptic orbit of semi-major axis $d$
448: and eccentricity $e$ in the Newtonian approximation. We take a
449: frame with the origin in the GC, $X$-$Y$ plane on the orbital
450: plane and the $X$ axis pointing toward the periastron of the
451: orbit. Hence, we can choose the Newtonian initial conditions to be
452: (see e.g. Smart (\citeyear{smart}))
453: \begin{eqnarray}
454: r_x^0&=&d(1+e)~, \nonumber \\
455: r_y^0&=&0~,\label{cond1}
456: \end{eqnarray}
457: and
458: \begin{eqnarray}
459: v_x^0&=&0,\nonumber \\
460: v_y^0&=&\sqrt{GM(r_x^0)\left[\frac{2}{d(1+e)}-\frac{1}{d}\right]}~.
461: \label{cond2}
462: \end{eqnarray}
463:
464: For the S2 star, $d$ and $e$ given in the literature are 919 AU
465: and 0.87 respectively. They yield the orbits of the S2 star for
466: different values of the black hole mass fraction $\lambda _{BH}$
467: shown in Figure \ref{orbite}. The Plummer model parameters are
468: $\alpha =5$, core radius $r_c\simeq 5.8$ mpc. Note that in the
469: case of $\lambda _{BH}=1$, the expected (prograde) periastron
470: shift is that given by eq. (\ref{schshift}), while the presence of
471: the stellar cluster leads to a retrograde periastron shift. For
472: comparison, the expected periastron shift for the S16 star is
473: given in Figure \ref{orbite2}. In the latter case, the binary
474: system orbital parameters were taken from Sch\"odel et al.
475: (\citeyear{Schoedel03}) assuming also for the S16 mass a
476: conservative value of $\simeq 10$ M$_{\odot}$.
477:
478: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% ARRAY DI FIGURE
479: \begin{figure*}[htbp]
480: \vspace{0.2cm}
481: \begin{center}
482: $\begin{array}{c@{\hspace{0.1in}}c@{\hspace{0.1in}}c}
483: \epsfxsize=2.0in \epsfysize=2.0in \epsffile{f4.eps} &
484: \epsfxsize=2.0in \epsfysize=2.0in \epsffile{f5.eps} &
485: \epsfxsize=2.0in \epsfysize=2.0in \epsffile{f6.eps} \\
486: %\\ [0.cm]
487: %\mbox{\bf a)} & \mbox{\bf b)}
488: \end{array}$
489: \end{center}
490: \caption{Post Newtonian orbits for different values of the black
491: hole mass fraction $\lambda_{BH}$ are shown for the S2 star
492: (upper panels). Here, we have assumed that the Galactic central
493: black hole is surrounded by a stellar cluster whose density
494: profile follows a Plummer model with $\alpha=5$ and a core radius
495: $r_c\simeq 5.8$ mpc. The periastron shift values in each panel is
496: given in arcseconds.} \label{orbite}
497: \end{figure*}
498: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
499:
500: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% ARRAY DI FIGURE
501: \begin{figure*}[htbp]
502: \vspace{0.2cm}
503: \begin{center}
504: $\begin{array}{c@{\hspace{0.1in}}c@{\hspace{0.1in}}c}
505: \epsfxsize=2.0in \epsfysize=2.0in \epsffile{f7.eps} &
506: \epsfxsize=2.0in \epsfysize=2.0in \epsffile{f8.eps} &
507: \epsfxsize=2.0in \epsfysize=2.0in \epsffile{f9.eps} \\
508: %\\ [0.cm]
509: %\mbox{\bf a)} & \mbox{\bf b)}
510: \end{array}$
511: \end{center}
512: \caption{The same as in Figure \ref{orbite} but for the S16--Sgr
513: A$^*$ binary system. In this case, the binary system orbital
514: parameters were taken from Ghez et al. (\citeyear{Ghez05})
515: assuming for the S16 mass a conservative value of $\simeq 10$
516: M$_{\odot}$.} \label{orbite2}
517: \end{figure*}
518: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
519:
520: In Figure \ref{shiftvscore} the S2 orbital shift $\Delta \Phi$ is
521: given as a function of the stellar cluster core radius $r_c$, for
522: different power law index values ($\alpha=4$ dashed line,
523: $\alpha=5$ dotted line and $\alpha=6$ solid line). In the left
524: panel, the black hole mass fraction is $\lambda_{BH}=0.8$ in order
525: to compare with Rubilar et al. (\citeyear{Rubilar01}) results,
526: while the right panel shows the $\lambda_{BH}=0.99$ case. Note
527: that for extremely compact clusters, $\Delta \Phi$ is quite small.
528: The same is true for large enough core radii, corresponding to
529: matter density profiles almost constant within the S2 orbit.
530: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% ARRAY DI FIGURE
531: \begin{figure*}[htbp]
532: \vspace{0.2cm}
533: \begin{center}
534: $\begin{array}{c@{\hspace{0.1in}}c@{\hspace{0.2in}}c}
535: \epsfxsize=2.8in \epsfysize=2.8in \epsffile{f10.eps} &
536: \epsfxsize=2.8in \epsfysize=2.8in \epsffile{f11.eps} \\
537: %\\ [0.cm]
538: %\mbox{\bf a)} & \mbox{\bf b)}
539: \end{array}$
540: \end{center}
541: \caption{The expected S2 periastron shift as a function of the
542: stellar cluster core radius is shown. Here we have assumed a
543: Plummer density profile for the stellar cluster. Dashed, solid and
544: dotted lines correspond to $\alpha=4,~5$ and $6$, respectively.
