1: %\documentclass[apj]{emulateapj}
2: \documentclass[12pt,preprint]{aastex}
3: %\usepackage{natbib}
4: \bibliographystyle{apj}
5: %\usepackage{graphicx}
6:
7: \newcommand{\msun}{M$_{\odot}$\,}
8: \newcommand{\Ms}{M$_s$}
9: \newcommand{\Lp}{L'}
10: \newcommand{\Kp}{K'}
11: \newcommand{\Hp}{H}
12:
13: % FIGSET-MACROS-BEGIN
14: \newcommand{\noprint}[1]{}
15: \newcommand{\figsetstart}{{\bf Fig. Set} }
16: \newcommand{\figsetend}{}
17: \newcommand{\figsetgrpstart}{}
18: \newcommand{\figsetgrpend}{}
19: \newcommand{\figsetnum}[1]{{\bf #1.}}
20: \newcommand{\figsettitle}[1]{ {\bf #1} }
21: \newcommand{\figsetgrpnum}[1]{\noprint{#1}}
22: \newcommand{\figsetgrptitle}[1]{\noprint{#1}}
23: \newcommand{\figsetplot}[1]{\noprint{#1}}
24: \newcommand{\figsetgrpnote}[1]{\noprint{#1}}
25: % FIGSET-MACROS-END
26:
27: \shorttitle{A Disk of Young Stars at the Galactic Center}
28: \shortauthors{Lu et al.}
29:
30: \begin{document}
31:
32: \title{A Disk of Young Stars at the Galactic Center as Determined by
33: Individual Stellar Orbits}
34:
35: \author{
36: J. R. Lu\altaffilmark{1},
37: A. M. Ghez\altaffilmark{1,2},
38: S. D. Hornstein\altaffilmark{1,3},
39: M. R. Morris\altaffilmark{1},
40: E. E. Becklin\altaffilmark{1},
41: K. Matthews\altaffilmark{4}
42: }
43: \email{jlu@astro.ucla.edu, ghez@astro.ucla.edu, seth.hornstein@colorado.edu,
44: morris@astro.ucla.edu, becklin@astro.ucla.edu, kym@caltech.edu}
45: \altaffiltext{1}{UCLA Department of Physics and Astronomy,
46: Los Angeles, CA 90095-1562}
47: \altaffiltext{2}{UCLA Institute of Geophysics and Planetary Physics,
48: Los Angeles, CA 90095-1565}
49: \altaffiltext{3}{Center for Astrophysics \& Space Astronomy,
50: Department of Astrophysical and Planetary Sciences,
51: University of Colorado, Boulder, CO 80309}
52: \altaffiltext{4}{Caltech Optical Observatories, California Institute of
53: Technology, MS 320-47, Pasadena, CA 91125}
54:
55: \begin{abstract}
56: % Limit is 250 words in the abstract
57: We present new proper motions from the 10 m Keck telescopes
58: for a puzzling population of massive, young stars located within 3\farcs5
59: (0.14 pc) of the supermassive black hole at the Galactic Center.
60: Our proper motion measurements have uncertainties of only
61: 0.07 mas yr$^{-1}$ (3 km s$^{-1}$), which is
62: $\gtrsim$ 7 times better than previous
63: proper motion measurements for these stars, and
64: enables us to measure accelerations as low as 0.2 mas yr$^{-2}$
65: (7 km s$^{-1}$ yr$^{-1}$).
66: Using these measurements, line-of-sight velocities from the literature,
67: and 3D velocities for additional young stars in the central parsec, we
68: constrain the true orbit of each individual star
69: and directly test the hypothesis that the massive stars
70: reside in two stellar disks as has been previously proposed.
71: Analysis of the stellar orbits reveals only one of the previously
72: proposed disks of young stars
73: using a method that is capable of detecting disks containing at least 7 stars.
74: The detected disk contains 50\% of the young stars, is inclined
75: by $\sim115^\circ$ from the plane of the sky, and is oriented at
76: a position angle of $\sim100^\circ$ East of North.
77: Additionally, the on-disk and off-disk populations have similar K-band
78: luminosity functions and radial distributions that decrease at larger
79: projected radii as $\propto r^{-2}$.
80: The disk has an out-of-the-disk velocity dispersion
81: of 28 $\pm$ 6 km s$^{-1}$, which corresponds to a half-opening angle of
82: $7^\circ \pm 2^\circ$, and several candidate disk members have
83: eccentricities greater than 0.2.
84: Our findings suggest that the young stars may have formed {\it in situ}
85: but in a more complex geometry than a simple, thin circular disk.
86: \end{abstract}
87:
88: \keywords{black hole physics -- Galaxy:center --
89: infrared:stars -- techniques:high angular resolution}
90:
91: \section{Introduction}
92: \label{sec:intro}
93:
94: The center of our Galaxy harbors not only a supermassive black hole
95: \citep[Sgr A*, $M_\bullet \sim 4 \times 10^6$ \msun;][]{
96: eckart96,genzel96,ghez98pm,ghez00nat,ghez03spec,ghez05orbits,
97: schodel02,schodel03,eisenhauer06}, but also
98: a population of massive (10-120 \msun), young ($\lesssim$10-100 Myr) stars
99: whose existence is a puzzle.
100: The origin of such young stars has been difficult to explain since
101: the gas densities observed today are orders of magnitude too low
102: for a gas clump to overcome the extreme tidal forces and
103: collapse to form stars
104: \citep[e.g.][for reviews]{sanders92,morris93,ghez05orbits,alexander05review}.
105: And yet, within the central parsec of our Galaxy, nearly 100 stars have
106: been classified as OB main-sequence stars,
107: more luminous OB giants and supergiants, and post-main-sequence
108: Wolf-Rayet stars \citep{allen90,krabbe91,blum95heI,krabbe95,tamblyn96,
109: najarro97,ghez03spec,paumard06}, with the more evolved massive stars having
110: ages as young as 6$\pm$2 Myr \citep{paumard06}.
111: Populations of young stars have also been observed in the nuclei of other
112: galaxies, such as M31 \citep{bender05}, suggesting that star formation
113: near a supermassive black hole may be a common, but not understood,
114: phenomenon in galaxy evolution. The close proximity
115: of the black hole at the center of the Milky Way provides a unique laboratory
116: for studying this ''paradox of youth''
117: \citep[e.g.][]{ghez03spec,ghez05orbits,schodel03,eisenhauer06}.
118:
119: Proposed resolutions to the paradox of youth
120: can be grouped into several broad categories, including
121: (1) rejuvenation of an older population such that older stars appear young,
122: (2) dynamical migration from larger radii, and
123: (3) {\it in situ} formation.
124: Rejuvenation scenarios include stripping \citep{davies98,davies05}
125: or tidal heating of the atmospheres of old stars \citep{alexander03},
126: or combining multiple low mass stars via collisional mergers
127: to form a higher-mass hot star akin to a ``blue straggler''
128: \citep{lee96gcmergers,morris93,genzel03cusp}.
129: Although these processes may be candidates for explaining the
130: closest young stars within the central arcsecond,
131: they cannot account for the OB giants, OB supergiants, and
132: Wolf-Rayet stars that are located at larger radii (1\arcsec-14\arcsec),
133: since the rate of collisions is too low to produce
134: the observed total numbers.
135: Thus, it appears that these massive young stars must have formed,
136: or were deposited, in the central region within the last 4-8 Myr.
137: Dynamical migration scenarios attempt to resolve the paradox of youth
138: with the formation of a massive star cluster at larger distances from
139: the black hole (3-30 pc).
140: Such a cluster would spiral in due to dynamical friction
141: and deposit stars at smaller radii where they
142: are observed today \citep{gerhard01}. However,
143: for a cluster to reach the central parsec in only a few million
144: years, it must be very massive and centrally concentrated
145: \citep{kim03,pzwart03irs16,mcmillan03,gurkan05},
146: and it may even require the existence of an intermediate-mass
147: black hole (IMBH) as an anchor in the cluster core \citep{hansen03,kim04}.
148: {\it In situ} star formation scenarios can resolve the paradox
149: of youth if a massive, self-gravitating gas disk
150: was once present around the black hole \citep{levin03}.
151: Such a disk would be sufficiently dense to overcome the strong tidal
152: forces, and gravitational instabilities would then lead to
153: fragmentation and the formation of stars,
154: as has been suggested in the context of both the Galactic Center
155: circumnuclear disk and AGN accretion disks in other galaxies
156: \citep[e.g.][]{kolykhalov80,shlosman89,morris96,sanders98,goodman03,nayakshinCuadra05}.
157:
158: Insight into the origins of the massive, young stars may be obtained
159: through observations of the spatial distribution and stellar dynamics of
160: this population.
161: Already, high-resolution infrared imaging and spectroscopy have shown
162: that the young stars between 0\farcs5 and 14\arcsec (0.02-0.6 pc)
163: exhibit coherent rotation \citep{genzel00}.
164: Analyses of the statistical properties of the
165: three-dimensional velocity vectors for these stars
166: suggest that they may reside in two disks.
167: The first proposed disk has a clockwise sense of rotation, as
168: projected onto the plane of the sky
169: \citep[][hereafter: clockwise-rotating or CW disk]{levin03},
170: while the second proposed disk is
171: counter-clockwise-rotating \citep[CCW][]{genzel03cusp} and is nearly
172: perpendicular to the first. The proposed disks extend
173: from $\sim$0\farcs8 to at least 7\arcsec \citep{paumard06}.
174: Other velocity vector analyses show that there are possible
175: co-moving groups or clusters of stars, including the IRS 13
176: cluster, which is proposed to lie within the putative CCW
177: disk \citep{maillard04irs13,schodel05}, and the IRS 16SW co-moving
178: group, which are also consistent with the proposed CW disk \citep{lu05irs16sw}.
179: The two proposed disks are inferred
180: to be oriented with an inclination and angle to the ascending node of
181: [$i_{CW}$=127$^\circ \pm$ 2$^\circ$, $\Omega_{CW}$=99$^\circ \pm$ 2$^\circ$] and
182: [$i_{CCW}$=24$^\circ \pm$ 4$^\circ$, $\Omega_{CCW}$=167$^\circ \pm$ 7$^\circ$] and
183: to have a finite angular thickness of
184: $\Delta\theta_{CW} \sim 14^\circ$ and
185: $\Delta\theta_{CW} \sim 19^\circ$ where $\Delta\theta$ is the
186: standard deviation of the orbital inclinations distributed
187: normally about the disk plane \citep{paumard06}.
188: The thickness of the stellar disks has been attributed to
189: thickening as a result of gravitational interactions between the two disks,
190: which provides an estimate of the disk masses \citep{nayakshin06thick}.
191: The derived mass is smaller than the mass inferred from the number of
192: observed young stars, assuming a Salpeter initial mass function (IMF);
193: accordingly, \citet{nayakshin06thick}
194: suggest that the disks have a top-heavy mass function.
195: Both {\it in situ} gas disk and in-spiraling star cluster
196: formation scenarios have been used to explain
197: the kinematics of this young star population and to
198: predict that the stars should lie in a common orbital plane.
199: However, the presence of two stellar disks with similarly aged
200: populations requires either two nearly concurrent gas disks or two
201: infalling star clusters; and both of these scenarios are difficult to produce.
202: Therefore, to understand the recent star formation history,
203: it is critical to measure the orbital planes of individual
204: stars in order to confirm the existence of the two stellar disks previously
205: derived from a statistical analysis of velocity vectors alone.
206:
207: The {\it in situ} gas disk and inspiraling star cluster
208: formation scenarios predict different structures and evolutions
209: for the resulting stellar disk, particularly with respect to the
210: eccentricities and radial distribution of stars within the disk.
211: Early models of a self-gravitating gas disk around the supermassive
212: black hole at the center of the Milky Way produce stars with
213: a steep radial profile in the disk surface density,
214: $\Sigma \propto r^{\alpha}$, with $\alpha \sim -2$ \citep{linPringle87,levin06}.
215: These models typically result in stars on
216: circular orbits as would be the case for the slow build up of a
217: gas disk that is circularized before there is sufficient mass for
218: gravitational instabilities to set in
219: \citep{milosav04,nayakshinCuadra05,levin06}.
220: The stellar eccentricities of an initially circular disk can relax
221: to higher eccentricities up to $e_{rms} = \sqrt{<e^2>} \sim $0.15 for a normal
222: IMF or $e_{rms} \sim $0.3 for a top-heavy IMF \citep{alexander07imf,cuadra08}.
223: More recent models have also shown that star formation can occur rapidly
224: before circularization in an initially eccentric
225: disk as might result from the infall of a single massive molecular cloud
226: or a cloud-cloud collision \citep{sanders98,nayakshin07sims,alexander08}.
227: These eccentric self-gravitating accretion disk models typically produce
228: a more top-heavy IMF than initially circular disks.
229: On the other hand, an inspiraling star cluster would dissolve into a
230: disk of stars with a flatter radial profile
231: \citep[$\Sigma \propto r^{-0.75}$;][]{berukoff06}
232: whose orbital eccentricities would reflect the
233: eccentricity of the cluster's orbit, which could be either circular
234: or eccentric \citep{pzwart03irs16,mcmillan03,kim03,kim04,gurkan05,berukoff06}.
235: Previous measurements of the radial distribution of young stars
236: yields a steep radial profile consistent with {\it in situ} formation
237: \citep{paumard06}.
238: Also, the eccentricities of the stars have previously been estimated from
239: observations by assuming that the stars orbit in a disk;
240: however, there are conflicting
241: results claiming that the stars in the clockwise-rotating disk are on
242: nearly circular orbits \citep{paumard06} or on eccentric orbits
243: \citep{beloborodov06}.
244: Determining the radial profile and stellar eccentricities of stars in a
245: disk may provide observational constraints on the origin of the young stars.
246:
247: We present an improved proper motion study that yields an order
248: of magnitude more precise proper motions and
249: the first measurement of accelerations in the
250: plane of the sky for stars outside the central arcsecond.
251: By combining the stellar positions, proper motions, radial velocities,
252: and accelerations, we estimate stellar orbital parameters
253: and test whether the young stars reside on one or two stellar disks in a more
254: direct manner than previous methods using only velocity information.
255: This provides a {\it direct} test of the existence, membership, and properties
256: of these disks. The observations are described in
257: \S\ref{sec:obs} and the astrometric analysis procedure and results are detailed
258: in \S\ref{sec:astrometry}. Orbit analysis and results are presented in
259: \S\ref{sec:orbitAnalysis} and \S\ref{sec:orbitResults} and a discussion
260: of the implications for the origin of the massive, young stars
261: at the Galactic Center is presented in \S\ref{sec:discussion}.
262:
263: \section{Observations}
264: \label{sec:obs}
265:
266: This study utilizes 29 epochs of high-resolution, infrared images of the
267: Galaxy's central stellar cluster, which were taken from 1995 to 2005
268: using both speckle and laser guide star adaptive optics (LGS AO) observing
269: techniques on the W.~M. Keck 10 m telescopes.
270: These data sets are listed in Table \ref{tab:obs} and all but the
271: additional LGS AO observation from 2005 are described in detail in earlier
272: papers \citep{ghez98pm,ghez00nat,ghez05orbits,lu05irs16sw,rafelski07}.
273: Columns 3 and 4 list the individual exposure times and the
274: total number of frames for each epoch of data.
275: All 27 speckle imaging observations were taken using the
276: facility near-infrared camera, NIRC \citep{NIRC,NIRCs}, which has a
277: plate scale of $\sim$20 mas per pixel, and
278: a 5\farcs22 $\times$ 5\farcs22 field of view.
279: The two adaptive optics imaging observations used the facility LGS AO
280: system \citep{wizinowich06,vanDam06} and the near-infrared camera,
281: NIRC2 (PI: K. Matthews) with a plate scale of
282: 9.963 $\pm$ 0.006 mas per pixel \citep{ghez08}
283: and a 10\farcs2 $\times$ 10\farcs2 field of view.
284: While the laser guide star is used to correct most of the
285: atmospheric aberrations, the low-order, tip-tilt terms were corrected
286: using visible observations of USNO 0600-28577051 (R = 13.7 mag and
287: $\Delta r_{SgrA*}$ = 19\arcsec).
288:
289: In addition to the 27 speckle observation and the 2004 LGS AO observations
290: described in previous works, a new LGS AO data set was obtained in
291: 2005 June. This data set was taken using two different
292: narrow-band filters,
293: K$_{CO}$ ($\lambda_{o}$=2.289 \micron, $\Delta\lambda$=0.027 \micron) and
294: K$_{cont}$ ($\lambda_{o}$=2.270 \micron, $\Delta\lambda$=0.030 \micron),
295: rather than the K' broadband filter used for the 2004 LGS AO observations.
296: For each filter, images were taken in a 5 position pattern around a
297: 4\farcs0 box with exposure times of 36 s
298: (t$_{exp}$ = 7.2 s, 5 coadds) and 59.5 s (t$_{exp}$ = 11.9 s, 5 coadds)
299: for the K$_{CO}$ and K$_{cont}$ filters, respectively.
300: The choice of narrow-band filters was driven by a different project
301: and the data sets from the two filters were combined together
302: for the present study (see \S\ref{sec:images}).
303: Resulting Strehl ratios were $\sim$0.25-0.35 in the individual frames.
304:
305:
306: \section{Astrometric Data Analysis and Results}
307: \label{sec:astrometry}
308:
309: The goal of this analysis is to obtain high precision astrometry for
310: a sample of young stars that are candidate disk members and
311: have existing radial velocity measurements.
312: Based on spectroscopic identification, there are currently 90 known
313: young stars with radial velocity measurements listed in \citet{paumard06}
314: based on high quality (``quality 1 or 2'') spectral classifications.
315: We define a {\it primary sample} that includes those known young stars found
316: in our astrometric data sets that
317: have projected radii between 0\farcs8 and 3\farcs5. The inner
318: radius is set by the proposed inner edge of the clockwise disk of young
319: stars and young stars interior to this radius are on more randomly
320: oriented orbits \citep{ghez05orbits,eisenhauer06}.
321: The outer radius is set by the field of view of the speckle data sets.
322: Over this region, \citet{paumard06} note that all young
323: stars brighter than K=13.5 should be identified, which includes
324: OB giants and supergiants.
325: A total of 32 such young stars are in our 11 year
326: astrometric data set and comprise the sample for this study. Of the
327: 32 stars in our sample, 23 are among the 36 stars thought to be part
328: of the clockwise disk, 2 are among the 12 candidate members of the
329: counter-clockwise disk, and the remaining 7 are among the 42
330: stars not assigned to either disk by \citet{paumard06}.
331:
332: We also define an {\it extended sample} that includes both the primary
333: sample of 32 stars and an additional 41 young stars found by
334: \citet{paumard06} at larger radii that are outside the field of view
335: of our astrometric measurements. The astrometry for the additional
336: 41 stars is taken from \citet{paumard06}\footnote{We note that there
337: are 4 additional young stars at larger radii that are not included
338: in our extended sample since they do not have proper motions
339: listed in \citet{paumard06}.},
340: which has an order of magnitude lower precision and lacks any constraints
341: on the accelerations.
342: However, we use the extended sample to explore the kinematics of the
343: young stars at larger radii with the same analysis techniques used
344: on the primary sample. We also note that the spectroscopic observations
345: used to identify the young stars at larger radii were taken in a different
346: setup than in the central regions, with lower spectral resolution
347: and lower Strehl; thus the
348: completeness limit may be somewhat brighter in this region.
349: However, any difference is statistically insignificant given that
350: a two-sample KS-test yields a 50\% probability that
351: the primary sample and those additional stars added to the extended
352: sample have the same K-band luminosity function. The extended sample
353: is used only to supplement our analysis; therefore, to avoid confusion,
354: all analysis and results are reported for the
355: primary sample, which has more precise proper motions and accelerations,
356: unless specifically noted otherwise.
357:
358: Astrometric positions for the young stars in the primary sample
359: are extracted from the imaging
360: data sets listed in Table \ref{tab:obs} using similar techniques to those
361: described in \citet{ghez98pm,ghez00nat}, \citet{lu05irs16sw},
362: and \citet{ghez05orbits}, with the
363: following key changes:
364: (1) geometric distortion is corrected in the speckle images using
365: an improved distortion solution (see \S\ref{sec:images},
366: Appendix \ref{app:speckDistort}),
367: (2) speckle images are combined with an improved algorithm developed
368: and implemented by \citet{sethThesis}, and
369: (3) image coordinates are transformed between data sets with more
370: degrees of freedom (see \S\ref{sec:coords}).
371: Sections \ref{sec:images} and \ref{sec:coords} describe the analysis in
372: detail and Section \ref{sec:astroResults} presents the astrometric results.
373:
374: \subsection{Image Processing}
375: \label{sec:images}
376:
377: To achieve precise astrometry, the basic image reduction steps,
378: particularly geometric distortion correction, must be carefully implemented.
379: First, both speckle and LGS AO individual exposures are processed using
380: standard techniques of sky subtraction, flat-fielding, and bad pixel correction.
381: Next, the images are transformed to correct for optical distortion.
382: For the LGS AO/NIRC2 images, optical distortions are well characterized at the
383: $\sim$2 milli-arcsecond level over 2''
384: \citep[][Appendix \ref{app:speckDistort}]{ghez08}
385: by the pre-ship review distortion
386: coefficients\footnote{http://www2.keck.hawaii.edu/inst/nirc2/}
387: and the distortions are removed from the images using the IRAF routine,
388: {\it Drizzle} \citep{drizzle}.
389: The speckle images, obtained with NIRC, have a known off-axis
390: distortion that can be corrected as described in \citet{ghez98pm}.
391: However, this distortion solution
392: does not account for any distortion introduced by the
393: additional optics in the NIRC reimager, which magnifies the image scale
394: by a factor of $\sim$7 from
395: seeing limited sampling to diffraction limited sampling.
396: Speckle data sets were acquired in such a way as to minimize
397: the effects of this residual distortion in the center of the field
398: of view and have resulting residual distortion errors that are
399: smaller than the typical centroiding error, which is $\sim$2 mas,
400: for stars at radii $<$ 0\farcs5.
401: However, astrometric uncertainties for stars outside this region
402: are dominated by the uncorrected distortion, which grows to
403: $\sim$6 mas near the field edge at a radius of 2\farcs5 \citep{ghez05orbits}.
404: In order to characterize the residual distortion in NIRC, simultaneous images
405: of the Galactic Center were obtained with both NIRC and NIRC2 with the NIRC2
406: images serving as a reference coordinate system
407: (see Appendix \ref{app:speckDistort}). The speckle image distortion
408: is mapped by comparing stars' positions in both NIRC and NIRC2 images.
409: As shown in Appendix \ref{app:speckDistort}, the resulting NIRC to NIRC2
410: transformation is characterized at the $\sim$2 mas level over the entire
411: field of view.
412:
413: After distortion correction, individual exposures are combined into a final
414: diffraction-limited image using different methods for speckle and LGS AO data
415: sets.
416: Speckle images are produced by first rejecting the low Strehl ratio frames
417: (typically 75\% of frames are rejected)
418: and then stacking the remaining frames using a
419: weighted shift-and-add (SAA) routine \citep{sethThesis}.
420: The resulting combined images have a point-spread function (PSF)
421: composed of a diffraction-limited core
422: (FWHM$\sim$0\farcs055) on top of a broad seeing halo (FWHM$\sim$0\farcs4).
423: The improved image combination algorithm attempts to maximize the
424: signal-to-noise ratio (SNR) of the final image while preserving
425: the highest spatial resolution. Quantitatively, the weighted SAA method
426: doubles the fraction of light contained in the diffraction-limited
427: core (from 3.5\% to 7.0\%) over the standard SAA scheme with no weighting
428: and no frame rejection \citep{sethThesis}.
429: The LGS AO individual exposures are all of similar
430: quality and are thus all averaged together, without weighting, in order
431: to produce the final high-resolution image for each data set.
432: Although the 2005 June data were taken in two different filters
433: (K$_{CO}$ and K$_{cont}$), all the images
434: were combined together to increase the final SNR.
435: While photometry from this epoch is marginally impacted,
436: the astrometry is comparable to other epochs.
437: Each data set was also sub-divided to produce three equivalent
438: quality (randomized in time) subsets to make three images used for determining
439: photometric and astrometric uncertainties.
440: The resulting images are summarized in Table \ref{tab:obs},
441: including the achieved spatial resolution (FWHM) and
442: the Strehl ratio.
443:
444: \subsection{Stellar Positions and Coordinate Transformations}
445: \label{sec:coords}
446:
447: In order to extract astrometric information for the sample of young stars, the
448: coordinate system from each data set is transformed into a common
449: reference frame using the stars in each image to determine the
450: transformation parameters. Since the
451: accuracy of this transformation relies on the assumption that there
452: is no net rotation of the sample, we use all stars
453: detected in each data set, not just the young stars, in this analysis.
454: The steps for (1) measuring stars' positions in each epoch,
455: (2) transforming to a common (relative) reference frame, and
456: (3) determining the absolute coordinate system are
457: described below and utilize all stars detected in the data sets;
458: then as a final step, the young star sample is extracted.
459:
460: In each data set, stars are identified and their positions measured
461: using the IDL point-spread function fitting routine ``StarFinder''
462: \citep{starfinder}.
463: StarFinder generates a PSF from several bright stars in the field
464: and cross-correlates the resulting PSF with the image.
465: The PSF was iteratively constructed using
466: IRS 16C and IRS 16NW for the speckle maps and
467: IRS 16C, 16NW, 16NE, 16SW, 33E, 33W, 7, 29N, and GEN+2.33+4.60
468: for the LGS AO images.
469: Candidate stars are those for which StarFinder correlation
470: peaks have a correlation value higher than 0.8 and positions
471: and fluxes are extracted by fitting the PSF to each correlation peak.
472: From the candidate star list, spurious detections are then eliminated by
473: requiring that each star be detected in all three of the subset-images
474: with a correlation of higher than 0.6.
475: The positional centroiding uncertainties for each candidate star are estimated
476: from the rms of their locations in the three subset-images, and an additional
477: systematic error term of 0.88 mas is added in quadrature to all stars in
478: LGS AO epochs to account for residual distortion in the central 5'' of
479: NIRC2 \citep{ghez08}.
480: The candidate stars are flux calibrated using the apparent
481: magnitudes of the non-variable stars,
482: IRS 16C, IRS 16SW-E, S2-17, S1-23, S1-3, S1-4, S2-22, S2-5, S1-68, S0-13,
483: and S1-25, as measured by \citet{rafelski07}.
484: The star detections from each epoch are cross-identified
485: with stars from all other epochs and those stars that are detected in
486: at least 16 out of 29 epochs are used to create a master star list.
487: The threshold of 16 or more epochs is used in order to insure
488: high astrometric precision; for a threshold of less than 16 epochs,
489: the number of detected stars rises dramatically as does the number of
490: sources showing significant ($\gtrsim$3$\sigma$) accelerations in
491: non-physical directions, indicating a high frequency of false detections
492: (see \S\ref{sec:astroResults} for further discussion).
493: Stars in the master list are also examined for source confusion, which may occur
494: when two stars pass close enough to each other such that StarFinder
495: only detects a single source with biased astrometry rather than detecting
496: both stars. Source measurements from individual epochs are rejected if two
497: stars pass within 55 mas ($\sim$1 spatial resolution element) of each other and
498: only one source is detected by StarFinder.
499: The results of this stage of the analysis are summarized in
500: Table \ref{tab:obs}, which provides for
501: each data set the average centroiding error for the brightest stars
502: (K$<$13; also see Figure \ref{fig:posError})
503: and the sensitivity as estimated
504: by the peak in a histogram of the K-band magnitudes (bins = 0.1 mag)
505: of all the stars in the data set.
506: Averaged over all stars in all maps, the centroiding uncertainties have a
507: mean value of 1.6 mas for the brightest stars (K $\leq$ 13 mag) and
508: 3.4 mas for the fainter stars (13 $<$ K $<$ 16 mag).
509: % Numbers generated in lu06yng.showEpochsPosError()
510: % Change the code line: idx = .... and rerun
511:
512: The coordinate system for each image is transformed to a common
513: local reference frame defined by the 2004 July LGS AO/NIRC2 image's
514: coordinates and pixel scale.
515: This particular LGS AO epoch was chosen as the reference because
516: the NIRC speckle distortion solution is tied to this epoch, thus
517: providing a smooth transition between speckle and LGS AO data sets.
518: The procedure for deriving the coordinate transformation for
519: all of the data sets is non-trivial, since the stars in the
520: images have detectable motions.
521: Optimal alignment is achieved by minimizing the error-weighted, net
522: displacement for all the stars as described by \citet{ghez98pm}
523: while allowing for translation, rotation, and two magnifications in
524: arbitrary, but perpendicular, directions. This is a higher order
525: transformation than was used in our earlier astrometric works, which
526: only allowed for translation and rotation.
527: The new transformation equations have the form
528: \begin{eqnarray}
529: x_{pix} = a_0 + a_1 x'_{pix} + a_2 y'_{pix} \\
530: y_{pix} = b_0 + b_1 y'_{pix} + b_2 x'_{pix}
531: \end{eqnarray}
532: where $x'_{pix}$ and $y'_{pix}$ are the input detector coordinates in pixels
533: and $x_{pix}$ and $y_{pix}$ are the output coordinates for each star,
534: and all other variables are free parameters that are common across all stars
535: in the alignment fit.
536: As in \citet{ghez05orbits}, stars within 0\farcs5 of Sgr A* are excluded
537: from the transformation as they exhibit large non-linear motions.
538: Additionally, all spectroscopically identified young stars are excluded
539: from the transformation as they have a known net rotation \citep{genzel00}.
540: Initially, each image is aligned to the reference image by assuming the
541: stars have no proper motions and finding the best-fit values for the
542: free parameters of the transformation, $a_0, a_1, a_2, b_0, b_1, b_2$, for
543: that image. However, after a first pass at the alignment of all the images,
544: proper motions are estimated and
545: used to refine the alignment solutions in a second pass.
546: Sources with estimated proper motions higher than
547: 1.5 mas yr$^{-1}$ (600 km s$^{-1}$)
548: are excluded from the transformation resulting in the elimination of
549: 2 sources that are near the edge of the speckle field-of-view and suffer
550: from edge effects.
551: Alignment uncertainties are estimated by a half-sample bootstrap method
552: \citep{astrostats,ghez05orbits}
553: and are small ($\sim$0.2 mas for
554: stars at $r<$2\arcsec) compared to the centroiding uncertainties
555: (see Figure \ref{fig:posError}).
556: % Number generated from lu06yng.plotPosError()
557: Alignment and centroiding
558: uncertainties are added in quadrature to produce a final relative
559: positional uncertainty for each star at each epoch.
560: The resulting astrometric data set contains stellar positions
561: and uncertainties for all epochs, transformed into the 2004 July NIRC2 pixel
562: coordinate system ($x_{pix}, y_{pix}$).
563:
564: The relative positions and uncertainties are transformed into J2000
565: absolute astrometric coordinates defined by radio observations of
566: SiO masers and Sgr A*. Using observations of the SiO masers in the infrared,
567: a set of {\it infrared absolute astrometric standards} are defined in
568: a process described in detail in an
569: appendix of \citet{ghez08}.
570: These astrometric standards are used to derive the transformation from
571: 2004 July NIRC2 pixel coordinates into absolute coordinates.
572: A statistically insignificant adjustment is made to place the origin at
573: the dynamical center of S0-2's orbit, which is known to high precision,
574: by offseting from the radio position of Sgr A* by 1 mas to the East
575: and 5 mas to the South. This offset is well within the absolute astrometric
576: uncertainty of $\sim$6 mas for Sgr A* \citep{ghez08}.
577: The stellar positions in all epochs are thus expressed in arcseconds
578: offset from the dynamical center with $+x$ increasing East and $+y$
579: increasing North and can be converted
580: into celestial coordinates using
581: (x, y) = ($\cos{\delta}\; \Delta\alpha$, $\Delta\delta$)
582: \footnote{When converting from (x, y) to ($\Delta\alpha$, $\Delta\delta$),
583: higher order terms are negligible (0.06 mas over 5\arcsec) because the
584: celestial sphere is sufficiently flat over our field of view.
585: }.
586: Positional uncertainties are taken as the quadratic sum of the relative
587: errors, which dominate, and the absolute error from uncertainties in the
588: plate scale and position angle.
589: Errors in the relative position of Sgr A* ($\sim$2 mas) are incorporated later
590: during the orbit analysis stage as a parameter of the potential of the
591: supermassive black hole
592: (see \S\ref{sec:orbitAnalysis}).
593: From the resulting absolute astrometric data set, the sample of young
594: stars is extracted.
595:
596:
597: \subsection{Proper Motions and Acceleration Results}
598: \label{sec:astroResults}
599:
600: For each of the young stars in the sample, positions, velocities, and
601: accelerations in the plane of the sky are derived by fitting second-order
602: polynomials to the star's
603: position as a function of time, weighted by the positional uncertainties.
604: The polynomials are fit independently in $x$ and $y$
605: coordinates and have the form
606: \begin{eqnarray}
607: x(t) = x_{ref} + v_{x,ref} (t - t_{ref}) + \frac{1}{2} a_{x,ref} (t - t_{ref})^2 \\
608: y(t) = y_{ref} + v_{y,ref} (t - t_{ref}) + \frac{1}{2} a_{y,ref} (t - t_{ref})^2
609: \end{eqnarray}
610: where $t$ is the time in years, $t_{ref}$ is a reference time
611: taken to be the mean of the time of all epochs weighted by the positional
612: uncertainties for each star,
613: $x_{ref}$ and $y_{ref}$ are the positions at the reference time, $v_{ref}$ is the
614: velocity at the reference time, and $a_{ref}$ is the acceleration at the
615: reference time. Uncertainties in the fit
616: parameters are determined from the covariance matrix.
617: Figures \ref{fig:posTime_S0-15} and \ref{fig:posTime_irs16NW} show the
618: polynomial fits for two example stars and
619: the resulting values for the kinematic variables for all stars are reported
620: in Table \ref{tab:yng_pm_table}.
621: Since the stars' motions
622: are assumed to be dominated by the central force from the black hole,
623: we convert $a_{x,ref}$ and $a_{y,ref}$ into radial and tangential
624: accelerations \footnote{This assumption
625: may not hold for stars in a gravitationally bound cluster, such as may be the
626: case for the 4 stars in the extended sample that make up the IRS 13
627: co-moving group; however, the deviations from the potential assumed above
628: should result in only $5-10$\% changes in the velocity vectors.}.
629: All tangential accelerations and
630: positive radial accelerations are non-physical and therefore provide
631: a check on the systematic errors of the acceleration measurements.
632: Figure \ref{fig:histAccel} shows a histogram of the
633: significance of the acceleration measurements both in the radial and
634: tangential directions for the young stars in our primary sample.
635: While the tangential and positive radial distributions are
636: slightly offset (0.6$\sigma$) from zero and
637: broader (1.5$\sigma$ vs. 1$\sigma$) than is expected
638: for a normal distribution,
639: any systematic errors appear to impact the results at the
640: $\lesssim 1\sigma$ level.
641:
642: The resulting velocity measurements for the young star sample outside
643: the central arcsecond are improved by at least a factor of 7 when compared
644: with our previous work \citep{ghez98pm,lu05irs16sw} and other recently
645: reported Galactic Center proper motions \citep[e.g.][]{genzel00,ott03thesis}.
646: The absolute uncertainties in our proper motions
647: are typically $\sim$0.06 mas yr$^{-1}$ ($\sim$2 km s$^{-1}$),
648: although stars detected in fewer epochs have somewhat higher values
649: (0.1 - 0.5 mas yr$^{-1}$; 4 - 20 km s$^{-1}$).
650: Figures \ref{fig:posTime_S0-15} and \ref{fig:posTime_irs16NW}
651: show examples of the measurements for two stars in our sample (S0-15
652: and IRS 16NW), and their corresponding proper motion fits
653: with 1$\sigma$ errorbars.
654:
655: In the young star sample, significant ($>$3$\sigma$) acceleration,
656: or curvature, in the plane of the sky is detected only for S0-15
657: (Figure \ref{fig:posTime_S0-15}).
658: This star has the second smallest projected separation from Sgr A*
659: in our sample,
660: at $\rho$ = 1\farcs0 (0.04 pc), and has a projected radial acceleration of
661: -0.21 $\pm$ 0.05 mas yr$^{-2}$ or, equivalently,
662: -9.6 $\pm$ 2.0 km s$^{-1}$ yr$^{-1}$ (see Figure \ref{fig:PMimage}).
663: S0-15 is more than twice as far from Sgr A*, in projection, than the seven
664: stars with previously detected accelerations,
665: which were all within a projected radius of less than 0\farcs4
666: \citep[0.016 pc, ][]{ghez00nat,eckart02,ghez05orbits,eisenhauer06}.
667:
668: The detection of acceleration is important in that it allows us to
669: solve for the line-of-sight distance, and thus the
670: three-dimensional position of a star relative to the black hole.
671: For a star in the gravitational potential well of a supermassive black hole,
672: the plane-of-the-sky acceleration, at a three-dimensional distance r,
673: in cylindrical coordinates is
674: \begin{equation}
675: \label{eqn:a2z}
676: a_{\rho} = \frac{-GM \rho}{r^3} = \frac{-GM \rho}{(\rho^2 + z^2)^{3/2}}
677: \end{equation}
678: where $\rho$ is the plane-of-the-sky radial coordinate and z is the
679: coordinate along the line of sight relative to Sgr A*.
680: The magnitude of the line-of-sight distance from Sgr A*, z,
681: can be solved for by adopting a black hole mass of
682: $M_{\bullet} = 4.4 \times 10^6$ \msun and a distance of
683: $R_{\circ} = 8.0$ kpc \citep[see \S\ref{sec:orbitAnalysis};][]{ghez08};
684: it is important to note that there is a remaining sign ambiguity for z.
685: The resulting line-of-sight distance from Sgr A*
686: for S0-15 is $|0.045 \pm 0.004|$ pc bringing the total separation between
687: S0-15 and Sgr A* to 0.060 pc.
688: % Numbers generated by python code:
689: % lu.calcZfromA('S0-15')
690:
691: The remaining stars in our sample have acceleration measurements that
692: constrain the line-of-sight distance.
693: While the lower limits of these acceleration magnitudes are not
694: significantly different
695: from zero at the 3$\sigma$ level, their upper limits are
696: smaller than the maximum allowed acceleration.
697: The maximum possible magnitude of the acceleration for a star at a
698: given $\rho$ occurs when z = 0. When the measured acceleration limits are
699: below this value, they provide a
700: lower limit on the star's line-of-sight distance to the SMBH.
701: Figure \ref{fig:accSignificance2} compares the measured acceleration limits
702: with the maximum possible acceleration for each star.
703: Any 3$\sigma$ acceleration limits below the maximum allowed value
704: gives useful constraints on the line-of-sight distances.
705: In addition to our explicit measurement for S0-15,
706: our high precision astrometric measurements are now yielding
707: 3$\sigma$ acceleration limits with a median value of
708: % Numbers from lu.plotPolyAccVsR()
709: -0.19 mas yr$^{-2}$ (-7.3 km s$^{-1}$ yr$^{-1}$)
710: that can significantly constrain the line-of-sight
711: distance for nine stars in our sample that are located as far as
712: 1\farcs7 (0.07 pc), in projection, away from the black hole.
713:
714: \section{Orbit Analysis}
715: \label{sec:orbitAnalysis}
716:
717: For a known point-mass Newtonian gravitational potential,
718: a star's orbital elements can be fully determined from
719: the measurement of only six kinematic variables.
720: For this analysis, we assume that the central point mass is a
721: black hole with characteristics determined by analysis of the
722: orbit of the star S0-2, which has been observed for nearly one
723: complete revolution
724: \citep{eisenhauer06,ghez08}.
725: Our proper motion analysis (\S\ref{sec:astroResults}) yields information
726: on five kinematic variables, including two positions, two velocities, and one
727: acceleration.
728: The sixth kinematic variable comes from radial velocities measured by
729: \citet{paumard06}. The reported radial velocities are averaged
730: over several years of observations; however, we adopt the same
731: reference epoch, $t_{ref}$, as for the proper motion analysis since
732: any change in the radial velocity due to acceleration along the
733: line-of-sight should be well
734: within the large measurement uncertainties in radial velocity
735: ($\sigma_{v_{z,ref}}\sim$20-100 km s$^{-1}$).
736: As described in \S\ref{sec:astroResults}, the plane-of-the-sky acceleration
737: can be converted into a line-of-sight distance that, when combined with the
738: projected distance, gives the full three-dimensional position for a star.
739: Although most of the stars in our sample have plane-of-the-sky accelerations
740: that are consistent with zero, the upper limits on the magnitude of the
741: acceleration provide valuable information by ruling out small
742: line-of-sight distances. We therefore use our best-fit accelerations
743: and uncertainties as formal measures of the acceleration when converting
744: to a line-of-sight distance.
745: Therefore the six measured quantities can be expressed as a
746: three-dimensional position and three-dimensional velocity at a certain
747: epoch ($t_{ref}$).
748: Given the properties of the black hole, these kinematic quantities can be
749: translated directly into 6 standard orbital elements
750: (see Appendix \ref{app:orbits}).
751:
752: A Monte Carlo simulation is carried out to transform
753: each star's six measured kinematic variables
754: ($x_{ref}$, $y_{ref}$, $v_{x,ref}$, $v_{y,ref}$, $v_{z,ref}$, $a_{\rho,ref}$)
755: into six orbital parameters ($i$, $\Omega$, $\omega$, $e$, $P$, $T_o$)
756: and their uncertainties.
757: A total of 10$^5$ Monte Carlo trials are run and, in each trial,
758: $4 + (6 \cdot 32)$ variables are randomly generated; four for the potential
759: parameters and six for each of the 32 stars' measured kinematic variables.
760: The four potential parameters are pulled from a four-dimensional probability
761: density function, PDF(M$_\bullet$, R$_o$, x$_o$, y$_o$),
762: based on the orbit of S0-2 derived by \citet{ghez08},
763: where the black hole's mass and line-of-sight distance are centered on
764: $M_{\bullet}=4.4 \times 10^6$ \msun and
765: $R_{\circ}=8.0$ kpc \footnote{These values correspond to a 12-parameter
766: orbit model for S0-2 (i.e. $v_z=0$ case) from an early version of
767: \citet{ghez08}. In this version, local distortions were
768: not corrected \citep[][Appendix B]{ghez08}; but the
769: resulting black hole mass and distance differ by $<1\sigma$ from the final
770: reported values.},
771: the dynamical center is adopted as the origin
772: with x$_o$ and y$_o$ defined as zero, and the projected one-dimensional
773: probability distributions' RMS errors are [1.0, 1.6] mas for [$x_o$, $y_o$],
774: $0.3 \times 10^6$ \msun for $M_{\bullet}$, and $0.3$ kpc for $R_\circ$
775: \footnote{Simulations were also performed using the lower black hole mass and
776: distance reported by \citep{eisenhauer06}. Our results on the detection of
777: only one stellar disk and on the properties of the disk are all consistent
778: within 1$\sigma$ error bars.}.
779: For each trial, all the stars' orbits are calculated using the same
780: potential parameters in order to preserve correlations between the
781: potential parameters and the orbital parameters such as eccentricity.
782: The kinematic variables for each star
783: are sampled from independent gaussian distributions,
784: each of which is centered at the best-fit value from Table
785: \ref{tab:yng_pm_table} and has a 1$\sigma$ width set to the measurement
786: uncertainty.
787: Any correlations between the measured kinematic variables are negligible
788: given the small uncertainties in the stars' relative angular positions
789: ($\lesssim$0.2\%) and velocities ($\lesssim$3\%) in the plane-of-the-sky
790: as compared to the uncertainties in the black hole mass ($\sim$10\%)
791: and the accelerations ($\sim$60\%).
792: The distribution for the acceleration, $a_\rho$, is truncated such that only
793: accelerations of bound orbits are allowed\footnote{The assumption that the
794: orbits are bound does not effect the results presented in this paper
795: discussed in \S\ref{sec:orbitResults} since
796: all unbound orbital solutions yield high inclination (edge-on) orbits
797: and large eccentricities (e > 1). Considering only bound orbits
798: simplifies the orbit analysis as we need only consider equations for
799: elliptical orbits rather than hyperbolic or parabolic orbits.
800: }, which follows from requiring a negative specific orbital energy,
801: \begin{equation}
802: E = \frac{v^2}{2} - \frac{GM}{r} < 0
803: \end{equation}
804: and substituting from Eq. \ref{eqn:a2z} to give the acceleration constraint
805: \begin{equation}
806: |a_\rho| > \frac{\rho v^6}{8(GM)^2}.
807: \end{equation}
808: For each trial and each star, the orbital parameters are computed and
809: the results of all the trials are combined into a
810: six-dimensional probability density function (PDF) by dividing up parameter
811: space into bins, summing the number of trials in each bin, and then
812: normalizing by the total number of trials.
813: This Monte Carlo method is a straight-forward way to combine
814: a star's six measurement PDFs and the
815: four-dimensional PDF for the central point mass, which shows
816: strong correlations between $M_\bullet$ and $R_o$,
817: to produce a six-dimensional PDF for each star's orbital elements,
818: PDF($i$, $\Omega$, $\omega$, $e$, $P$, $T_o$), which has strong correlations
819: between the orbital parameters.
820: The results of these simulations are plotted for an example star, IRS 16SW,
821: in Figure \ref{fig:pdfParams_irs16SW}
822: to show that $i$ and $\Omega$ are generally well determined and that
823: $e$, in some cases, can be usefully constrained.
824: Similar figures of the orbital parameters for every star are
825: shown in Figure Set 7, which is available online in the electronic edition
826: of this manuscript.
827:
828: The resulting stellar orbital parameters are constrained by several different
829: factors.
830: First, a measured acceleration that is significantly different from zero,
831: such as for S0-15, yields the best
832: determined orbit since the line-of-sight distance is confined to a small
833: range of values (Figure \ref{fig:pdfParams_example}, {\it top}).
834: Secondly, each star has a maximum allowed acceleration,
835: $a_{{\rho}, max} = |-GM/\rho^2|$, at the closest
836: possible distance set by the observed projected radius.
837: Stars with measured accelerations more than 3$\sigma$ below
838: the maximum allowed acceleration, such as IRS 16NW,
839: have strong lower limits on their line-of-sight distances, which translate
840: into significant constraints on the direction of the angular momentum
841: vector, $\vec{L}$, and can be equivalently expressed as constraints on
842: inclination, $i$, and on the angle to the ascending node, $\Omega$
843: (Figure \ref{fig:pdfParams_example}, {\it middle}).
844: Finally, even stars without significant limits on their
845: line-of-sight distance from accelerations have some well
846: constrained orbital elements. In particular, $i$ and $\Omega$
847: are well constrained as a result of the precise measurements for the
848: stellar velocities and potential parameters.
849: Furthermore, if the star's total velocity is higher than the circular
850: velocity at the two-dimensional projected radius, then it is higher than the
851: circular velocity at {\it all} distances and only non-zero eccentricity
852: orbits are allowed (Figure \ref{fig:pdfParams_example}, {\it bottom}).
853:
854: The Monte Carlo analysis described above assumes that, in the absence
855: of an acceleration measurement, the acceleration should be drawn from a
856: uniform probability distribution; or, in other words, we adopt a
857: uniform acceleration prior. For those stars that are
858: only in the extended sample, the Monte Carlo
859: orbit analysis samples from this uniform acceleration prior
860: ranging from the largest allowed acceleration by the
861: projected radius to the smallest allowed for the orbit to remain bound.
862: For these stars and for stars in the primary sample with acceleration
863: limits that are not significantly
864: smaller than the maximum physically allowed acceleration,
865: the uniform-$a_\rho$ prior is an important assumption.
866: To test how sensitive our results are to this assumption, we performed
867: the same Monte Carlo analysis as detailed above using an
868: alternative assumption that the prior acceleration distribution is
869: uniform in z,
870: which shifts the line-of-sight distance PDF to larger values when
871: compared with a uniform-$a_\rho$ prior. On a star-by-star basis, the
872: resulting orbital parameters are consistent within 1$\sigma$
873: for both priors, with one exception. The young star S0-14 has an
874: eccentricity that is constrained to be higher than 0.93 (3$\sigma$)
875: with a uniform-$a_\rho$ prior, while with a uniform-z prior, all
876: eccentricities are allowed within 3$\sigma$. S0-14 is distinguishable
877: from all other stars in our sample in that it has a total velocity of
878: only 50 km s$^{-1}$, as compared to 160-640 km s$^{-1}$ for the rest of
879: the sample.
880: Such a small velocity translates into a very large range of allowed
881: line-of-sight distances which are not well sampled by a uniform-$a_\rho$
882: prior. S0-14's range of $i$ and $\Omega$ are not largely affected by the
883: choice of prior; therefore, we exclude S0-14 from our eccentricity analysis,
884: but we keep it in all other orbital analyses.
885:
886: To distinguish between these two possible priors, we examine the resulting
887: distribution of orbital phases.
888: For a set of stars whose motion is dominated by the supermassive black hole
889: and that have been orbiting for more than a few
890: orbital time scales, the distribution of orbital phases should be
891: uniform.
892: The distribution of orbital phases for our sample is constructed by
893: summing the orbital phase PDFs for all the stars.
894: Figure \ref{fig:comparePriorsPhase} shows that
895: while the uniform-$a_\rho$ prior produces a population that is
896: uniformly distributed in orbital phase, the uniform-z prior produces
897: a distribution that is strongly peaked at 0 (periapse) due to the
898: higher occurence of large line-of-sight distances that, for a given
899: velocity, creates an artificial bias towards periapse.
900: Such a strong bias towards periapse is unlikely to occur even if
901: some of the young stars reside in a gravitationally bound cluster,
902: such as IRS 13, where all the cluster members would have a similar
903: orbital phase.
904: Based on our assumption that the distribution of orbital phases
905: should be roughly uniform, we adopt a uniform-$a_\rho$ prior instead of
906: the uniform-z prior in the following sections.
907:
908: \section{Orbit Results}
909: \label{sec:orbitResults}
910:
911: \subsection{Detection of the Clockwise Disk}
912: \label{sec:diskDetect}
913:
914: A large number of stars appear to share a common orbital plane
915: based on our analysis, which has no prior assumption about the
916: existence of a disk.
917: The orientation of a star's orbital plane can be described by
918: a unit vector originating at Sgr A*'s position and pointing
919: normal to the orbital plane ($\vec{n}$); and, this normal vector's
920: direction can be expressed by the inclination angle ($i$) and the angle to the
921: ascending node ($\Omega$) using
922: \begin{equation}
923: \vec{n} = \left( \begin{array}{c} n_x \\ n_y \\ n_z \end{array} \right) =
924: \left( \begin{array}{c}
925: \sin\, i\; \cos\, \Omega \\
926: -\sin\, i\; \sin\, \Omega \\
927: -\cos\, i\; \end{array} \right).
928: \end{equation}
929: The direction of each star's orbital plane normal vector is
930: determined from the joint two-dimensional probability density
931: function of $i$ and $\Omega$,
932: PDF($i$, $\Omega$),
933: which is constructed by
934: binning the resulting $i$ and $\Omega$ values from the Monte Carlo
935: simulation in a two-dimensional histogram with equal solid angle bins using
936: the HEALpix framework \citep{healpix}.
937: Figure \ref{fig:iomap} shows PDF($i$, $\Omega$)
938: projected onto the sky as viewed from Sgr A*
939: for the same three example stars shown in Figure \ref{fig:pdfParams_example}.
940: % Numbers from:
941: % papers.lu06yng.diskMembers()
942: % papers.lu06yng.histSolidAngles()
943: Figure \ref{fig:allPlanePDF} shows, for all stars,
944: the contours for the 68\% confidence region, which,
945: on average, covers a solid angle of $SA_{\vec{n}} \sim$0.2 steradian (sr)
946: for the primary sample and 0.6 sr for stars found only in the extended sample,
947: which have larger proper motion uncertainties.
948: Table \ref{tab:eccDisk} \& \ref{tab:eccDiskExtended}
949: list this solid angle, $SA_{\vec{n}}$, for each star in the primary and
950: extended samples.
951: The bound orbit assumption does not greatly impact the size of the
952: $SA_{\vec{n}}$ because the orbital
953: parameters $i$ and $\Omega$ asymptote at large line-of-sight distances
954: as can be seen in Figure \ref{fig:pdfParams_example}.
955: Stars with acceleration limits significantly smaller than $a_{{\rho}, max}$
956: have two isolated solutions because small line-of-sight distances (z)
957: are not permitted and at large line-of-sight distances the positive-z
958: and negative-z solutions asymptote to
959: two different values of $\Omega$ (see Figure \ref{fig:pdfParams_example}).
960: Despite this degeneracy, the clockwise ($i$=90$^\circ$-180$^\circ$)
961: stars' normal vectors appear to cluster around a common point indicating
962: that many of these stars lie on a common orbital plane.
963:
964: The directions of the stars' normal vectors show a
965: statistically significant clustering as measured by the
966: the density of normal vectors in the sky as viewed from Sgr A*.
967: To quantify the density of normal vector directions, we use
968: a nearest neighbor density estimate, which is commonly used to identify
969: galaxy clusters \citep[e.g.,][]{nearestNeighbor}, and take the density
970: at each point on the sky to be
971: \begin{equation}
972: \Sigma = \frac{k}{2\pi (1-\cos{\theta_k})} \textrm{stars sr}^{-1}
973: \end{equation}
974: where $\theta_k$ is the angle to the $k^{th}$ nearest star
975: and $k$ is taken to be 6.
976: We calculate the expectation value for the density of normal vectors at
977: each point on the sky using the Monte Carlo simulation discussed earlier.
978: For each Monte Carlo trial, the sky is divided into 12288 equal area pixels
979: (0.001 sr) using a HEALpix grid and the
980: density of normal vectors is calculated for each pixel.
981: These estimates are then averaged together over all the trials to provide
982: an average density per pixel on the sky.
983: The resulting average density of normal vectors is nearly the same
984: for a choice of 4th, 5th, or 7th nearest neighbor.
985: Additionally, a similar analysis using a fixed
986: aperture to calculate the density of normal vectors at each point on the
987: sky produced similar, but less smooth, results as the nearest neighbor
988: approach we adopt here.
989: A peak in the density of normal vectors is detected at
990: $i = 115^\circ \pm 3^\circ$ and $\Omega = 100^\circ \pm 3^\circ$,
991: which provides direct evidence of a common orbital plane without any
992: prior assumptions (see Figure \ref{fig:orbitPlane}).
993: The uncertainty on the peak position is taken as the
994: half-width at half-maximum of the peak divided by the square-root of the
995: number of stars that are candidate disk members,
996: $\sqrt{N_{disk-stars}}$ (see below).
997: We also note that an analysis of the entire extended
998: sample produces a peak at the exact same position.
999: The mean density of normal vectors at the peak is
1000: 0.016 stars deg$^{-2}$ with a negligible uncertainty on the mean value
1001: ($< 10^{-4}$ stars deg$^{-2}$).
1002: The significance of the peak is determined by comparing the
1003: background density of normal vectors, which is defined by the
1004: average (0.001 stars deg$^{-2}$) and standard deviation (0.0008 stars deg$^{-2}$)
1005: of all other pixels on the sky
1006: after first rejecting those pixels ($\sim$0.25 sr)
1007: that are high outliers (more than three standard deviations).
1008: % Numbers generated by python code:
1009: % lu.plotIOcontours()
1010: The density peak is $\sim$19$\sigma$ above the observed
1011: background density.
1012: A second comparison can be made to the density expected
1013: if the 32 stars in our sample were isotropically distributed over
1014: 4$\pi$ steradians.
1015: The observed peak in the density is
1016: $\gtrsim$20 times higher than this isotropic density. Thus we conclude
1017: that there is a statistically significant common orbital plane of young stars.
1018:
1019: % Numbers generated by python code:
1020: % from gcwork import analyticOrbits as aorb
1021: % aorb.diskWidthCDF('aorb_efit/disk.neighbor.dat', 2)
1022: The majority of the young stars that are orbiting in the clockwise direction
1023: are likely to be orbiting in this common plane.
1024: A comparison of each star's normal vector to the common plane's normal vector
1025: allows us to determine which stars are {\it not} on the common plane with
1026: high statistical significance.
1027: All other stars are then considered {\it candidate} members.
1028: First, a preliminary estimate of the thickness of the common plane is
1029: determined by defining the solid angle extent of the
1030: plane, $SA_{plane}$, encompassed by the contour at which the
1031: density drops to half of the peak value.
1032: % Numbers generated by python code:
1033: % lu.plotIOcontours()
1034: This corresponds to a region
1035: with a solid angle of $SA_{plane} \sim$0.1 sr, which gives a half-opening
1036: angle of 0.2 radians (10$^\circ$) for a cone with the same $SA_{plane}$.
1037: Then, each star's probability density function, PDF($i$, $\Omega$),
1038: is integrated over this region to determine the probability that the
1039: star is a disk member.
1040: The orientation of the stars' normal vectors have a wide range of
1041: uncertainties as expressed by the total solid angle covered by each star,
1042: so it is
1043: necessary to distinguish between those stars that have a low probability
1044: due to a large $\vec{n}$-uncertainty (i.e. large solid angle)
1045: vs. those stars that have a low probability
1046: because they are significantly offset from the common plane.
1047: Therefore, we normalize the above integrated probability by
1048: the probability at the peak of the star's PDF integrated over a region
1049: that has the same total area as the common plane
1050: \begin{eqnarray}
1051: \mathrm{L}(\mathrm{not\;on\;plane}) & = & 1 -
1052: \frac{\int_{\mathrm{plane}} \mathrm{PDF}(i, \Omega)\;\; d\mathrm{SA}}
1053: {\int_{\mathrm{peak}} \mathrm{PDF}(i, \Omega)\;\; d\mathrm{SA}} \\
1054: \int_{\mathrm{plane}} d\mathrm{SA} & = & \int_{\mathrm{peak}} d\mathrm{SA}
1055: \end{eqnarray}
1056: where SA is the solid angle and L(not on plane)
1057: is the likelihood that the star is {\it not} on the common plane.
1058: Those stars with likelihoods, L(not on plane),
1059: of greater than 0.9973 (equivalent to 3$\sigma$ for a gaussian distribution)
1060: are flagged as non-members of the common plane.
1061: The remaining set of stars are considered candidate members of the common plane.
1062: Table \ref{tab:eccDisk} \& \ref{tab:eccDiskExtended}
1063: list [1 - L(not on disk)] for each
1064: star and Figure \ref{fig:PMimage} shows the positions of
1065: candidate members of the common plane in red and non-members in blue.
1066: Of the primary sample of 32 stars, 26 of which are orbiting in a clockwise
1067: sense on the plane-of-the-sky, we find that 22 are possible members of the
1068: common plane ($N_{disk-stars} = 22$).
1069:
1070: The clockwise common plane that we measure is slightly offset from the
1071: clockwise planes proposed in earlier works.
1072: Over-plotted in black on Figure \ref{fig:orbitPlane} is the candidate
1073: orbital plane proposed by \citet{levin03} with
1074: updated values from \citet{paumard06} for the candidate plane normal
1075: vector (solid black) and thickness (dashed black).
1076: The previously proposed plane was derived by minimizing a statistical metric,
1077: K, in order to find the best-fit common orbital plane from the velocity
1078: vectors of a sample of stars (see Appendix \ref{app:kmetric}).
1079: However, some stars are not members of the common plane and including
1080: them in the fit biases the result since they have extremely well
1081: measured velocities (S0-15, IRS 16C, S3-19).
1082: % Use young.loadYoungStars('./', verbose=True) to find Paumard names.
1083: For example, using the K metric
1084: approach of \citet{levin03}, fitting all 26 clockwise stars in our
1085: primary sample gives i = 128$^\circ$ and $\Omega$ = 102$^\circ$ with K = 0.7,
1086: which is closer to the disk found by \citet{paumard06} at
1087: i = 127$^\circ$ and $\Omega$ = 99$^\circ$.
1088: While fitting only the 22 stars that are consistent with the clockwise disk
1089: based on our orbit analysis gives i = 117$^\circ$ and $\Omega$ = 98$^\circ$ with
1090: K = 0.2. Therefore, using the K metric to determine the common plane can
1091: produce biased results due to the inclusion of non-members.
1092: By combining position, velocity, and acceleration information in order to
1093: determine the orbital plane for each star, the direction of a common
1094: orbital plane can be estimated more robustly.
1095:
1096: The detected common orbital plane is composed of stars dispersed in a
1097: disk rather than in a single cluster as can be seen from the
1098: stars' positions within the common plane shown in
1099: Figure \ref{fig:diskVelRadius}.
1100: In this figure, the stars' positions have been
1101: converted into a disk coordinate system defined as
1102: [$\hat{p}$, $\hat{q}$, $\hat{n}$] where $\hat{n}$ is perpendicular
1103: to the disk plane,
1104: $\hat{p}$ is along the line of ascending nodes (where the plane of the sky
1105: intersects the disk plane), and $\hat{q} = \hat{n}\times\hat{p}$.
1106: For each star, all orbital solutions that fall within 10$^\circ$ of the
1107: common orbital plane are combined to create
1108: a probability distribution for the star's position in the disk,
1109: PDF($p$, $q$), which is shown in Figure \ref{fig:diskVelRadius} ({\it left}).
1110: Each stars probability distribution is elongated in the $q$-direction
1111: due to the large range of line-of-sight distances, $z$, that
1112: are possible within the small range of possible disk inclinations for this
1113: nearly edge-on plane of the disk.
1114: The thickness in the $p$ direction is largely set by the uncertainties
1115: in the potential parameters ($M_\bullet$, R$_o$) and velocities.
1116: The distribution of young stars within the plane shows a range of
1117: position angles on the plane, consistent with a stellar disk rather
1118: than a stellar cluster.
1119:
1120: The CW stellar disk is detected both in our analysis of the primary sample
1121: and in a similar analysis of the entire extended sample.
1122: The additional young stars in the extended sample have larger velocity
1123: uncertainties and no acceleration information, therefore the Monte Carlo
1124: orbit analysis samples from a prior probability distribution that is
1125: uniform in acceleration ranging from the largest allowed by the
1126: projected radius to the smallest allowed for the orbit to
1127: remain bound.
1128: We note that even if we ignore the acceleration measurements for our primary
1129: sample analysis, the CW stellar disk is still detected, although the
1130: significance is lowered from $\sim$19$\sigma$ to $\sim$8$\sigma$ above the
1131: background density of normal vectors.
1132: % Run plot_disk_healpix on both
1133: % aorb_efit/disk.neighbor.dat
1134: % aorb_efit_noaccel/disk.neighbor.dat
1135: Thus the additional stars' orbits in the extended
1136: sample are still constrained (see Figure \ref{fig:histSolidAngle}),
1137: even though they have larger uncertainties as compared to the stars in
1138: just the primary sample.
1139: The density of normal vectors from the extended sample analysis shows a
1140: peak within 1$^\circ$ of the disk's position from the primary sample.
1141:
1142: The analysis of the extended sample shows that $\sim$50\% of the young
1143: stars reside on the CW disk and there is no statistically significant
1144: change ($>3\sigma$) in the fractional number of disk stars at different radii.
1145: For reference, the 73 young stars in the extended sample are
1146: distributed on the plane of the sky with a surface density
1147: that decreases with radius as $\rho^{-2.1 \pm 0.4}$.
1148: % lu.membersVsRadius()
1149: Within a projected radius of 3'', the fraction of candidate
1150: disk members is 72\% $\pm$ 9\% (18 out of 25) and
1151: at projected radii larger than 3'', the fraction of candidate disk members
1152: is 42\% $\pm$ 7\% (20 out of 48).
1153: % lu.radialDistribution(extended=True)
1154: Given the small number of known young stars, Poisson statistics indicate
1155: that this change in the fraction of candidate disk members is only marginally
1156: statistically significant at the 2.6$\sigma$ level.
1157: Likewise, the projected surface density for the on-disk and off-disk
1158: populations shown no significant difference from each other or from that of
1159: the total population.
1160: Thus the number of candidate disk members does not change with radius and
1161: roughly half of the young stars reside on the CW disk.
1162:
1163: The K-band luminosity function (KLF) of the young stars does not change
1164: significantly with radius or when considering stars on and off the
1165: disk. To compare the KLF as a function of radius, the entire extended
1166: sample of young stars is divided into a near sample (r $<$ 3\farcs5)
1167: and a far sample (r $\geq$ 3\farcs5) and the KLF is constructed for
1168: each. A two-sample KS test yields a probability of 46\% that the
1169: near and far samples have the same KLF. Similarly, the KLF is constructed
1170: for stars on and off the disk and a two-sample KS test yields
1171: a probability of 74\% that the on-disk and off-disk samples have the
1172: same KLF. Finding more young stars will allow for a more detailed
1173: comparison of the KLF for different subsets within the young stars
1174: population.
1175:
1176: \subsection{Limits on Additional Stellar Disks}
1177:
1178: In our primary sample, no common orbital plane is detected for the
1179: counter-clockwise population of stars; however, our sample is limited to
1180: six counter-clockwise
1181: orbiting stars, only two of which (IRS 16NE, IRS 16NW) are claimed by
1182: \citet{paumard06} to
1183: reside on the counter-clockwise disk.
1184: Out of the 6 counter-clockwise stars in our primary sample, we find that only
1185: IRS 16NE and S2-66 could be consistent with the previously proposed
1186: counter-clockwise disk.
1187: The proposed counter-clockwise disk may have a larger
1188: radial extent than is covered by our observations,
1189: so in order to fully explore whether
1190: our lack of detection of a 2nd disk is due to our limited field-of-view,
1191: it is necessary to analyze the extended sample.
1192: As discussed in \S\ref{sec:orbitAnalysis},
1193: the uniform acceleration prior adopted for this analysis tends to
1194: overemphasize face-on orbital planes, making it easier to detect the
1195: proposed CCW disk, as \citet{paumard06} suggest it has an inclination
1196: of 24$^\circ$.
1197:
1198: % See aorb_efit_all/disk.neighbor.dat.ps
1199: % Expected density of ~ 0.002 stars/sq. deg.
1200: % lu.checkCCWdisk()
1201: % lu.diskMembers(extended=True)
1202: % lu.diskMembersCCW()
1203: Using the extended sample, our analysis of the density of normal vectors,
1204: in the region of the proposed counter-clockwise disk, reveals no
1205: significant over-density.
1206: Of the 73 stars in the extended sample, at least 34 are not on the clockwise
1207: disk and thus we compare the density observed in the region of the
1208: proposed counter-clockwise
1209: disk to that expected for an isotropic distribution of 34 stars.
1210: The observed density of normal vectors in the region of the
1211: counter-clockwise disk
1212: is 2.4$\times$10$^{-3}$ stars deg$^{-2}$, which is only a factor of 3 above
1213: what is expected for an isotropic distribution and is less than
1214: 1$\sigma$ above the background over the rest of the sky (excluding
1215: the clockwise peak).
1216: This density of normal vectors corresponds to only 3 stars
1217: within 19$^\circ$ of the putative CCW disk,
1218: where 19$^\circ$ is the disk thickness proposed
1219: by \citet{paumard06}, and is consistent
1220: with random fluctuations of an isotropic distribution having the
1221: $\vec{n}$-uncertainties shown in Figure \ref{fig:histSolidAngle}.
1222: % Code (python)
1223: % for 34 stars with average solid angle 0.5 sr.
1224: %
1225: % lu.isotropicError(35, 0.5)
1226: %
1227: % Isotropic density is (0.82 +/- 2.4) x 10^-3 stars/deg^2
1228: We estimate that this analysis is capable of revealing, at the 3$\sigma$ level,
1229: a stellar disk with more than 7 stars within a solid angle cone of radius =
1230: 19$^\circ$ at the location of the proposed CCW disk;
1231: thus the proposed CCW disk containing 17 stars as suggested by
1232: \citet{paumard06} should have been detected with this approach.
1233:
1234: There are several principle differences between our analysis and
1235: that in earlier works. First, previous works make the {\it a priori}
1236: assumption that a disk exists through the use of the statistical metric,
1237: K, and the results were not compared to a null hypothesis (i.e. no disk)
1238: to establish the statistical significance of a disk detection.
1239: Furthermore, the K metric used in previous works suffers from a bias which
1240: is described in Appendix \ref{app:kmetric}.
1241: The primary goal of our methodology
1242: is to minimize the number of {\it a priori} assumptions and to
1243: fully quantify the significance of any disk detected as compared to the null
1244: hypothesis that there is no disk. Therefore, we choose to search for disks
1245: using all the young stars rather than first trimming out stars based on
1246: projected angular momentum criteria or radii.
1247: Also, we determine the range of allowed orbital orientations for each
1248: star individually rather than searching for a disk from a statistical
1249: sample of young stars. In this fashion, we utilize not only the
1250: direction information for a velocity vector, as has been used previously,
1251: but also the physical relationships between the magnitude of the velocity
1252: and the positional information. This method allows for no disk to be
1253: detected, while the previously used statistical tests assumed a disk
1254: model and, therefore, must be compared
1255: to the no-disk hypothesis using simulations of isotropic populations.
1256: Without the simulations, the significance of any disk detection via
1257: the K metric cannot be fully quantified. The resulting distribution
1258: of orbits from our analysis is consistent with the
1259: hypothesis of a single, clockwise disk plus a more randomly
1260: distributed population.
1261:
1262: \subsection{Properties of the Clockwise Disk}
1263: \label{sec:diskProperties}
1264:
1265: We now examine, in detail, the properties of the detected clockwise disk.
1266: With the identification of a single stellar disk and a
1267: candidate list of disk members, we investigate the following:
1268: (1) the thickness of the disk,
1269: (2) the radial profile of the disk,
1270: (3) the azimuthal isotropy of the disk,
1271: (4) the eccentricities of stars in the disk,
1272: and (5) the luminosity function of the stars in the disk.
1273: These properties are critical for distinguishing between {\it in situ}
1274: and infalling cluster formation scenarios, as well as for understanding the
1275: dynamical evolution of the young stars both on and off the disk.
1276:
1277: % Numbers from lu.diskThickness() and lu.velDispAnalysis()
1278: The observed disk of young stars has a significant intrinsic thickness;
1279: however, the vertical velocity dispersion is less than previously determined.
1280: To measure the thickness of the disk,
1281: the dispersion of the velocities out of the plane (along the $\vec{n}$
1282: direction) is calculated from all candidate disk members by projecting
1283: each star's three-dimensional velocity vector along the
1284: disk's normal vector to give $v_{\vec{n}}$.
1285: The measurement uncertainties in both $\vec{v}$ and $\vec{n}$
1286: are propagated through this coordinate transformation.
1287: The intrinsic velocity dispersion is calculated using
1288: \begin{eqnarray}
1289: \sigma^2_{\vec{n},intrinsic} & \;=\; &
1290: \sigma^2_{\vec{n},measured} \;-\; \sigma^2_{\vec{n},bias} \\
1291: \sigma^2_{\vec{n},intrinsic} & \;=\; &
1292: \left( \frac{1}{N_{disk-stars} - 1} \right) \left(
1293: \displaystyle\sum_{i=0}^{N_{disk-stars}} v^2_{\vec{n},i} \;-\;
1294: \displaystyle\sum_{i=0}^{N_{disk-stars}} error^2(v_{\vec{n},i})
1295: \right)
1296: \end{eqnarray}
1297: where the bias term, $\sigma_{\vec{v},bias}$, is 19 km s$^{-1}$ and accounts for
1298: added dispersion as a result of uncertainties in the measurements.
1299: The resulting intrinsic velocity dispersion is 28 $\pm$ 6 km s$^{-1}$, which
1300: is significantly different from zero, thus a finite thickness is required.
1301: However, this velocity dispersion is a factor of 2 smaller than
1302: that found using the previously proposed disk solution of \citet{paumard06}
1303: and is slightly smaller than the value reported in \citet{beloborodov06}
1304: due to our improved identification of candidate disk members.
1305: The disk's
1306: thickness can be expressed as the ratio of the vertical scale height to
1307: radius, $h/r = \sigma_{\vec{n},intrinsic} / <|\vec{v}|>$, and is 0.08 $\pm$ 0.02.
1308: Following a similar analysis to \citet{beloborodov06}, but with the above
1309: relationship between $h/r$ and the velocity dispersion, the disk thickness
1310: can also be described using a gaussian distribution of inclination angles
1311: about the disk plane with a standard deviation of
1312: $\Delta\theta$ and is related to the scale height of the disk by
1313: $h/r \sim \sqrt{1/2} \Delta\theta$. This yields a dispersion angle of
1314: $\Delta\theta = 7^\circ \pm 2^\circ$ for the young stellar disk. This
1315: more rigorous determination of the disk thickness is consistent with the
1316: thickness we derived in \S\ref{sec:diskDetect} from the half-width at
1317: half-maximum
1318: of the peak in the density of normal vectors; thus the selection of the
1319: candidate disk members is likely robust.
1320: In comparison, the previously proposed disk solutions
1321: yield a disk thickness of $h/r = 0.2$ ($\Delta\theta = 14^\circ$)
1322: and $h/r = 0.1$ ($\Delta\theta = 9^\circ$)
1323: for \citet{paumard06} and \citet{beloborodov06}, respectively.
1324: We caution that all of these conversions from velocity dispersion to disk
1325: scale height and dispersion angle assume circular orbits and an isothermal
1326: disk structure.
1327: From our analysis, we note that the out-of-the-plane velocity dispersion
1328: shows no statistically significant variation with radius in the disk both
1329: for the primary (difference is $1\sigma \sim$ 7 km s$^{-1}$) and
1330: the full extended samples (difference is $1\sigma \sim$ 14 km s$^{-1}$).
1331: Therefore, the observations are consistent with a thin disk of
1332: uniform velocity dispersion at all radii.
1333:
1334: The surface density of stars in the disk falls off rapidly as a
1335: function of radius. In order to extend the radial coverage, we
1336: consider the entire extended sample in this analysis.
1337: The young stars that are candidate disk members have constraints
1338: on their three-dimensional radii if we limit their orbital solutions
1339: to those close to the disk plane. Thus the disk's surface
1340: density can be determined as a function of three-dimensional radius
1341: rather than just the projected two-dimensional radius as discussed at the
1342: end of \S\ref{sec:diskDetect}.
1343: The distribution for each star's position within the disk plane,
1344: PDF($p$, $q$), is constructed from orbits that are within 10$^\circ$
1345: of the disk and is shown in Figure \ref{fig:diskVelRadius}.
1346: Then the disk's surface density at each radius is computed
1347: numerically by sampling the PDF($p$, $q$) 10$^5$ times for all the candidate
1348: disk members and constructing a radial histogram for each trial.
1349: The radial histograms are combined for all the trials to find the peak
1350: and 68\% confidence bounds for the expected number of stars at each radius.
1351: This is converted into an azimuthally integrated
1352: surface density by dividing by the area of a ring
1353: at each radius. This method of constructing the surface density captures
1354: both the measurement error in the individual stars and the finite
1355: thickness of the disk, which has not been incorporated into previous
1356: estimates. The resulting azimuthally averaged surface density on the
1357: disk is shown for the extended sample in Figure \ref{fig:diskRadialDist}
1358: and has a best-fit power-law profile of $r^{-2.3\pm0.7}$.
1359: This is consistent with the previous results \citep{paumard06},
1360: but our analysis accounts for the uncertainty in each stars
1361: line-of-sight distance due to the finite disk thickness and, therefore
1362: yields a larger uncertainty on the power-law index.
1363:
1364: Visual examination of the stars' positions in the disk plane
1365: (Figure \ref{fig:diskVelRadius}) suggests there may be
1366: some anisotropy as evidenced by the clustering of stars on the lower
1367: part of the disk; however, this over-density is only marginally
1368: statistically significant based on the following analysis.
1369: In order to search for non-uniformities,
1370: we compare the observed stellar surface density of the
1371: extended sample within the disk plane
1372: with the surface density expected for an azimuthally symmetric disk.
1373: The observed stellar surface density is measured by
1374: sampling from all stars' PDF($p$, $q$) for 10$^5$ trials and calculating the
1375: stellar surface density over a grid of points in the disk plane
1376: for each trial. For each point on the
1377: disk plane, the surface densities from all trials are combined, yielding
1378: the most probable surface density with uncertainties.
1379: The resulting two-dimensional map of observed surface densities is then
1380: compared to the expected surface densities for an azimuthally symmetric
1381: disk by subtracting the two values and dividing by the uncertainties.
1382: This produces a surface density excess map that shows the significance
1383: of any excess.
1384: The disk shows a marginally significant ($\sim 3\sigma$) over-density
1385: on the front side (q $<$ 0) of the disk and a corresponding under-density on
1386: the back side (q $>$ 0).
1387:
1388: A few candidate disk stars show evidence for eccentric orbits.
1389: To determine whether any of the stars' eccentricities are consistent
1390: with a circular orbit, the six-dimensional probability density function for
1391: the orbital parameters is marginalized and re-expressed as a PDF for
1392: the eccentricity vector (see Appendix \ref{app:orbits}),
1393: %or Runge-Lenz;
1394: PDF($e_x$, $e_y$, $e_z$). The magnitude of this
1395: vector is the orbital eccentricity and the direction
1396: points along the semi-major axis towards the periapse position.
1397: The PDF for the eccentricity vector cannot be further marginalized to
1398: produce a PDF of the eccentricity magnitude without introducing a bias
1399: due to the positive, definite nature of a vector magnitude. This is the
1400: same bias term as described in the velocity dispersion analysis; however,
1401: unlike the velocities, the eccentricity distributions are strongly
1402: non-gaussian and the bias term cannot be easily accounted for in the
1403: marginalization.
1404: The peak of PDF($e_x$, $e_y$, $e_z$) gives the unbiased orbital
1405: eccentricity and the 99.7\% confidence interval of the three-dimensional
1406: distribution is used to determine the range for the
1407: one-dimensional eccentricity.
1408: Tables \ref{tab:eccDisk} and \ref{tab:eccDiskExtended}
1409: show the 99.7\% confidence range of the eccentricities for all stars
1410: in the primary and extended samples. Also,
1411: Figure \ref{fig:eccentricity} shows the eccentricity 99.7\% confidence
1412: lower limit
1413: for the candidate disk members in red, non-disk members in blue, and excludes
1414: S0-14 (see \S\ref{sec:orbitAnalysis}).
1415: When considering all possible orbital solutions, the resulting eccentricity
1416: ranges show that 2 candidate disk members from the primary sample
1417: have 99.7\% confidence eccentricity lower limits of greater than 0.2.
1418: Restricting the possible orbital solutions to only those
1419: having normal vectors oriented within 10$^\circ$ of the disk normal
1420: vector increases the number to 8 candidate disk members
1421: with 99.7\% confidence eccentricity lower limits larger than 0.2.
1422:
1423: We find high-eccentricity
1424: stars in the disk, similar to the analysis of \citet{beloborodov06} in
1425: which they assumed an infinitely thin disk. However, our analysis
1426: incorporates the finite thickness of the disk and places statistical
1427: errors on the eccentricities for individual stars.
1428: % Old Way -- lu.plotEccAnalytic()
1429: % New Way -- lu.eccVector()
1430: % Table -- lu.tableEccDiskProb()
1431:
1432: The average eccentricity of the entire population is not yet well
1433: constrained.
1434: The eccentricity for the stellar disk is determined using the
1435: eccentricity vector.
1436: For each candidate disk member, orbital solutions are selected
1437: whose normal vectors point within 10$^\circ$ of the disk normal vector.
1438: These orbital solutions are combined for all the disk stars by
1439: averaging their PDFs to create
1440: a combined probability distribution for all stars' eccentricity vectors,
1441: which is then projected into the disk plane and plotted in two dimensions
1442: (Figure \ref{fig:eccVector}).
1443: This two-dimensional probability distribution gives an unbiased estimate
1444: of the eccentricity magnitude
1445: and shows that while the characteristic disk eccentricity peaks
1446: at e=0.22, it is consistent with e=0.0 $-$ 0.8 at the 1$\sigma$ level,
1447: reflecting the large eccentricity uncertainties for the majority of the
1448: candidate disk members.
1449: % lu.diskEccVector()
1450: % lu.eccDistribution()
1451:
1452: \section{Discussion}
1453: \label{sec:discussion}
1454:
1455: The kinematic analysis of the young stars in the central parsec
1456: around our Galaxy's supermassive black hole has implications for the recent
1457: star formation history in this region.
1458: Our first attempt at determining individual orbits for young stars that
1459: reside outside the central arcsecond shows definitive evidence for
1460: the clockwise-rotating disk that was suggested
1461: by \citet{levin03} and was subsequently refined by
1462: \citet{genzel03cusp} and \citet{paumard06}.
1463: Our results do not show a statistically significant second disk.
1464: The presence of a single stellar disk eliminates the need to
1465: invoke two distinct starburst events occuring roughly 6 Myr ago and
1466: greatly simplifies the demands on both {\it in situ} and infalling
1467: cluster scenarios. For instance, in the self-gravitating gas disk scenario,
1468: the detection of only a single stellar disk lifts the requirement
1469: for a second disk to rapidly build up gas, fragment,
1470: and form stars within 1-2 Myr of the formation of the first disk.
1471: Likewise, for the infalling cluster
1472: scenario, the presence of only one stellar disk means that the
1473: frequency of such infall events is half that required for
1474: the existence of two disks.
1475: On the strength of our confirming only one stellar disk, we consider
1476: whether all of the young stars within the central parsec may have formed
1477: in a single burst of star formation.
1478:
1479: Such a scenario must explain not only the observed
1480: clockwise stellar disk, oriented at
1481: $i\sim$115$^\circ$ and $\Omega\sim$100$^\circ$, but also the
1482: presence of roughly half of the young stars from our extended
1483: sample on more isotropically distributed orbits out of the disk.
1484: In the single starburst scenario, the out-of-the-disk stars could either
1485: be generated during the formation process or could intially be in
1486: the disk and then perturbed through subsequent dynamical evolution.
1487: Self-relaxation of the disk has not had sufficient time to produce
1488: the out-of-the-plane population \citep{alexander07imf,cuadra08},
1489: but other mechanisms have been
1490: proposed such as scattering by an inward-migrating IMBH \citep{yu07}.
1491: Currently, our results show that the
1492: on-disk and off-disk populations of young stars look very similar
1493: outside the central arcsecond (0.04 pc)
1494: both in terms of the K-band luminosity function and the surface density
1495: profiles that decreases at larger projected radii as $\propto r^{-2}$.
1496: However, the number of young stars in the disk drops at radii smaller
1497: than 0.08 pc; and at radii of $\lesssim$0.04 pc, none of the observed young
1498: S-stars are in the disk \citep{ghez05orbits,eisenhauer06}.
1499: This drop in the number of disk stars at
1500: small radii may be the result of resonant relaxation or other dynamical
1501: processes if the central arcsecond S-stars are a continuation of the
1502: disk population \citep{hopman06}.
1503: Thus, if dynamical evolution produced the off-disk population, then
1504: the dynamical process must not be a strong function of radius
1505: beyond 0.08 pc.
1506:
1507: Our distributions show that a potential problem with the single starburst
1508: scenario is the presence of the apparent massive star cluster, IRS 13, located
1509: $\sim$4'' from the supermassive black hole \citep{maillard04irs13,schodel05}.
1510: The cluster's orbit is not in the disk plane
1511: and, given the proposed mass of IRS 13 ($>$10$^3$ \msun), it is
1512: unlikely that it could have been ejected from the disk.
1513: However, the definition of IRS 13 as a cluster and the derived
1514: mass is based on observations of only 3-4 bright stars and
1515: is complicated by enhanced dust and gas emission in the vicinity.
1516: More data are needed to determine the total mass of IRS 13 and its
1517: relationship to the disk stars.
1518:
1519: Our results also have implications for the star formation mechanism.
1520: For both infalling cluster and {\it in situ} formation
1521: scenarios, we consider whether the observed characteristics
1522: of the young stellar disk can be explained.
1523: We observe a stellar disk with an out-of-the-disk velocity dispersion
1524: of 28 $\pm$ 6 km s$^{-1}$. Additionally, if we consider only orbital
1525: solutions within the disk (disk prior), we find that at least 8 of the 22
1526: candidate disk stars have 99.7\% confidence lower limits on the
1527: eccentricity of greater than 0.2.
1528: Therefore, any formation scenario should explain not only a single thin
1529: stellar disk but also allow for non-circular stellar orbits of
1530: some stars in the disk.
1531:
1532: First, for the infalling star cluster formation scenario,
1533: some of the disk properties we observe are well explained
1534: and others appear difficult to reconcile with this model.
1535: For instance, eccentric orbits are easily produced.
1536: Stars that are stripped from a cluster as
1537: it spirals in should have a similar inclination and
1538: eccentricity as the cluster itself. Therefore, an infalling cluster
1539: with an initially eccentric orbit will produce a disk of stars with
1540: similarly eccentric orbits \citep{berukoff06}.
1541: Previous studies have observed co-moving clumps of stars,
1542: such as IRS 16SW \citep{lu05irs16sw} and IRS 13 \citep{schodel05},
1543: that appeared to support the infalling cluster formation scenario as
1544: they could be the remnant core of the dissipated cluster.
1545: We tentatively observe evidence for an over-density of stars on the
1546: front half of the disk at the position of the IRS 16SW co-moving group.
1547: However this over-density may be explained by the effects of
1548: extinction that reduces the number of young stars identified
1549: on the back half of the disk at a given magnitude.
1550: The extinction is highly variable throughout the region
1551: and the back half of the disk is behind a patch of higher exctinction
1552: \citep[$\Delta$A$_K$ = 0.3 - 1.4; ][]{scoville03,schodel07}.
1553: Thus the apparent overdensity on the front half of the disk,
1554: corresponding to the IRS 16SW co-moving group, can perhaps be ascribed to
1555: differential extinction.
1556: More data are needed to confirm the observed disk asymmetry and to determine
1557: whether the cause is extinction.
1558: Our results yield a steep radial profile for the young stars in
1559: the disk, as also found by \citet{paumard06}, which appears to be
1560: inconsistent with the flatter profile expected for an infalling cluster
1561: \citep[$r^{-0.75}$, ][]{berukoff06}.
1562: We note that mass segregation is observed in massive star
1563: clusters that are only a few million years old
1564: \citep{hillenbrand98,fischer98,stolte06}. Any mass segregation that existed
1565: prior to the cluster's dissolution may impact the observed radial profile
1566: as the massive stars would have resided preferentially in the cluster core
1567: and would therefore have been deposited at the smallest radii.
1568: Thus, the massive stars O stars that we observe today
1569: may have a steeper radial profile than the entire young star population.
1570: Additionally, the lack of X-ray emission from pre-main-sequence stars
1571: \citep{nayakshinSunyaev06} is not well explained by an infalling cluster
1572: model.
1573: A larger and deeper survey for young stars over the central $\sim$5 pc could
1574: definitively rule out this scenario if the tidal tails of the disrupted
1575: clusters are not detected.
1576:
1577: Some theories of {\it in situ} star formation
1578: take place in a circular gas disk. Such a gas disk
1579: can be built up from a steady inward migration of
1580: material or from many small cloud-infall events and the disk
1581: would circularize prior to becoming massive enough to form stars from
1582: self-gravity ($>10^4$ \msun). Such a formation scenario would most
1583: likely produce a steep radial profile in agreement with our observations.
1584: Our observations of over 30\% of the
1585: candidate disk members with eccentricities greater than 0.2 appears to be
1586: inconsistent with an initially circular disk of stars and a normal
1587: initial mass function.
1588: A disk of stars on initially circular orbits and with a normal IMF
1589: will relax over 6 Myr and produce a thermal distribution of
1590: eccentricities with an rms eccentricity of 0.15 or less \citep{alexander07imf}.
1591: For such a disk, only 4 out of 22 stars should have eccentricities higher than
1592: 0.2, compared with the 8 out of 22 observed when a disk prior
1593: is imposed on the primary sample.
1594: Therefore, in order for the disk to have been initially circular with
1595: a normal-IMF, some additional dynamical processing other than self-relaxation
1596: is needed. Other possibilities are that the initial mass function may
1597: have been top-heavy, the binary fraction may have been extremely high,
1598: or IMBHs could have formed, all resulting in faster relaxation
1599: to higher eccentricities, but these are not
1600: sufficient to explain the out-of-the-disk population of young stars
1601: \citep{alexander07imf,cuadra08}. The gas disk formation scenario may be
1602: modified \citep{alexander07imf,cuadra08} to accommodate the observed high
1603: stellar eccentricities and out-of-the plane population by
1604: building up a massive gas disk in a single cloud infall or a cloud-cloud
1605: collision event, in which the clouds are on eccentric orbits
1606: \citep{sanders98,vollmerDuschl01,nayakshin07sims}.
1607: The gas disk would then have a high eccentricity for a short period
1608: of time during which stars might form \citep{alexander08,bonnell08}.
1609: The cloud-cloud collision scenario may yield both a thin stellar disk and
1610: a more distributed population of stars at larger radii
1611: with a range of angular momenta as a result of the complex interactions
1612: and shocks during the collision.
1613: It is also conceivable that a cloud-cloud
1614: collision scenario might give rise to out-of-the-disk clumps of gas
1615: that could form a cluster such as IRS 13.
1616: Refined estimates of the eccentricity and inclination distributions
1617: of the young stars and more detailed theoretical
1618: analysis are needed to investigate the viability of this scenario.
1619:
1620:
1621: \section{Conclusions}
1622:
1623: In summary, the advent of laser guide star adaptive optics has allowed
1624: us to retroactively improve our 11 year astrometric data set used for
1625: monitoring stars orbiting our Galactic Center. This has increased our
1626: proper motion precision, with resulting uncertainties of $\sim$3 km s$^{-1}$,
1627: and allowed us, for the first time, to make measurements of and place
1628: limits on accelerations for stars outside the central arcsecond
1629: out to a radius of 3\farcs5, with typical 3$\sigma$ acceleration
1630: limits of -0.19 mas yr$^{-2}$.
1631: By combining our improved stellar positions and proper motions with radial
1632: velocity information from the literature, we compute orbits
1633: for individual young stars proposed to lie in stellar disks orbiting
1634: the supermassive black hole.
1635: The orbits for the young stars confirm only a single disk
1636: of young stars at a high inclination rotating in a clockwise sense and
1637: there is no statistically significant evidence for a second disk.
1638: Stars within the well-defined, clockwise disk
1639: have an out-of-the-disk velocity dispersion of 28 $\pm$ 6 km s$^{-1}$
1640: and several stars have high eccentricities. These disk properties
1641: suggest that star formation may have occurred in a single event, rather
1642: than the two events previously needed to explain two stellar disks; however,
1643: there are open questions as to how $\sim$50\% of all young stars can be
1644: perturbed out of the disk plane and whether
1645: the apparent compact cluster, IRS 13, which is not part of the stellar
1646: disk, requires a separate star formation or dynamical event.
1647: Future directions include (1) obtaining new LGSAO
1648: data sets with improved astrometry to measure accelerations for the
1649: young stars at all radii and (2) identifying new young stars within
1650: the central parsec in order to better constrain the orbital
1651: properties of these stars and to study in detail the
1652: distribution of eccentricities and semi-major axes for stars both in and
1653: out of the disk.
1654:
1655: \acknowledgements
1656:
1657: Support for this work was provided by NSF grant AST-0406816, and
1658: the NSF Science \& Technology Center for AO, managed by UCSC
1659: (AST-9876783).
1660: Additional support for J.R.L. was provided by
1661: a NSF Graduate Research Fellowship.
1662: We would like to thank Brad Hansen
1663: and the anonymous referee for helpful comments.
1664: The W.M.~Keck Observatory is operated as a scientific
1665: partnership among the California Institute of Technology, the University
1666: of California and the National Aeronautics and Space Administration.
1667: The Observatory was made possible by the generous financial support of
1668: the W.M.~Keck Foundation.
1669:
1670: {\it Facilities:} \facility{Keck:II (NIRC2)}, \facility{Keck:I (NIRC)}
1671:
1672:
1673: %---------------
1674: %
1675: % Appendices
1676: %
1677: %---------------
1678: \begin{appendix}
1679:
1680: \section{NIRC Speckle Distortion}
1681: \label{app:speckDistort}
1682:
1683: In the speckle data sets, optical distortions, introduced by the NIRC
1684: reimager, are small near the center of the field-of-view where
1685: Sgr A* was positioned, but grow to dominate the positional
1686: uncertainties for stars located more than $\sim$0\farcs5 from Sgr A*
1687: (see Figure \ref{fig:distort} and \S\ref{sec:images}).
1688: Now, utilizing images of the Galactic Center obtained with NIRC2,
1689: which has optical distortions characterized at the $\sim$2 mas level
1690: \citep{ghez08},
1691: we can, for the first time, similarly quantify and correct
1692: the optical distortions in the NIRC reimager speckle data sets.
1693:
1694: Images of the Galactic Center were obtained with both NIRC and NIRC2 on
1695: consecutive nights during July 2004
1696: and the NIRC2 images were used as a reference coordinate system.
1697: The individual NIRC speckle exposure times are only
1698: 0.1 seconds and have insufficient signal-to-noise to detect more than the
1699: brightest 5 stars. Exposures were obtained in sets of 100 and each set is
1700: combined to produce a single image in which approximatly 100 stars are
1701: detected. It is assumed that the images are mostly stationary on the
1702: NIRC detector during each set of exposures. For each stacked image, the stars'
1703: positions are compared to those in the NIRC2 image and the offsets are
1704: mapped into NIRC detector coordinates
1705: (see Figure \ref{fig:distort2d}, {\it left}).
1706: In this fashion, a distortion map is built up from many stacks of images
1707: which are dithered and rotated such that stars fall on many different
1708: positions on the detector.
1709: The distortion solution was obtained by fitting the distortion map with
1710: polynomials of the form
1711: \begin{equation}
1712: (x^\prime + 128) = a_0 + a_1(x - 128) + a_2(y - 128)
1713: \end{equation}
1714: \begin{equation}
1715: (y^\prime + 128) = b_0 + b_1(x - 128) + b_2(y - 128)
1716: \end{equation}
1717: where the best-fit distortion parameters are listed in Table
1718: \ref{tab:distort}.
1719: The new distortion solution improves the RMS residual errors
1720: per stack by a factof of 3 to $\sim$3 mas
1721: (Figure \ref{fig:speck_distort}), which is further reduced in the
1722: final image by averaging the dithered stacks.
1723: % The fit had a reduced $\chi^2$=0.9 and was computed
1724: % by assuming a centroiding uncertainty in a single stack
1725: % of 0.5 pixels (10 mas) as was measured for the sources
1726: % in the central 50$\times$50 pixel box where distortion is minimal.
1727: Higher-order polynomial terms did not sufficiently improve the
1728: fit to warrant inclusion. The above solution is applied
1729: after the initial application of the standard NIRC distortion correction.
1730: The map of positional differences between stars in the NIRC and NIRC2
1731: images before and after the NIRC-reimager distortion correction is shown
1732: in Figure \ref{fig:distort2d} ({\it right}).
1733: The resulting radial dependence on the RMS
1734: positional uncertainty is greatly improved and is shown in
1735: Figure \ref{fig:distort}, which plots many stars' RMS residual offset from
1736: their best-fit proper motions across all epochs. In the final analysis
1737: of the speckle data, the relative astrometric uncertainty is $\sim$2 mas.
1738:
1739: \section{Analytic Orbit Equations}
1740: \label{app:orbits}
1741:
1742: The orbit of a star in a known point source potential can be derived from a
1743: single measurement of a star's orbital state vector.
1744: At epoch t$_{ref}$, the orbital state vector is usually
1745: described by the star's position, $\vec{r}$, and velocity, $\vec{v}$,
1746: relative to the central mass. For the analysis in this paper,
1747: the state vector is estimated using measurements of the three-dimensional
1748: velocity, $\vec{v} = [v_x, v_y, v_z]$, and the
1749: projected position, $\vec{r}_{2D} = [x, y]$, and $z$ is derived from
1750: the radial acceleration on the plane of the sky. For brevity, we have
1751: removed the $ref$ subscript notation and all of the above variables are
1752: measured at t$_{ref}$.
1753: Orbital trajectories are then inferred from conservation of
1754: energy, specific angular momentum, and eccentricity
1755: ($\epsilon$, $\vec{h}$, $\vec{e}$),
1756: which are related by $\vec{e} \cdot \vec{h} = 0$ and
1757: $|e|^2 - 1 = 2\,\epsilon\,h^2 / GM$ giving 5 constants of motion
1758: plus an undetermined reference time.
1759: %% Above holds for ANY conserved 1/r^2 potential.
1760: %% In a perturbed potential where E and \vec{L} are still
1761: %% conserved, then only e^2 is conserved rather than \vec{e}
1762: Equivalently, the orbital trajectory can
1763: be expressed using the standard Keplerian orbital elements:
1764: period ($P$), eccentricity ($e$), time of periapse passage (T$_{\circ}$),
1765: inclination ($i$), position angle of the ascending node ($\Omega$),
1766: and the longitude of periapse
1767: \citep[$\omega$; see Equations \ref{eqn:i}, \ref{eqn:e}, \ref{eqn:w},
1768: \ref{eqn:o}, \ref{eqn:p}, \ref{eqn:t0} and][for detailed descriptions of these orbital parameters]{ghez05orbits}.
1769: The 3D position and velocity state vectors can be used to calculate
1770: the orbit of the star around the black hole (by algebraic manipulation of
1771: Kepler's Laws).
1772: Here we present the analytic expressions used to compute the orbital elements
1773: from the state vectors.
1774:
1775: Orbit determination for the young stars in our sample is tractable because
1776: the mass and position of the black hole are determined by independent means,
1777: namely the well determined orbits of stars much closer to the black hole.
1778: The coordinate system is set such that Sgr A* resides at the origin,
1779: $\hat{x}$ and $\hat{y}$ increase with right ascension and declination,
1780: and $\hat{z}$ increases with the line-of-sight distance from the Earth
1781: to Sgr A* with z=0 at the location of the black hole.
1782: Combining the two state vectors, $\vec{r}$ and $\vec{v}$, and the
1783: black hole mass, there are three intermediate vectors that describe the
1784: geometry of the orbit both in three-dimensions and projected onto the plane
1785: of the sky. These are (1) the specific angular momentum vector,
1786: $\vec{h}$, which points normal to the plane of the orbit, (2) the
1787: eccentricity vector, $\vec{e}$, which points in the direction of periapse,
1788: and (3) the ascending node vector, $\vec{\Omega}$, which points to where
1789: the star passes through the plane of the sky moving away from us,
1790: and are given by
1791: \begin{eqnarray}
1792: \vec{h} & = & \vec{r} \times \vec{v} \\
1793: \vec{e} & = & \frac{\vec{v} \times \vec{h}}{GM} - \frac{\vec{r}}{|\vec{r}|} \\
1794: \vec{\Omega} & = & \vec{h} \times \hat{z}.
1795: \end{eqnarray}
1796: The semi-major axis can also be calculated as an intermediate quantity
1797: \begin{eqnarray}
1798: a & = & \left ( \frac{2}{|\vec{r}|} - \frac{|\vec{v}|^2}{GM} \right )^{-1}.
1799: \end{eqnarray}
1800: Then the five standard orbital parameters that describe the shape and
1801: period of the orbit are then
1802: \begin{eqnarray}
1803: i & = & \arccos{ \left ( \frac{-\vec{h} \cdot \hat{z}}{|\vec{h}|} \right )}
1804: \label{eqn:i} \\
1805: e & = & |\vec{e}|
1806: \label{eqn:e} \\
1807: \omega & = & \arccos \left ( \frac{(\hat{z} \times \vec{h}) \cdot \vec{e}}
1808: {|\hat{z} \times \vec{h}||\vec{e}|} \right ) \quad
1809: (\textrm{if } \vec{e} \cdot \hat{z} < 0 \textrm{ then }
1810: \omega = 2\pi - \omega)
1811: \label{eqn:w} \\
1812: \Omega & = & \arctan \left ( \frac{\vec{\Omega} \cdot \hat{x}}
1813: {\vec{\Omega} \cdot \hat{y}} \right )
1814: \label{eqn:o} \\
1815: \left ( \frac{P}{[yr]} \right ) & = & \sqrt{ \left ( \frac{a}{[AU]} \right )^3
1816: \left ( \frac{[M_\sun]}{M} \right ) }
1817: \label{eqn:p} \\
1818: \end{eqnarray}
1819: where i = 0 if the orbit is in the plane of the sky and $\Omega$ is
1820: measured East ($\hat{x}$) of North ($\hat{y}$).
1821: The remaining orbital parameter is the epoch of
1822: periapse passage and can be computed in a number of different ways.
1823: We first compute several intermediate quantities of interest such as the
1824: Thiele-Innes constants (A,B,C,F,G,H), and the eccentric anomaly as shown below:
1825: \begin{eqnarray}
1826: A & = & a(\cos \omega \cos \Omega - \sin \omega \sin \Omega \cos i) \\
1827: B & = & a(\cos \omega \sin \Omega + \sin \omega \cos \Omega \cos i) \\
1828: F & = & a(-\sin \omega \cos \Omega - \cos \omega \sin \Omega \cos i) \\
1829: G & = & a (-\sin \omega \sin \Omega + \cos \omega \cos \Omega \cos i) \\
1830: \cos E & = & \frac{Gr_y - Fr_x}{AG - BF} + e \\
1831: \sin E & = & \frac{Ar_x - Br_y}{AG - BF} \frac{1}{\sqrt{ 1 - e^2 }} \\
1832: E & = & \arctan \left( \frac{\sin E}{\cos E} \right ).
1833: \end{eqnarray}
1834: %C & = & a( \sin \omega \sin i) \\
1835: %H & = & a(\cos \omega \sin i) \\
1836: And finally the epoch of periapse passage are calculated from
1837: these intermediate quantities using
1838: \begin{eqnarray}
1839: T_o & = & t_{ref} - \frac{P}{2\pi} (E - e \sin E).
1840: \label{eqn:t0}
1841: \end{eqnarray}
1842:
1843: \section{K Metric}
1844: \label{app:kmetric}
1845:
1846: The previously proposed planes were derived by minimizing a metric that
1847: \citet{levin03} call $\chi^2$, but we call K, and which is defined as
1848: \begin{equation}
1849: K = \frac{1}{N-1}\displaystyle\sum^N_{i=1}
1850: \frac{(\vec{n}\cdot\vec{v_i})^2}{(n_x\sigma_{v_{x,i}})^2 +
1851: (n_y\sigma_{v_{y,i}})^2 + (n_z\sigma_{v_{z,i}})^2}
1852: \end{equation}
1853: where $N$ is the number of stars, $\vec{v}_i$ is the velocity of each star,
1854: $\sigma_{v_{x,i}}$, $\sigma_{v_{y,i}}$, $\sigma_{v_{z,i}}$ are the velocity
1855: uncertainties for each star, and $\vec{n}$ is the normal vector to the disk
1856: plane that is found in the fitting process.
1857: This metric is used to find, statistically, the best-fit common orbital
1858: plane from the velocity vectors of a sample of stars.
1859: The K metric suffers from several shortcomings.
1860: First, the K metric is described as a $\chi^2$ metric; however,
1861: standard $\chi^2$ minimization takes the form of
1862: (data - model)$^2$/(data errors)$^2$ where the data errors have no
1863: dependency on the model parameters. The K metric includes the model
1864: parameters in the data-error term and does not necessarily have an
1865: expectation value of 1 for normal errors. The appropriate function
1866: to minimize in order to find the best-fit common orbital plane
1867: can be derived from maximum likelihood theory if we assume that the
1868: likelihood function is given by
1869: \begin{equation}
1870: L = \prod_{i=1}^N \frac{1}{\sqrt{2 \pi \sigma_i^2}}
1871: \exp{ \left [ - \frac{(\vec{n} \cdot \vec{v}_i)^2}{2 \sigma_i^2} \right ]}
1872: \end{equation}
1873: where $\sigma_i$ depends on the disk model parameters that are
1874: being sought by
1875: \begin{equation}
1876: \sigma_i^2 = (n_x \sigma_{v_{x,i}})^2 +
1877: (n_y\sigma_{v_{y,i}})^2 + (n_z\sigma_{v_{z,i}})^2.
1878: \end{equation}
1879: Standard practice is then to take the logarithm of the likelihood, $L$, and
1880: minimize the resulting function in Equation \ref{eq:minfunc} in order to find
1881: the best fit disk model parameters. The above likelihood function then becomes
1882: \begin{eqnarray}
1883: \ln L & = & -\frac{N}{2} \ln (2\pi) - \sum_{i=1}^N \ln \sigma_i
1884: + \sum_{i=1}^N -\frac{ (\vec{n} \cdot \vec{v}_i)^2 }{ 2 \sigma_i^2 } \\
1885: -2 \ln L & = & N \ln (2 \pi) + 2 \sum_{i=1}^N \ln \sigma_i +
1886: \sum_{i=1}^N \frac{ (\vec{n} \cdot \vec{v}_i)^2 }{\sigma_i^2} \label{eq:minfunc}
1887: \end{eqnarray}
1888: and the first two terms are constant and do not factor into finding an
1889: extremum in the above equation. The third term on the right-hand side is the
1890: K metric previously used to determing the disk parameters. However, the
1891: second term on the right-hand side also depends on the free parameters
1892: in $\vec{n}$ and must be included in the minimization process.
1893: This extra term that has not
1894: previously been included in the disk fitting process has the full form
1895: \begin{equation}
1896: \ln{ \sqrt{(n_x\sigma_{v_{x,i}})^2 + (n_y\sigma_{v_{y,i}})^2 +
1897: (n_z\sigma_{v_{z,i}})^2}}
1898: \end{equation}
1899: and standard chi-squared probability functions cannot be applied.
1900: Second, even when accounting for the extra term, the metric can still
1901: introduce substantial bias. In particular, radial
1902: velocity uncertainties, $\sigma_{v_{z,i}}$, are larger than the proper motion
1903: errors by a factor of 2 on average in previous publications. During
1904: K-minimization, this over-weights solutions with a larger $n_z$ resulting
1905: in a bias against edge-on planes. Finally, in order to properly
1906: evaluate the probability of obtaining a given
1907: value of the K-metric by random chance, one must perform simulations of
1908: an isotropic distribution of stars. However, such simulations are extremely
1909: sensitive to the input distribution of semi-major axes and eccentricities
1910: which are not yet well constrained by observations.
1911: Thus, when utilizing such statistical tests for finding a common orbital
1912: plane, it is difficult to compare to the null hypothesis -- an isotropic
1913: distribution of stars -- and to quantify the significance of a disk.
1914:
1915: \end{appendix}
1916:
1917: \begin{thebibliography}{78}
1918: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
1919:
1920: \bibitem[{{Alexander} {et~al.}(2008){Alexander}, {Armitage}, {Cuadra}, \&
1921: {Begelman}}]{alexander08}
1922: {Alexander}, R.~D., {Armitage}, P.~J., {Cuadra}, J., \& {Begelman}, M.~C. 2008,
1923: \apj, 674, 927
1924:
1925: \bibitem[{{Alexander} {et~al.}(2007){Alexander}, {Begelman}, \&
1926: {Armitage}}]{alexander07imf}
1927: {Alexander}, R.~D., {Begelman}, M.~C., \& {Armitage}, P.~J. 2007, \apj, 654,
1928: 907
1929:
1930: \bibitem[{{Alexander}(2005)}]{alexander05review}
1931: {Alexander}, T. 2005, \physrep, 419, 65
1932:
1933: \bibitem[{{Alexander} \& {Morris}(2003)}]{alexander03}
1934: {Alexander}, T., \& {Morris}, M. 2003, \apjl, 590, L25
1935:
1936: \bibitem[{{Allen} {et~al.}(1990){Allen}, {Hyland}, \& {Hillier}}]{allen90}
1937: {Allen}, D.~A., {Hyland}, A.~R., \& {Hillier}, D.~J. 1990, \mnras, 244, 706
1938:
1939: \bibitem[{{Babu} \& {Feigelson}(1996)}]{astrostats}
1940: {Babu}, G.~J., \& {Feigelson}, E.~D. 1996, {Astrostatistics} (Astrostatistics
1941: by G.J.~Babu and E.D.~Feigelson.~London: Chapman and Hall, 1996.)
1942:
1943: \bibitem[{{Beloborodov} {et~al.}(2006){Beloborodov}, {Levin}, {Eisenhauer},
1944: {Genzel}, {Paumard}, {Gillessen}, \& {Ott}}]{beloborodov06}
1945: {Beloborodov}, A.~M., {Levin}, Y., {Eisenhauer}, F., {Genzel}, R., {Paumard},
1946: T., {Gillessen}, S., \& {Ott}, T. 2006, \apj, 648, 405
1947:
1948: \bibitem[{{Bender} {et~al.}(2005){Bender}, {Kormendy}, {Bower}, {Green},
1949: {Thomas}, {Danks}, {Gull}, {Hutchings}, {Joseph}, {Kaiser}, {Lauer},
1950: {Nelson}, {Richstone}, {Weistrop}, \& {Woodgate}}]{bender05}
1951: {Bender}, R., {Kormendy}, J., {Bower}, G., {Green}, R., {Thomas}, J., {Danks},
1952: A.~C., {Gull}, T., {Hutchings}, J.~B., {Joseph}, C.~L., {Kaiser}, M.~E.,
1953: {Lauer}, T.~R., {Nelson}, C.~H., {Richstone}, D., {Weistrop}, D., \&
1954: {Woodgate}, B. 2005, \apj, 631, 280
1955:
1956: \bibitem[{{Berukoff} \& {Hansen}(2006)}]{berukoff06}
1957: {Berukoff}, S.~J., \& {Hansen}, B.~M.~S. 2006, \apj, 650, 901
1958:
1959: \bibitem[{{Blum} {et~al.}(1995){Blum}, {Depoy}, \& {Sellgren}}]{blum95heI}
1960: {Blum}, R.~D., {Depoy}, D.~L., \& {Sellgren}, K. 1995, \apj, 441, 603
1961:
1962: \bibitem[{{Bonnell} \& {Rice}(2008)}]{bonnell08}
1963: {Bonnell}, I.~A., \& {Rice}, W.~K.~M. 2008, Science, 321 (5892), 1060
1964:
1965: \bibitem[{{Cuadra} {et~al.}(2008){Cuadra}, {Armitage}, \&
1966: {Alexander}}]{cuadra08}
1967: {Cuadra}, J., {Armitage}, P.~J., \& {Alexander}, R.~D. 2008, \mnras, 388, L64
1968:
1969: \bibitem[{{Davies} {et~al.}(1998){Davies}, {Blackwell}, {Bailey}, \&
1970: {Sigurdsson}}]{davies98}
1971: {Davies}, M.~B., {Blackwell}, R., {Bailey}, V.~C., \& {Sigurdsson}, S. 1998,
1972: \mnras, 301, 745
1973:
1974: \bibitem[{{Davies} \& {King}(2005)}]{davies05}
1975: {Davies}, M.~B., \& {King}, A. 2005, \apjl, 624, L25
1976:
1977: \bibitem[{{Diolaiti} {et~al.}(2000){Diolaiti}, {Bendinelli}, {Bonaccini},
1978: {Close}, {Currie}, \& {Parmeggiani}}]{starfinder}
1979: {Diolaiti}, E., {Bendinelli}, O., {Bonaccini}, D., {Close}, L., {Currie}, D.,
1980: \& {Parmeggiani}, G. 2000, \aaps, 147, 335
1981:
1982: \bibitem[{{Dressler}(1980)}]{nearestNeighbor}
1983: {Dressler}, A. 1980, \apj, 236, 351
1984:
1985: \bibitem[{{Eckart} \& {Genzel}(1996)}]{eckart96}
1986: {Eckart}, A., \& {Genzel}, R. 1996, \nat, 383, 415
1987:
1988: \bibitem[{{Eckart} {et~al.}(2002){Eckart}, {Genzel}, {Ott}, \&
1989: {Sch{\"o}del}}]{eckart02}
1990: {Eckart}, A., {Genzel}, R., {Ott}, T., \& {Sch{\"o}del}, R. 2002, \mnras, 331,
1991: 917
1992:
1993: \bibitem[{{Eisenhauer} {et~al.}(2005){Eisenhauer}, {Genzel}, {Alexander},
1994: {Abuter}, {Paumard}, {Ott}, {Gilbert}, {Gillessen}, {Horrobin}, {Trippe},
1995: {Bonnet}, {Dumas}, {Hubin}, {Kaufer}, {Kissler-Patig}, {Monnet},
1996: {Str{\"o}bele}, {Szeifert}, {Eckart}, {Sch{\"o}del}, \&
1997: {Zucker}}]{eisenhauer06}
1998: {Eisenhauer}, F., {Genzel}, R., {Alexander}, T., {Abuter}, R., {Paumard}, T.,
1999: {Ott}, T., {Gilbert}, A., {Gillessen}, S., {Horrobin}, M., {Trippe}, S.,
2000: {Bonnet}, H., {Dumas}, C., {Hubin}, N., {Kaufer}, A., {Kissler-Patig}, M.,
2001: {Monnet}, G., {Str{\"o}bele}, S., {Szeifert}, T., {Eckart}, A.,
2002: {Sch{\"o}del}, R., \& {Zucker}, S. 2005, \apj, 628, 246
2003:
2004: \bibitem[{{Fischer} {et~al.}(1998){Fischer}, {Pryor}, {Murray}, {Mateo}, \&
2005: {Richtler}}]{fischer98}
2006: {Fischer}, P., {Pryor}, C., {Murray}, S., {Mateo}, M., \& {Richtler}, T. 1998,
2007: \aj, 115, 592
2008:
2009: \bibitem[{{Fruchter} \& {Hook}(2002)}]{drizzle}
2010: {Fruchter}, A.~S., \& {Hook}, R.~N. 2002, \pasp, 114, 144
2011:
2012: \bibitem[{{Genzel} {et~al.}(2000){Genzel}, {Pichon}, {Eckart}, {Gerhard}, \&
2013: {Ott}}]{genzel00}
2014: {Genzel}, R., {Pichon}, C., {Eckart}, A., {Gerhard}, O.~E., \& {Ott}, T. 2000,
2015: \mnras, 317, 348
2016:
2017: \bibitem[{{Genzel} {et~al.}(2003){Genzel}, {Sch{\" o}del}, {Ott}, {Eisenhauer},
2018: {Hofmann}, {Lehnert}, {Eckart}, {Alexander}, {Sternberg}, {Lenzen}, {Cl{\'
2019: e}net}, {Lacombe}, {Rouan}, {Renzini}, \& {Tacconi-Garman}}]{genzel03cusp}
2020: {Genzel}, R., {Sch{\" o}del}, R., {Ott}, T., {Eisenhauer}, F., {Hofmann}, R.,
2021: {Lehnert}, M., {Eckart}, A., {Alexander}, T., {Sternberg}, A., {Lenzen}, R.,
2022: {Cl{\' e}net}, Y., {Lacombe}, F., {Rouan}, D., {Renzini}, A., \&
2023: {Tacconi-Garman}, L.~E. 2003, \apj, 594, 812
2024:
2025: \bibitem[{{Genzel} {et~al.}(1996){Genzel}, {Thatte}, {Krabbe}, {Kroker}, \&
2026: {Tacconi-Garman}}]{genzel96}
2027: {Genzel}, R., {Thatte}, N., {Krabbe}, A., {Kroker}, H., \& {Tacconi-Garman},
2028: L.~E. 1996, \apj, 472, 153
2029:
2030: \bibitem[{{Gerhard}(2001)}]{gerhard01}
2031: {Gerhard}, O. 2001, \apjl, 546, L39
2032:
2033: \bibitem[{{Ghez} {et~al.}(2003){Ghez}, {Duch{\^ e}ne}, {Matthews}, {Hornstein},
2034: {Tanner}, {Larkin}, {Morris}, {Becklin}, {Salim}, {Kremenek}, {Thompson},
2035: {Soifer}, {Neugebauer}, \& {McLean}}]{ghez03spec}
2036: {Ghez}, A.~M., {Duch{\^ e}ne}, G., {Matthews}, K., {Hornstein}, S.~D.,
2037: {Tanner}, A., {Larkin}, J., {Morris}, M., {Becklin}, E.~E., {Salim}, S.,
2038: {Kremenek}, T., {Thompson}, D., {Soifer}, B.~T., {Neugebauer}, G., \&
2039: {McLean}, I. 2003, \apjl, 586, L127
2040:
2041: \bibitem[{{Ghez} {et~al.}(2005{\natexlab{a}}){Ghez}, {Hornstein}, {Lu},
2042: {Bouchez}, {Le Mignant}, {van Dam}, {Wizinowich}, {Matthews}, {Morris},
2043: {Becklin}, {Campbell}, {Chin}, {Hartman}, {Johansson}, {Lafon}, {Stomski}, \&
2044: {Summers}}]{ghez05lgs}
2045: {Ghez}, A.~M., {Hornstein}, S.~D., {Lu}, J.~R., {Bouchez}, A., {Le Mignant},
2046: D., {van Dam}, M.~A., {Wizinowich}, P., {Matthews}, K., {Morris}, M.,
2047: {Becklin}, E.~E., {Campbell}, R.~D., {Chin}, J.~C.~Y., {Hartman}, S.~K.,
2048: {Johansson}, E.~M., {Lafon}, R.~E., {Stomski}, P.~J., \& {Summers}, D.~M.
2049: 2005{\natexlab{a}}, \apj, 635, 1087
2050:
2051: \bibitem[{{Ghez} {et~al.}(1998){Ghez}, {Klein}, {Morris}, \&
2052: {Becklin}}]{ghez98pm}
2053: {Ghez}, A.~M., {Klein}, B.~L., {Morris}, M., \& {Becklin}, E.~E. 1998, \apj,
2054: 509, 678
2055:
2056: \bibitem[{{Ghez} {et~al.}(2000){Ghez}, {Morris}, {Becklin}, {Tanner}, \&
2057: {Kremenek}}]{ghez00nat}
2058: {Ghez}, A.~M., {Morris}, M., {Becklin}, E.~E., {Tanner}, A., \& {Kremenek}, T.
2059: 2000, \nat, 407, 349
2060:
2061: \bibitem[{{Ghez} {et~al.}(2005{\natexlab{b}}){Ghez}, {Salim}, {Hornstein},
2062: {Tanner}, {Lu}, {Morris}, {Becklin}, \& {Duch{\^e}ne}}]{ghez05orbits}
2063: {Ghez}, A.~M., {Salim}, S., {Hornstein}, S.~D., {Tanner}, A., {Lu}, J.~R.,
2064: {Morris}, M., {Becklin}, E.~E., \& {Duch{\^e}ne}, G. 2005{\natexlab{b}},
2065: \apj, 620, 744
2066:
2067: \bibitem[{{Ghez} {et~al.}(2008){Ghez}, {Salim}, {Weinberg}, {Lu}, {Do}, {Dunn},
2068: {Matthews}, {Morris}, {Yelda}, {Becklin}, {Kremenek}, {Milosavljevic}, \&
2069: {Naiman}}]{ghez08}
2070: {Ghez}, A.~M., {Salim}, S., {Weinberg}, N.~N., {Lu}, J.~R., {Do}, T., {Dunn},
2071: J.~K., {Matthews}, K., {Morris}, M., {Yelda}, S., {Becklin}, E.~E.,
2072: {Kremenek}, T., {Milosavljevic}, M., \& {Naiman}, J. 2008, \apj, submitted
2073:
2074: \bibitem[{{Goodman}(2003)}]{goodman03}
2075: {Goodman}, J. 2003, \mnras, 339, 937
2076:
2077: \bibitem[{{G{\'o}rski} {et~al.}(2005){G{\'o}rski}, {Hivon}, {Banday},
2078: {Wandelt}, {Hansen}, {Reinecke}, \& {Bartelmann}}]{healpix}
2079: {G{\'o}rski}, K.~M., {Hivon}, E., {Banday}, A.~J., {Wandelt}, B.~D., {Hansen},
2080: F.~K., {Reinecke}, M., \& {Bartelmann}, M. 2005, \apj, 622, 759
2081:
2082: \bibitem[{{G{\"u}rkan} \& {Rasio}(2005)}]{gurkan05}
2083: {G{\"u}rkan}, M.~A., \& {Rasio}, F.~A. 2005, \apj, 628, 236
2084:
2085: \bibitem[{{Hansen} \& {Milosavljevi{\' c}}(2003)}]{hansen03}
2086: {Hansen}, B.~M.~S., \& {Milosavljevi{\' c}}, M. 2003, \apjl, 593, L77
2087:
2088: \bibitem[{{Hillenbrand} \& {Hartmann}(1998)}]{hillenbrand98}
2089: {Hillenbrand}, L.~A., \& {Hartmann}, L.~W. 1998, \apj, 492, 540
2090:
2091: \bibitem[{{Hopman} \& {Alexander}(2006)}]{hopman06}
2092: {Hopman}, C., \& {Alexander}, T. 2006, \apj, 645, 1152
2093:
2094: \bibitem[{{Hornstein}(2007)}]{sethThesis}
2095: {Hornstein}, S.~D. 2007, PhD thesis, UCLA
2096:
2097: \bibitem[{{Kim} {et~al.}(2004){Kim}, {Figer}, \& {Morris}}]{kim04}
2098: {Kim}, S.~S., {Figer}, D.~F., \& {Morris}, M. 2004, \apjl, 607, L123
2099:
2100: \bibitem[{{Kim} \& {Morris}(2003)}]{kim03}
2101: {Kim}, S.~S., \& {Morris}, M. 2003, \apj, 597, 312
2102:
2103: \bibitem[{{Kolykhalov} \& {Syunyaev}(1980)}]{kolykhalov80}
2104: {Kolykhalov}, P.~I., \& {Syunyaev}, R.~A. 1980, Soviet Astronomy Letters, 6,
2105: 357
2106:
2107: \bibitem[{{Krabbe} {et~al.}(1991){Krabbe}, {Genzel}, {Drapatz}, \&
2108: {Rotaciuc}}]{krabbe91}
2109: {Krabbe}, A., {Genzel}, R., {Drapatz}, S., \& {Rotaciuc}, V. 1991, \apjl, 382,
2110: L19
2111:
2112: \bibitem[{{Krabbe} {et~al.}(1995){Krabbe}, {Genzel}, {Eckart}, {Najarro},
2113: {Lutz}, {Cameron}, {Kroker}, {Tacconi-Garman}, {Thatte}, {Weitzel},
2114: {Drapatz}, {Geballe}, {Sternberg}, \& {Kudritzki}}]{krabbe95}
2115: {Krabbe}, A., {Genzel}, R., {Eckart}, A., {Najarro}, F., {Lutz}, D., {Cameron},
2116: M., {Kroker}, H., {Tacconi-Garman}, L.~E., {Thatte}, N., {Weitzel}, L.,
2117: {Drapatz}, S., {Geballe}, T., {Sternberg}, A., \& {Kudritzki}, R. 1995,
2118: \apjl, 447, L95
2119:
2120: \bibitem[{{Lee}(1996)}]{lee96gcmergers}
2121: {Lee}, H.~M. 1996, in IAU Symposium, Vol. 169, Unsolved Problems of the Milky
2122: Way, ed. L.~{Blitz} \& P.~J. {Teuben}, 215
2123:
2124: \bibitem[{{Levin}(2007)}]{levin06}
2125: {Levin}, Y. 2007, \mnras, 374, 515
2126:
2127: \bibitem[{{Levin} \& {Beloborodov}(2003)}]{levin03}
2128: {Levin}, Y., \& {Beloborodov}, A.~M. 2003, \apjl, 590, L33
2129:
2130: \bibitem[{{Lin} \& {Pringle}(1987)}]{linPringle87}
2131: {Lin}, D.~N.~C., \& {Pringle}, J.~E. 1987, \mnras, 225, 607
2132:
2133: \bibitem[{{Lu} {et~al.}(2005){Lu}, {Ghez}, {Hornstein}, {Morris}, \&
2134: {Becklin}}]{lu05irs16sw}
2135: {Lu}, J.~R., {Ghez}, A.~M., {Hornstein}, S.~D., {Morris}, M., \& {Becklin},
2136: E.~E. 2005, \apjl, 625, L51
2137:
2138: \bibitem[{{Maillard} {et~al.}(2004){Maillard}, {Paumard}, {Stolovy}, \&
2139: {Rigaut}}]{maillard04irs13}
2140: {Maillard}, J.~P., {Paumard}, T., {Stolovy}, S.~R., \& {Rigaut}, F. 2004, \aap,
2141: 423, 155
2142:
2143: \bibitem[{{Matthews} {et~al.}(1996){Matthews}, {Ghez}, {Weinberger}, \&
2144: {Neugebauer}}]{NIRCs}
2145: {Matthews}, K., {Ghez}, A.~M., {Weinberger}, A.~J., \& {Neugebauer}, G. 1996,
2146: \pasp, 108, 615
2147:
2148: \bibitem[{{Matthews} \& {Soifer}(1994)}]{NIRC}
2149: {Matthews}, K., \& {Soifer}, B.~T. 1994, Experimental Astronomy, 3, 77
2150:
2151: \bibitem[{{McMillan} \& {Portegies Zwart}(2003)}]{mcmillan03}
2152: {McMillan}, S.~L.~W., \& {Portegies Zwart}, S.~F. 2003, \apj, 596, 314
2153:
2154: \bibitem[{{Milosavljevi{\'c}} \& {Loeb}(2004)}]{milosav04}
2155: {Milosavljevi{\'c}}, M., \& {Loeb}, A. 2004, \apjl, 604, L45
2156:
2157: \bibitem[{{Morris}(1993)}]{morris93}
2158: {Morris}, M. 1993, \apj, 408, 496
2159:
2160: \bibitem[{{Morris} \& {Serabyn}(1996)}]{morris96}
2161: {Morris}, M., \& {Serabyn}, E. 1996, \araa, 34, 645
2162:
2163: \bibitem[{{Najarro} {et~al.}(1997){Najarro}, {Krabbe}, {Genzel}, {Lutz},
2164: {Kudritzki}, \& {Hillier}}]{najarro97}
2165: {Najarro}, F., {Krabbe}, A., {Genzel}, R., {Lutz}, D., {Kudritzki}, R.~P., \&
2166: {Hillier}, D.~J. 1997, \aap, 325, 700
2167:
2168: \bibitem[{{Nayakshin} \& {Cuadra}(2005)}]{nayakshinCuadra05}
2169: {Nayakshin}, S., \& {Cuadra}, J. 2005, \aap, 437, 437
2170:
2171: \bibitem[{{Nayakshin} {et~al.}(2007){Nayakshin}, {Cuadra}, \&
2172: {Springel}}]{nayakshin07sims}
2173: {Nayakshin}, S., {Cuadra}, J., \& {Springel}, V. 2007, \mnras, 379, 21
2174:
2175: \bibitem[{{Nayakshin} {et~al.}(2006){Nayakshin}, {Dehnen}, {Cuadra}, \&
2176: {Genzel}}]{nayakshin06thick}
2177: {Nayakshin}, S., {Dehnen}, W., {Cuadra}, J., \& {Genzel}, R. 2006, \mnras, 366,
2178: 1410
2179:
2180: \bibitem[{{Nayakshin} \& {Sunyaev}(2005)}]{nayakshinSunyaev06}
2181: {Nayakshin}, S., \& {Sunyaev}, R. 2005, \mnras, 364, L23
2182:
2183: \bibitem[{{Ott}(2003)}]{ott03thesis}
2184: {Ott}, T. 2003, PhD thesis, Max-Planck-Institut f{\" u}r Extraterrestrische
2185: Physik
2186:
2187: \bibitem[{{Paumard} {et~al.}(2006){Paumard}, {Genzel}, {Martins}, {Nayakshin},
2188: {Beloborodov}, {Levin}, {Trippe}, {Eisenhauer}, {Ott}, {Gillessen}, {Abuter},
2189: {Cuadra}, {Alexander}, \& {Sternberg}}]{paumard06}
2190: {Paumard}, T., {Genzel}, R., {Martins}, F., {Nayakshin}, S., {Beloborodov},
2191: A.~M., {Levin}, Y., {Trippe}, S., {Eisenhauer}, F., {Ott}, T., {Gillessen},
2192: S., {Abuter}, R., {Cuadra}, J., {Alexander}, T., \& {Sternberg}, A. 2006,
2193: \apj, 643, 1011
2194:
2195: \bibitem[{{Portegies Zwart} {et~al.}(2003){Portegies Zwart}, {McMillan}, \&
2196: {Gerhard}}]{pzwart03irs16}
2197: {Portegies Zwart}, S.~F., {McMillan}, S.~L.~W., \& {Gerhard}, O. 2003, \apj,
2198: 593, 352
2199:
2200: \bibitem[{{Rafelski} {et~al.}(2007){Rafelski}, {Ghez}, {Hornstein}, {Lu}, \&
2201: {Morris}}]{rafelski07}
2202: {Rafelski}, M., {Ghez}, A.~M., {Hornstein}, S.~D., {Lu}, J.~R., \& {Morris}, M.
2203: 2007, \apj, 659, 1241
2204:
2205: \bibitem[{{Sanders}(1992)}]{sanders92}
2206: {Sanders}, R.~H. 1992, \nat, 359, 131
2207:
2208: \bibitem[{{Sanders}(1998)}]{sanders98}
2209: ---. 1998, \mnras, 294, 35
2210:
2211: \bibitem[{{Sch{\" o}del} {et~al.}(2005){Sch{\" o}del}, {Eckart}, {Iserlohe},
2212: {Genzel}, \& {Ott}}]{schodel05}
2213: {Sch{\" o}del}, R., {Eckart}, A., {Iserlohe}, C., {Genzel}, R., \& {Ott}, T.
2214: 2005, \apjl, 625, L111
2215:
2216: \bibitem[{{Sch{\" o}del} {et~al.}(2003){Sch{\" o}del}, {Ott}, {Genzel},
2217: {Eckart}, {Mouawad}, \& {Alexander}}]{schodel03}
2218: {Sch{\" o}del}, R., {Ott}, T., {Genzel}, R., {Eckart}, A., {Mouawad}, N., \&
2219: {Alexander}, T. 2003, \apj, 596, 1015
2220:
2221: \bibitem[{{Sch{\"o}del} {et~al.}(2007){Sch{\"o}del}, {Eckart}, {Alexander},
2222: {Merritt}, {Genzel}, {Sternberg}, {Meyer}, {Kul}, {Moultaka}, {Ott}, \&
2223: {Straubmeier}}]{schodel07}
2224: {Sch{\"o}del}, R., {Eckart}, A., {Alexander}, T., {Merritt}, D., {Genzel}, R.,
2225: {Sternberg}, A., {Meyer}, L., {Kul}, F., {Moultaka}, J., {Ott}, T., \&
2226: {Straubmeier}, C. 2007, \aap, 469, 125
2227:
2228: \bibitem[{{Sch{\"o}del} {et~al.}(2002){Sch{\"o}del}, {Ott}, {Genzel},
2229: {Hofmann}, {Lehnert}, {Eckart}, {Mouawad}, {Alexander}, {Reid}, {Lenzen},
2230: {Hartung}, {Lacombe}, {Rouan}, {Gendron}, {Rousset}, {Lagrange}, {Brandner},
2231: {Ageorges}, {Lidman}, {Moorwood}, {Spyromilio}, {Hubin}, \&
2232: {Menten}}]{schodel02}
2233: {Sch{\"o}del}, R., {Ott}, T., {Genzel}, R., {Hofmann}, R., {Lehnert}, M.,
2234: {Eckart}, A., {Mouawad}, N., {Alexander}, T., {Reid}, M.~J., {Lenzen}, R.,
2235: {Hartung}, M., {Lacombe}, F., {Rouan}, D., {Gendron}, E., {Rousset}, G.,
2236: {Lagrange}, A.-M., {Brandner}, W., {Ageorges}, N., {Lidman}, C., {Moorwood},
2237: A.~F.~M., {Spyromilio}, J., {Hubin}, N., \& {Menten}, K.~M. 2002, \nat, 419,
2238: 694
2239:
2240: \bibitem[{{Scoville} {et~al.}(2003){Scoville}, {Stolovy}, {Rieke},
2241: {Christopher}, \& {Yusef-Zadeh}}]{scoville03}
2242: {Scoville}, N.~Z., {Stolovy}, S.~R., {Rieke}, M., {Christopher}, M., \&
2243: {Yusef-Zadeh}, F. 2003, \apj, 594, 294
2244:
2245: \bibitem[{{Shlosman} \& {Begelman}(1989)}]{shlosman89}
2246: {Shlosman}, I., \& {Begelman}, M.~C. 1989, \apj, 341, 685
2247:
2248: \bibitem[{{Stolte} {et~al.}(2006){Stolte}, {Brandner}, {Brandl}, \&
2249: {Zinnecker}}]{stolte06}
2250: {Stolte}, A., {Brandner}, W., {Brandl}, B., \& {Zinnecker}, H. 2006, \aj, 132,
2251: 253
2252:
2253: \bibitem[{{Tamblyn} {et~al.}(1996){Tamblyn}, {Rieke}, {Hanson}, {Close},
2254: {McCarthy}, \& {Rieke}}]{tamblyn96}
2255: {Tamblyn}, P., {Rieke}, G.~H., {Hanson}, M.~M., {Close}, L.~M., {McCarthy},
2256: D.~W., \& {Rieke}, M.~J. 1996, \apj, 456, 206
2257:
2258: \bibitem[{{van Dam} {et~al.}(2006){van Dam}, {Bouchez}, {Le Mignant},
2259: {Johansson}, {Wizinowich}, {Campbell}, {Chin}, {Hartman}, {Lafon}, {Stomski},
2260: \& {Summers}}]{vanDam06}
2261: {van Dam}, M.~A., {Bouchez}, A.~H., {Le Mignant}, D., {Johansson}, E.~M.,
2262: {Wizinowich}, P.~L., {Campbell}, R.~D., {Chin}, J.~C.~Y., {Hartman}, S.~K.,
2263: {Lafon}, R.~E., {Stomski}, Jr., P.~J., \& {Summers}, D.~M. 2006, \pasp, 118,
2264: 310
2265:
2266: \bibitem[{{Vollmer} \& {Duschl}(2001)}]{vollmerDuschl01}
2267: {Vollmer}, B., \& {Duschl}, W.~J. 2001, \aap, 377, 1016
2268:
2269: \bibitem[{{Wizinowich} {et~al.}(2006){Wizinowich}, {Le Mignant}, {Bouchez},
2270: {Campbell}, {Chin}, {Contos}, {van Dam}, {Hartman}, {Johansson}, {Lafon},
2271: {Lewis}, {Stomski}, {Summers}, {Brown}, {Danforth}, {Max}, \&
2272: {Pennington}}]{wizinowich06}
2273: {Wizinowich}, P.~L., {Le Mignant}, D., {Bouchez}, A.~H., {Campbell}, R.~D.,
2274: {Chin}, J.~C.~Y., {Contos}, A.~R., {van Dam}, M.~A., {Hartman}, S.~K.,
2275: {Johansson}, E.~M., {Lafon}, R.~E., {Lewis}, H., {Stomski}, P.~J., {Summers},
2276: D.~M., {Brown}, C.~G., {Danforth}, P.~M., {Max}, C.~E., \& {Pennington},
2277: D.~M. 2006, \pasp, 118, 297
2278:
2279: \bibitem[{{Yu} {et~al.}(2007){Yu}, {Lu}, \& {Lin}}]{yu07}
2280: {Yu}, Q., {Lu}, Y., \& {Lin}, D.~N.~C. 2007, \apj, 666, 919
2281:
2282: \end{thebibliography}
2283:
2284: \clearpage
2285:
2286: %---------------
2287: %
2288: % Tables
2289: %
2290: %---------------
2291:
2292: % Generated by hand
2293: \input{tab1}
2294:
2295: % Generated from python:
2296: % papers.lu06yng.tableAstrometry()
2297: \input{tab2}
2298:
2299: % Generated from python:
2300: % papers.lu06yng.tableEccDiskProb()
2301: \input{tab3}
2302:
2303: % Generated from python:
2304: % papers.lu06yng.tableEccDiskProb(extended=True)
2305: \input{tab4}
2306:
2307: % Generated from IDL> yng_pm_table_lu06yng
2308: % \input{lu06yng_tab_orbits}
2309:
2310: % Table for app_speckDistort
2311: \input{tab5}
2312:
2313: \clearpage
2314:
2315: %---------------
2316: %
2317: % Figures
2318: %
2319: %---------------
2320:
2321:
2322: % Generated from python:
2323: % papers.lu06yng.plotPosError()
2324: \begin{figure}
2325: \begin{center}
2326: \plotone{f1.eps}
2327: \end{center}
2328: \caption{
2329: Positional uncertainties for stars as a function of stellar brightness
2330: {\it (top)} and distance from the black hole, Sgr A*, which is
2331: near the center of the field of view {\it (bottom)}.
2332: To show the full range of possible values, the
2333: centroiding {\it (solid)} and the alignment {\it (dashed)} uncertainties
2334: are shown for the best (1999.559) and
2335: worst (1996.485) speckle epochs and one of the LGS AO epochs (2004.567).
2336: The uncertainties are the median values of all stars within
2337: magnitude bins of $\Delta$K = 1 or radius bins of $\Delta r$ = 0\farcs3.
2338: Note that alignment uncertainties are small compared to centroid
2339: uncertainties.
2340: }
2341: \label{fig:posError}
2342: \end{figure}
2343:
2344: % Generated by python code:
2345: % lu.plotPosVsTime()
2346: \begin{figure}
2347: \epsscale{0.6}
2348: \plotone{f2.eps}
2349: \caption{
2350: Measured positions and residuals as a function of time for S0-15,
2351: a source with a significant, non-zero acceleration measurement,
2352: in X and Y ({\it top}). Positions are reported relative
2353: to Sgr A* and do not include the uncertainties in the
2354: transformation to the absolute coordinate system (i.e. plate scale,
2355: position angle, and position of Sgr A*).
2356: The best fit quadratic polynomial modeling the velocity and acceleration
2357: of the source is shown ({\it green solid}) with the 1$\sigma$ errorbars
2358: ({\it green dashed}).
2359: Also plotted are the X and Y residuals after subtracting off the best fit
2360: velocity ({\it middle}) and the best fit acceleration
2361: curve ({\it bottom}).
2362: The X (East-West) and Y (North-South) position plots ({\it top}) have
2363: a (y-axis) range of 0\farcs16 and residual plots ({\it middle, bottom}) have
2364: a (y-axis) range of $\pm$8 mas.
2365: }
2366: \label{fig:posTime_S0-15}
2367: \end{figure}
2368:
2369: % Generated by python code:
2370: % lu.plotPosVsTime()
2371: \begin{figure}
2372: \epsscale{0.6}
2373: \plotone{f3.eps}
2374: \caption{
2375: Measured positions and residuals as a function of time for
2376: IRS 16NW, a source that has an acceleration consistent with zero,
2377: but significantly below the maximum possible acceleration.
2378: See the caption in Figure \ref{fig:posTime_S0-15} for more information.
2379: }
2380: \label{fig:posTime_irs16NW}
2381: \end{figure}
2382:
2383: % Generated by python code:
2384: % lu.histAccel()
2385: \begin{figure}
2386: \epsscale{1.0}
2387: \plotone{f4.eps}
2388: \caption{Histograms of $a_\rho$/$\sigma_{a_\rho}$ showing the significance of the
2389: acceleration measurements in both the radial ({\it top}) and tangential
2390: ({\it bottom}) directions. While, the distributions show an offset from zero
2391: indicating a possible bias due to systematic errors, such as residual
2392: distortion, that are not well characterized, it appears that any biases
2393: are limited to the $\sim$1$\sigma$ level.
2394: The only star with significant negative radial acceleration
2395: ($\gtrsim$4$\sigma$) is S0-15 and it is assumed to be a real acceleration due
2396: to the gravity of the supermassive black hole.
2397: }
2398: \label{fig:histAccel}
2399: \end{figure}
2400:
2401: % Generated with python code:
2402: % papers.lu06yng.plotImagePM()
2403: % papers.lu06yng.plotImagePM(extended=True)
2404: \begin{figure}
2405: \epsscale{1.0}
2406: \plottwo{f5a.eps}{f5b.eps}
2407: \caption{
2408: Positions and proper motion vectors of the young stars in our sample.
2409: Candidate disk members are shown in red and non-disk members are
2410: shown in blue over-plotted on an LGS AO image in grey-scale. The names of
2411: the stars in the primary sample are shown in the {\it left} panel and
2412: the complete extended sample is shown
2413: in a zoomed-out view in the {\it right} panel. The position of Sgr A*
2414: is marked with a black cross.
2415: }
2416: \label{fig:PMimage}
2417: \end{figure}
2418:
2419: % Generated by python code:
2420: % papers.lu06yng.plotPolyAccVsR()
2421: \begin{figure}
2422: \epsscale{1.0}
2423: \plotone{f6.eps}
2424: \caption{The significance of observed limits for the plane-of-the-sky
2425: acceleration. For a given projected radius, there is a maximum
2426: allowed acceleration (dashed line).
2427: If the measured accelerations from polynomial fitting, shown as
2428: 3$\sigma$ upper limits on the y-axis, are less than the maximum allowed
2429: acceleration (below the dashed line), then significant constraints can be
2430: placed on the line-of-sight distance, z, and subsequently the orbital
2431: parameters of the star. S0-15 has a significant detection of non-zero
2432: acceleration and is plotted with its 1$\sigma$ errorbars.}
2433: \label{fig:accSignificance2}
2434: \end{figure}
2435:
2436: \figsetstart
2437: \figsetnum{7}
2438: \figsettitle{Orbital Parameters}
2439:
2440: \figsetgrpstart
2441: \figsetgrpnum{7.1}
2442: \figsetgrptitle{S0-15}
2443: \figsetplot{f7_1.eps}
2444: \figsetgrpnote{Orbit parameters for S0-15.}
2445: \figsetgrpend
2446:
2447: \figsetgrpstart
2448: \figsetgrpnum{7.2}
2449: \figsetgrptitle{S0-14}
2450: \figsetplot{f7_2.eps}
2451: \figsetgrpnote{Orbit parameters for S0-14}
2452: \figsetgrpend
2453:
2454: \figsetgrpstart
2455: \figsetgrpnum{7.3}
2456: \figsetgrptitle{S1-2}
2457: \figsetplot{f7_3.eps}
2458: \figsetgrpnote{Orbit parameters for S1-2.}
2459: \figsetgrpend
2460:
2461: \figsetgrpstart
2462: \figsetgrpnum{7.4}
2463: \figsetgrptitle{S1-3}
2464: \figsetplot{f7_4.eps}
2465: \figsetgrpnote{Orbit Parameters for S1-3.}
2466: \figsetgrpend
2467:
2468: \figsetgrpstart
2469: \figsetgrpnum{7.5}
2470: \figsetgrptitle{S1-8}
2471: \figsetplot{f7_5.eps}
2472: \figsetgrpnote{Orbit Parameters for S1-8.}
2473: \figsetgrpend
2474:
2475: \figsetgrpstart
2476: \figsetgrpnum{7.6}
2477: \figsetgrptitle{IRS 16NW}
2478: \figsetplot{f7_6.eps}
2479: \figsetgrpnote{Orbit Parameters for IRS 16NW.}
2480: \figsetgrpend
2481:
2482: \figsetgrpstart
2483: \figsetgrpnum{7.7}
2484: \figsetgrptitle{IRS 16C}
2485: \figsetplot{f7_7.eps}
2486: \figsetgrpnote{Orbit Parameters for IRS 16C.}
2487: \figsetgrpend
2488:
2489: \figsetgrpstart
2490: \figsetgrpnum{7.8}
2491: \figsetgrptitle{S1-12}
2492: \figsetplot{f7_8.eps}
2493: \figsetgrpnote{Orbit Parameters for S1-12.}
2494: \figsetgrpend
2495:
2496: \figsetgrpstart
2497: \figsetgrpnum{7.9}
2498: \figsetgrptitle{S1-14}
2499: \figsetplot{f7_9.eps}
2500: \figsetgrpnote{Orbit Parameters for S1-14.}
2501: \figsetgrpend
2502:
2503: \figsetgrpstart
2504: \figsetgrpnum{7.10}
2505: \figsetgrptitle{IRS 16SW}
2506: \figsetplot{f7_10.eps}
2507: \figsetgrpnote{Orbit Parameters for IRS 16SW.}
2508: \figsetgrpend
2509:
2510: \figsetgrpstart
2511: \figsetgrpnum{7.11}
2512: \figsetgrptitle{S1-21}
2513: \figsetplot{f7_11.eps}
2514: \figsetgrpnote{Orbit Parameters for S1-21.}
2515: \figsetgrpend
2516:
2517: \figsetgrpstart
2518: \figsetgrpnum{7.12}
2519: \figsetgrptitle{S1-22}
2520: \figsetplot{f7_12.eps}
2521: \figsetgrpnote{Orbit Parameters for S1-22.}
2522: \figsetgrpend
2523:
2524: \figsetgrpstart
2525: \figsetgrpnum{7.13}
2526: \figsetgrptitle{S1-24}
2527: \figsetplot{f7_13.eps}
2528: \figsetgrpnote{Orbit Parameters for S1-24.}
2529: \figsetgrpend
2530:
2531: \figsetgrpstart
2532: \figsetgrpnum{7.14}
2533: \figsetgrptitle{IRS 16CC}
2534: \figsetplot{f7_14.eps}
2535: \figsetgrpnote{Orbit Parameters for IRS 16CC.}
2536: \figsetgrpend
2537:
2538: \figsetgrpstart
2539: \figsetgrpnum{7.15}
2540: \figsetgrptitle{S2-4}
2541: \figsetplot{f7_15.eps}
2542: \figsetgrpnote{Orbit Parameters for S2-4.}
2543: \figsetgrpend
2544:
2545: \figsetgrpstart
2546: \figsetgrpnum{7.16}
2547: \figsetgrptitle{S2-6}
2548: \figsetplot{f7_16.eps}
2549: \figsetgrpnote{Orbit Parameters for S2-6.}
2550: \figsetgrpend
2551:
2552: \figsetgrpstart
2553: \figsetgrpnum{7.17}
2554: \figsetgrptitle{S2-7}
2555: \figsetplot{f7_17.eps}
2556: \figsetgrpnote{Orbit Parameters for S2-7.}
2557: \figsetgrpend
2558:
2559: \figsetgrpstart
2560: \figsetgrpnum{7.18}
2561: \figsetgrptitle{IRS 29N}
2562: \figsetplot{f7_18.eps}
2563: \figsetgrpnote{Orbit Parameters for IRS 29N.}
2564: \figsetgrpend
2565:
2566: \figsetgrpstart
2567: \figsetgrpnum{7.19}
2568: \figsetgrptitle{IRS 16SW-E}
2569: \figsetplot{f7_19.eps}
2570: \figsetgrpnote{Orbit Parameters for IRS 16SW-E.}
2571: \figsetgrpend
2572:
2573: \figsetgrpstart
2574: \figsetgrpnum{7.20}
2575: \figsetgrptitle{IRS 33N}
2576: \figsetplot{f7_20.eps}
2577: \figsetgrpnote{Orbit Parameters for IRS 33N.}
2578: \figsetgrpend
2579:
2580: \figsetgrpstart
2581: \figsetgrpnum{7.21}
2582: \figsetgrptitle{S2-17}
2583: \figsetplot{f7_21.eps}
2584: \figsetgrpnote{Orbit Parameters for S2-17.}
2585: \figsetgrpend
2586:
2587: \figsetgrpstart
2588: \figsetgrpnum{7.22}
2589: \figsetgrptitle{S2-16}
2590: \figsetplot{f7_22.eps}
2591: \figsetgrpnote{Orbit Parameters for S2-16.}
2592: \figsetgrpend
2593:
2594: \figsetgrpstart
2595: \figsetgrpnum{7.23}
2596: \figsetgrptitle{S2-19}
2597: \figsetplot{f7_23.eps}
2598: \figsetgrpnote{Orbit Parameters for S2-19.}
2599: \figsetgrpend
2600:
2601: \figsetgrpstart
2602: \figsetgrpnum{7.24}
2603: \figsetgrptitle{S2-66}
2604: \figsetplot{f7_24.eps}
2605: \figsetgrpnote{Orbit Parameters for S2-66.}
2606: \figsetgrpend
2607:
2608: \figsetgrpstart
2609: \figsetgrpnum{7.25}
2610: \figsetgrptitle{S2-74}
2611: \figsetplot{f7_25.eps}
2612: \figsetgrpnote{Orbit Parameters for S2-74.}
2613: \figsetgrpend
2614:
2615: \figsetgrpstart
2616: \figsetgrpnum{7.26}
2617: \figsetgrptitle{IRS 16NE}
2618: \figsetplot{f7_26.eps}
2619: \figsetgrpnote{Orbit Parameters for IRS 16NE.}
2620: \figsetgrpend
2621:
2622: \figsetgrpstart
2623: \figsetgrpnum{7.27}
2624: \figsetgrptitle{S3-5}
2625: \figsetplot{f7_27.eps}
2626: \figsetgrpnote{Orbit Parameters for S3-5.}
2627: \figsetgrpend
2628:
2629: \figsetgrpstart
2630: \figsetgrpnum{7.28}
2631: \figsetgrptitle{IRS 33E}
2632: \figsetplot{f7_28.eps}
2633: \figsetgrpnote{Orbit Parameters for IRS 33E.}
2634: \figsetgrpend
2635:
2636: \figsetgrpstart
2637: \figsetgrpnum{7.29}
2638: \figsetgrptitle{S3-19}
2639: \figsetplot{f7_29.eps}
2640: \figsetgrpnote{Orbit Parameters for S3-19.}
2641: \figsetgrpend
2642:
2643: \figsetgrpstart
2644: \figsetgrpnum{7.30}
2645: \figsetgrptitle{S3-25}
2646: \figsetplot{f7_30.eps}
2647: \figsetgrpnote{Orbit Parameters for S3-25.}
2648: \figsetgrpend
2649:
2650: \figsetgrpstart
2651: \figsetgrpnum{7.31}
2652: \figsetgrptitle{S3-30}
2653: \figsetplot{f7_31.eps}
2654: \figsetgrpnote{Orbit Parameters for S3-30.}
2655: \figsetgrpend
2656:
2657: \figsetgrpstart
2658: \figsetgrpnum{7.32}
2659: \figsetgrptitle{S3-10}
2660: \figsetplot{f7_32.eps}
2661: \figsetgrpnote{Orbit Parameters for S3-10.}
2662: \figsetgrpend
2663:
2664: \figsetend
2665:
2666: \begin{figure*}
2667: %\figurenum{7}
2668: \epsscale{1.0}
2669: \plotone{f7.eps}
2670: \caption{
2671: The range of allowed orbital parameters for IRS 16SW
2672: as determined from the observed
2673: two-dimensional position in the plane of the sky,
2674: the three-dimensional velocity,
2675: and the acceleration. The probability distribution for
2676: each orbital parameter is determined
2677: by sampling from a gaussian distribution for each of the
2678: observed quantities and analytically converting to the standard
2679: orbital elements.
2680: High density (dark) regions represent the most probable values for each
2681: orbital parameter and the resulting 1$\sigma$ and 2$\sigma$ contours
2682: are shown as black lines.
2683: }
2684: \label{fig:pdfParams_irs16SW}
2685: \end{figure*}
2686:
2687:
2688: %%% --- PDF PARAMS PLOTS ---%%%
2689: % Generated from python code:
2690: % papers.lu06yng.plotOrbitPdfExample()
2691: \begin{figure}
2692: \epsscale{0.8}
2693: \plotone{f8.eps}
2694: \caption{The range of allowed eccentricities ($e$), inclinations ($i$),
2695: and angles to ascending nodes ($\Omega$)
2696: as determined by our orbit analysis for three example stars.
2697: The range of z values (horizontal axis) extends to all possible
2698: bound orbits for the star. The probability density function is shown
2699: in color with the 1$\sigma$ and 2$\sigma$ contours drawn shown as black lines.
2700: S0-15 has a measured acceleration that is significantly different from zero.
2701: IRS 16C has an acceleration upper limit that is less
2702: than the maximum allowed acceleration, and thus a lower limit
2703: on the line-of-sight distance, $|z|$. IRS 16CC has no significant
2704: acceleration limit, but has a high velocity that is always larger than
2705: the circular velocity, thus prohibiting circular orbits. Also, by assuming
2706: the star is bound, the direction of the normal vector to
2707: IRS 16CC's orbital plane
2708: is restricted to a low inclination.
2709: }
2710: \label{fig:pdfParams_example}
2711: \end{figure}
2712:
2713: % Generated by python code:
2714: % lu.comparePriorsPhase()
2715: \begin{figure}
2716: \epsscale{1.0}
2717: \plotone{f9.eps}
2718: \caption{The resulting distribution of orbital phases for all stars
2719: when assuming either a uniform acceleration prior ({\it black}) or
2720: a uniform z prior ({\it gray}) and then imposing the measured
2721: accelerations. The uniform z prior shows a strong bias
2722: towards an orbital phase of 0, which corresponds to periapse; while
2723: the uniform acceleration prior shows a more uniform distribution.
2724: }
2725: \label{fig:comparePriorsPhase}
2726: \end{figure}
2727:
2728: % Generated by IDL code:
2729: % .r lu06yng
2730: % iomapSample, save=1
2731: \begin{figure}
2732: \centering
2733: \includegraphics[scale=0.3,angle=90]{f10a.ps} \\
2734: \includegraphics[scale=0.3,angle=90]{f10b.ps} \\
2735: \includegraphics[scale=0.3,angle=90]{f10c.ps}
2736: \caption{
2737: The orientation of three stars' orbital planes as described by the
2738: probability distribution of the planes' normal vector projected
2739: onto the sky as viewed from Sgr A*.
2740: Colors indicate the probability density for a star's
2741: normal vector to point at each pixel on the sky.
2742: The constraint on the stars' normal vectors are set
2743: by ({\it top}: S0-15) a measured acceleration; ({\it middle}: IRS 16C)
2744: a significant acceleration limit; ({\it bottom}: IRS 16CC) the star's high
2745: velocity and assuming the orbit is bound.
2746: }
2747: \label{fig:iomap}
2748: \end{figure}
2749:
2750: % Generated by IDL code:
2751: % .r lu06yng
2752: % mapAll
2753: \begin{figure}
2754: \begin{center}
2755: \includegraphics[scale=0.5,angle=90]{f11a_new.ps} \\
2756: \includegraphics[scale=0.5,angle=90]{f11b_new.ps}
2757: \caption{The 1$\sigma$ contours of all stars'
2758: probability distribution functions for
2759: the orientation of their orbital planes. This shows the distribution of
2760: stellar orbit orientations around the sky.
2761: The primary sample is plotted on the {\it top} and
2762: if there are degenerate solutions for a given star,
2763: then one solution is plotted with a solid line and the other with a dashed
2764: line. Additional sources found only in the secondary sample are plotted on the
2765: {\it bottom} and are plotted with dashed lines as there are no acceleration
2766: constraints and each star has a single solution with large uncertainties.
2767: We note that the orientation of the projection shown in this figure is
2768: rotated by 180$^\circ$ with respect to that shown in earlier publications
2769: \citep[e.g.][]{eisenhauer06,nayakshin06thick} in order to more easily
2770: see the region around the proposed disks.
2771: }
2772: \label{fig:allPlanePDF}
2773: \end{center}
2774: \end{figure}
2775:
2776: % Generated by IDL code:
2777: % plot_disk_healpix, 'aorb_efit/disk.neighbor.dat', 12288, 4, plottrial=2, /olddisks
2778: % cp aorb_efit/disk.neighbor.dat.ps plots/lu06yng_planes.ps
2779: \begin{figure*}
2780: \begin{center}
2781: \includegraphics[scale=0.5,angle=90]{f12a.ps} \\
2782: \includegraphics[scale=0.5,angle=90]{f12b.ps}
2783: \caption{The density of normal vectors to the orbital planes of the stars
2784: in our primary ({\it top}) and extended ({\it bottom}) samples.
2785: Densities are indicated in colors (stars deg$^{-2}$) on a linear scale and
2786: the peak indicates an over-density of stars with similar orbital planes.
2787: Over-plotted in black are the candidate orbital
2788: planes as proposed by \citet{levin03} and \citet{genzel03cusp} with
2789: updated values from \citet{paumard06} for the candidate plane normal
2790: vector and uncertainties (solid black) and the disk thickness (dashed black)
2791: shown as solid angles of 0.05 sr and 0.09 sr for the clockwise and
2792: counter-clockwise disks respectively.
2793: }
2794: \label{fig:orbitPlane}
2795: \end{center}
2796: \end{figure*}
2797:
2798: % Generated by python code
2799: % ludisk.plotDiskPositionMC()
2800: \begin{figure}
2801: \epsscale{1.0}
2802: \plottwo{f13a_new.eps}{f13b_new.eps}
2803: \caption{
2804: Positions for all candidate disk members in the disk plane from
2805: our primary ({\it left}) and extended ({\it right}) samples.
2806: The field of view for this study is projected onto the disk plane
2807: and shows the outer ({\it dash}) and inner ({\it dot-dash}) boundaries.
2808: For each disk candidate, a contour shows the star's position within
2809: the disk for all orbital solutions within 10$^\circ$ of the disk plane.
2810: The color scale shows the probability density function for each star's
2811: position in the disk, normalized by the likelihood of disk membership.
2812: This normalization shows stars with a higher and lower likelihood of
2813: disk membership as {\it darker red} or {\it lighter yellow}, respectively.
2814: }
2815: \label{fig:diskVelRadius}
2816: \end{figure}
2817:
2818: % Generated by python code
2819: % lu.histSolidAngles()
2820: \begin{figure}
2821: \epsscale{1.0}
2822: \plotone{f14.eps}
2823: \caption{The distribution of $\vec{n}$-uncertainties as expressed by
2824: the area of the 1$\sigma$ region in which a star's normal vector can point.
2825: The uncertainties are shown both for the sample in this work
2826: ({\it black}) and for the stars with only three-dimensional position
2827: and two-dimensional velocity
2828: information extracted from \citet{paumard06} that are used in the
2829: search for a second disk ({\it gray}).}
2830: \label{fig:histSolidAngle}
2831: \end{figure}
2832:
2833: % Generated by python code
2834: % ludisk.calcDiskRadialDist2()
2835: % ludisk.plotDiskRadialDist()
2836: \begin{figure}
2837: \epsscale{1.0}
2838: \plotone{f15.eps}
2839: \caption{
2840: The radial distribution of stars within the disk plane for the
2841: extended sample. The best fit line is shown ({\it dashed}) and was
2842: constructed by excluding the first data point and the last three data
2843: points where field of view limitations may affect the distribution.
2844: }
2845: \label{fig:diskRadialDist}
2846: \end{figure}
2847:
2848: % Generated by python code
2849: % lu.plotEccVectorLimits()
2850: \begin{figure}
2851: \epsscale{0.8}
2852: \plotone{f16.eps}
2853: \caption{The distribution of eccentricity lower limits
2854: as determined from individual stellar
2855: orbits, excluding S0-14. The top panel shows
2856: the 99.7\% confidence lower limit from all possible orbital solutions
2857: for candidate disk members ({\it red circles}) and non-disk members
2858: ({\it blue squares}). Stars from the primary sample ({\it filled}) and
2859: stars added in the extended sample ({\it unfilled}) are both shown.
2860: Sources in only the
2861: extended sample have less constrained eccentricities due to their larger
2862: velocity uncertainties. The bottom panel
2863: shows the candidate disk members 99.7\% confidence lower limits after
2864: restricting the orbital solutions to those with
2865: normal vectors within 10$^\circ$ of the disk. By assuming disk membership,
2866: the range of eccentricities is more restricted for the candidate disk
2867: members.
2868: }
2869: \label{fig:eccentricity}
2870: \end{figure}
2871:
2872: % Generated py python code
2873: % lu.diskEccVector()
2874: \begin{figure}
2875: \epsscale{1.0}
2876: \plotone{f17.eps}
2877: \caption{Combined probability distribution for the candidate
2878: disk stars' eccentricity vectors. The eccentricity vectors for
2879: orbital solutions with normal vectors within 10$^\circ$ of
2880: the disk's normal vector are projected onto the disk plane.
2881: The 1$\sigma$ and 2$\sigma$ confidence-level contours are shown in
2882: black.
2883: }
2884: \label{fig:eccVector}
2885: \end{figure}
2886:
2887: \clearpage
2888:
2889: % Figure for app_speckDistort
2890: \begin{figure}
2891: \epsscale{0.7}
2892: \plotone{f18.eps}
2893: \caption{
2894: The improvement in positional accuracy at large radii as a result
2895: of correcting geometric distortion in speckle data sets.
2896: To characterize the systematic positional uncertainty, we take
2897: each star at each epoch and calculate the residual positional offset,
2898: which is defined as the difference between the
2899: measured position and the position as determined by the best fit velocity
2900: ($x = x_o + v*\Delta t$). Then the RMS of the residuals is calculated
2901: across all epochs for each star. All stars' resulting RMS values are sorted
2902: by the distance between the star and Sgr A* (which was at the center of
2903: the images) and then averaged over radius bins of 0\farcs3.
2904: The radial trend is shown for
2905: data prior to the new distortion correction (top) and after the new
2906: distortion correction (bottom).
2907: }
2908: \label{fig:distort}
2909: \end{figure}
2910:
2911: % Figure for app_speckDistort
2912: \begin{figure}
2913: \begin{center}
2914: \epsscale{1.0}
2915: \plottwo{f19a.eps}{f19b.eps}
2916: \caption{Map of the positional differences between stars
2917: observed near-simultaneously with NIRC and NIRC2.
2918: The maps are plotted in the original NIRC detector coordinates
2919: and show residuals before {\it (left; a)} and after {\it (right; b)}
2920: the NIRC-reimager distortion solution.
2921: }
2922: \label{fig:distort2d}
2923: \end{center}
2924: \end{figure}
2925:
2926: % Figure for app_speckDistort
2927: % lu.speckDistortMode()
2928: \begin{figure}
2929: \epsscale{1.0}
2930: \plotone{f20.eps}
2931: \caption{Distribution of the residuals before ({\it gray}) and
2932: after ({\it black}) correcting for the NIRC image converter
2933: distortion. Residuals are calculated by comparing a star's position in
2934: each NIRC image stack to the position in the LGS AO/NIRC2 image.
2935: These residuals are further reduced in the final image because the
2936: stacks are dithered small amounts on the detector and residual
2937: distortion can be averaged out if it is randomly oriented over
2938: the scale of the dither.
2939: }
2940: \label{fig:speck_distort}
2941: \end{figure}
2942:
2943:
2944: \end{document}
2945: