------------------------------------------------------------------------ From: Andrea Ghez ghez@astro.ucla.edu To: gcnews@aoc.nrao.edu Subject: gcnews contribution %arXiv:0808.2870 \documentclass[12pt,preprint]{aastex} %\usepackage{emulateapj5} %\usepackage{lscape} \newcommand{\brg}{Br${\gamma}$ } \begin{document} \title{Measuring Distance and Properties of the Milky Way's Central Supermassive Black Hole with Stellar Orbits} %Measuring the Distance to the Galactic Center and the Extended Mass Distribution Through Orbital Motion} \author{ A. M. Ghez\altaffilmark{1,2}, S. Salim\altaffilmark{1,4}, N. N. Weinberg\altaffilmark{3,5}, J. R. Lu\altaffilmark{1}, J. K. Dunn\altaffilmark{1}, M. Morris\altaffilmark{1}, E. E. Becklin\altaffilmark{1}, M. Milosavljevic\altaffilmark{6}, %B. T. Soifer\altaffilmark{3}, %D. Thompson\altaffilmark{3} } \altaffiltext{1}{UCLA Department of Physics and Astronomy, Los Angeles, CA 90095 -1547; ghez, jlu, tdo, jkdunn, morris, syelda, becklin@astro.ucla.edu} \altaffiltext{2}{UCLA Institute of Geophysics and Planetary Physics, Los Angeles, CA 90095-1565} \altaffiltext{3}{California Institute of Technology, Division of Mathematics, Physics and Astronomy, Pasadena, CA 91125; kym@caltech.edu} \altaffiltext{4}{NOAO, 950 N Cherry Ave, Tucson, AZ 85719, samir@noao.edu} \altaffiltext{5}{University of California Berkeley, Department of Astronomy\\ Berkeley, CA 94720-3411 nnw@astron.berkeley.edu} \altaffiltext{6}{University of Texas, Department of Astronomy, Austin, TX 78712 milos@astro.as.utexas.edu} \altaffiltext{7}{UCSC, Department of Astronomy \& Astrophysics, Santa Cruz, CA 95064, jnaiman@astro.ucsc.edu} \begin{abstract} We report new precision measurements of the properties of our Galaxy's supermassive black hole. Based on astrometric (1995-2007) and radial velocity (2000-2007) measurements from the W. M. Keck 10-meter telescopes, a fully unconstrained Keplerian orbit for the short period star S0-2 provides values for the distance (R$_0$) of 8.0 $\pm$ 0.6 kpc, the enclosed mass (M$_{bh}$) of 4.1 $\pm$ 0.6 $\times$ $10^6 M_{\odot}$, and the black hole's radial velocity, which is consistent with zero with 30 km/s uncertainty. If the black hole is assumed to be at rest with respect to the Galaxy (e.g., has no massive companion to induce motion), we can further constrain the fit and obtain R$_0$ = 8.4 $\pm$ 0.4 kpc and M$_{bh}$ = 4.5 $\pm$ 0.4 $\times$ $10^6 M_{\odot}$. More complex models constrain the extended dark mass distribution to be less than 3-4 $\times$ $10^5 M_{\odot}$ within 0.01 pc, $\sim$100x higher than predictions from stellar and stellar remnant models. For all models, we identify transient astrometric shifts from source confusion (up to 5x the astrometric error) and the assumptions regarding the black hole's radial motion as previously unrecognized limitations on orbital accuracy and the usefulness of fainter stars. Future astrometric and RV observations will remedy these effects. Our estimates of R$_0$ and the Galaxy's local rotation speed, which it is derived from combining R$_0$ with the apparent proper motion of Sgr A*, ($\theta_0$ = 229 $\pm$ 18 km s$^{-1}$), are compatible with measurements made using other methods. The increased black hole mass found in this study, compared to that determined using projected mass estimators, implies a longer period for the innermost stable orbit, longer resonant relaxation timescales for stars in the vicinity of the black hole and a better agreement with the M$_{bh}$-$\sigma$ relation. \end{abstract} \keywords{black hole physics -- Galaxy:center --- Galaxy:kinematics and dynamics --- infrared:stars -- techniques:high angular resolution} \section{INTRODUCTION} \label{sec:intro} Ever since the discovery of fast moving (v $>$ 1000 km s$^{-1}$) stars within 0.$\tt''$3 (0.01 pc) of our Galaxy's central supermassive black hole (Eckart \& Genzel 1997; Ghez et al. 1998), the prospect of using stellar orbits to make precision measurements of the black hole's mass (M$_{bh}$) and kinematics, the distance to the Galactic center (R$_0$) and, more ambitiously, to measure post-Newtonian effects has been anticipated (Jaroszynski 1998, 1999; Salim \& Gould 1999; Fragile \& Mathews 2000; Rubilar \& Eckart 2001; Weinberg, Milosavlejic \& Ghez 2005; Zucker \& Alexander 2007; Kraniotis 2007; Will 2008). An accurate measurement of the Galaxy's central black hole mass is useful for putting the Milky Way in context with other galaxies through the apparent relationship between the mass of the central black hole and the velocity dispersion, $\sigma$, of the host galaxy (e.g., Ferrarese \& Merrit 2000; Gebhardt et al. 2000; Tremaine et al. 2002). It can also be used as a test of this scaling, as the Milky Way has the most convincing case for a supermassive black hole of any galaxy used to define this relationship. %Unfortunately, recent %estimates of the Milky Way's black hole mass have varied by more than %a factor of 2 with the use of different mass determination methods %and different assumed or inferred distances %(e.g., Eckart \& Genzel 1997; Ghez et al. 1998; %Genzel et al. 2000; Chakrabarty \& Saha 2001; Eisenhaer et al. 2005). Accurate estimates of R$_0$ impact a wide range of issues associated with the mass and structure of the Milky Way, including possible constraints on the shape of the dark matter halo and the possibility that the Milky Way is a lopsided spiral (e.g., Reid 1993; Olling \& Merrifield 2000; Majewski et al. 2006). Furthermore, if measured with sufficient accuracy ($\sim$1\%), the distance to the Galactic center could influence the calibration of standard candles, such as RR Lyrae stars, Cepheid variables and giants, used in establishing the extragalactic distance scale. In addition to estimates of $M_{bh}$ and R$_0$, precision measurements of stellar kinematics offer the exciting possibility of detecting deviations from a Keplerian orbit. This would allow an exploration of a possible cluster of stellar remnants surrounding the central black hole, suggested by Morris (1993), Miralda-Escud{\'e} \& Gould(2000), and Freitag et al. (2006). Estimates for the mass of the remnant cluster range from $10^4 - 10^5 M_{\odot}$ within a few tenths of a parsec of the central black hole. Absence of such a remnant cluster would be interesting in view of the hypothesis that the inspiral of intermediate-mass black holes by dynamical friction could deplete any centrally concentrated cluster of remnants. Likewise, measurements of post-newtonian effects would provide a test of general relativity, and, ultimately, could probe the spin of the central black hole. Tremendous observational progress has been made over the last decade towards obtaining accurate estimates of the orbital parameters of the fast moving stars at the Galactic center. Patience alone permitted new astrometric measurements that yielded the first accelerations (Ghez et al. 2000; Eckart et al. 2002), which suggested that the orbital period of the best characterized star, S0-2, could be as short as 15 years. The passage of more time then led to full astrometric orbital solutions (Sch\"odel et al. 2002, 2003; Ghez et al. 2003, 2005a), which increased the implied dark mass densities by a factor of $10^4$ compared to earlier velocity dispersion work and thereby solidified the case for a supermassive black hole. The advent of adaptive optics enabled radial velocity measurements of these stars (Ghez et al. 2003), which permitted the first estimates of the distance to the Galactic center from stellar orbits (Eisenhauer et al. 2003, 2005). % might also introduce the youth problem here and the interest in gaining % insight from their orbits In this paper, we present new orbital models for S0-2. These provide the first estimates of the distance to the Galactic center and limits on the extended mass distribution based on data collected with the W. M. Keck telescopes. The ability to probe the properties of the Galaxy's central supermassive black hole has benefitted from several advancments since our previous report (Ghez et al. 2005). First, new astrometric and radial velocity measurements have been collected between 2004 and 2007, increasing the quantity of kinematic data available. Second, the majority of the new data was obtained with the laser guide star adaptive optics system at Keck, improving the quality of the measurements (Ghez et al. 2005b; Hornstein et al. 2007). These new data sets are presented in \S\ref{sec:obs}. Lastly, new data analysis has improved our ability to extract radial velocity estimates from past spectroscopic measurements, allowing us to extend the radial velocity curve back in time by two years, as described in \S\ref{sec:data_analysis}. The orbital analysis, described in \S\ref{sec:orbit}, identifies several sources of previously unrecognized biases and the implications of our results are discussed in \S\ref{sec:disc}. \section{OBSERVATIONS \& DATA SETS} \label{sec:obs} \subsection{High Angular Resolution Imaging: Speckle and Adaptive Optics} For the first eleven years of this experiment (1995-2005), the proper motions of stars orbiting the center of our Galaxy were obtained from $K$[2.2 $\mu m$]-band speckle observations of the central stellar cluster with the W. M. Keck I 10-meter telescope and its facility near-infrared camera, NIRC (Matthews \& Soifer 1994; Matthews et al.\ 1996). A total of 27 epochs of speckle observations are included in the analysis conducted in this paper, of which 22 have been reported in earlier papers by our group (Ghez et al. 1998, 2000, 2005a). Five new speckle observations, between 2004 April and 2005 June, were conducted in a similar manner. In summary, during each observing run, $\sim$10,000 short ($t_{exp}$ = 0.1 sec) exposure frames were obtained with NIRC in its fine plate scale mode, which has a scale of 20.46 $\pm$ 0.01 mas pixel$^{-1}$ (see Appendix B) and a corresponding field of view of 5\farcs 2 $\times$ 5\farcs 2. Interleaved with these observations were similar sequences on a dark patch of sky. From these data, we produce images that are diffraction-limited ($\theta$ = 0\farcs 05) and have Strehl ratios of $\sim$0.05. With the advent of laser guide star adaptive optics (LGSAO) in 2004 on the 10 m W. M. Keck II telescope (Wizinowich et al. 2006; van Dam et al. 2006), we have made measurements of the Galaxy's central stellar cluster with much higher Strehl ratios (Ghez et al. 2005b). Between 2004 and 2007, nine LGSAO data sets were taken using the W. M. Keck II facility near-infrared camera, NIRC2 (P.I. K. Matthews), which has an average plate scale of 9.963 $\pm$ 0.006 mas pixel$^{-1}$ (see Appendix C) and a field of view of 10\farcs 2 $\times$ 10\farcs 2. All but one of the observations were obtained through a K' ($\lambda_0$=2.12 $\mu$m, $\Delta \lambda$=0.35 $\mu$m) band-pass filter, with the remaining one obtained through narrow band filters (CO: $\lambda_0$ = 2.278 $\mu$m, $\Delta \lambda$ = 0.048 $\mu$m and Kcont: $\lambda_0$ = 2.27 $\mu$m, $\Delta \lambda$ = 0.030 $\mu$m). During these observations, the laser guide star's position was fixed to the center of the camera's field of view and therefore moved when the telescope was dithered. While the laser guide star is used to correct most of the important atmospheric aberrations, it does not provide information on the tip-tilt term, which, for all our LGSAO observations (imaging and spectroscopy), was obtained from visible observations of USNO 0600-28577051 (R = 13.7 mag and $\Delta r_{SgrA*}$ = 19$\arcsec$). Details of the observing setup for 2004 July 26, 2005 June 30, and 2005 July 31 are described in detail in Ghez et al. (2005b), Lu et al. (2008), and Hornstein et al. (2007), respectively. While each of these early LGSAO observations had a slightly different setup and dither pattern, the more recent, deeper, LGSAO measurements (2006-2007) were obtained with nearly identical setups. Specifically, we used a 20 position dither pattern with randomly distributed (but repeatable) positions in a 0\farcs 7 $\times$ 0\farcs 7 box and an initial position that placed IRS 16NE on pixel (229, 720) at a sky PA set to 0.0. This setup keeps the brightest star in the region, IRS 7 (K=6.4), off the field of view at all times. At each position, three exposures, each composed of 10 coadded 2.8 sec integrations, were obtained; the integration time was set with the aim of keeping the detector's response linear beyond the full width at half maximum (FWHM) point for the brightest (K=9.0) star in the field of view; the number of images per position was chosen to provide the minimum elapsed time needed to allow the LGSAO system's optimization algorithm to converge ($\sim$3 min.) before dithering. Table \ref{tbl_img} summarizes all the new imaging data sets. %, and resulted in a total on-sky exposure time %of 108 sec in 2004 at K', 96 sec in 2005 June through narrow-band filters (2.3 microns) and 688 sec and 868 sec at H and K', %respectively in 2005. % In 2005 June, data was taken through %two narrow-bandpass filters %at dither positions centered at the middle and the corners %of 4\arcsec $\times$ 4\arcsec box. With exposure times for the frames %obtained at each of the five dither positions of 36 s and 59.5 s for the CO %and Kcont filters, respectively, these images have comparable signal to noise %ratios. While all the 2 $\mu$m LGSAO %images are diffraction-limited ($\theta$ $\sim$ 0\farcs 05) %and have Strehl ratios of $\sim$0.3, %the H-band images have resolutions that are slightly larger than %the diffraction limit ($\theta$=60 mas) and somewhat lower Strehl ratios %of $\sim$0.2. \subsection{Adaptive Optics Spectroscopy} To monitor the line-of-sight motions of stars orbiting the center of our Galaxy between the years 2000 and 2007, high angular resolution spectroscopic observations of stars in the Sgr A* stellar cluster were taken with both the natural guide star adaptive optics (NGSAO; Wizinowich et al. 2000) system (2000-2004) and the LGSAO system (2005-2007) on the W. M. Keck II 10 m telescope. The NGSAO atmospheric corrections and the LGSAO tip-tilt corrections were made on the basis of visible observations of USNO 0600-28579500 (R = 13.2 mag and $\Delta r$ $\sim$ 30$\arcsec$) and USNO 0600-28577051 (R = 13.7 mag and $\Delta r$ $\sim$ 19$\arcsec$), respectively. While the angular resolution of the NGSAO spectra was typically 2-3 times the diffraction limit ($\theta_{diff}$ = 54 mas), a point spread function (PSF) FWHM of $\sim$ 70 mas at 2 $\micron$ was achieved for the LGSAO long exposure spectra. Three different spectrometers have been used over the course of this study. Our earliest measurements were obtained in 2000 June with NIRSPEC (McLean et al. 1998, 2000) in its low resolution slit spectrometer mode (R $\sim$ 2600). It was not originally designed to go behind the adaptive optics system and therefore had inefficient throughput in its AO mode; it was, however, the only spectrometer available behind the AO system in 2000. While the resulting low signal to noise data set yielded no line detections in the initial analysis of S0-2 (Gezari et al. 2002), we now have the advantage of knowing what type of lines are present in the spectra and have therefore included this data set in our analysis by retroactively identifying the Br$\gamma$ line, which is used to measure radial velocities (see \S3.2) Between 2002 and 2005, NIRC2 (P.I. K. Matthews) was used in its spectroscopic R $\sim$ 4000 mode, which is generated with a 20 mas pixel scale, a medium-resolution grism and a 2 pixel slit. In 2002, this produced the first line detection in S0-2 (Ghez et al. 2003) and, since then, three new NIRC2 measurements (2 with NGSAO and 1 with LGSAO) have been obtained. Since 2005, OSIRIS, which is an integral field spectrograph with a 2 $\micron$ spectral resolution of $\sim$ 3600 (Larkin et al. 2006), has been used. The field of view of this spectrograph depends on the pixel scale and filter. Most of the OSIRIS observations were taken using the 35 mas pixel scale and the narrow band filter Kn3 (2.121 to 2.229 $\micron$; includes Br$\gamma$), which results in a field of view of $1.\arcsec 12 \times 2.\arcsec 24$, and were centered on S0-2. %, but contained %many of the other short period stars, including S0-16. All of the OSIRIS observations were obtained with the LGSAO system. Table \ref{tbl_spec} summarizes the details of the 10 new spectroscopic measurements of S0-2 that were made between the years 2003 and 2007 (see Gezari et al 2002 \& Ghez et al. 2003 for details of the 2000-2002 measurements). \section{DATA EXTRACTION} \label{sec:data_analysis} \subsection{Image Analysis \& Astrometry} The individual speckle and adaptive optics data frames are processed in two steps to create a final average image for each of the 34 imaging observing runs. First, each frame is sky-subtracted, flat fielded, bad-pixel-corrected, corrected for distortion effects and, in the case of the speckle data, resampled by a factor of two; the distortion correction applied to the NIRC2/LGSAO data is from the NIRC2 pre-ship review results (\url{http://www2.keck.hawaii.edu/inst/nirc2/preship\_testing.pdf}) and those applied to the speckle data sets are the combined transformations given in Ghez et al. (1998) and Lu et al. (2008). The frames are then registered on the basis of the position of IRS 16C, for the speckle images, and a crosss-correlation of the entire image, for the LGSAO image, and combined. For the adaptive optics data sets, the frames whose PSF has a FWHM $<$ 1.25 x FWHM$_{min}$, where FWHM$_{min}$ is the minimum observed FWHM for each epoch and which typically includes $\sim$70\% of the measured frames, are combined with a weighted average with weights set equal to their strehl ratios. To increase the signal to noise ratio of the 2005 June data set, the data taken through the two narrow-band filters are averaged together. For the speckle data set, only the best $\sim$ 2,000 frames from each observing run are combined using a weighted ``Shift-and-Add" technique described by Hornstein (2007). The selected frames from each observing run (speckle and LGSAO) are also divided into three independent subsets from which three subset images are created in a similar manner to the average images; these subset images are used to assess photometric and astrometric measurement uncertainties. Figure \ref{fig_aoVsp} shows examples of the final average LGSAO and speckle images. While all the images sets have point spread function (PSF) cores that are nearly diffraction-limited ($\theta$ $\sim$ 0\farcs 06 vs. $\theta_{diff.~lim}$ = 0\farcs05), the LGSAO images have much higher image quality than the speckle images, with median Strehl ratios of $\sim$0.3 and 0.07, for the LGSAO and speckle images, respectively. %and can achieve as %much as a factor of 7 higher astrometric precision (0.2 mas for K$\sim$14 mag %stars) Point sources are identified and characterized in each of the images %and cross-identified between images using an evolving strategy, which has been %developed in Ghez et al. (1998, 2000, 2005a) and Lu et al. (2007) with only %minor variations here and which %entails four basic steps. In the first step, we generate a conservative %list of sources for each image to help minimize spurious source detections. %This is done by running the point spread function (PSF) fitting program using the PSF fitting program StarFinder (Diolaiti et al. 2000) on both the average images and the subset images. StarFinder iteratively generates a PSF based on user selected point sources\footnote {In this analysis, the stars that are input into the PSF construction are IRS 16C, 16NW, and S2-17 for the speckle images and IRS 16C, 16NW, 16NE, 16SW, 33E, 33W, 7, 29N, and GEN+2.33+4.60 for the LGSAO images.} in the image and identifies additional sources in the image by cross-correlating the resulting PSF with the image. The initial source list for each image is composed only of sources detected in the average images with correlation values above 0.8 and in all three subset images with correlation values above 0.6. Eleven bright (K$<$14 mag), non-variable sources establish the photometric zero points for each list based on measurements made by Rafelski et al. (2007; IRS 16C, IRS 16SW-E, S2-17, S1-23, S1-3, S1-4, S2-22, S2-5, S1-68, S0-13, S1-25). As shown in Figure \ref{speck_ao_klimit}, the deep LGSAO images (K$_{lim} \sim $19 mag) are three magnitudes more sensitive than the speckle images (K$_{lim} \sim $16 mag), which results in roughly three times more sources being detected in the LGSAO images than the speckle images over a comparable region. Because of the higher signal to noise, as shown in Figure \ref{speck_ao_astro}, the centroiding uncertainties ($\delta X', \delta Y'$), which are estimated from the RMS error of the measurements in the three subset images, are a factor of 6 more precise for the deep LGSAO data sets (0.17 mas) than the speckle data sets (1.1 mas), for bright stars (K$<$13 mag); the plateau observed in the relative centroiding uncertainties for the brighter stars (K$<$13) in the LGSAO images is likely caused by the combined effects of differential tip-tilt jitter and residual optical distortions across the field of view. The sources identified each night are matched across multiple epochs and their positions are transformed to a common coordinate system that will be referred to as the {\it cluster reference frame}. As detailed in Appendix A, the transformation for each epoch is derived by minimizing the net displacement of a set of ``coordinate reference" stars, allowing for proper motions, relative to their positions in a common reference image, which, in this case, is the 2004 July LGSAO image. This procedure attempts to ensure that in the cluster reference frame the coordinate reference stars are at rest (i.e., no net translation, rotation, expansion, or skew). A total of $\sim$470 and $\sim$120 stars serve as coordinate reference stars in the LGSAO and speckle epochs, respectively. These stars are selected based on the following criteria: (1) high detection correlations ($>$0.9), ensuring good positional accuracy, (2) located more than 0\farcs 5 from Sgr A* to avoid sources with measurable non-linear motions (i.e., accelerations in the plane of the sky $> \sim$8 km/s/yr), (3) low velocities ($<$ 15 mas/yr, or equivalently $\sim$600 km s$^{-1}$), which eliminates possible coordinate reference sources that have been mismatched across epochs, and (4) lack of spectroscopic identification as a young star from Paumard et al. (2006) to eliminate the known net rotation of the young stars in the cluster reference frame. Positional uncertainties from this transformation process, which are characterized by a half sample bootstrap applied to the coordinate reference stars, are a factor of $\sim$1.5 (speckle) to 6 (LGSAO) smaller than the centroiding uncertainties and grow by less than a factor of 2 between the center of the field of view (minimum) and a radius of 3\arcsec. An additional source of positional error originates from residual optical distortion in NIRC2. While the residual distortion in NIRC2 is small, the extremely precise centroid measurements in the deep LGSAO images make it a significant effect. The presence of such a systematic error is established by examining the distribution of positional residuals, normalized by measurement (centroiding plus alignment) uncertainties, to the linear proper motion fits for the coordinate reference stars. The speckle data sets do not show large, measurable biases; the speckle measurements, on average, are only 1$\sigma$ off from the linear proper motion fit. In contrast, the much more precise deep-LGSAO astrometric measurements are, on average, 5$\sigma$ off from these fits. As described in Appendix B, we account for this effect at two stages of our analysis. First, 0.88 mas is added in quadrature to the positional uncertainties of the coordinate reference stars to account for systematic errors in the coordinate transformations. Second, a local correction, in the coordinate reference frame, is derived and applied to the positions of the short period stars that were made with LGSAO setups that differ from that of the reference image. This procedure ensures that residuals from both linear proper motion fits to the coordinate reference stars (see Appendix A \& B) and from orbit fits to S0-2 (see \S4) are consistent with a normal distribution. Source confusion can introduce positional biases that can be comparable to and, at certain times, larger than the statistical errors caused by background or detector noise. This occurs when two stars are sufficiently close to each other that only one source, rather than two, is identified in our analysis with a brightness that includes flux from both sources and a position that corresponds roughly to the photocenter of the two stars. We divide the problem of handling source confusion in our data set into the following two cases: (1) the impact of unresolved, underlying stars that are known sources, because they were sufficiently well separated at other times, and bright enough, to be independently detected, and (2) the impact of unresolved, underlying stars that are not identified by this study at another time. Because the sources are moving so rapidly, instances of the former case are easily identified and are typically blended for one year. An underlying source that is comparably bright to the source of interest can have a significant impact on the astrometry; to quantify this effect, we examine the idealized, noise-free case of a perfectly known PSF by using our empirical PSFs to generate idealized binary stars and running StarFinder on these simulated images, inputting the known PSF. In this case, the astrometric bias is zero once the two components are detected. As Figure \ref{bias_binary} shows, when the sources are blended, the resulting astrometric biases can be easily as large as 10 mas, which is much larger than our centroiding uncertainties. Such a large astrometric bias occurs when the underlying source is at least half as bright as the primary source and has a projected, although unresolved, separation of $\sim$ 40 mas. We conservatively choose to eliminate all astrometric measurements that are known to be the blend of two sources from the orbital analysis; specifically, if the predicted positions of two known sources are separated by less than 60 mas and only one of them is detected, then that measurement is removed from our analysis. For S0-2 (K=14.2 mag), the eliminated data points are those made in 1998, due to confusion with S0-19 (K=15.6 mag), in 2002, due to overlap with SgrA*-IR (K$_{median}$ = 16.4 mag, but can be as bright as 14 mag; see Do et al. 2008), and in May 2007, due to superposition with S0-20 (K=15.9). The impact of these overlapping sources, in the first two cases, can be seen in the photometric measurements (see Figure \ref{s02_phot}). Source confusion from unknown sources is a smaller effect than that from known sources, since the unknown stars, in general, are fainter than the known sources. Given the long time-baseline of the speckle imaging experiment, knowledge of sources in this region is most likely complete down to K= 16.0 mag. While sources as faint as K = 19 mag have been detected in this region with LGSAO, crowding and the short time baseline of these deeper observations limit the census of these sources. Therefore, source confusion from unknown sources can give rise to astrometric biases for S0-2 as large as 3 mas (from a K=16 mag source), but are typically significantly smaller since underlying sources will generally be fainter than K = 16 mag. To characterize the expected astrometric bias from the undetected source distribution, a Monte Carlo simulation was performed by generating multiple images with all known stars plus a random stellar distribution that, in total, follows the K luminosity function and radial profile from Sch\"odel et al. (2007). By running these simulated images through our data analysis prodedure, we estimate that the astrometric error from unknown sources for S0-2 is, on average, 0.5 mas and 1.2 mas for the LGSAO and speckle images, respectively, and that it scales roughly with the photometric bias and galacto-centric distance. However, it should be noted that the exact value of this bias is model dependent. While the photometric bias may be detected in the speckle data toward closest approach (see Figure \ref{s02_phot}), the estimated astrometric biases are smaller than other sources of positional uncertainty already included for the majority of the S0-2 data points. We therefore do not incorpate them into the reported positional uncertainties. Confusion with unknown sources gives rise to larger astrometric biases for S0-16, S0-19, and S0-20, since these sources are fainter than S0-2. Given the velocity dispersion in this region and the angular resolution of the data sets, the expected timescale associated with biases from source confusion is $\sim$1-2 years. As a final step, the relative astrometric positions are placed in an absolute coordinate reference frame using the positions of seven SiO masers (Reid et al. 2003, 2007). Infrared observations of these masers with the Keck II LGSAO/NIRC2 system between 2005 and 2007 were obtained with the same camera (i.e., plate scale) used for the precision astrometry measurements described above, but with a nine position box pattern and a 6$\arcsec$ dither offset to create a 22$\arcsec \times$22$\arcsec$ mosaic of these masers (see Appendix C for details). A comparison of the maser positions measured in this infrared mosaic to the predicted radio positions at this epoch from Reid et al. (2003) establishes that the mosaic has an average pixel scale of 9.963 $\pm$ 0.005 mas/pixel and a position angle of north with respect to the NIRC2 columns of 0.$^o$13 $\pm$ 0.$^o$02. This same analysis localizes the radio position of Sgr A* in the infrared mosaic to within 5 mas in the east-west and north-south directions. By aligning the infrared stars detected in both the larger infrared mosaic and the precision astrometry image taken during the same observing run, we have the necessary coordinate transformations to convert our relative astrometric position measurements into an absolute reference frame. For the orbit analysis described in \S4, the uncertainties in this transformation are applied only after model orbits have been fit to the relative astrometry and are a negligible source of uncertainty in the final mass and R$_o$ estimates. \subsection{Spectral Analysis \& Radial Velocities} In the analysis of the spectral data, we accomplish the initial basic data processing steps using standard IRAF procedures, for NIRC2 and NIRSPEC, and a facility IDL data extraction pipeline for OSIRIS. Specifically, each data set is first (1) flat fielded, (2) dark subtracted, (3) bad pixel and cosmic ray corrected, (4) spatially dewarped, and (5) wavelength calibrated. Wavelength calibration is performed by identifying OH emission lines from sky spectra and fitting a low-order polynomial function to the location of the lines. For the NIRSPEC spectra, neon emission lines from arc lamps provide the wavelength calibration. The accuracy of the wavelength calibration is $\sim$9 km s$^{-1}$ or less for NIRC2 and OSIRIS as measured by the dispersion of the residuals to the fit. Next, the one-dimensional stellar spectra are extracted using a spatial window that covers $\sim 0\farcs 1$ for the two dimensional spectral data sets from NIRC2 and NIRSPEC. For the three dimensional spectral data set from OSIRIS, an extraction box 0\farcs 14 $\times$ 0\farcs 14 was used. To correct for atmospheric telluric absorption features, each spectrum is divided by the spectrum of an A-type star. Prior to this step, the A-type star's strong intrinsic Br${\gamma}$ feature is removed. In the case of the NIRC2 and OSIRIS observations, this correction is done with observations of a G2V star, which is divided by a model solar spectrum. The Br$\gamma$ corrected region in the G star is then substituted into the same region of the A star (Hanson et al. 1996). In the case of the NIRSPEC observations, the A-type star's \brg feature is corrected with a model spectrum of Vega\footnote{Model taken from the 1993 Kurucz Stellar Atmospheres Atlas (\url{ftp://ftp.stsci.edu/cdbs/cdbs2/grid/k93models/standards/vega\_c95.fits})} rebinned to the resolution of the A-type star's spectrum and convolved with a Gaussian to match the spectral resolution of the observations. The resulting stellar spectra are corrected for all telluric absorption features; however, they are still contaminated by background emission due to the gas around the Galactic center. The local background is estimated and removed by subtracting spectra extracted from regions that are $\sim$0\farcs 1 away. Finally, all the spectra within each night of observation are combined in an average, weighted by the signal to noise ratio. Radial velocity estimates are determined for each spectrum on the basis of the location of the Br$\gamma$ line. While a few of our spectra with broader spectral coverage also show a weaker He I triplet at 2.116 $\micron$, we do not incorporate measurements from this line, as it is a blend of transitions that can bias the resulting radial velocities (see Figure \ref{spec_avg}). A Gaussian model is fit to each of the Br$\gamma$ line profiles and the wavelength of the best fit peak, is compared to the rest wavelength of $\lambda_{vacuum}$ = 2.1661 $\micron$ to derive an observed radial velocity. %The final S0-2 spectrum for %the 2000 NIRSPEC data was cross-correlated with a %gaussian profile of the %same FWHM and equivalent width as the 2003 identified \brg line. % - ie there were less free parameters in this fit. To obtain radial velocities in the local standard of rest (LSR) reference frame, %, which are reported in Table XXX, each observed radial velocity is corrected for the Earth's rotation, its motion around the Sun, and the Sun's peculiar motion with respect to the LSR (U = 10 km s$^{-1}$, radially inwards; Dehnen \& Binney 1998). Since the LSR is defined as the velocity of an object in circular orbit at the radius of the sun, the Sun's peculiar motion with respect to the average velocity of stars in its vicinity should give the Sun's motion toward the center of the Galaxy. The uncertainties in the final radial velocities are obtained from the rms of the fits to the line profile measurements from at least three independent subsets of the original data set. Figure \ref{spec_brg} shows how S0-2's Br$\gamma$ line has shifted over time and how the measurement of this line has improved by a factor of 5 with improved instrumentation. For the deep LGSAO spectroscopic observations, the radial velocity uncertainties for S0-2 are typically $\sim$20-25 km s$^{-1}$. \section{ORBITAL ANALYSIS \& RESULTS} \label{sec:orbit} \subsection{Point Mass Only Analysis} To derive the black hole's properties, we assume that the stars are responding to the gravitational potential of a point mass. %which allows multiple stars to contribute %simultaneously to the solution for the %central object's properties. In this analysis, the 7 properties of the central black hole that are fitted are its mass ($M$), distance (R$_0$), location on the plane of the sky ($X_0$, $Y_0$) and motion ($V_x$, $V_y$, $V_z$). %. Initially, we assume that the %black hole's position is fixed relative to the central stellar cluster. In addition to these common free parameters, there are the following 6 additional free parameters for each star: period ($P$), eccentricity ($e$), time of periapse passage ($T_0$), inclination (i), position angle of the ascending node ($\Omega$), and the longitude of periapse ($\omega$). Using a conjugate gradient $\chi^2$ minimization routine that simultaneously fits the astrometric and radial velocity measurements, we fit this model to measurements that are given in Tables \ref{tbl_pos} \& \ref{tbl_rv}, %described in \ref{sec:obs} \& \ref{sec:data_analysis}, which includes 27 epochs of astrometric measurements and 11 epochs of radial velocity (RV) measurements, as well as 5 additional epochs of radial velocity measurements reported in the literature (Eisenhauer et al. 2003, 2005). This excludes all the astrometric measurements of S0-2 that are confused with another known source (see \S3.1). While the 2002 astrometric data are eliminated due to confusion with SgrA*, the 2002 RV points are not, since SgrA* is featureless and therefore does not bias the measurement of RV from S0-2's Br$\gamma$ absorption line. In total, there are 38 astrometric data points and 5 RV measurements. All values reported for each parameter are the best fit values obtained from minimizing the total $\chi^2$, which is the sum of the $\chi^2$ from each data type (i.e., $\chi^2_{tot} = \chi^2_{ast} + \chi^2_{RV}$). The uncertainties on the fitted parameters are estimated using a Monte Carlo simulation, which is a robust approach when performing a fit with many correlated parameters. We created $10^5$ artificial datasets ($N_{sim}$) containing as many points as the observed dataset (astrometry and radial velocities), in which each point is randomly drawn from a Gaussian distribution centered on the actual measurement and whose 1 $\sigma$ width is given by the associated uncertainty, and run the $\chi^2$ minimization routine for each realization. $N_{sim}$ was set to $10^5$ in order to achieve $\sim$6\% accuracy in the resulting estimates of the a 99.73\% confidence limits (3$\sigma$ equivalent for a gaussian distribution) of the orbital parameters. Because the $\chi^2$ function contains many local minima, each realization of the data is fit 1000 times ($N_{seed}$) with different seeds to find the global minimum. The resulting distribution of $10^5$ values of the fitted parameters from the Monte Carlo simulation, once normalized, is the a joint probability distribution function of the orbital parameters ($PDF({\vec O})$, where $\vec O$ is a vector containing all the orbital parameters,$O_i$). For each orbital parameter, $PDF({\vec O})$ is marginalized against all other orbital parameters to generate a $PDF(O_i)$. The confidence limits for each parameter are obtained by integrating each $PDF(O_i)$ from its peak\footnote{While the best values from minimizing $\chi^2$ can differ slightly (but well within the uncertainties) from the peak of the $PDF(O_i)$ values, this has negligible impact on the reported uncertainties.} outwards to a probability of 68%. Compared to all other stars at the center of the Milky Way, S0-2 dominates our knowledge of the central black hole's properties. Two facts contribute to this effect. Most importantly, it has the shortest known orbital period (P = 15 yr; Sch\"odel et al. 2002, 2003; Ghez et al. 2003, 2005a). Furthermore, among the known short-period stars, it is the brightest star and therefore the least affected by stellar confusion (see Figure 1). Several other stars, in principle, also offer constraints on the black hole's properties. In particular, S0-16 is the next most kinematically important star, as it is the only other star that yields an independent solution for the black hole's properties. %The resulting uncertainties for the black hole parameters %are at least a factor of 5 smaller for those %derived from S0-2's measurements compared to those obtained %independently from any other star's measurements. However, independent solutions for the black hole's position from fits to S0-2 and S0-16 measurements differ by more than 5 $\sigma$ (see Figure \ref{bhPosition}). While S0-16's measurements in 2000 have already been omitted due to overlap with the position of SgrA*, three independent lines of reasoning lead us to believe that some of S0-16's remaining astrometric measurements must be significantly biased by radiation from unrecognized, underlying stars. % First, as shown in Figure \ref{bias_binary}, unknown sources can introduce astrometric biases as large as 9 mas for S0-16 (K=15), in contrast with only 3 mas for S0-2 (K = 14), because it is only 1 mag above the completeness limit for detection in the speckle data set (K$\sim$16 mag; see \S3.1). % Second, a comparison of the solution for the position of the black hole ($X_0, Y_0$) based on both the astrometric and radial velocity measurements to that based on astrometry alone (fixing the distance, which cannot be solved for without radial velocities) yields a consistent position from modeling the two cases for S0-2, but produces different results for the two cases from modeling S0-16's measurements, with the inferred $X_0$ and $Y_0$ from astrometry alone shifting further away from that obtained from modeling S0-2's orbit prediction and thereby increasing the discrepancy to 10$\sigma$. % Third and last, while the position of the dynamical center from S0-2's orbit is statistically consistent with SgrA*-Radio/IR, which is the emissive source associated with the central black hole (e.g., Melia \& Heino 2001; Genzel et al. 2003a; Ghez et al. 2004; 2005b; Hornstein et al. 2007), the solution from S0-16 is not (see Figure \ref{bhPosition}); this difference cannot be explained by allowing the black hole to move with time or by introducing an extended mass distribution. We therefore restrict our remaining analysis to S0-2. As shown in Figures \ref{s02_orb} \& \ref{s02_orb_resid}, the astrometric and radial velocity measurements for S0-2 are well fit by a simple Keplerian model. For a 13 parameter model (right-hand side of figures), the best fit to the data % need to introduce somewhere that VLT RVs are being used! produces a total $\chi^2$ of 54.8 for 57 degrees of freedom ($dof$) and a $\chi^2/dof$ of 0.961. From the Monte Carlo simulation, we derive probability distributions for the central black hole's properties, which are shown in Figure \ref{massRo} and characterized % % should add a plot on Xo, Yo % in Table \ref{tbl_orb}. These distributions give a best fit for the central black hole's mass of M$_{bh}$ = $4.1 \pm 0.6 \times 10^6 M_{\odot}$ and distance of R$_0$ = 8.0 $\pm$ 0.6 kpc (all quoted uncertainties are 68\% confidence values). The position of the black hole is confined to within $\pm$ 1 mas ($\sim$100 Schwarzschild Radii). As can be seen in Figure \ref{massRo}, the inferred black hole's mass is highly correlated with its distance. Estimates from orbital modeling are expected to have a power law relationship of the form $Mass \propto M_{\odot} ~ Distance^{\alpha}$ with $\alpha$ between 1 and 3. For the case of astrometric data only, $\alpha$ should be 3 and, for the case of radial velocity data only, $\alpha$ is expected to be 1. Currently, the relationship is $M = (4.1 \pm 0.1 \times 10^6 M_{\odot})(R_0/8.0 ~ kpc)^{1.8}$, which suggests that the astrometric and radial velocity data sets are having roughly equal affect in the model fits for mass\footnote{The uncertainty in the mass scaling relationship is obtained for the case in which R$_0$ is fixed to 8.0 kpc and therefore does not include the uncertainty in R$_0$.}. %For comparison with other recent mass estimates, a model %with a fixed distance of 8.0 kpc is also fit to the data %and this results in a mass estimate of $4.0 \times 10^6 M_{\odot}$, consistent with the %relation above. A fit that includes the biased astrometric data points significantly alters the best fit solution for S0-2. Including both the 1998 and 2002 data points, which correspond to confusion with S0-19 and SgrA*-IR respectively, results in a higher mass ($5.7 \times 10^6 M_{\odot}$), distance (9.4 kpc), and $\chi^2/dof$ (1.7). % 673.7, 635.9 Including the 2002 but not the 1998 data points also produces elevated values ($5.2 \times 10^6 M_{\odot}$ and 9.1 kpc) and $\chi^2/dof$ (1.1). % 673.7, 635.7) % % orig % (673.6, 635.4) % so this doesn't change enough to explain the offset This demonstrates that it is important to account for the astrometric biases introduced by unresolved sources. Formal uncertainties in mass and distance estimates from orbital fits can be reduced by adding {\it a priori} information. In particular, it is, in principle, possible to constrain the dynamical center to be at the position of SgrA*-IR. However, as shown in Figure \ref{bhPosition}, the six measurements of SgrA*-IR's position in the deep LGSAO images (2005-2007), which have the most precise astrometric measurements, have an average value that differs from the position of the black hole inferred from S0-2's orbit by 9.3 mas and a variance of 3 mas, which is a factor of 4 larger than expected from the measured positional uncertainties (0.7 mas). SgrA*-IR is located where the underlying sources are expected to have the highest number density and velocity dispersion, which should induce time variable positional biases. SgrA*-IR's average K magnitude in these deep LGSAO images is 16.4, which is comparable to the completeness limit for sources in this region (see \S3.1) and which is, consequently, potentially subject to large astrometric biases (see Figure \ref{bias_binary}). We therefore suspect that the measured positions of SgrA*-IR suffer from astrometric biases from underlying sources and do not use its positions to constrain the model fits. Another prior, which has been imposed in earlier orbital analyses of S0-2 for R$_0$ (Eisenhauer et al. 2003; 2005), is on the black hole's motion relative to the measurements' reference frame. Setting the three dimensional velocity to zero and fitting a 10 parameter model ($\chi^2/dof$ = 1.3; see left-hand side of Figures \ref{s02_orb} \& \ref{s02_orb_resid}) yeilds uncertainties in the black hole's properties that are a factor of 2 smaller (R$_0$ = 8.0 $\pm$ 0.3 kpc and M$_{bh}$ = $4.4 \pm 0.3 \times 10^6 M_{\odot}$). However this assumption is not justified (see, e.g., Salim \& Gould 1999; Nikiforov 2008). Introducing $V_x$ and $V_y$ (defined such that positive numbers are motions in the E and N directions, respectively) into the fit allows the dynamical center to move linearly in time in the plane of the sky with respect to the cluster reference frame. Such an apparent motion can arise from either a physical or a data analysis effect. In the case of a physical effect, the black hole could be moving with respect to the stellar cluster under the gravitational influence of a massive companion or the black hole and the cluster could be participating in a mutually opposing sloshing mode. In the case of a data analysis effect, the reference frame could be non-stationary with respect to the position of the dynamical center, which might arise if there was a systematic problem in our alignment of images. Introducing these two parameters therefore provides a way of examining possible systematic reference frame problems. Fits to a 12 parameter model ($V_z$ fixed to zero) to the data have a minimum $\chi^2/dof$ of 0.95, uncertainties in the black hole's properties that are larger than the 10-parameter model, but smaller than the 13-parameter model (R$_0$ 8.4 $\pm$ 0.4 kpc and M$_{bh}$ = $4.5 \pm 0.4 \times 10^6 M_{\odot}$), and an estimate for the black hole's motion relative to the central stellar cluster of $V_x$ = -0.40 $\pm$ 0.25 mas/yr (17 $\pm$ 11 km/sec) and $V_y$ = 0.39 $\pm$ 0.14 mas/yr (16 $\pm$ 6 km/sec). Since these relative velocities are comparable to the constraints on the IR reference frame's motion with respect to SgrA*-Radio (i.e., an absolute reference frame in which the black hole's position is known; see Appendix C), it is important to leave $V_x$ and $V_y$ as free parameters, even for the case in which one assumes that the black hole has no intrinsic motion motion with respect to the cluster. Because the black hole is so often assumed to be at rest, we report the complete solution for the 12 parameter fit (V$_z$ fixed to zero) in Table \ref{tbl_orb}. As Figure \ref{RoVz} shows, the black hole's motion along the line of sight with respect to our assumed local standard of rest ($V_z$) dominates the uncertainties in R$_0$ in our 13 parameter model. Priors on $V_z$ therefore have a signficant impact on the resulting uncertainties. Unlike the plane of the sky, the reference frame along the line of sight is unlikely to have an instrumental systematic drift, since each of the spectra are calibrated against OH lines (see \S3.2). However, it is possible that there is a residual radial velocity offset between the LSR and the S0-2 dynamical center. The Sun's peculiar motion with respect to the LSR along the line of sight might differ from the assumed 10 km s$^{-1}$; that is, the practical realization of the LSR is not on a circular orbit around the Galactic center as might occur due the bar potential or to the spiral perturbations, so that the average velocity of stars in the solar vicinity might have a (small) net radial component. Alternatively, the dynamical center of S0-2 could differ from the dynamical center of the Galaxy as determined at the Sun's (i.e., LSR's) distance, as might result from the presence of an intermediate mass black hole companion. From the model fit, the implied motion of the LSR along the line-of-sight with respect to S0-2's dynamical center is -20 $\pm$ 33 km/sec, which is consistent with no net motion. While no significant motion is detected in $V_x$, $V_y$, or $V_z$, the 3$\sigma$ upper limits for the magnitudes of all three are comparable to one another in our 13 parameter model (48, 30, and 119 km/sec, respectively). Since there are no direct contraints on these quantities that can improve these limits, we have allowed them to be fully free parameters. However, if we {\it assume} that the black hole is stationary with respect to the Galaxy, we also need to consider the case of Vz set to zero\footnote{Allowing for the uncertainty in the LSR in $V_z$ ($\pm$ 2 km/sec; Gould 2004) produces results that are not distinguishable from those reported for the $V_z$ = 0 case.}. \subsection{Point Mass Plus Extended Mass Distribution Analysis} Limits on an extended mass distribution within S0-2's orbit are derived by assuming that the gravitational potential consists of a point mass and an extended mass distribution, and allowing for a Newtonian precession of the orbits (see, e.g., Rubilar \& Eckart 2001). In order to do this, we use the orbit fitting procedure described in Weinberg et al. (2005), and adopt an extended mass distribution that has a power-law density profile $\rho(r)=\rho_0(r/r_0)^{-\gamma}$. This introduces two additional parameters to the model: the normalization of the profile and its slope $\gamma$. The total enclosed mass is then given by \begin{equation} M($0\farcs 5), 3) have velocities less than 15 mas/yr and velocity errors less than 5 mas/yr, 4) have reasonable velocity fits ($\chi^2/dof < 4$), and 5) are brighter than K=15. With absolute kinematics for 158 stars within 5'', we solve for a 4 parameter transformation model by comparing the relative positions in the reference epoch image, which are in instrumental pixel coordinates, and the estimated absolute coordinates for that epoch, which are in arcsec relative to the position of SgrA*. Since all other epochs are aligned to this reference epoch, positional measurements for all stars in all epochs are easily transformed into absolute coordinates. While uncertainties in the absolute infrared reference frame dominate the final absolute positional uncertainties relative to SgrA*-Radio, they are a negligible source of uncertainty for the orbital analysis. \begin{thebibliography}{} \bibitem[B{\'e}langer et al. (2006)]{2006JPhCS..54..420B} B{\'e}langer, G., Terrier, R., de Jager, O.~C., Goldwurm, A., \& Melia, F.\ 2006, Journal of Physics Conference Series, 54, 420 \bibitem[Chakrabarty \& Saha (2001)]{2001AJ....122..232C} Chakrabarty, D., \& Saha, P.\ 2001, \aj, 122, 232 \bibitem[Cox (2000)]{2000asqu.book.....C} Cox, A.~N.\ 2000, Allen's astrophysical quantities, 4th ed.~Publisher: New York: AIP Press; Springer, 2000.~Editedy by Arthur N.~Cox.~ ISBN: 0387987460, \bibitem[Dehnen \& Binney (1998)]{1998MNRAS.298..387D} Dehnen, W., \& Binney, J.~J.\ 1998, \mnras, 298, 387 \bibitem[Diolaiti et al. (2000)]{2000A&AS..147..335D} Diolaiti, E., Bendinelli, O., Bonaccini, D., Close, L., Currie, D., \& Parmeggiani, G.\ 2000, \aaps, 147, 335 \bibitem[Do et al. (2008)]{} Do, T., Ghez, A.~.M, Morris, M.~R., Yelda, S., Lu, J.~R., Hornstein, S., Matthews, K. 2008, \apj, submitted \bibitem[Eckart \& Genzel (1997)]{1997MNRAS.284..576E} Eckart, A., \& Genzel, R.\ 1997, \mnras, 284, 576 \bibitem[Eckart et al. (2002)]{bib5} Eckart, A., Genzel, R., Ott, T., \& Sch\"odel, R. 2002, \mnras, 331, 917 \bibitem[Eckart et al. (2006)]{2006A&A...455....1E} Eckart, A., Sch{\"o}del, R., Meyer, L., Trippe, S., Ott, T., \& Genzel, R.\ 2006, \aap, 455, 1 \bibitem[Eisenhauer et al. (2003)]{eisenhauer03} Eisenhauer, F., Sch\"odel, R., Genzel, R., Ott, T., Tecza, M., Abuter, R., Eckart, A., \& Alexander, T. 2003, \apj, 597, L121 \bibitem[Eisenhauer et al. (2005)]{2005ApJ...628..246E} Eisenhauer, F., et al.\ 2005, \apj, 628, 246 \bibitem[Ferrarese \& Merritt (2000)]{2000ApJ...539L...9F} Ferrarese, L., \& Merritt, D.\ 2000, \apjl, 539, L9 \bibitem[Figer et al. (2003)]{2003ApJ...599.1139F} Figer, D.~F., et al.\ 2003, \apj, 599, 1139 \bibitem[Fragile \& Mathews (2000)]{2000ApJ...542..328F} Fragile, P.~C., \& Mathews, G.~J.\ 2000, \apj, 542, 328 \bibitem[Freitag et al. (2006)]{2006ApJ...649...91F} Freitag, M., Amaro-Seoane, P., \& Kalogera, V.\ 2006, \apj, 649, 91 \bibitem[Gebhardt et al. (2000)]{2000ApJ...539L..13G} Gebhardt, K., et al.\ 2000, \apjl, 539, L13 \bibitem[Genzel et al. (1997)]{1997MNRAS.291..219G} Genzel, R., Eckart, A., Ott, T., \& Eisenhauer, F.\ 1997, \mnras, 291, 219 \bibitem[Genzel et al. (2003a)]{2003Natur.425..934G} Genzel, R., Sch{\"o}del, R., Ott, T., Eckart, A., Alexander, T., Lacombe, F., Rouan, D., \& Aschenbach, B.\ 2003a, \nat, 425, 934 \bibitem[Genzel et al. (2003b)]{2003ApJ...594..812G} Genzel, R., et al.\ 2003b, \apj, 594, 812 \bibitem[Genzel et al. (2000)]{2000MNRAS.317..348G} Genzel, R., Pichon, C., Eckart, A., Gerhard, O.~E., \& Ott, T.\ 2000, \mnras, 317, 348 \bibitem[Gezari et al. (2002)]{gezari02} Gezari, S., Ghez, A.~M., Becklin, E.~E., Larkin, J., McLean, I.~S., \& Morris, M. 2002, \apj, 576, 790 \bibitem[Ghez et al. (2003)]{ghez03} Ghez, A.~M., Duch\^ene, G., Matthews, K., Hornstein, S.~D., Tanner, A., Larkin, J., Morris, M., Becklin, E.~E., Salim, S., Kremenek, T., Thompson, D., Soifer, B.T., Neugebauer, G., McLean, I. 2003, \apj, 586, L127 \bibitem[Ghez et al. (2005b)]{2005ApJ...635.1087G} Ghez, A.~M., et al.\ 2005b, \apj, 635, 1087 \bibitem[Ghez et al. (1998)]{1998ApJ...509..678G} Ghez, A.~M., Klein, B.~L., Morris, M., \& Becklin, E.~E.\ 1998, \apj, 509, 678 \bibitem[Ghez et al. (2000)]{bib13} Ghez, A.~M., Morris, M., Becklin, E.~E., Tanner, A., \& Kremenek, T. 2000, \nat, 407, 349 \bibitem[Ghez et al. (2005a)]{2005ApJ...620..744G} Ghez, A.~M., Salim, S., Hornstein, S.~D., Tanner, A., Lu, J.~R., Morris, M., Becklin, E.~E., \& Duch{\^e}ne, G.\ 2005a, \apj, 620, 744 \bibitem[Ghez et al. (2004)]{2004ApJ...601L.159G} Ghez, A.~M., et al.\ 2004, \apjl, 601, L159 \bibitem[Gnedin \& Primack (2004)]{2004PhRvL..93f1302G} Gnedin, O.~Y., \& Primack, J.~R.\ 2004, Physical Review Letters, 93, 061302 \bibitem[Gondolo \& Silk (1999)]{1999PhRvL..83.1719G} Gondolo, P., \& Silk, J.\ 1999, Physical Review Letters, 83, 1719 \bibitem[Gould(2004)]{2004ApJ...607..653G} Gould, A.\ 2004, \apj, 607, 653 \bibitem[Hanson, Conti, \& Rieke (1996)]{hanson96} Hanson, M.~M., Conti, P.~S., \& Rieke, M.~J.\ 1996, \apjs, 107, 281 \bibitem[Hall \& Gondolo (2006)]{2006PhRvD..74f3511H} Hall, J., \& Gondolo, P.\ 2006, \prd, 74, 063511 \bibitem[Haller \& Melia (1996)]{1996ApJ...464..774H} Haller, J.~W., \& Melia, F.\ 1996, \apj, 464, 774 \bibitem[Hopman \& Alexander (2006)]{2006ApJ...645.1152H} Hopman, C., \& Alexander, T.\ 2006, \apj, 645, 1152 \bibitem[Hornstein (2007)]{2007PhDT.........6H} Hornstein, S.~D.\ 2007, Ph.D.~Thesis, UCLA \bibitem[Hornstein et al. (2007)]{2007ApJ...667..900H} Hornstein, S.~D., Matthews, K., Ghez, A.~M., Lu, J.~R., Morris, M., Becklin, E.~E., Rafelski, M., \& Baganoff, F.~K.\ 2007, \apj, 667, 900 \bibitem[Jaroszynski (1998)]{1998AcA....48..653J} Jaroszynski, M.\ 1998, Acta Astronomica, 48, 653 \bibitem[Jaroszy{\'n}ski (1999)]{1999ApJ...521..591J} Jaroszy{\'n}ski, M.\ 1999, \apj, 521, 591 \bibitem[Kerr \& Lynden-Bell (1986)]{1986MNRAS.221.1023K} Kerr, F.~J., \& Lynden-Bell, D.\ 1986, \mnras, 221, 1023 \bibitem[Kraniotis 2007]{} Kraniotis, G. V. 2007, Class. Quantum Grav., 24, 1775 \bibitem[Larkin et al. (2006)]{2006NewAR..50..362L} Larkin, J., et al.\ 2006, New Astronomy Review, 50, 362 \bibitem[Levin \& Beloborodov (2003)]{2003ApJ...590L..33L} Levin, Y., \& Beloborodov, A.~M.\ 2003, \apjl, 590, L33 \bibitem[Lu et al. (2006)]{2006JPhCS..54..279L} Lu, J.~R., Ghez, A.~M., Hornstein, S.~D., Morris, M., Matthews, K., Thompson, D.~J., \& Becklin, E.~E.\ 2006, Journal of Physics Conference Series, 54, 279 \bibitem[Lu et al. (2008)]{2006JPhCS..54..279L} Lu, J.~R., Ghez, A.~M., Hornstein, S.~D., Morris, M., Matthews, K., Thompson, D.~J., \& Becklin, E.~E.\ 2008, \apj, submitted \bibitem[Majewski et al. (2006)]{2006ApJ...637L..25M} Majewski, S.~R., Law, D.~R., Polak, A.~A., \& Patterson, R.~J.\ 2006, \apjl, 637, L25 \bibitem[Matthews et al. (1996)]{} Matthews, K., Ghez, A. M., Weinberger, A. J., and Neugebauer, G. 1996, \pasp, 108, 615 \bibitem[Matthews \& Soifer (1994)]{} Matthews,~K. and Soifer,~B.~T. 1994, Astronomy with Infrared Arrays: The Next Generation, ed. I. McLean, Kluwer Academic Publications (Astrophysics and Space Science, v. 190, p. 239) \bibitem[McLean et al. (1998)]{mclean98} McLean, I.~S.~et al.\ 1998, \procspie, 3354, 566 \bibitem[McLean et al. (2000)]{mclean00} McLean, I.~S., Graham, J.~R., Becklin, E.~E., Figer, D.~F., Larkin, J.~E., Levenson, N.~A., \& Teplitz, H.~I.\ 2000, \procspie, 4008, 1048 \bibitem[Melia \& Falcke(2001)]{2001ARA&A..39..309M} Melia, F., \& Falcke, H.\ 2001, \araa, 39, 309 \bibitem[M{\'e}ndez et al. (1999)]{1999ApJ...524L..39M} M{\'e}ndez, R.~A., Platais, I., Girard, T.~M., Kozhurina-Platais, V., \& van Altena, W.~F.\ 1999, \apjl, 524, L39 \bibitem[Merritt et al. (2002)]{2002PhRvL..88s1301M} Merritt, D., Milosavljevi{\'c}, M., Verde, L., \& Jimenez, R.\ 2002, Physical Review Letters, 88, 191301 \bibitem[Miralda-Escud{\'e} \& Gould (2000)]{2000ApJ...545..847M} Miralda-Escud{\'e}, J., \& Gould, A.\ 2000, \apj, 545, 847 \bibitem[Morris (1993)]{1993ApJ...408..496M} Morris, M.\ 1993, \apj, 408, 496 \bibitem[Mouawad et al. (2005)]{2005AN....326...83M} Mouawad, N., Eckart, A., Pfalzner, S., Sch{\"o}del, R., Moultaka, J., \& Spurzem, R.\ 2005, Astronomische Nachrichten, 326, 83 \bibitem[Nikiforov, I. I. et al. (2008)]{} Nikiforov, I. I. 2008, Proceedings of the International Conference on ``Dynamics of Galaxies," in press (arXiv:0803.0825) \bibitem[Olling \& Merrifield (2000)]{2000MNRAS.311..361O} Olling, R.~P., \& Merrifield, M.~R.\ 2000, \mnras, 311, 361 \bibitem[Olling \& Merrifield (2001)]{2001MNRAS.326..164O} Olling, R.~P., \& Merrifield, M.~R.\ 2001, \mnras, 326, 164 \bibitem[Rafelski et al. (2007)]{2007ApJ...659.1241R} Rafelski, M., Ghez, A.~M., Hornstein, S.~D., Lu, J.~R., \& Morris, M.\ 2007, \apj, 659, 1241 \bibitem[Rauch \& Tremaine (1996)]{1996NewA....1..149R} Rauch, K.~P., \& Tremaine, S.\ 1996, New Astronomy, 1, 149 \bibitem[Reid (1993)]{1993ARA&A..31..345R} Reid, M.~J.\ 1993, \araa, 31, 345 \bibitem[Reid \& Brunthaler (2004)]{2004ApJ...616..872R} Reid, M.~J., \& Brunthaler, A.\ 2004, \apj, 616, 872 \bibitem[Reid et al.(2003)]{2003ApJ...587..208R} Reid, M.~J., Menten, K.~M., Genzel, R., Ott, T., Sch{\"o}del, R., \& Eckart, A.\ 2003, \apj, 587, 208 \bibitem[Rubilar \& Eckart (2001)]{2001A&A...374...95R} Rubilar, G.~F., \& Eckart, A.\ 2001, \aap, 374, 95 \bibitem[Salim \& Gould (1999)]{1999ApJ...523..633S} Salim, S., \& Gould, A.\ 1999, \apj, 523, 633 \bibitem[Sch\"odel et al. (2002)]{bib31} Sch\"odel, R. et al.\ 2002, \nat, 419, 694 \bibitem[Sch\"odel et al. (2003)]{} Sch\"odel, R., Ott, T., Genzel, R., Eckart, A., Mouawad, N., \& Alexander, T. 2003, \apj, 596, 1015 \bibitem[Sch{\"o}del et al.(2007)]{2007A&A...469..125S} Sch{\"o}del, R., et al.\ 2007, \aap, 469, 125 \bibitem[Scoville et al. 2003]{} Scoville, N. Z., Stolovy, S. R., Rieke, M., Christopher, M. H., Yusef-Zadeh, F. 2003, \apj, 594, 294 \bibitem[Tagger \& Melia (2006)]{2006ApJ...636L..33T} Tagger, M., \& Melia, F.\ 2006, \apjl, 636, L33 \bibitem[Tremaine et al. (2002)]{2002ApJ...574..740T} Tremaine, S., et al.\ 2002, \apj, 574, 740 \bibitem[Ullio et al. (2001)]{2001PhRvD..64d3504U} Ullio, P., Zhao, H., \& Kamionkowski, M.\ 2001, \prd, 64, 043504 \bibitem[van Dam et al. (2006)]{} van Dam, M. A. et al. 2006, \pasp, 118, 310 \bibitem[Weinberg (1972)]{} Weinberg, S. 1972 Gravitation and Cosmology: Principles and Applications of the General Theory of Relativity (New York: Wiley) \bibitem[Weinberg et al. (2005)]{2005ApJ...622..878W} Weinberg, N.~N., Milosavljevi{\'c}, M., \& Ghez, A.~M.\ 2005, \apj, 622, 878 \bibitem[Will(2008)]{2008ApJ...674L..25W} Will, C.~M.\ 2008, \apjl, 674, L25 \bibitem[Taylor \& Weisberg (1989)]{1989ApJ...345..434T} Taylor, J.~H., \& Weisberg, J.~M.\ 1989, \apj, 345, 434 \bibitem[Wizinowich et al. (2000)]{wizin00} Wizinowich, P.~L., Acton, D.~S., Lai, O., Gathright, J., Lupton, W., \& Stomski, P.~J.\ 2000, \procspie, 4007, 2 \bibitem[Wizinowich et al. (2006)]{} Wizinowich, P. L. et al. 2006, \pasp, 118, 297 \bibitem[Zakharov et al. (2007)]{2007PhRvD..76f2001Z} Zakharov, A.~F., Nucita, A.~A., de Paolis, F., \& Ingrosso, G.\ 2007, \prd, 76, 062001 \bibitem[Zucker \& Alexander (2007)]{2007ApJ...654L..83Z} Zucker, S., \& Alexander, T.\ 2007, \apjl, 654, L83 \end{thebibliography} \clearpage % % python code % from papers import r0 % r0.plotTableObs() % \input{tab1.tex} \clearpage \input{tab2.tex} \clearpage %\input{tab_data.tex} \input{tab3.tex} \clearpage \input{tab4.tex} \clearpage %\input{tab_orb.tex} \input{tab5.tex} \clearpage %\input{r0_masers.tex} \input{tab6.tex} \clearpage %\input{r0_nirc2_astrometry.tex} \input{tab7.tex} \clearpage \begin{figure} \epsscale{1.0} %\plotone{/u/ghez/doc/paper/Ro/figures/image_central_blowup.eps} \plotone{f1.eps} \figcaption{ A comparison of raw images obtained with LGSAO and speckle imaging with the Keck 10 m telescopes. The large scale image is an LGSAO image obtained in 2005. The inset LGSAO image (top right) and speckle image (bottom right) are centered on the black hole, SgrA* (marked with a cross), with a field of view of 1\farcs 0 $\times$ 1\farcs 0, also obtained in 2005. The image quality, depth, and astrometric precision have all been greatly improved with the advent of LGSAO. \label{fig_aoVsp} } \end{figure} % % python code % from papers import r0 % r0.plotTableObs() % makes speck_ao_klimit.eps % \begin{figure} \epsscale{1.0} \plotone{f2.eps} \figcaption{ Comparison of the sensitivity of the average images from each epoch. The recent LGSAO images, with significantly longer on-sky integration times (t$_{tot}$ $\sim$ 50 vs.\ 3 min) and much higher strehl ratios, are three magnitudes more sensitive than any of the speckle images. \label{speck_ao_klimit} } \end{figure} \begin{figure} \epsscale{1.0} \plotone{f3.eps} \figcaption{ Comparison of the centroid uncertainties as a function of brightness. Because the very brightest stars (K$\sim$9) are saturated in their cores in the LGSAO images, there is a slight rise in their centroid uncertainties compared to somewhat fainter sources. Overall, however, for bright sources (K$<$13), the long exposure LGSAO images achieve a centroiding uncertainty of just 0.17 mas, a factor of $\sim$6 better than the earlier work done with speckle imaging. \label{speck_ao_astro} } \end{figure} \begin{figure} \epsscale{1.0} \plotone{f4.ps} \figcaption{ Astrometric bias introduced by an unresolved source in the case of a binary star generated and analyzed with a known PSF. Two cases are shown: (solid line) PSF from LGSAO image in 2006 May and (dotted line) PSF from speckle image in 1998 July. The contour lines show the amount of bias (in mas) introduced by an underlying source of the indicated flux ratio and separation. Once the neighboring source is detected, which happens at separations of $\sim$60 mas, the astrometric bias drops to zero in this idealized case. For 1:1 binaries, pairs with smaller separations can be resolved. This figure shows that biases well above the positional uncertainties ($\sim$1 mas) can occur due to underlying sources. \label{bias_binary} } \end{figure} \begin{figure} \epsscale{1.0} %\plotone{/u/ghez/doc/paper/Ro/figures/s02_phot.ps} \plotone{f5.ps} \figcaption{ Photometric measurements vs. time for S0-2 (top) and S0-16 (bottom). Measurements that were made when these sources coincided with another known source are plotted as unfilled points and excluded from the model fitting procedure. S0-16 is more affected by underlying sources, because it is fainter. Even without {\it a priori} knowledge of the underlying sources, their effect is clearly visible in photometric measurements made in 1998 and 2002, for S0-2, and 1996-1999 \& 2000, for S0-16. \label{s02_phot} } \end{figure} \begin{figure} \epsscale{1.0} %\plotone{/u/ghez/doc/paper/Ro/figures/S0-2_total_ave.eps} \plotone{f6.eps} \figcaption{The weighted average of all S0-2 spectra obtained with the W. M. Keck II telescope. Since only some of the data sets contain the shorter wavelengths, the signal to noise ratio is lower at wavelengths shortward of 2.13 $\mu$m. While \brg and He I lines are clearly detected, only the \brg line, which is stronger and not the blend of multiple lines, is used to measure the radial velocity of S0-2 as a function of time. \label{spec_avg} } \end{figure} \begin{figure} \epsscale{0.8} %\plotone{/u/ghez/doc/paper/Ro/figures/S0-2_timeline_v2.ps} \plotone{f7.ps} \figcaption{ Measurements of S0-2's Br$\gamma$ line . These three measurements show that, over time, S0-2's radial velocity has changed by more than 2,600 km s$^{-1}$. In order to improve the line detection in the low SNR NIRSPEC observation, the emission from the local gas was not removed, which leaves a large Br$\gamma$ emission feature centered at small radial velocities compared to that of the star at this time. With improvements in the adaptive optics system and instrumentation (from NIRSPEC/NGSAO [bottom], to NIRC2/NGSAO [middle] and finally to OSIRIS/LGSAO [top]), the precision with which the Br$\gamma$ absoprtion line can be measured in S0-2 has improved by a factor of 5. \label{spec_brg} } \end{figure} \begin{figure} \epsscale{1.0} %\plotone{/u/ghez/doc/paper/Ro/figures/bhComparePos.eps} \plotone{f8.eps} \figcaption{ Comparison of estimates of the black hole's location. Colored contours represent the estimates of the dynamical center from model fits to kinematic measurements of S0-2 (K = 14.0; blue) and S0-16 (K = 15.0; red). Black contours show the SgrA*-Radio position. All contours are plotted at the 68\%, 95\%, and 99.7\% confidence levels (equivalent to 1, 2, and 3$\sigma$ for a Gaussian distribution). The solid black points are all the measurements of SgrA*-IR (K$\sim$16) in the maps used for the astrometric analysis. The discrepancy in the black hole's location from S0-16's positional measurements appear to be a consequence of biases from unrecognized, underlying stars and thus only S0-2's measurements are used to infer the properties of the central black hole. Likewise, the astrometric positions of SgrA*-IR, which is even fainter than S0-16, also may be biased (see discussion in \S4.1) and are therefore not used to constrain the orbital model used to fit S0-2. \label{bhPosition} } \end{figure} % % sm % /net/uni/Groups/ghez/ghez/analysis/Ro/08_02_16/plots % s02.sm % \begin{figure} \epsscale{0.8} %\plotone{/u/ghez/doc/paper/Ro/figures/s02.ps} \plotone{f9.ps} \figcaption{ The best fit to the astrometric and radial velocity data, assuming a Keplerian orbital model. The filled points were included in the formal fit, while the unfilled points are measurements that are excluded due to source confusion. Uncertainties are plotted on all points, except the unfilled/excluded points (here the uncertainties are comparable to the size of the points) for clarity. ({\it Left}) To compare with what has been done in the past to estimate R$_0$, we show the fit to the data with a 10 parameter model, which includes the black hole's mass (M$_{bh}$), distance (R$_0$), and location in the plane of the sky ($X_0, Y_0$) as free parameters, and which fixes the black hole's three dimensional velocity ($V_x, V_y, V_z$) to zero. This results in a $\chi^2/dof \sim 1.4$. ({\it Right}) The data are better reproduced by a 13 parameter model, which includes the black hole's mass (M$_{bh}$), distance (R$_0$), location in the plane of the sky ($X_0, Y_0$), and three dimensional velocity ($V_x, V_y, V_z$) as free parameters, and results in a $\chi^2/dof \sim 0.97$. Adding these extra free parameters, and in particular $V_z$, increases the uncertainties in the black hole's properties by a factor of two. \label{s02_orb} } % % sm % /net/uni/Groups/ghez/ghez/analysis/Ro/08_02_16/plots % s02_resid.sm % \end{figure} \begin{figure} \epsscale{0.8} %\plotone{/u/ghez/doc/paper/Ro/figures/s02_resid.ps} \plotone{f10.ps} \figcaption{ The residuals to the best fit Keplerian orbital models shown in Figure 9. The filled points were included in the formal fit, while the unfilled points are measurements that are excluded due to source confusion. \label{s02_orb_resid} } \end{figure} \begin{figure} \epsscale{1.0} %\plotone{/u/ghez/doc/paper/Ro/figures/massRo.08_02_16.vel3d.1e2.eps} \plotone{f11.eps} \figcaption{ The correlation of the estimated black hole's mass and distance. The density of solutions from the MC simulations are shown as a color image, with the contours marking the 68\%, 95\%, and 99.7\% confidence limits. While mass and distance are well determined from the orbit of S0-2, they are not independent quantities. The exact scaling depends on the relative impact of the astrometric and radial data on the model fits. Currently, the inferred mass scales with the inferred distance as M $\propto$ R$_0$$^{1.8}$. \label{massRo} } \end{figure} \begin{figure} \epsscale{1.0} %\plotone{/u/ghez/doc/paper/Ro/figures/RoVz.eps} \plotone{f12.eps} \figcaption{ Correlation of the estimated black hole's distance and line-of-sight velocity ($V_z$) from our 13 parameter model fit. $V_z$ dominates the uncertainties in R$_0$ and consequently M$_{bh}$. Priors on $V_z$ can reduce the uncertainties in R$_0$ by a factor of two. All previous studies have set $V_z$ to zero, which implicitly assumes that there are no massive companions to our Galaxy's central supermassive black hole and that the local standard of rest is perfectly known. \label{RoVz} } \end{figure} \begin{figure} \epsscale{1.0} %\plotone{/u/ghez/doc/paper/Ro/draft/mass_extended.ps} \plotone{f13.ps} \figcaption{ Limits on the amount of mass in an extended distribution contained within S0-2's apoapse distance. The three lines correspond to the 68.3\%, 95.4\%, and 99.7\% upper-bound confidence limits. The 99.7\% confidence upper-bound of 3-4 $\times$ 10$^5$ M$_{\odot}$ is a fairly weak function of the slope of the assumed power-law mass profile. Simple models of the stellar distribution suggest M$_{ext}(< 0.01 \textrm{pc}) \sim 10^3 M_{\odot}$, a factor of $\sim$100 smaller than the current measurement uncertainty. \label{extended} } \end{figure} % Figure for app_absoluteAstro % Uses data in: % 08_03_26_all_test/ % 08_03_26_all_test_noyng % in ipython (no -pylab flag): % from gcwork import astrometry as ast % ast.plotTransVsTime('align/align_d_rms_1000) %\begin{figure} %\epsscale{1.0} %\plottwo{../figures/trans_stability_yng.eps}{../figures/trans_stability_noyng.eps} %\caption{ %The plate scale {\it (top)} and the position angle {\it (bottom)} %over time for all data sets aligned using a set of stars that %includes {\it (left)} and excludes {\it (right)} the known young stars. %The plate scale is relative to the %plate scale in the reference epoch of 2004 July LGS. The position %angle is the absolute position angle offset from the value reported %in the NIRC and NIRC2 instrument headers. The plate scale appears to %be stable over multiple years for both NIRC ({\it red}) and NIRC2 ({\it blue}) %once the young stars are excluded. %NIRC shows several systematic jumps in the position angle relative to %the value reported in the image headers, which is most likely a result %of instrument or telescope changes. %} %\label{fig:trans_stability} %\end{figure} \begin{figure} \plottwo{f14a.eps}{f14b.eps} \figcaption{ The plate scale {\it (top)} and the position angle {\it (bottom)} over time for all data sets aligned using a set of stars that excludes {\it (left)} and includes {\it (right)} the known young stars. The plate scale is relative to the plate scale in the reference epoch of 2004 July LGS. The position angle is the absolute position angle offset from the value reported in the NIRC and NIRC2 instrument headers. Once the young stars are excluded, the estimated plate scales for NIRC ({\it squares}) and NIRC2 ({\it diamonds}) are very stable, approximately 0.05\% and 0.03\% (rms), respectively, over multiple years. NIRC shows several systematic jumps in the position angle relative to the value reported in the image headers, which is most likely a result of instrument or telescope changes. \label{trans} } \end{figure} \begin{figure} \epsscale{1.0} %\plotone{/u/ghez/doc/paper/Ro/figures/imageMasers.eps} \plotone{f15.eps} \figcaption{ Infrared mosaic measuring the positions of the SiO masers. The 7 masers, whose radio positions are well measured by Reid et al. 2007 and which used to establish an absolute reference frame, are circled. Dotted lines depict the outline of example LGSAO (green) and speckle (blue) images in which the short period stars are measured and placed in the cluster reference frame. Since the masers sparsely sample the area of interest, only low order polynomials are used to calibrate the cluster reference frame (i.e., pixel scale, orientation, and position of SgrA*-Radio). \label{fig:masers} } \end{figure} \end{document}