545: The black hole mass fraction has been fixed to $\lambda_{BH}=0.8$
546: (left panel) and $\lambda_{BH}=0.99$ (right panel), respectively.
547: Note the existence of a maximum approximately corresponding to the
548: S2 semi-major axis.} \label{shiftvscore}
549: \end{figure*}
550: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
551: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
552:
553: Figures \ref{orbite} and \ref{shiftvscore} show that the expected
554: S2 periastron shift depends strongly on the total mass of the
555: cluster. In particular, the shift due to the cluster is opposite
556: in sign (retrograde motion) to the relativistic effect due to the
557: black hole in the GC. Moreover, for each value of the cluster mass
558: and power law index, there exist two density profiles
559: (corresponding to two particular core radii) which have total
560: shift almost zero, implying that the periastron advance due to the
561: cluster is equal (in magnitude) to the periastron shift due to the
562: black hole. A numerical analysis shows that the transition from a
563: prograde shift (due to the black hole) to retrograde shift (due to
564: the extended mass) occurs at $\lambda_{BH} \simeq 0.9976$,
565: $0.9986$ and $0.9990$ for $\alpha =4$, $5$ and $6$, respectively.
566: This means that a small fraction of mass in the cluster
567: drastically changes the overall shift.
568:
569: We would like to note that the assumption of the Plummer model to
570: describe the mass density distribution of the stellar cluster
571: around the central black hole is an oversimplification. Indeed,
572: one expects that in presence of a central black hole, the stellar
573: profile should follow a Bachall-Wolf law with density distribution
574: $\rho_c(r) \propto r^{-7/4}$ \citep{bw,binneytremaine} at least up
575: to $\tilde{r}_H\ll r_H$, where $r_H=GM_{BH}/\sigma_*^2\simeq 0.5$
576: pc is the radius of the black hole influence sphere. In the
577: following, we call $\tilde{r}_H$ the distance ($\ll r_H$) up to
578: which the cluster mass density follows the Bachall-Wolf law.
579:
580: In order to study the effect of such a cusp on the expected S2
581: periastron shift, we consider three different cases {\it a)} the
582: cusp is entirely contained within the S2 periastron distance
583: $R_{S2}$ (i.e. $\tilde{r}_H\le R_{S2}$), {\it b)} the cusp extends
584: beyond the S2 periastron distance (thus making the S2 star move in
585: a mass gradient) and {\it c)} the stellar density profile follows
586: a cusp law up to the distance $\tilde{r}_H$ from the center and a
587: Plummer law for $r\geq \tilde{r}_H$. In cases {\it a)} and {\it
588: b)} all stars are in a cusp density profile. In any case we
589: require that the total mass enclosed within $4.87\times 10^{-3}$
590: pc is $M\simeq 3.67\times 10^6 ~M_{\odot}$.
591:
592: In case {\it a)}, the total S2 periastron shift is just the sum of
593: the shift due to the black hole and the shift caused by the
594: stellar cusp (that contributes with the same sign). Hence, the S2
595: shift turns out to be $\Delta \Phi \simeq 0.17$ degree per
596: revolution.
597:
598: In case {\it b)}, by requiring that the total mass enclosed within
599: $4.87\times 10^{-3}$ pc is $M\simeq 3.67\times 10^6 ~M_{\odot}$,
600: we find that the dependence of the cusp mass and the induced S2
601: periastron shift on $\tilde{r}_H$ vanishes. Indeed, in Figure
602: \ref{cusp1}, we give the mass enclosed within the distance $r$ for
603: different values of $\lambda_{BH}$. Solid, dotted and dashed lines
604: correspond to $\lambda_{BH}= 0.1$, $\lambda_{BH}= 0.5$ and
605: $\lambda_{BH}=0.9$, respectively. Figure \ref{cusp2} shows the
606: expected S2 periastron shift as a function of $\lambda_{BH}$. As
607: noted before, the shift due to the cluster is opposite in sign
608: with respect to that due to the black hole. Moreover, for
609: $\lambda_{BH}\simeq 0.998$ the total shift turns out to be zero
610: since the contributions of the black hole and the cluster cancel
611: out. It is noticing that, since in the case of cusp profiles the
612: density gradient is larger than in the case of a usual
613: ($\alpha=5$) Plummer model, the value of the S2 periastron shift
614: gets generally larger values. Only if the Plummer core radius is
615: around $2$ mpc the resulting S2 periastron shifts are comparable
616: in both cases.
617: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
618: % figure Density
619: \begin{figure}[h]
620: \begin{center}
621: \includegraphics[scale=0.50]{f16.eps}\qquad
622: \end{center}
623: \caption{The mass enclosed within the distance $r$ is shown for
624: different fractions $\lambda_{BH}$ of the total mass $M$ contained
625: within the S2 orbit. Solid, dotted and dashed lines correspond to
626: $\lambda_{BH}= 0.1$, $\lambda_{BH}= 0.5$ and $\lambda_{BH}=0.9$,
627: respectively. The stellar cluster is assumed to follow an
628: $r^{-7/4}$ density profile.} \label{cusp1}
629: \end{figure}
630: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
631: % figure Density
632: \begin{figure}[h]
633: \begin{center}
634: \includegraphics[scale=0.50]{f17.eps}\qquad
635: \end{center}
636: \caption{The expected S2 periastron shift as a function of the
637: mass ratio parameter $\lambda_{BH}$. Solid and dashed lines
638: correspond to the S2 shift due to the black hole and to the
639: stellar cusp, respectively. We note that the shift due to the
640: stellar cusp is independent on the $\tilde{r}_H$ value, that, in
641: this case, has been assumed to be larger than the S2 semi-major
642: axis (case {\it b}).} \label{cusp2}
643: \end{figure}
644: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
645: We have then considered the superposition of a Plummer model and a
646: Bahcall-Wolf profile (case {\it c}) extended up to $\tilde{r}_H$
647: such as the cusp density at $\tilde{r}_H$ equals that of the
648: Plummer model at the same distance. Here, if $\tilde{r}_H\ll
649: R_{S2}$ the S2 periastron shift will be practically equal to that
650: caused by the Plummer model (see right panel in Fig. 6) since in
651: this case the cusp will have a minor influence. On the contrary,
652: for an extended cusp ($\tilde{r}_H\gg R_{S2}$), the cusp effect on
653: the S2 periastron shift will dominate reconciling with case $b$.
654:
655: As a last point, we mention that we have also considered the
656: effect due to an extrapolation of the observed stellar density
657: profile - the innermost point of which is the S2 star at a
658: distance of 0.1\arcsec - within $R_{S2}$. Following
659: \citet{Genzel03b} and assuming a cusp stellar density profile, we
660: find that the enclosed mass is in the range 30-300 $M_{\odot}$
661: (for a constant mass density or a power law with index
662: $\gamma=1.4$). Therefore, the cusp effect on the S2 periastron
663: shift is negligible since the corresponding $\lambda_{BH}$ is
664: always greater than 0.99992. However, we caution that the case
665: under investigation in the present paper is different with respect
666: to \citet{Genzel03b} since we are assuming that a fraction of the
667: mass contained within $R_{S2}$ may be in a stellar cluster. Hence,
668: the cluster mass content may be larger, thus providing a stronger
669: effect on the S2 periastron shift.
670:
671:
672:
673: \section{Tightening mass limits of Sgr~A$^*$}
674:
675: We know that the mass of Sgr~A$^*$ within the S2 orbit is
676: $3.67\times 10^6$ M$_{\odot}$ to a high accuracy. Though there is
677: nothing definite known about the mass distribution, there is
678: strong reason to believe that there is a black hole of several
679: solar masses, possibly surrounded by a significant cluster. In
680: principle the cluster mass could dominate over the black hole, be
681: comparable to it or be dominated by it. That there {\it is} a
682: cluster is highly likely on account of the large number of stars
683: observed near Sgr~A$^*$. Though these lie outside the S2 orbit,
684: many stars so far unseen probably do lie within the orbit as well.
685: In this section we use current data on the Brownian motion of
686: Sgr~A$^*$ and the evaporation time for the putative cluster to put
687: limits on the cluster mass and hence on the black hole mass.
688:
689: Chatterjee, Hernquist and Loeb (\citeyear{chatterje}) have
690: developed a simple model to describe the dynamics of a massive
691: black hole surrounded by a dense stellar cluster. The total force
692: acting on the black hole is separated into two independent parts,
693: one of which is the slowly varying force due to the stellar
694: ensemble and the other the rapid stochastic force due to close
695: stellar encounters. In the case of a stellar system with a Plummer
696: distribution, the motion of the black hole is similar to that of a
697: Brownian particle in a harmonic potential. Thus the black hole
698: one-dimensional mean-square velocity is given by
699: \begin{equation}
700: <v_x^2>=\frac{2}{9}\frac{GM_{CL}m_*}{r_cM_{BH}}~, \label{eqvel1}
701: \end{equation}
702: where it has been assumed that the cluster is composed of objects
703: with equal mass, $m_{*}$. For a Plummer ($\alpha$ = 5) stellar
704: cluster, the total mass within $R$ is
705: \begin{equation}
706: M(R)=M_{BH}+\frac{M_{CL}R^3}{(R^2+r_c^2)^{3/2}}~. \label{eqmass1}
707: \end{equation}
708:
709: Since $<v_x^2>$ is less than a certain maximum value
710: $<v_x^2>_{max}$, from eqs. (\ref{eqvel1}) and (\ref{eqmass1}) one
711: obtains
712: \begin{equation}
713: M_{BH} > M(R) \left\{1+\frac{9}{2} \left[ <v_x^2>_{max} \frac{r_c
714: R^3}{G (R^2+r_c^2)^{3/2}m_* } \right]\right\}^{-1}~,\label{eqlim1}
715: \end{equation}
716: the right hand side corresponding to a minimum black hole mass, as
717: constrained by the Brownian motion of the central black hole. In
718: Figure \ref{figlimiti1} the minimum black hole mass allowed by the
719: Brownian motion of Sgr~A$^*$ is given as a function of the stellar
720: cluster core radius, for two different proper motion velocities of
721: the black hole: 1.3 km s$^{-1}$ (dashed lines), and 2 km s$^{-1}$
722: (dotted lines). The total mass contained within $R=0.01$ pc of
723: Sgr~A$^*$ has been taken to be $M\simeq 3.67\times 10^6$
724: M$_{\odot}$.
725: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
726: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
727: \begin{figure}[h]
728: \begin{center}
729: \includegraphics[scale=0.55]{f12.eps}\qquad
730: \end{center}
731: \caption{The minimum black hole mass allowed by the Brownian
732: motion of Sgr~A$^*$ is given as a function of the stellar cluster
733: core radius for the different black hole proper motion velocities.
734: We assume that a total mass $M\simeq 3.67\times 10^6$ M$_{\odot}$
735: is contained within $R_{S2}=4.87$ mpc. Dashed and dotted lines
736: have been obtained for velocities of $1.3$ km ${\rm s}^{-1}$ and
737: $2$ km $\rm{s}^{-1}$, respectively. For each given curve only the
738: region above it is allowed.}\label{figlimiti1}
739: \end{figure}
740:
741: Chatterjee, Hernquist and Loeb (\citeyear{chatterje}) derived an
742: evaporation time for a cluster, but concentrated on the large
743: scale cluster $r_c\simeq$ 10 pc about Sgr~A$^*$, and hence assumed
744: that M$_{CL}$ $\gg$ M$_{BH}$. On the other hand Rauch and Tremaine
745: (\citeyear{rauch}) and Mouawad et al. (\citeyear{moawad}) consider
746: only the region interior to the orbit of S2 and assume M$_{CL}$
747: $\ll$ M$_{BH}$.
748:
749: We need to allow for all possibilities while considering the
750: cluster interior to the orbit of S2, including M$_{CL}\simeq$
751: M$_{BH}$. For this purpose we consider a cluster of core radius
752: $r_c$ and mass M$_{CL}$ = M $-$ M$_{BH}$. We now need to obtain
753: the generalization of the formula of Chatterjee, Hernquist and
754: Loeb (\citeyear{chatterje}) for the median relaxation time in this
755: more general situation. For this purpose, as usual, we assume that
756: the cluster consists of components of the same mass $m_*$ and
757: evaluate the crossing time in the usual way to obtain the general
758: median relaxation time
759: \begin{equation}
760: T_{r}=\frac{0.14(1.3~r_cM)^{3/2}}{\sqrt{G}M_{CL}m_*log(0.4M/m_*)}.
761: \end{equation}
762: It is easy to verify that in the approximation M$_{CL}$ $\gg$
763: M$_{BH}$ we recover the formula of Chatterjee, Hernquist and Loeb
764: (\citeyear{chatterje}) and in the approximation M$_{CL}$ $\ll$
765: M$_{BH}$ we recover the formula of Rauch and Tremaine
766: (\citeyear{rauch}).The evaporation time is then $T_{evap}\simeq
767: 300~T_{r}$ (\citealt{binneytremaine}, p.525).
768:
769: One can assume different ``reasonable" values of the time that the
770: cluster would have been in existence and hence use the evaporation
771: time to further limit the black hole mass in the GC. It is clear
772: that 10$^8$ years = 0.1 Gyr is less than the minimum value that
773: could be regarded as reasonable, 1 Gyr is more reasonable and 10
774: Gyr is likely to be a good value to take. The results are given in
775: Table \ref{tab3} for $m_*$ = 1 M$_{\odot}$. Note that the tightest
776: bound gives a very stringent upper limit of
777: $9\times10^4$M$_{\odot}$ on the cluster mass. Also note that the
778: value decreases if the average $m_*$ is taken to be larger.
779: \begin{table*}
780: \begin{center}
781: \begin{tabular}{|c|c|c|c|c|}
782: \hline \rule{0pt}{3ex} $\rm{ T_{evap} (Gyr)}$ & ${\rm r_c^{1.3}
783: (mpc)}$ & ${\rm \lambda_{BH}^{1.3}}$ & ${\rm r_c^{2.0}
784: (mpc)}$ & ${\rm \lambda_{BH}^{2.0}}$ \\
785: \hline
786: \rule{0pt}{3ex}$0.1$ & $0.87$ & $0.762$ & $1.28$ & $0.645$ \\
787: \hline
788: \rule{0pt}{3ex}$1$ & $2.11$ & $0.919$ & $2.61$ & $0.876$ \\
789: \hline
790: \rule{0pt}{3ex}$10$ & $4.12$ & $0.975$ & $5.27$ & $0.964$ \\
791: \hline
792: \end{tabular}
793: \end{center}
794: \caption{The cluster core radius $r_c$ and minimum black hole mass
795: fraction $\lambda_{BH}$ for the limits obtained by
796: $<v>_{max}^2=1.3$ and 2.0, for $T_{evap}=0.1$, 1 and 10 Gyr.}
797: \label{tab3}
798: \end{table*}
799:
800: \section{The spin of the black hole}
801:
802: The periastron shift is the net contribution of the relativistic
803: retrograde shift due to the black hole and the Newtonian prograde
804: shift due to the surrounding cluster. Obviously, if the periastron
805: advance due to the stellar cluster were known, the contribution of
806: periastron advance due to the black hole could be obtained by
807: subtracting from the measured quantity. The question arises
808: whether the information obtained would be adequate to obtain both
809: the black hole mass and spin parameters. Though we can put
810: reasonably sharp bounds on the stellar cluster about the black
811: hole, is it good good enough for our purpose? If so, we could use
812: eq.(\ref{kershift}) to obtain the spin of the black hole for
813: different values in the possible range for the periastron shift.
814: It is easy to see from Fig.\ref{shiftvscore} that for
815: $\lambda_{BH} = 0.99$ and allowing for the maximum range of
816: unknown values of $\alpha$ and $r_{c}$ the $1.8\times10^3<-\Delta
817: \phi<4.7\times10^3$ or $1.9\times10^{-3}<-\Delta
818: \phi_E<4.7\times10^{-3}$. For the sharpest limit obtained,
819: $\alpha=5$, and Brownian motion 1.3 km s$^{-1}$, $\lambda_{BH} =
820: 0.975$, we find that $\Delta \phi \simeq -4.47\times10^3$ or
821: $\Delta \phi_E \simeq -4.7\times10^{-3}$. For Brownian motion 2.0
822: km s$^{-1}$, $\lambda_{BH} = 0.964$, $\Delta \phi \simeq
823: -5.75\times10^3$ or $\Delta \phi_E \simeq -6.0\times10^{-3}$. This
824: is a factor of 5 {\it less} than the effect of the spin. Hence
825: this method {\it cannot} be used to determine the spin. For this
826: we need the cluster parameter values, rather than upper limits for
827: them. Alternatively, one would need to rely on the retrolensing
828: method suggested earlier (\citealt{depaolis4,depaolis5}).
829:
830:
831: \section{Determination of cluster parameters}
832:
833: Using the stronger (1$\sigma$) limit of 1.3 km $\rm{s}^{-1}$ and
834: the weaker (2$\sigma$) limit of 2.0 km $\rm{s}^{-1}$ to limit the
835: Brownian motion of Sgr A$^*$ for our calculations and evaporation
836: times of 1 and 10 Gyr for the cluster, we obtained the minimum
837: black hole mass. For the stronger limits on the Brownian motion
838: and the evaporation time, it is $3.579\times 10^6$ M$_{\odot}$
839: corresponding to a $\lambda_{BH} \simeq 0.975$ for $\alpha =5$.
840: Our numerical analysis shows that the transition from a prograde
841: shift (due to the black hole) to a retrograde shift (due to the
842: extended mass assumed to be distributed with a Plummer density
843: profile) occurs at $\lambda_{BH} \simeq 0.9976$, $0.9986$ and
844: $0.9990$ for $\alpha =4$, $5$ and $6$, respectively. Hence, even a
845: small cluster around the central massive black hole limits the
846: possibility to observe and use the periastron shift of the S2
847: star.
848:
849: Since we have modeled the star cluster density profile by a
850: Plummer model, the periastron shift contribution due to the
851: stellar cluster depends on three parameters: the central density
852: $\rho _0$ (or equivalently $\lambda _{BH}$); the core radius
853: $r_c$; and the power-law index $\alpha$. This degeneracy in the
854: determination of the stellar cluster parameters is due to the
855: measurement of the periastron shift of a single star. This is
856: easily seen by inspecting Figure \ref{fig_par0}, which has been
857: obtained for illustrative purposes for the S2 star by setting
858: $\lambda _{BH}=0.99$ and varying both the core radius $r_c$ and
859: power-law index $\alpha$ for the star cluster density profile.
860: Each contour line corresponds to a given S2 periastron shift in
861: units of degrees. To solve the parameter degeneracy and determine
862: the stellar cluster parameters (by studying the periastron advance
863: effect), the periastron shifts for at least three different stars
864: have to be measured with sufficient accuracy. Consider, for
865: example, the S16 star having an orbital period of $\simeq 36$ yr
866: and eccentricity $e\simeq 0.97$. Measuring its periastron shift
867: and comparing with the S2 result will give much tighter
868: information about the stellar cluster parameters. From Figure
869: \ref{fig_par1} it is evident that there are regions (intersections
870: between dashed and solid lines) in the $\alpha$-$r_c$ plane for
871: which one measures values of the periastron shift for the S2 and
872: S16 stars. Obviously, there could be (as yet unobserved) stars
873: with orbit apocenters comparable to S2, but with different
874: eccentricities (for example larger than 0.87) or stars closer to
875: the GC black hole than S2 or S16 stars. Monitoring their orbits
876: and measuring their periastron shifts will be extremely helpful in
877: reconstructing the cluster density profile. As an example, in
878: Figure \ref{fig_par2} we compare the expected S2 periastron shift
879: (solid lines obtained for $\lambda _{BH}=0.99$) with the
880: periastron shift of a star whose orbit has an eccentricity of
881: $\simeq 0.87$ and semi-major axis 3 times smaller than that of S2.
882:
883: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
884: \begin{figure}[h]
885: \begin{center}
886: \includegraphics[scale=0.55]{f13.eps}\qquad
887: \end{center}
888: \caption{The expected S2 periastron shift for $\lambda _{BH}=0.99$
889: for different values of both the core radius $r_c$ and power-law
890: index $\alpha$ for the considered Plummer density profile. Each
891: contour line corresponds to a given S2 periastron shift in degree
892: units. Note that a degeneracy occurs since there exist different
893: values of the power law index and core radius corresponding to the
894: same periastron shift.} \label{fig_par0}
895: \end{figure}
896: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
897:
898: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
899: \begin{figure}[h]
900: \begin{center}
901: \includegraphics[scale=0.55]{f14.eps}\qquad
902: \end{center}
903: \caption{The same as in Figure 8. Dotted lines show contours for
904: the S16 star. If the periastron shifts of both stars will be
905: measured in the future the intersection between the corresponding
906: contour lines will give information about the central stellar
907: density profile.} \label{fig_par1}
908: \end{figure}
909: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
910:
911: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
912: \begin{figure}[h]
913: \begin{center}
914: \includegraphics[scale=0.55]{f15.eps}\qquad
915: \end{center}
916: \caption{The same as in Figure 8. Dotted lines show contours for a
917: star with orbit of $\simeq 0.87$ (the same as the S2 star) but
918: semi-major axis 3 times smaller with respect to the S2 one. If the
919: periastron shifts of both stars will be observed in the future the
920: intersection between the corresponding contour lines gives
921: information about the central stellar density profile.}
922: \label{fig_par2}
923: \end{figure}
924: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
925:
926: As is evident from Figures 8 - 10, one can obtain estimates of the
927: $r_c$, $\alpha$ and $\lambda _{BH}$, provided that three stars
928: have been observed to sufficient accuracy. Assume that we have
929: adequate accuracy of observation to see periastron shifts of
930: 10$^{-2.5}$ mas, which is the value required to see the
931: relativistic periastron shift. To what accuracy have we limited
932: the cluster parameters? To determine this, we could just vary
933: $\lambda_{BH}$ for a given $r_c$. The effect of this change would
934: be less than the effect of changing $r_c$ {\it and}
935: $\lambda_{BH}$. As such, if we want to know how accurately the
936: cluster parameters are determined, we need to calculate the
937: maximum change in $r_c$ {\it along with} the change in
938: $\lambda_{BH}$, as allowed by the Brownian motion limit. By
939: varying $\lambda_{BH}$ by 10$^{-2}$ and $r_c$ maximally we find
940: that we get the required accuracy, for evaporation rates from 1 to
941: 10 Gyr and Brownian motion of 1.3 to 2.0 km s$^{-1}$. With this
942: accuracy we should also be able to separate out the classical
943: periastron shift of the stellar cluster and the relativistic
944: effect of a maximal (and even slightly less than maximal) Kerr
945: black hole. With better accuracy we should be able to get an
946: estimate, or at least an upper bound, for the black hole spin as
947: well. The question now is, what is required to achieve this
948: accuracy of observation of the periastron shift of three stars?
949: This point is discussed in the next section.
950:
951: \section{Observational requirements for determination of black
952: hole spin and cluster parameters}
953:
954: In the near future, observations using large diameter telescopes
955: in combination with adaptive optics may allow us to reach the
956: angular resolution needed to measure the periastron shift of stars
957: close to the GC back hole. Consider for example an instrument with
958: an angular resolution $\Delta \Phi _A$ and assume that the
959: relative position of stars can be determined to about $1/\epsilon$
960: of the achieved angular resolution, i.e. the position accuracy is
961: $\Delta \Phi _P\simeq \Delta \Phi _A/\epsilon$. The positional
962: accuracy can be increased by a factor $\sqrt{N}$, if $N$ reference
963: stars are used. In this case, the maximum positional accuracy is
964: simply given by (\citealt{Rubilar01})
965: \begin{equation}
966: \Delta \phi _{P} = \frac{\Delta \Phi _P}{\sqrt{N}}~.
967: \end{equation}
968: It follows that if the periastron position of a star shifts by an
969: amount $\Delta \Phi _E$ (as observed from Earth), to obtain the
970: desired accuracy we need at least that $\Delta \phi _{P} \simeq
971: \Delta \Phi _E$. In this case, the minimum number of reference
972: stars can be determined once both the instrument angular and
973: positional accuracies are known, i.e.
974: \begin{equation}
975: N_{max}=\left(\frac{\Delta \Phi _P}{\Delta \Phi _E}\right)^2~.
976: \end{equation}
977:
978: As an example, the LBT interferometer has angular resolution
979: $\Delta \Phi _A\simeq 30$ mas and the relative position of stars
980: is conservatively estimated to be about 1/30 of that value
981: (\citealt{Rubilar01}). Therefore, to measure the periastron shift
982: with adequate accuracy to see the least shift for the cluster
983: parameters allowed, and thereby detect the relativistic shift, we
984: need $N$ to be $\simeq (6.6\times10^{-1}/2\times10^{-3})^2$ or
985: 10$^5$ reference stars, which automatically provides the accuracy
986: required to see the maximal Kerr (spin) effect. For PRIMA (see
987: http://www.eso.org/projects/vlti/instru/prima/index\_prima.html)
988: the relative positional accuracy is planned to be $\simeq
989: 10~\mu$as. As such, we only need a single reference star.
990:
991:
992: %An even better relative stellar position accuracy of $\simeq
993: %1~\mu$as is intended for the SIM mission (see
994: %http://planetquest.jpl.nasa.gov/SIM). In this case we may even be
995: %able to put bounds on the black hole spin. In fact for a single
996: %reference star $a/M_{BH}\geq$ 0.02 could (in principle) be seen!
997:
998:
999: \section{Concluding remarks}
1000:
1001: We have used the fact that the stellar cluster close to the
1002: central black hole seems spherically symmetric to limit the
1003: Brownian motion of Sgr A$^*$ to be the observed proper motion. We
1004: have taken the stronger (1$\sigma$) limit of 1.3 km $\rm{s}^{-1}$
1005: and the weaker (2$\sigma$) limit of 2.0 km $\rm{s}^{-1}$ for our
1006: calculations. We also used evaporation times of 1 to 10 Gyr for
1007: the cluster, appropriately modified to incorporate the
1008: gravitational well due to the black hole, to put further
1009: constraints on the cluster mass. The results of our calculations
1010: show that the stellar periastron shifts due to the cluster, even
1011: limited to the extent considered, may totally swamp not only the
1012: Kerr (spin) effect but also the Schwarzschild effect. However, the
1013: discussion focused on the observations for a single star, S2. By
1014: modelling the star cluster density profile with a Plummer low,
1015: the periastron shift contribution due to the stellar cluster
1016: depends on three parameters: the central density $\rho _0$ (or
1017: equivalently $\lambda _{BH}$), the core radius $r_c$, and the
1018: power-law index $\alpha$. Consequently, with observations of three
1019: stars we should be able to determine the cluster parameters
1020: adequately. \footnote{Note that, wether we would have known that
1021: the star cluster follows a Bahcall-Wolf profile, by measuring the
1022: periastron advance of only one star we may be able to calculate
1023: the only parameter: $\lambda_{BH}$ (from Fig. 8).} We have
1024: addressed the question of what is required to obtain the desired
1025: accuracy for observing the relativistic effect. It turns out that
1026: we need about 10$^5$ reference stars with the LBT interferometer.
1027: With the accuracy expected of PRIMA, it should be enough to use
1028: only one reference star.
1029: %With SIM we may even be able to put limits on the black hole
1030: %spin to an accuracy of about 2$\%$ of the black hole mass with a
1031: %{\it single} reference star! For 100 reference stars we could
1032: %increase the accuracy to 0.2$\%$.
1033:
1034:
1035: \acknowledgements
1036:
1037: This work has been partially supported by MIUR (Programmi di
1038: Ricerca Scientifica di Rilevante Interesse Nazionale (PRIN04) -
1039: prot. 2004020323$\_$004). We would like to thank an unknown
1040: referee for very useful suggestions and comments that have
1041: improved our paper. Two of us (AQ and AFZ) would like to thank the
1042: Department of Physics of University of Lecce and {\it INFN}
1043: (Italy) where this work has been initiated. FDP and AAN would like
1044: to thank the $30^{th}$ International Nathiagali Summer College
1045: (Pakistan), where the original version of this work has been
1046: completed. AFZ is also grateful to the National Natural Science
1047: Foundation of China (NNSFC) (Grant \# 10233050) and National Key
1048: Basic Research Foundation (Grant \# TG 2000078404) for a partial
1049: financial support of the work.
1050:
1051:
1052: \begin{thebibliography}{}
1053: %\bibitem[Auri\`ere(1982)]{aur82} Auri\`ere, M. 1982, \aap,
1054: % 109, 301
1055:
1056: \bibitem[\protect\citeauthoryear{Bahcall \& Wolf}{1977}]{bw}
1057: Bachall J.N. and Wolf R.A. 1977, \apj, 216, 883
1058:
1059: \bibitem[\protect\citeauthoryear{Bini et al.}{2005}]{bini2005}
1060: Bini D. et al. 2005, Gen. Rel. Grav., 37, 1263
1061:
1062: \bibitem[\protect\citeauthoryear{Binney \& Tremaine}{1987}]{binneytremaine}
1063: Binney J. and Tremaine S., {\it Galactic Dynamics}, Princeton
1064: University Press, Princeton, New Jersey, 1987
1065:
1066: \bibitem[\protect\citeauthoryear{Boyer \& Price}{1965}]{boyerprice}
1067: Boyer R.H. and Price T.G. 1965, Proc. Camb. Phil. Soc. 61, 531
1068:
1069: \bibitem[\protect\citeauthoryear{Chatterjee et al.}{2002}]{chatterje}
1070: Chatterjee P., Hernquist L. and Loeb A. 2002, \apj, 572, 371
1071:
1072: \bibitem[\protect\citeauthoryear{Delplancke et al.}{2003}]{Delplancke03}
1073: Delplancke F. et al. 2003, \aap, 286, 99
1074:
1075: \bibitem[\protect\citeauthoryear{De Paolis et al.}{2003}]{depaolis1}
1076: De Paolis F. et al. 2003, \aap, 409, 809
1077:
1078: \bibitem[\protect\citeauthoryear{De Paolis et al.}{2004}]{depaolis2}
1079: De Paolis F. et al. 2004, \aap, 415, 1
1080:
1081: \bibitem[\protect\citeauthoryear{De Paolis et al.}{2005}]{depaolis4}
1082: De Paolis F. et al. 2005, in {\it Proc. Eleventh Regional Conf. on
1083: Math. Phys.} eds. Rahvar S, Sadooghi N, and Shojai F, World
1084: Scientific
1085:
1086: %\bibitem[\protect\citeauthoryear{De Paolis et al.}{2006}]{depaolis5}
1087: %De Paolis F. et al., 2006 ``Spectral shift in retro-lensing by
1088: %spinning black holes", submitted to {\it Class. and Quantum Grav.}
1089:
1090: \bibitem[\protect\citeauthoryear{Fabian}{2005}]{Fabian04}
1091: Fabian A.C. 2005, Ap\&SS, 300, 97
1092:
1093: \bibitem[\protect\citeauthoryear{Fabian et al.}{2000}]{Fabi00}
1094: Fabian A.C. et al. 2000, \pasp, 112, 1145
1095:
1096: \bibitem[\protect\citeauthoryear{Fabian et al.}{1995}]{fabian1}
1097: Fabian A.C. et al. 1995, \mnras, 277, L11
1098:
1099: \bibitem[\protect\citeauthoryear{Fragile \& Mathews}{2000}]{Fragile00}
1100: Fragile P.C. \& Mathews G.J. 2000, \apj, 542, 328
1101:
1102: \bibitem[\protect\citeauthoryear{Genzel et al.}{2003 a}]{Genzel03}
1103: Genzel R. et al. 2003 a, Nature, 425, 934
1104:
1105: \bibitem[\protect\citeauthoryear{Genzel et al.}{2003 b}]{Genzel03b}
1106: Genzel R. et al. 2003 b, ApJ, 594, 812
1107:
1108: \bibitem[\protect\citeauthoryear{Ghez et al.}{2003}]{Ghez03}
1109: Ghez A.M. et al. 2003, \apjl, 586, L127
1110:
1111: \bibitem[\protect\citeauthoryear{Ghez et al.}{2004}]{Ghez04}
1112: Ghez A.M., et al. 2004, \apjl, 601, L159
1113:
1114: \bibitem[\protect\citeauthoryear{Ghez et al.}{2005}]{Ghez05}
1115: Ghez A.M., et al. 2005, \apj, 620, 744
1116:
1117: \bibitem[\protect\citeauthoryear{Jaroszynski}{1998}]{Jaroszynski98}
1118: Jaroszynski M. 1998, Acta Astron., 48, 653
1119:
1120: \bibitem[\protect\citeauthoryear{Jaroszynski}{1999}]{Jaroszynski99}
1121: Jaroszynski M. 1999, \apj, 521, 591
1122:
1123: \bibitem[\protect\citeauthoryear{Jaroszynski}{2000}]{Jaroszynski00}
1124: Jaroszynski M. 2000, Acta Astron., 50, 67
1125:
1126: \bibitem[\protect\citeauthoryear{Mouawad et al.}{2005}]{moawad}
1127: Mouawad N. et al. 2005, Astron. Nachr., 326, 83
1128:
1129: \bibitem[\protect\citeauthoryear{Quirrenbach}{2003}]{Quirrenbach03}
1130: Quirrenbach A. 2003, Ap\& SS, 286, 277
1131:
1132: \bibitem[\protect\citeauthoryear{rauch}{1996}]{rauch}
1133: Rauch K.P. \& Tremaine S. 1996, NewA 1, 149
1134:
1135: \bibitem[\protect\citeauthoryear{Reid and Bruthaler}{2004}]{reid2004}
1136: Reid M.J. and Brunthaler A. 2004, \apj, 616, 872
1137:
1138: \bibitem[\protect\citeauthoryear{Reid et al.}{1999}]{reid1999}
1139: Reid M.J. et al. 1999, \apj, 524, 816
1140:
1141: \bibitem[\protect\citeauthoryear{R\"{o}ttgering et al.}{2003}]{Rottgering03}
1142: R\"{o}ttgering H.J.A. et al. 2003, astro-ph/0308538
1143:
1144: \bibitem[\protect\citeauthoryear{Rubilar \& Eckart}{2001}]{Rubilar01}
1145: Rubilar G.F. \& Eckart A. 2001, \aap, 374, 95
1146:
1147: \bibitem[\protect\citeauthoryear{Sch\"odel et
1148: al.}{2003}]{Schoedel03} Sch\"odel R. et al. 2003, \apj, 596, 1015
1149:
1150: \bibitem[\protect\citeauthoryear{Shen et
1151: al.}{2005}]{shen} Shen Z.-Q. et al. 2005, Nature, 438, 62
1152:
1153:
1154: \bibitem[\protect\citeauthoryear{Smart}{1977}]{smart}
1155: Smart W.M. 1977, {\it Textbook on Spherical Astronomy}, Cambridge
1156: University Press
1157:
1158: \bibitem[\protect\citeauthoryear{Tanaka et al.}{1995}]{tanaka1}
1159: Tanaka Y. et al. 1995, Nature, 375, 659
1160:
1161: \bibitem[\protect\citeauthoryear{Weinberg, Miloslavljevi\'{c} \& Ghez}{2005}]{Weinberg05}
1162: Weinberg N.N., Miloslavljevi\'{c} M.\& Ghez A.M. 2005, \apj, 622,
1163: 878
1164:
1165: \bibitem[\protect\citeauthoryear{Weinberg}{1972}]{Weinberg72}
1166: Weinberg S. 1972, Gravitation and Cosmology: Principles and
1167: Applications of the General Theory of Relativity, Wiley, New York
1168:
1169: \bibitem[\protect\citeauthoryear{Zakharov et al.}{2003a}]{ZKLR02}
1170: Zakharov A.F.et al. 2003a, \mnras, 342, 1325
1171:
1172: \bibitem[\protect\citeauthoryear{Zakharov et al.}{2005a}]{ZNDI_04}
1173: Zakharov A.F.et al. 2005a, New Astronomy, 10, 479
1174:
1175: \bibitem[\protect\citeauthoryear{Zakharov et al.}{2005b}]{depaolis3}
1176: Zakharov A.F. et al. 2005b, \aap, 442, 795
1177:
1178: \bibitem[\protect\citeauthoryear{Zakharov \& Repin }{2003b}]{zak_rep03_aa}
1179: Zakharov A.F. \& Repin S.V. 2003b, \aap, 406, 7
1180:
1181: \bibitem[\protect\citeauthoryear{Zakharov \& Repin }{2003c}]{Zak_rep03_AR}
1182: Zakharov A.F. \& Repin S.V. 2003c, Astron. Rep., 47, 733
1183:
1184: \bibitem[\protect\citeauthoryear{Zakharov \& Repin }{2004}]{ZR_ASR04}
1185: Zakharov A.F. \& Repin S.V. 2004, Adv. Space Res., 34, 1837
1186:
1187: \end{thebibliography}
1188:
1189: %\clearpage
1190:
1191:
1192: \end{document}
1193:
1194: %%
1195: %% End of file `sample.tex'.
1196: