0808.2282/ms.tex
1: \documentclass[numberedappendix]{emulateapj}
2: %%\documentclass[manuscript]{aastex}
3: %%\documentclass[preprint2]{aastex}
4: 
5: \begin{document}  
6: 
7: \title{Candidate Disk Wide Binaries in the Sloan Digital Sky Survey}
8: 
9: \author{
10: Branimir Sesar\altaffilmark{1},
11: \v{Z}eljko Ivezi\'{c}\altaffilmark{1},
12: Mario Juri\'{c}\altaffilmark{2}
13: }
14: 
15: \altaffiltext{1}{University of Washington, Dept.~of Astronomy, Box
16:                            351580, Seattle, WA 98195-1580}
17: \altaffiltext{2}{Institute for Advanced Study, 1 Einstein Drive,
18:                            Princeton, NJ 08540}
19: 
20: \begin{abstract}
21: 
22: Using SDSS Data Release 6, we construct two independent samples of candidate
23: stellar wide binaries selected as i) pairs of unresolved sources with angular
24: separation in the range $3\arcsec - 16\arcsec$, ii) common proper motion pairs
25: with $5\arcsec - 30\arcsec$ angular separation, and make them publicly
26: available. These samples are dominated by disk stars, and we use them to
27: constrain the shape of the main-sequence photometric parallax relation
28: $M_r(r-i)$, and to study the properties of wide binary systems. We estimate
29: $M_r(r-i)$ by searching for a relation that minimizes the difference between
30: distance moduli of primary and secondary components of wide binary candidates.
31: We model $M_r(r-i)$ by a fourth degree polynomial and determine the coefficients
32: using Markov Chain Monte Carlo fitting, independently for each sample. Both
33: samples yield similar relations, with the largest systematic difference of 0.25
34: mag for F0 to M5 stars, and a root-mean-square scatter of 0.13 mag. A similar
35: level of agreement is obtained with photometric parallax relations recently
36: proposed by \citet{jur08}. The measurements show a root-mean-square scatter of
37: $\sim0.30$ mag around the best fit $M_r(r-i)$ relation, and a mildly
38: non-Gaussian distribution. We attribute this scatter to metallicity effects and
39: additional unresolved multiplicity of wide binary components. Aided by the
40: derived photometric parallax relation, we construct a series of high-quality
41: catalogs of candidate main-sequence binary stars. These range from a sample of
42: $\sim17,000$ candidates with the probability of each pair to be a physical
43: binary (the ``efficiency'') of $\sim65\%$, to a volume-limited sample of
44: $\sim1,800$ candidates with an efficiency of $\sim90\%$. Using these catalogs,
45: we study the distribution of semi-major axes of wide binaries, $a$, in the
46: $2,000 < a < 47,000$~AU range. We find the observations to be well described by
47: the \"Opik distribution, $f(a)\propto 1/a$, for $a<a_{break}$, where $a_{break}$
48: increases roughly linearly with the height $Z$ above the Galactic plane
49: ($a_{break}\propto12,300\,Z{\rm [kpc]}^{0.7}$~AU). The number of wide binary
50: systems with $100 \, {\rm AU} < a < a_{break}$, as a fraction of the total
51: number of stars, decreases from 0.9\% at $Z=0.5$ kpc to 0.5\% at $Z=3$ kpc. The
52: probability for a star to be in a wide binary system is independent of its
53: color. Given this color, the companions of red components seem to be drawn
54: randomly from the stellar luminosity function, while blue components have a
55: larger blue-to-red companion ratio than expected from luminosity function.
56: \end{abstract}
57: 
58: \keywords{binaries: visual --- stars: distances --- Hertzsprung-Russell diagram}
59: 
60: \section{Introduction}
61: \label{introduction}
62: 
63: Binary systems can be roughly divided into close (semi-major axes $a\la10$ AU)
64: and wide (semi-major axes $a\ga100$ AU, \citealt{cha07}) pairs. Close binary
65: systems have long been recognized as useful tools for studies of stellar
66: properties. For example, the stellar parameters such as the masses and radii of
67: individual stars are readily determined to high confidence using eclipsing
68: binaries \citep{and91}. Wide binary systems have proven to be a tool for studies
69: of star formation processes, as well as an exceptionally useful tracer of local
70: potential and tidal fields through which they traverse. Specifically, they were
71: used to place the constraints on the nature of halo dark matter \citep{ycg04}
72: and to explore the dynamical history of the Galaxy \citep{aph07}. A further
73: comprehensive list of current applications of wide binaries can be found in
74: \citet{cha07}.
75: 
76: Close binaries, owing to their relatively short orbital periods and equally
77: short timescales of brightness or spectrum fluctuations, are fairly easy to
78: detect. Unambiguous identification of wide binary systems, on the other hand,
79: requires accurate astrometry on much longer timescales, as these systems have
80: orbital periods $\ga10,000$ years. However, instead of requiring unambiguous
81: identification, large samples of {\em candidate} wide binaries can be selected
82: by simply assuming that pairs of stars with small angular separation are also
83: gravitationally bound \citep{bs81,gou95}, or by searching for common proper
84: motion pairs \citep{luy79,pov94,aph00,gs03,cg04,lb07}. The angular separation
85: method is simple to apply, but it also introduces a relatively large number of
86: false candidates due to chance association of nearby pairs. The contamination by
87: random associations can be reduced by imposing constraints, such as the common
88: proper motion, or by requiring that the stars are at similar distances. The
89: distances can be inferred through a variety of means, one of which is the use of
90: an appropriate photometric parallax\footnote{Also known as ``color-luminosity
91: relation''.} relation.
92: 
93: The photometric parallax relation provides the absolute magnitude of a star
94: given that star's color and metallicity. There are a number of proposed
95: photometric parallax relations for main sequence stars in the literature that
96: differ in the methodology used to derive them, photometric systems, and the
97: absolute magnitude and metallicity range in which they are applicable. Not all
98: of them are mutually consistent, and most exhibit significant intrinsic scatter
99: of order a half a magnitude or more (see Figure~2 in
100: \citealt[hereafter J08]{jur08}).
101: 
102: Instead of using an existing relation to select wide binaries, we propose a
103: novel method that {\em simultaneously} derives the photometric parallax relation
104: {\em and} selects a sample of wide binary candidates. The method relies on the
105: fact that components of a physical binary have equal distance moduli
106: ($m_1 - M_1 = m_2 - M_1$) and therefore
107: $\delta \equiv \Delta M - \Delta m \equiv (M_2 - M_1) - (m_2 - m_1)  = 0$.
108: Assuming that both stars are on the main sequence, and the {\em shape} of the
109: adopted photometric parallax relation is correct, the difference in absolute
110: magnitudes $\Delta M=M_2-M_1$ calculated from the parallax relation must equal
111: the measured difference of apparent magnitudes, $\Delta m=m_2-m_1$. The
112: $\Delta M = \Delta m$ equality for binaries must be valid irrespective of color,
113: and therefore represents a test of the validity of the adopted photometric
114: parallax relation or, alternatively, a way to estimate the parallax relation.
115: 
116: In practice, the distribution of $\delta$ will not be a delta-function both due
117: to instrumental (finite photometric precision) and physical effects (true
118: vs.~apparent pairs). However, for {\em true} wide binaries, the distribution of
119: $\delta$ is expected to be narrow, strongly peaked at zero, and the individual
120: $\delta$ values are expected to be uncorrelated with color. In contrast, the
121: distribution of $\delta$ values for randomly associated stellar pairs (hereafter
122: random pairs) should be much broader even when the correct photometric parallax
123: relation is adopted, reflecting the different distances of components of
124: projected binary pairs. This dichotomy can be used to assign a probability to
125: each candidate, of whether it is a true physical binary or a result of chance
126: projection on the sky.
127: 
128: The paper is organized as follows. In Section~\ref{data}, we give an overview
129: of the SDSS imaging data, and describe the selection, completeness and
130: population composition of two initial, independent samples of candidate
131: binaries. In Section~\ref{method} we describe the photometric parallax
132: estimation method, compare the best-fit photometric parallax relations to the
133: J08 relation, and analyze the scatter in predicted absolute magnitudes. The
134: properties of wide binaries, such as the color and spatial distributions, are
135: analyzed in Section~\ref{properties}. Finally, the results and their
136: implications for future surveys are discussed in Section~\ref{discussion}.
137: 
138: \section{The Data}
139: \label{data}
140: 
141: \subsection{Overview of the SDSS Imaging Data}
142: \label{sdss_overview}
143: 
144: Thanks to the quality of its photometry and astrometry, as well as the large
145: sky coverage, the SDSS stands out among available optical sky surveys. The SDSS
146: provides homogeneous and deep ($r<22.5$) photometry in five bandpasses ($u$,
147: $g$, $r$, $i$, and $z$, \citealt{gun98,hog02,smi02,gun06,tuc06}) accurate to
148: 0.02 mag (rms scatter) for unresolved sources not limited by photon statistics
149: \citep{scr02,ive03}, and with a zeropoint uncertainty of 0.02 mag \citep{ive04}.
150: The survey sky coverage of 10,000 deg$^2$ in the northern Galactic cap and 300
151: deg$^2$ in the southern Galactic cap results in photometric measurements for
152: well over 100 million stars and a similar number of galaxies \citep{sto02}. The
153: recent Data Release 6 \citep{amc08} lists\footnote{See
154: \url[HREF]{http://www.sdss.org/dr6}} photometric data for 287 million unique
155: objects observed in 9583 deg$^2$ of sky, and can be accessed through the Catalog
156: Archive Server\footnote{\url[HREF]{http://cas.sdss.org}} (CAS) 
157: CasJobs\footnote{\url[HREF]{http://casjobs.sdss.org/CasJobs/}}
158: interface. Astrometric positions are accurate to better than $0.1\arcsec$ per
159: coordinate (rms) for sources with $r<20.5$ \citep{pie03}, and the morphological
160: information from the images allows reliable star-galaxy separation to
161: $r\sim21.5$ \citep{lup02}.
162: 
163: The five-band SDSS photometry can be used for very detailed source
164: classification, e.g., separation of quasars and stars \citep{ric02}, spectral
165: classification of stars to within one to two spectral subtypes
166: \citep{len98,fin00,haw02,cov07}, identification of horizontal-branch and RR
167: Lyrae stars \citep{yan00,sir04,ive05,ses07}, and low-metallicity G and K giants
168: \citep{hel03}.
169: 
170: Proper motion data exist for SDSS sources matched to the USNO-B1.0 catalog
171: \citep{mon03}. We take proper motion measurements from the \citet{mun04} catalog
172: based on astrometric measurements from the SDSS and Palomar Observatory Sky
173: Surveys (POSS-I; \citealt{ma63}, POSS-II; \citealt{rei91}). Despite the sizable
174: random and systematic astrometric errors in the Schmidt surveys, the combination
175: of a long baseline (50 years for the POSS-I survey) and a recalibration of the
176: photographic data using positions of SDSS galaxies, results in median random
177: errors for proper motions of only 3 mas yr$^{-1}$ for $r < 19.5$ (per
178: coordinate), with substantially smaller systematic errors \citep{mun04}.
179: Following a recommendation by Munn et al., when using their catalog we select
180: SDSS stars with only one USNO-B match within $1\arcsec$, and require proper
181: motion rms fit residuals to be less than 350 mas in both coordinates. We note
182: that the proper motion measurements publicly available as a part of SDSS Data
183: Release 6 are known to have significant systematic errors (Munn et al., 
184: in prep.). Here we use a revised set of proper motion measurements which will
185: become publicly available as a part of SDSS Data Release 7.
186: 
187: \subsection{The Initial Sample of Close Resolved Stellar Pairs}
188: \label{initial_select}
189: 
190: For objects in the SDSS catalog, the photometric pipeline \citep{lup02} sets a
191: number of flags that indicate the status of each object, warn of possible
192: problems with the image itself, and warn of possible problems in the measurement
193: of various quantities associated with the object. These flags can be used to
194: remove duplicate detections (in software) of the same object, and to select
195: samples of unresolved sources with good photometry.
196: 
197: According to the SDSS Catalog Archive Server ``Algorithms'' webpage\footnote{
198: \url[HREF]{http://cas.sdss.org/dr6/en/help/docs/algorithm.asp?key=flags}},
199: duplicate detections of the same objects can be removed by considering only
200: those which have the ``status'' flag set to PRIMARY. We consider only PRIMARY
201: objects, and select those with good photometry by requiring that the BINNED1
202: flag is set to 1, and PSF\_FLUX\_INTERP, DEBLEND\_NOPEAK, INTERP\_CENTER,
203: BAD\_COUNTS\_ERROR, NOTCHECKED, NOPROFILE, PEAKCENTER, and EDGE image processing
204: flags are set to 0 in the $gri$ bands. The moving unresolved sources, such as
205: asteroids, are avoided by selecting sources with the DEBLENDED\_AS\_MOVING flag
206: set to 0.
207: 
208: Good photometric accuracy (mean PSF magnitude errors $<0.03$ mag, see Figure~1
209: in \citealt{ses07}) is obtained by selecting sources with $14<r'<20.5$, where
210: $r'$ is the $r$ band PSF magnitude uncorrected for ISM extinction. The PSF
211: magnitudes corrected for ISM extinction (using maps by \citealt{SFD98}), and
212: used throughout this work, are noted as $u$, $g$, $r$, $i$, and $z$.
213: 
214: To create the initial sample of resolved stellar pairs, we
215: query\footnote{SQL queries are listed in Appendix~\ref{appendix}} the CAS
216: ``Neighbors'' table (lists all SDSS pairs within $30\arcsec$) for pairs of
217: sources that pass the above criteria, and that have
218: \begin{equation}
219: (r_1-r_2)[(g-i)_1-(g-i)_2]>0\label{Teff},
220: \end{equation}
221: where the subscript 1 is hereafter assigned to the brighter component. With this
222: condition we require that the component with bluer $g-i$ color is brighter in
223: the $r$ band. About $40\%$ of random pairs are rejected with this condition. We
224: estimate that about $3\%$ of true binary systems might be excluded by this cut
225: (due to uncertainties in the $g-i$ color caused by photometric errors), but
226: their exclusion does not significantly influence our results.
227: 
228: We select $\sim4.2$ million pairs for the initial sample of resolved stellar
229: pairs, and plot the observed distribution of angular separation $\theta$,
230: $f_{obs}(\theta)$, in Figure~\ref{ang_sep} ({\em top}). For a uniform (random)
231: distribution of stars, the number of neighboring stars within an annulus
232: $\Delta \theta$ increases linearly with $\theta$, and therefore, the number
233: of random pairs also increases with $\theta$. To find the number of random pairs
234: as a function of $\theta$, we fit $f_{rnd}(\theta)= C \, \theta$ to the
235: $f_{obs}$ histogram (in the $\theta > 15\arcsec$ region), and find $C=9043$
236: arcsec$^{-1}$. For large separation angles ($\theta > 15\arcsec$) the two
237: distributions closely match, indicating that the majority of observed pairs are
238: simply random associations, and are not physically related. At separation
239: angles smaller than $\sim15\arcsec$ the frequency of observed pairs shows an
240: excess, suggesting the presence of true, gravitationally bound systems. However,
241: even at small separation angles, the selected pairs include a non-negligible
242: fraction of random pairs and require further refinement, or careful statistical
243: accounting for random contamination.
244: 
245: Throughout this work we use samples of random pairs (random samples, hereafter)
246: to account for random contamination in candidate binaries. We define the random
247: sample as a sample of pairs with $20\arcsec<\theta<30\arcsec$ taken from the
248: initial pool of stellar pairs. Since pairs in the random sample pass the same
249: data quality selection as candidate binaries, and since virtually all of them
250: are chance associations ($99.75\%$; see Section~\ref{samples} and
251: Figure~\ref{ang_sep}), the random sample is a fair representation of the
252: population of randomly associated stars in candidate binary samples.
253: 
254: \subsection{The Geometric Selection}
255: \label{geo_select}
256: 
257: The excess of pairs with $\theta<15\arcsec$ in Figure~\ref{ang_sep}
258: ({\em top}) likely indicates a presence of true binaries, and the angular
259: separation provides a simple, {\em geometric} criterion to select candidate
260: binary systems. This excess, shown as the ratio $f_{obs}/f_{rnd}$ in
261: Figure~\ref{ang_sep} ({\em bottom}), increases for $\theta<15\arcsec$, reaches a
262: relatively flat peak of $\sim1.45$ for $3\arcsec<\theta<4\arcsec$, and sharply
263: decreases for $\theta<2\arcsec$ due to finite seeing and inability to resolve
264: close pairs of sources. This excess is related to the fraction of true binaries,
265: $\epsilon(\theta)$, as
266: \begin{equation}
267: \epsilon(\theta) = 1 - f_{rnd}(\theta)/f_{obs}(\theta)\label{eff2}.
268: \end{equation}
269: Using Figure~\ref{ang_sep} ({\em bottom}), we choose $3\arcsec<\theta<4\arcsec$
270: for our geometric selection criterion, since the fraction of true binaries is
271: expected to reach a maximum of $\sim35\%$ in this range.
272: 
273: The interpretation of the excess of close stellar pairs as gravitationally bound
274: binary pairs implies that the components are at similar distances. If this is
275: true, and if it is possible to constrain the distance via a photometric parallax
276: relation, than their distribution in the color-magnitude diagram should be
277: different than for a sample of randomly associated stars.
278: 
279: To test this hypothesis, we select 51,753 candidate binaries with
280: $3\arcsec<\theta<4\arcsec$. We compare their distribution in the
281: $\Delta r=r_2-r_1$ vs.~$\Delta(g-i)=(g-i)_2 - (g-i)_1$ diagram to the
282: distribution of pairs from the random sample, as shown in Figure~\ref{counts}.
283: The number of pairs in this random sample is restricted to 51,753. Were the
284: selection a random process, the selected candidates would have the same
285: distribution in this diagram as the random sample, and the average
286: candidate-to-random ratio would be $\sim1$. However, in the region where
287: \begin{equation}
288: 4.33\Delta(g-i) - \Delta r + 0.4 > 0\label{color_cut1},
289: \end{equation}
290: and
291: \begin{equation}
292: 2.31\Delta(g-i) - \Delta r - 0.46 < 0\label{color_cut2}
293: \end{equation}
294: the two distributions are different (average candidate-to-random ratio of
295: $\sim1.7$), implying that $>40\%$ of candidates are found at similar
296: distances. In principle, a selection cut using Equations~\ref{color_cut1}
297: and~\ref{color_cut2} could be made to increase the fraction of true binaries in
298: the candidate sample. We do not make such a cut a priori, but instead develop a
299: method (described in Section~\ref{method}) that robustly ``ignores'' random
300: pairs while estimating the photometric parallax relation. After a best-fit
301: photometric parallax relation is obtained, the contamination can be minimized by
302: selecting only pairs where both components are at similar distances, as
303: described in Section~\ref{samples}.
304: 
305: The $r$ vs.~$g-i$ distributions of brighter and fainter components of candidate
306: binaries are shown in Figure~\ref{r_vs_gi}. We find that the brighter components
307: in the candidate sample are mostly disk G to M dwarfs, while the fainter
308: components are mostly M dwarfs.
309: 
310: \subsection{The Kinematic Selection}
311: \label{kin_select}
312: 
313: As seen from Figure~\ref{ang_sep} ({\em top}), candidate binaries with
314: $\theta>15\arcsec$ cannot be efficiently selected using angular distance only,
315: as nearly all pairs in this range are most likely chance associations. In this
316: regime, a {\em kinematic} selection based on common proper motion should be more
317: efficient, as random pairs have a small probability ($\sim0.005$ determined
318: using Monte Carlo simulations) to be common proper motion pairs (using selection
319: criteria listed below).
320:  
321: We therefore select a second sample of 14,148 candidate binaries by searching
322: for common proper motion pairs with proper motion difference
323: $\Delta \mu=|\mathbf{\mu_2}-\mathbf{\mu_1}|<5$ mas $yr^{-1}$, and with absolute
324: proper motion in the range 15 mas
325: $yr^{-1}<|\mathbf{\mu}|_{max}<$ 400 mas $yr^{-1}$, where
326: $|\mathbf{\mu}|_{max}=max(|\mathbf{\mu_1}|,|\mathbf{\mu_2}|)$. These criteria
327: require that the directions of two proper motion vectors agree at a $1\sigma$
328: level, and that the proper motion is detected at a $5\sigma$ level or higher.
329: The common proper motion pairs with orbital motion $\ga1\arcsec$ over 50 years
330: are not selected because their USNO-B and SDSS positions place them outside the
331: $1\arcsec$ search radius used by Munn et al. The angular separation of common
332: proper motion pairs is limited to $9\arcsec<\theta<30\arcsec$. Pairs of sources
333: with $\theta<9\arcsec$ are usually blended in the USNO-B data and may not have
334: reliable proper motion measurements (see Section~\ref{limitations}), while the
335: maximum angular separation between sources in the CAS ``Neighbors'' table
336: defines the upper limit of $\theta<30\arcsec$. However, for purposes of
337: Section~\ref{limitations}, we have created a sample of common proper motion
338: pairs that extends to $\theta=500\arcsec$. We have done so by matching
339: SDSS sources (that pass the quality flags from Section~\ref{initial_select})
340: within a $500\arcsec$ search radius into common proper motion pairs. Since this
341: matching is computationally expensive, we have done this only for one sample.
342: The $r$ vs.~$g-i$ distributions of brighter and fainter components of
343: kinematically-selected candidate binaries are similar to those shown in
344: Figure~\ref{r_vs_gi}.
345: 
346: \subsection{The Sample Completeness}
347: \label{composition}
348: 
349: Before proceeding with the determination of photometric parallax relations and
350: discussion of the properties of wide binary systems, we summarize the
351: completeness of geometric and kinematic samples, and estimate their expected
352: fractions of disk and halo populations. The samples are selected from a
353: highly-dimensional space of measured parameters and an understanding of the
354: selection effects is a prerequisite for determining the limitations of various
355: derived statistical properties. For example, the geometric sample is selected
356: using five parameters: the $g-i$ color of the two components, $(g-i)_1$ and
357: $(g-i)_2$, their apparent magnitudes, $r_1$ and $r_2$, and their angular
358: separation on the sky, $\theta$. The latter three can be transformed with the
359: aid of a photometric parallax relation into a difference of their apparent
360: magnitudes, $\Delta m =r_2-r_1$, distance $D$, and the projected physical
361: separation, $a$. We seek to constrain the photometric parallax relation by
362: minimizing the difference $\delta = \Delta M - \Delta m$, where $\Delta M$ is a
363: two-dimensional function of $(g-i)_1$ and $(g-i)_2$ (Section~\ref{method}), and
364: at the same time derive constraints on the two-dimensional color distribution of
365: wide binaries, on their $a$ distribution, and on any variation of these
366: distributions with position in the Galaxy (Section~\ref{properties}). Not all of
367: these constraints can be derived independently of each other, and most are
368: subject to severe selection effects. By judiciously selecting data subsets and
369: projections of this five-dimensional parameter space, these effects can be
370: understood and controled, as described below. 
371: 
372: To illustrate the most important selection effects, we employ the photometric
373: parallax relation and its dependence on metallicity derived by
374: \citet[hereafter I08a]{ive08a}. The quantitative differences between their
375: photometric parallax relation and the ones derived here have negligible impact
376: on the conclusions derived in this Section. For simplicity, we select a sample
377: of $\sim$2.8 million stars with $r<21.5$ observed towards the north Galactic
378: pole ($b>70\arcdeg$), and study their counts as a function of distance and the
379: $g-i$ color. Due to this choice of field position, the distance to each star is
380: approximately equal to its distance from the Galactic plane (for a detailed
381: study of the dependence of stellar number density on position within the Milky
382: Way, see J08). Figure~\ref{Fig:DMvsColor} illustrates several important
383: selection effects. 
384: 
385: First, for any $g-i$ color there is a minimum and maximum distance corresponding
386: to the SDSS saturation limit at $r\sim14$ and the adopted faint limit at
387: $r=21.5$; the probed distance range extends from 100 pc to 25 kpc. Within the
388: distance limits appropriate for a given color, the sample is essentially
389: complete ($\sim$98\%, \citealt{fin00}). Second, these limits are strongly
390: dependent on color: the bluest stars saturate at a distance of about 1 kpc,
391: while the reddest stars are too faint to be detected even at a few hundred pc.
392: Equivalently, due to the finite dynamic range of SDSS apparent magnitudes, there
393: is no distance range where the entire color range from the blue disk turn-off
394: edge to the red edge of luminosity function is completely covered. At best, at
395: distances of about 1 kpc the color completeness extends from the blue edge to
396: the peak of luminosity function at $g-i \sim 2.7$. Third, when pairing stars
397: into candidate binary systems, their color distribution at a given distance (the
398: requirement that the differences of apparent and absolute magnitudes are similar
399: places the two stars from a candidate pair into a narrow horizontal strip in the
400: distance modulus (DM) vs.~$g-i$ diagram shown in Figure~\ref{Fig:DMvsColor})
401: will be clipped: the ratio of the number of candidate binaries and the number of
402: all single stars in the sample decreases at distances significantly different
403: from $\sim1$ kpc because of a bias against blue-red pairs.
404: 
405: The binary samples selected from the $\sim$1 kpc distance range can be used to
406: measure the two-dimensional color distribution of wide binaries, as well as to
407: gauge the dependence of their $a$ distribution on color. The dependence of the
408: $a$ distribution on distance from the Galactic plane can also be studied over a
409: substantial distance range, but {\it only under the assumption that it is
410: independent of color.} 
411: 
412: The imposed $\theta$ range ($3\arcsec$ to $30\arcsec$) limits the range of
413: probed physical separation to values proportional to distance, and ranging from
414: 3,000 AU to 30,000 AU at a distance of 1 kpc. We discuss and account for these
415: effects in more detail in Section~\ref{spatial}.
416: 
417: \subsection{The separation of disk and halo populations}
418: \label{separation}
419: 
420: The counts of main-sequence stars shown in Figure~\ref{Fig:DMvsColor} include
421: both disk and halo populations. With the available data, there are three methods
422: that might be used for separating stars (including candidate binary systems)
423: into disk and halo populations (Juri\'{c} et al., in prep.):
424: \begin{enumerate}
425: \item
426: A statistical method based on the stellar number density profiles (J08): beyond
427: about 3 kpc from the plane, halo stars begin to dominate. However, as shown in
428: Figure~\ref{Fig:DMvsColor}, only stars bluer than $g-i=2$ are detected at such
429: distances. The stellar number density profiles suggest that the fraction of halo
430: stars is below $\sim$20\% closer than 1.5 kpc from the Galactic plane (see
431: Figure~6 in I08a).
432: \item
433: Classification based on metallicity into low-metallicity
434: ($\left[Fe/H\right]<-1$) halo stars and higher metallicity stars. As shown by
435: I08a, this is a robust and accurate method even when using photometric
436: metallicity estimator, but it works only for stars with $g-i \la 0.7$ due to the
437: limitations of the photometric metallicity method, and the SDSS spectroscopic
438: metallicity is available only for a small fraction of stars in the candidate
439: samples.
440: \item 
441: Kinematic selection based on proper motion measurements, and implemented via a
442: reduced proper motion diagram (e.g., \citealt{sg03}; \citealt{mun04}, and
443: references therein). However, as discussed in detail in Appendix B, this method
444: is robust only closer than 2-3 kpc from the Galactic plane due to a rotational
445: velocity gradient of disk stars which diminishes kinematic differences between
446: halo and disk stars further away from the plane. 
447: \end{enumerate}
448: 
449: Given the limitations of these methods, it is not possible to reliably separate
450: disk and halo populations throughout the explored parameter space, and in both
451: geometric and kinematic samples. For geometric sample, the third method is not
452: applicable because SDSS-POSS proper motions are not reliable at small angular
453: distances ($\theta\la9\arcsec$; see Section~\ref{limitations}). The requirement
454: $g-i \la 0.7$ required for the second method results in a subsample with too
455: narrow a color range to constrain the photometric parallax relation.
456: Nevertheless, the analysis of this subsample based on results from I08a
457: indicates that fewer than 10\% of stars in geometric sample belong to halo
458: population (this fraction increases with the distance from the Galactic plane;
459: see Figure 6 in I08a), and thus we expect that halo contamination plays only a
460: minor role in the geometric sample. 
461: 
462: The kinematic sample is expected to include a non-negligible fraction of halo
463: stars due to the selection of stars with substantial proper motions. We use the
464: reduced proper motion diagram to estimate the fraction of halo candidate binary
465: stars in this sample. The reduced proper motion for an arbitrary photometric
466: bandpass, here $r$, is defined as
467: \begin{equation}
468: \label{Eq:rpm}
469:                    r_{RPM} =  r + 5\log{(\mu)},
470: \end{equation}
471: where $\mu$ is proper motion in arcsec $yr^{-1}$ (sometimes an additional offset
472: of 5 mag is added). Using a relationship between proper motion, distance and
473: tangential velocity,
474: \begin{equation}
475:                   v_t = 4.47 \, \mu \, D 
476: \end{equation}
477: and 
478: \begin{equation} 
479:                   r-M_r = 5\log{(D)}-5, 
480: \end{equation}
481: Equation~\ref{Eq:rpm} can be rewritten as 
482: \begin{equation}
483: \label{Eq:rpm2}
484:                r_{RPM} =  M_r + 5\log{(v_t)} + C,
485: \end{equation}
486: where $D$ is distance in parsec, $M_r$ is the absolute magnitude, and $v_t$ is
487: the heliocentric tangential velocity (the projection of the heliocentric
488: velocity on the plane of the sky), and $C$ is a constant ($C=-8.25$ if $v_t$ is
489: expressed in km $s^{-1}$). Therefore, for a population of stars with the same
490: $v_t$, the reduced proper motion is a measure of their absolute magnitude. As
491: shown using similar data as discussed here, halo and disk stars form two
492: well-defined and separated  sequences in the reduced proper motion vs.~color
493: diagram (e.g., \citealt{sg03}; \citealt{mun04}; and references therein). We
494: discuss the impact of different metallicity and velocity distributions of halo
495: and disk stars on their reduced proper motion distributions in more detail in
496: Appendix B.
497: 
498: Figure~\ref{Fig:1wd} shows reduced proper motion diagrams for stars observed 
499: towards the north Galactic pole, constructed for two ranges of observed proper 
500: motion: 15-50 mas $yr^{-1}$ and 50-400 mas $yr^{-1}$. The choice of the proper
501: motion range, together with unavoidable apparent magnitude limits, strongly
502: affects the probed distance range: the larger is the proper motion, the closer
503: is the distance range over which the selection fraction is non-negligible. We
504: find that the two sequences closely follow the expectations based on the
505: analysis of metallicity and velocity distributions from I08a. The halo sequence
506: can be efficiently separated by selecting stars with reduced proper motion
507: larger than a boundary generated using the photometric parallax relation from
508: I08a, evaluated for the median halo metallicity ($\left[Fe/H\right]=-1.5$) and
509: with $v_t$=180 km $s^{-1}$ (see Equation~\ref{Eq:rpm2}). This separation method
510: is conceptually identical to the $\eta$ separator discussed by \citet{sg03}.
511: They also proposed to account for a shift of the reduced proper motion sequences
512: with galactic latitude, an effect which we discuss in more detail in Appendix B.
513: For the reasons described there, to account for the variation of the reduced
514: proper motion sequences away from the Galactic pole, we simply offset the $v_t$
515: value from 180 km $s^{-1}$ to 110 km $s^{-1}$ (i.e., the separator moves upwards
516: in Figure~\ref{Fig:1wd} by 1 mag). While this selection removes some disk
517: binaries, it is designed to exclude most of halo binaries from the sample.
518: 
519: With the aid of reduced proper motion separator, we separate kinematic sample
520: into candidate halo (1,336 pairs) and disk binaries (10,112 pairs). This
521: fraction of halo systems is consistent with the above estimate obtained for the
522: geometric sample. To assess selection effects, we first investigate the sample
523: of single stars. The top left panel in Figure~\ref{Fig:2wd} shows the fraction
524: of all the stars shown in Figure~\ref{Fig:1wd} that have proper motion larger
525: than 15 mas $yr^{-1}$ and $r<19.5$ (the latter limit ensures the SDSS-POSS
526: proper motion catalog completeness above $\sim$90\%). The selection efficiency
527: is a strong function of distance, and falls from its maximum of $\sim$95\% for
528: nearby stars to below 50\% at a distance of about 1 kpc. The candidate disk
529: stars are detected in significant numbers to $\sim$3 kpc, and halo stars beyond
530: $\sim$1 kpc. The fraction of selected stars that are classified as halo stars is
531: below 20\% closer than $\sim$1.5 kpc from the Galactic plane, and becomes
532: essentially 100\% beyond 3 kpc. 
533: 
534: The kinematic difference between halo and disk stars is blurred at distances
535: beyond 2-3 kpc (see Appendix B), and the majority of disk stars at such
536: distances are misidentified as halo stars (the metallicity distribution implies
537: that disk stars do exist at distances as large as 7 kpc from the Galactic plane,
538: see Figure~10 in I08a). To demonstrate this effect, we use subsamples of
539: candidate disk and halo binaries identified using the reduced proper motion
540: diagram that have $0.2<(g-r)_1<0.4$. For these pairs it is possible to estimate
541: photometric metallicity (I08a) and use it as an independent population
542: classifier. Figure~\ref{plot_ug} shows that practically all candidate binaries
543: with $\left[Fe/H\right]>-1$ further than $\sim$2 kpc from the Galactic plane are
544: misclassified as halo stars when using reduced proper motion diagram. 
545: 
546: In summary, geometric sample is heavily dominated by disk binaries, with halo
547: contamination all but negligible closer than about 2 kpc from the plane.
548: Kinematic sample becomes severely incomplete ($<$50\%) further than $\sim$2 kpc
549: from the plane, and has a higher fraction of halo binaries than geometric
550: sample, at a given distance from the plane. However, this halo contamination can
551: be efficiently removed using the reduced proper motion diagram. Unfortunately,
552: the number of selected halo binaries is insufficient in number (1,336 in
553: kinematic and 5,556 in geometric sample), and spans too narrow a color range to
554: robustly constrain the photometric parallax relation. Therefore, both samples of
555: candidate binaries are supposed to yield similar photometric parallax relations,
556: because both are dominated by disk stars.
557: 
558: \section{The Photometric Parallax Estimation Method}
559: \label{method}
560: 
561: In principle, both the normalization and the shape of the photometric parallax 
562: relation (i.e., the shape of the main sequence in the Hertzsprung-Russell
563: diagram) vary as a function of color and metallicity \citep{lcl88, sie02}. Since
564: our data do not allow a reliable estimate of metallicity over the entire range
565: of observed colors, we can only estimate the ``mean'' shape of the photometric
566: parallax relation as a function of color, for all metallicities present in the
567: sample. Such a mean shape is approximately an average of individual
568: $\left[Fe/H\right]$-dependent relations, weighted by the sample metallicity
569: distribution. J08 derived such ``mean'' photometric parallax relations
570: appropriate at the red end for the nearby, metal-rich stars, and at the blue end
571: for distant, metal-poor stars. I08a discuss the offset of photometric parallax
572: relation as a function of metallicity (see their Figure~20), and derived the
573: metallicity range implied by ``mean'' photometric parallax relations from J08.
574: The derived metallicity range is consistent with the spatial distribution of
575: metallicity derived by I08a and the color-magnitude limits of the SDSS survey.
576: 
577: \subsection{The Photometric Parallax Parametrization}
578: \label{parametrization}
579: 
580: We adopt the J08 polynomial $r-i$ parametrization of the photometric parallax
581: relation
582: \begin{equation}
583: M_r(r-i|\mathbf{p})=A+B(r-i)+C(r-i)^2+D(r-i)^3+E(r-i)^4\label{photo_parallax},
584: \end{equation}
585: where $\mathbf{p}=(A,B,C,D,E)$ are the parameters we wish to estimate. To
586: improve their accuracy, Juri\'c et al.~used a maximum likelihood technique to
587: estimate the $r-i$ color from the observed $g-r$ and $r-i$ colors. Because of
588: the brighter flux limit employed here, we use the measured $g-i$ color to derive
589: a best estimate of the $r-i$ color via a stellar locus relation (J08):
590: \begin{equation}
591: g-i = 1.39(1-exp[-4.9(r-i)^3-2.45(r-i)^2-1.68(r-i)-0.050]) + r - i\label{interp}
592: \end{equation}
593: The $r-i$ color estimate obtained with this method has several times smaller
594: noise than the measured $r-i$ color. This is because the observed dynamic range
595: for the $g-i$ color is much larger than of the $r-i$ color ($\sim3$ mag
596: vs.~$\sim1$ mag), while their measurement errors are similar.
597: 
598: \subsection{The Parameter Estimation Algorithm}
599: \label{algorithm}
600: 
601: The goal of parameter estimation algorithm is to determine the photometric
602: parallax relation, $M_r(r-i|\mathbf{p})$, that minimizes the width of the
603: distribution of $\delta$ values for {\em true} binary systems, where
604: $\delta = (M_{r2}-M_{r1})-(r_2-r_1)$. The $\chi^2$ minimization cannot be used
605: for this purpose because random pairs, if not removed from the sample, will
606: strongly bias the best-fit $M_r$. The available color, angular separation, and
607: proper motion information are insufficient to separate the random pairs from the
608: true binaries. Therefore, we need to design a fitting algorithm that will be
609: least affected as possible by random pairs.
610: 
611: We begin by studying the behavior of $\delta$ values in mock catalogs. The first
612: step in creating a mock catalog is the selection of 51,753 (random) pairs from
613: the random sample. Note that the fraction of true binaries in the random sample
614: is only $\sim0.25\%$ (see Section~\ref{samples}). True binaries are then
615: ``created'' in the mock catalog by replacing the observed $r_2$ magnitudes for
616: $20\%$ of pairs with
617: \begin{equation}
618: r_2 = r_1 + (M_{r2}-M_{r1}) + N(0,0.1)\label{r2_mag},
619: \end{equation}
620: where $M_r=M_r(r-i|\mathbf{p_0})$ and
621: $\mathbf{p_0}=(3.2,13.30,-11.50,5.40,-0.70)$ (Equation 2 coefficients from J08).
622: The $N(0,0.1)$ is a Gaussian random variate added to account for the intrinsic
623: scatter around the photometric parallax relation. The result of this process is
624: a mock sample of candidates where $20\%$ of pairs are ``true'' binaries, and the
625: rest ($80\%$) is the contamination made of random pairs. The distribution of
626: $\delta$ values for ``true'' binaries is, by definition, a 0.1 mag wide Gaussian
627: centered on zero when $M_r=M_r(r-i|\mathbf{p_0})$.
628: 
629: Figure~\ref{delta_hist_mock} ({\em top}) shows the distribution of $\delta$
630: values for the mock sample evaluated with the ``true'' [$M_r(r-i|\mathbf{p_0})$]
631: photometric parallax relation. The observed $\delta$ distribution can be
632: described as a sum of a Gaussian and a non-Gaussian component. The non-Gaussian
633: component is due to random pairs (the contamination), while the Gaussian
634: component (0.1 mag wide and centered on zero) is due to the true binaries.
635: 
636: When an $M_r$ relation different from the ``true'' (or best-fit) $M_r$ is
637: adopted, the Gaussian component becomes wider and {\em the peak height of the
638: $\delta$ distribution decreases}, as shown in Figure~\ref{delta_hist_mock}
639: ({\em bottom}). At the same time, the peak height of the $\delta$ distribution
640: of the contamination changes much less since the distribution is much wider
641: ($\sim2.3$ mag wide). Therefore, {\em minimizing} the width of the $\delta$
642: distribution of true binaries, is equivalent to {\em maximizing} the peak height
643: of the entire $\delta$ distribution. We quantify this peak height as the number
644: of candidate binaries in the most populous $\delta$ bin.
645: 
646: \subsection{The Algorithm Implementation}
647: \label{implementation}
648: 
649: To robustly explore the parameter space that defines the photometric parallax
650: relation, and to find the best-fit coefficients $\mathbf{p}$, we implement our
651: algorithm as a Markov chain Monte Carlo (MCMC) process. The MCMC description
652: given here and our implementation of the algorithm are based on examples given
653: by \citet{teg04}, \citet{for05}, and \citet{cro06}.
654: 
655: The basic idea of the MCMC aproach is to take an {\em n}-step intelligent random
656: walk around the parameter space while recording the point in parameter space for
657: each step. Each successive step is allowed to be some small distance in
658: parameter space from the previous position. A step is always accepted if it
659: improves the fit, and is sometimes accepted on a random basis even if the fit is
660: worse, where the goodness of the fit is quantified by some parameter (usually
661: with $\chi^2$). The random acceptance of a bad fit ensures that the MCMC does
662: not become stuck in a local minimum, and allows the MCMC to fully explore the
663: surrounding parameter space.
664: 
665: We start a Monte Carlo Markov chain by setting all coefficients from
666: Equation~\ref{photo_parallax} to zero ($\mathbf{p_i}=0$). Using this initial set
667: of coefficients we evaluate $\delta=(M_{r2}-M_{r1})-(r_2-r_1)$ for all candidate
668: binaries assuming $M_r(r-i|\mathbf{p_i})$, and bin $\delta$ values in 0.1 mag
669: wide bins. The number of candidate binaries in the most populous bin, $P_i$, is
670: used to quantify the relative goodness of the fit.
671: 
672: Given $\mathbf{p_i}$, a new candidate step,
673: $\mathbf{p_n}=\mathbf{p_i}+\Delta\mathbf{p}$, is generated, where the step size,
674: $\Delta \mathbf{p}$, is a vector of independent Gaussian random variates with
675: initial widths, $\sigma$, set to 1. Using the candidate set of coefficients,
676: $\mathbf{p_n}$, $\delta$ values are evaluated, binned, and the height of the
677: $\delta$ distribution is assigned to parameter $P_n$.
678: 
679: Following the Metropolis-Hastings rule \citep{met53,has70} the candidate step is
680: accepted ($\mathbf{p_{i+1}}=\mathbf{p_n}$, $P_{i+1}=P_n$) if $P_n > P_i$ or if
681: $exp(P_n - P_i)>\xi$, where $\xi$ is a random number between 0 and 1
682: ($\xi\in[0,1]$). Otherwise, the candidate step is rejected.
683: 
684: While the Metropolis-Hastings rule guarantees that the chain will converge, it
685: does not specify {\em when} the convergence is achieved. The speed of the
686: convergence depends on the Gaussian scatter $\sigma$ used to calculate the step
687: size $\Delta \mathbf{p}$. If the scatter is too large, a large fraction of
688: candidate steps is rejected, causing the chain to converge very slowly. If the
689: scatter is too small, the chain behaves like a random walk, and the number of
690: steps required to traverse some short distance in the parameter space scales as
691: $1/\sigma^2$. The choice of optimal Gaussian scatter $\sigma$ (for each fitted
692: coefficient), as a function of the position in the parameter space, is not
693: trivial and it can be very complicated even if the fitted coefficients are
694: uncorrelated.
695: 
696: To determine the optimal $\sigma$ values we follow the \citet{teg04}
697: prescription (see their Appendix A). After every 100 accepted steps we compute
698: the coefficient covariance matrix
699: $\mathbf{C} = \langle \mathbf{p} \mathbf{p}^t \rangle - \langle \mathbf{p} \rangle \langle \mathbf{p}^t \rangle$
700: from the chain itself, diagonalize it as
701: $\mathbf{C}=\mathbf{R} \Lambda \mathbf{R}^t$, and use it to calculate a new step
702: size $\Delta \mathbf{p'}=\mathbf{R}^t \mathbf{\Lambda}^{1/2} \Delta \mathbf{p}$
703: for each coefficient separately. We find that this transformation greatly
704: accelerates the convergence of a chain.
705: 
706: Due to the stochastic nature of the MCMC, the best-fit relations (coefficients
707: with the highest $P_i$ value in a chain) from different chains will not
708: necessarily be the same. To quantify the intrinsic scatter between different
709: best-fit relations, we run fifty 10,000-element long chains, and select the
710: best-fit coefficients from each chain for subsequent comparison (see
711: Section~\ref{robustness}). The proper mixing and convergence of chains is
712: confirmed using the Gelman \& Rubin $R$ statistic \citep{gr92}. Gelman \& Rubin
713: suggest running the chains until $R<1.2$ for all fitted coefficients. With
714: 10,000 elements in each chain, we obtain $R<1.01$ for all fitted coefficients.
715: 
716: In the end, we select $\mathbf{p}=(A,B,C,D,E)$ with the highest $P_i$ value
717: among all chains as our best-fit relation. The constant term $A$ is not
718: constrained with our algorithm, because $A$ (from $M_{r2}$ and $M_{r1}$) cancel
719: out when evaluating $\delta$. Instead, we constrain $A$ by requiring $M_r=10.07$
720: at $r-i=1.1$, obtained from trigonometric parallaxes of nearby M dwarfs
721: \citep{wwh05}.
722: 
723: \subsection{Algorithm Robustness Test}
724: \label{robustness}
725: 
726: To test the robustness of our algorithm, we apply it to the mock sample
727: described in Section~\ref{algorithm}. The best-fit relations (obtained from
728: Markov chains) are compared on a $0.1\leqslant r-i\leqslant1.5$ grid in 0.01 mag
729: steps. We find an rms scatter of 0.05 mag between Markov chains, and 0.05 mag
730: rms scatter between the true and the best-fit relation with the highest $P_i$
731: value.
732: 
733: We repeat this test with a mock sample containing $30\%$ of true binaries. The
734: rms scatter between the best-fit relations decreases to 0.03 mag, and the rms
735: scatter between the true and the best-fit relation with the highest $P_i$ value
736: decreases to 0.01 mag.
737: 
738: Even when only $20\%$ of sources are true binaries (i.e., contamination by
739: random pairs is $80\%$) our algorithm recovers the ``true'' photometric parallax
740: relation at the 0.05 mag (rms) level. The accuracy of the fit increases (to 0.01
741: mag rms) as the contamination decreases (from $80\%$ to $70\%$).
742: 
743: \subsection{Best-fit Photometric Parallax Relations}\label{analysis}
744: 
745: We apply the method described in Section~\ref{implementation} to two samples of
746: candidate binaries and obtain the best-fit photometric parallax relations
747: \begin{equation}
748: M_r = 3.32 + 15.02(r-i) - 18.58(r-i)^2 + 13.28(r-i)^3 - 3.39(r-i)^4
749: \label{Mr_geo}
750: \end{equation}
751: \begin{equation}
752: M_r = 3.42 + 13.75(r-i) - 15.50(r-i)^2 + 10.40(r-i)^3 - 2.43(r-i)^4
753: \label{Mr_kin}
754: \end{equation}
755: for the geometrically- and kinematically-selected samples, respectively.
756: Candidate halo binaries were removed from the kinematically-selected sample
757: using reduced proper motion diagrams (Section~\ref{separation}) before the
758: Equation~\ref{Mr_kin} was derived. The photometric parallax relations for halo
759: stars cannot be robustly constrained using geometrically- or
760: kinematically-selected halo binaries because the color range they span is too
761: narrow ($g-i < 1.0$ at 3-4 kpc, see Figures~\ref{Fig:DMvsColor}
762: and~\ref{Fig:2wd}).
763: 
764: We test the correctness of the shape by studying the dependence of median
765: $\delta$ values on the $g-i$ colors of the brighter and the fainter components.
766: If the {\em shape} of these photometric parallax relations is correct, the
767: distribution of $\delta$ values will be centered on zero, and the individual
768: $\delta$ values will not correlate with color. The medians are used because they
769: are more robust to outliers (random pairs in the sample). We start by
770: calculating $\delta$ values for each candidate binary sample (using the
771: appropriate $M_r$ relation), and then select candidates with $|\delta|<0.4$.
772: This cut reduces the contamination by random pairs, as demonstrated in
773: Section~\ref{scatter}. The selected candidate binaries are binned in $g-i$
774: colors of the brighter and the fainter component, and the median $\delta$ values
775: are shown in Figure~\ref{medians}.
776: 
777: The distributions of the median $\delta$ for each pixel are fairly narrow
778: (0.07 mag), and centered on zero. Irrespective of color and the choice of the
779: two best-fit photometric parallax relations, the deviations are confined to the
780: 0.25 mag range, placing an upper limit on the errors in the mean shape of the
781: adopted relations.
782: 
783: In Figure~\ref{Mr_ri} we compare the adopted photometric parallax relations 
784: to J08 ``faint''
785: \begin{equation}
786: M_r = 4.0 + 11.86(r-i) - 10.74(r-i)^2 + 5.99(r-i)^3 - 1.20(r-i)^4
787: \label{Mr_faint}
788: \end{equation}
789: and ``bright''
790: \begin{equation}
791: M_r = 3.2 + 13.30(r-i) - 11.50(r-i)^2 + 5.40(r-i)^3 - 0.70(r-i)^4
792: \label{Mr_bright}
793: \end{equation}
794: photometric parallax relations. The rms difference between
795: Equations~\ref{Mr_geo} and~\ref{Mr_kin}, and Equation~\ref{Mr_bright} is
796: $\sim0.13$ mag, comparable to the rms difference between our
797: Equations~\ref{Mr_geo} and~\ref{Mr_kin} ($\sim0.13$ mag). The maximum difference
798: between Equations~\ref{Mr_geo} and~\ref{Mr_kin}, and Equation~\ref{Mr_bright} is
799: $\sim0.25$ mag, again comparable to the maximum difference between our
800: Equations~\ref{Mr_geo} and~\ref{Mr_kin} ($\sim0.25$ mag). The different color
801: distributions of the two samples, shown in Figure~\ref{geo_kin_gi1_gi2_comp},
802: together with metallicity effects, is the most likely explanation for
803: differences between the two photometric parallax relations.
804: 
805: \subsection{The Analysis of the Scatter in Predicted Absolute Magnitudes}
806: \label{scatter}
807: 
808: The scatter in $\delta$ values can be expressed as
809: \begin{equation}
810: \langle \delta^2 \rangle = \langle (\Delta M - \Delta m)^2 \rangle \approx \langle \Delta M^2 \rangle + \langle \Delta r^2 \rangle \label{delta_scatter},
811: \end{equation}
812: where $\langle \Delta M^2 \rangle$ is the scatter in predicted absolute
813: magnitudes, and $\langle \Delta m^2 \rangle$ is the scatter in measured apparent
814: magnitudes. Since the photometric uncertainties of SDSS are well understood, the
815: intrinsic scatter around the $M_r(r-i)$ relation is possible to measure and
816: characterize.
817: 
818: In Figure~\ref{delta_hists} we plot the observed distributions of $\delta$
819: values for the geometrically- and kinematically-selected binaries, and overplot
820: the $\delta$ distribution of the random sample. The $\delta$ values for the
821: random sample were calculated with Equations~\ref{Mr_geo} and~\ref{Mr_kin},
822: respectively. The $\delta$ distribution of the random sample was fitted to the
823: observed $\delta$ distribution in the $|\delta|>1$ range using the
824: Kolmogorov-Smirnov test.
825: 
826: By comparing the random and the observed $\delta$ distributions, we find that
827: the two match well for $|\delta|>1$ (the Kolmogorov-Smirnov test reports
828: $P\sim0.95$), indicating that candidate binaries with $|\delta|>1$ are almost
829: certainly random pairs. On the other hand, as $\delta$ approaches zero, the two
830: distributions become remarkably different ($P\sim10^{-7}$ for $|\delta|<1$),
831: indicating that these candidate binaries are dominated by true binary systems,
832: and not by random pairs.
833: 
834: The $\delta$ distribution for true binaries (Figure~\ref{delta_hists}, dashed
835: line), obtained by subtracting the random from the observed $\delta$
836: distribution, is clearly not Gaussian. It can be modeled as a sum of two
837: Gaussian distributions (``narrow'' and ``wide'') centered close to zero, and
838: about 0.1 mag and 0.55 mag wide. The centers, widths, and areas for the best-fit
839: Gaussian distributions are given in Table~\ref{gauss}.
840: 
841: To determine the consistency of the observed scatter with photometric errors, we
842: normalize the $\delta$ values for the kinematically-selected sample with
843: expected formal errors,
844: \begin{equation}
845: \sigma_{\delta} = (\sigma_{M_{r2}}^2 + \sigma_{M_{r1}}^2 + \sigma_{r_2}^2 + \sigma_{r_1}^2)^{1/2},
846: \end{equation}
847: and plot the $\delta/\sigma_{\delta}$ distribution in Figure~\ref{delta_norm}.
848: The $\delta/\sigma_{\delta}$ distribution for true binaries is not a Gaussian
849: with a width of 1, as we would expect if the scatter in the $\delta$
850: distribution was only due to photometric errors in the $gri$ bands (note that
851: the expected random error in $M_r$ is about 5-10 times larger than the random
852: error of the $g-i$ color because $dM_r/d(g-i)$ varies from $\sim$10 at the blue
853: edge to $\sim$5 at the red edge).
854: 
855: The width of $\delta/\sigma_{\delta}$ distribution for the
856: geometrically-selected candidate binaries is about 3 times smaller than in the
857: kinematically-selected sample. We find that this is due to {\em overestimated}
858: photometric errors in the geometrically-selected sample, as shown in
859: Figure~\ref{photo_errors}. The overestimated photometric errors in the $gri$
860: bands overestimate the expected formal error $\sigma_{\delta}$, and the overall
861: $\delta/\sigma_{\delta}$ distribution is too narrow. We speculate that the small
862: angular separation ($\sim3\arcsec$) between the components is the cause of
863: overestimated photometric errors (perhaps due to sky background estimates). The
864: small angular separation of components does not affect the magnitudes of stars
865: in the geometrically-selected sample. If it did, the two $\delta$ distributions
866: would be significantly different which, as shown in Figure~\ref{delta_hists}, is
867: not the case.
868: 
869: The observed non-Gaussian scatter in predicted absolute magnitudes may be due to
870: photometric parallax variation as a function of metallicity. As noted at the
871: beginning of Section~\ref{method}, we can only estimate the ``mean'' shape of
872: the photometric parallax relation. Since the intrinsic photometric parallax for
873: a given wide binary system is different from the mean relation, $\Delta M$ (the
874: difference of predicted absolute magnitudes) and $\Delta m$ (the measured
875: difference of apparent magnitudes) will differ. This discrepancy will increase
876: for systems where the components have significantly different colors.
877: 
878: To test the assumption that the shape of photometric parallax relation increases
879: the scatter in predicted absolute magnitudes, we use the mock sample constructed
880: in Section~\ref{algorithm} and add a color-dependent offset to apparent
881: magnitudes
882: \begin{eqnarray}
883: r'_1 = r_1 + \xi (g-i)_1 \\
884: r'_2 = r_2 + \xi (g-i)_2
885: \end{eqnarray}
886: where $\xi$ is a random number between zero and one (the same for both
887: components). These color-dependent offsets simulate the change in the shape of
888: the photometric parallax relation due to metallicity. We apply the algorithm
889: described in Section~\ref{implementation} to the mock sample, and obtain a
890: revised photometric parallax relation. Using this relation, we analyze the
891: distribution of $\delta$ values and find that it can be modeled as a sum of two
892: Gaussians centered on zero, with widths of 0.1 and 0.3 mag. This result suggests
893: that the non-Gaussian scatter observed in candidate samples may be caused by
894: the difference between the shapes of the mean photometric parallax relation
895: and a true relation for a given metallicity (and perhaps other effects, such as
896: age).
897: 
898: This model-based conclusion is consistent with a direct comparison of relations
899: derived here and the relations from I08a evaluated for the median halo
900: metallicity ($\left[Fe/H\right]=-1.5$) and the median disk metallicity
901: ($\left[Fe/H\right]=-0.7$ for distances probed by our sample; see Figure~5 in
902: the above paper). The two relations corresponding to halo and disk stars are
903: offset by 0.6 mag due to metallicity difference. Our relations match the
904: low-metallicity relation at the blue end and the high-metallicity relation at
905: the red end. Therefore, in the worst case scenario of extremely blue ($r-i=0.3$)
906: and red ($r-i=1.4$) disk stars, the maximum error in the difference of their
907: absolute magnitudes is 0.6 mag. When convolved with the observed color
908: distribution of pairs, the expected scatter is about 0.2-0.3 mag, consistent
909: with the observed and simulated widths of the $\delta$ distributions.
910: 
911: Unresolved binarity of components in candidate samples may also contribute to
912: the non-Gaussian scatter in predicted absolute magnitudes. The multiplicity
913: studies of G dwarfs \citep{dm91} and M dwarfs \citep{fm92} find that a
914: significant fraction of G and M dwarf stars ($40-60\%$) are unresolved binary
915: systems. If a component of a wide binary system is an unresolved binary system,
916: its luminosity will be underestimated (with the magnitude of the offset
917: depending on the actual composition of the binary) and the $\delta$ value for
918: the wide binary system will systematically deviate from zero. In
919: Appendix~\ref{model} we model the presence of unresolved binaries in wide binary
920: systems, and find that the model can explain the observed $\delta$ scatter.
921: 
922: Therefore, both the intrinsic variations of the photometric parallax relation
923: and unresolved binaries can explain the observed non-Gaussian scatter of
924: $\delta$. The data discussed here are insufficient to disentangle these two
925: effects.
926: 
927: Finally, the uncertainty in predicted absolute magnitudes (error distribution 
928: for photometric parallax method) can be obtained by drawing random values, $x$,
929: from a non-Gaussian distribution
930: \begin{equation}
931: f(x) = A_1 \, N(x|\mu_1, \sigma_1/\sqrt{2}) + A_2 \, N(x|\mu_2, \sigma_2/\sqrt{2}), 
932: \label{Mr_scatter}
933: \end{equation}
934: where $N(x|\mu,\sigma)$ are Gaussian distributions, and the best-fit parameters
935: are listed in Table~\ref{gauss}.
936: 
937: \section{The Properties of Wide Binaries}
938: \label{properties}
939: 
940: The best-fit photometric parallax relation can be utilized to further refine the
941: samples of candidate binaries and to address questions about their dynamical and
942: physical properties such as
943: \begin{itemize}
944: \item Do wide binaries have the same spatial distribution as single stars?
945: \item Do wide binaries have the same color distribution as single stars?
946: \item Are the color distributions of components in wide binary systems 
947:       consistent with random pairings? 
948: \item What is the distribution of semi-major axis for wide binaries?
949: \item Does the distribution of semi-major axis vary with the position in
950:       the Galaxy?
951: \end{itemize}
952: 
953: \subsection{High-Efficiency Samples of Candidate Binaries}
954: \label{samples}
955: 
956: We use the best-fit photometric parallax relations to select samples of
957: candidate binaries with high selection efficiency (high fraction of true
958: binaries) by imposing further constraints on $\delta$ values in geometric and
959: kinematic sample.
960: 
961: As shown in Figure~\ref{delta_hists}, the fraction of random pairs in the
962: candidate sample is simply $A_{random}/A_{observed}$, where $A_{random}$ and
963: $A_{observed}$ are the integrals of the random (triangles) and total (thick
964: solid line) $\delta$ distributions. The fraction of true binaries, or the
965: {\em selection efficiency}, is then
966: \begin{equation}
967: \epsilon = 1 - A_{random}/A_{observed}\label{eff}
968: \end{equation}
969: Without a cut on $\delta$, the fraction of true binaries (the selection
970: efficiency) in the geometrically- and kinematically-selected samples is $34\%$
971: and $35\%$, respectively. It is reassuring to find that the $\epsilon$ value for
972: the geometrically-selected sample obtained here, and the one measured in
973: Section~\ref{geo_select} match so well (at a $1\%$ level), even though the two
974: methods for estimating $\epsilon$ are independent.
975: 
976: The selection efficiency of $35\%$ for the kinematically-selected sample may
977: seem low, given that only $0.5\%$ of random pairs pass the common proper motion
978: criteria. This points to a low fraction of true binaries with angular separation
979: greater than $15\arcsec$. If this fraction is about 1/400 ($0.25\%$), the common
980: proper motion criteria will select 2 random pairs (0.5\% out of a 400), and only
981: 1 true binary system. Therefore, $66\%$ of the sample (2 out of 3) will be
982: random pairs, and $34\%$ (1 out of 3) will be true binary systems, similar to
983: what we find for the kinematically-selected sample. The result that only 1/400
984: pairs with $\theta>15\arcsec$ are true binaries puts the fraction of random
985: pairs in the random sample at $99.75\%$.
986: 
987: Figure~\ref{delta_hists} shows that the true binaries have a much smaller
988: range of $\delta$ values than the random pairs. Therefore, a cut on $\delta$
989: would reduce the contamination, and increase the fraction of true binaries in a
990: sample. By requiring $|\delta|<0.4$, we construct samples where $63\%$ and
991: $64\%$ of candidates are true binaries. The numbers of candidate binaries in
992: these cleaner samples are 16,575 (geometrically-selected) and 5,157 candidates
993: (kinematically-selected), with the expected total number of true binaries about
994: 13,743. The sample efficiency for the geometric sample can be further increased
995: to 90\% by requiring $|\delta|<0.2$ and $Z < 0.3$ kpc, where $Z$ is the height
996: above the Galactic plane. Compared to the existing catalogs of wide binaries by
997: \citet{cg04} and \citet{lb07}, our samples represent a 20-fold increase in the
998: number of candidate binaries and probe much deeper into the halo (to $\sim4$
999: kpc). Although a non-negligible fraction of candidate pairs are due to random
1000: pairings ($\sim35\%$), the increase in the number of potential physical pairs is
1001: substantial. 
1002: 
1003: We emphasize that our method only selects candidates where both components are
1004: {\em main-sequence} stars, while rejecting systems where one of the components
1005: has evolved off the main sequence. This is due to the photometric parallax
1006: relation, as defined here, being correct for main-sequence stars only. Together
1007: with the small expected fraction of giant stars in our sample due to faint
1008: apparent magnitudes (1-2\%, \citealt{fin00}; I08a), this bias results in
1009: practically pure main-sequence sample. We note that the application of a
1010: photometric parallax relation that corresponds to some mean metallicity
1011: distribution introduces systematic errors in estimated $M_r$. We partially
1012: mitigate this problem by averaging distances determined for each binary
1013: component (using Equation~\ref{Mr_geo}). Based on the behavior of photometric
1014: parallax relations and $\delta$ distribution discussed in Section~\ref{scatter},
1015: the systematic uncertainty in obtained distances is most likely not larger than
1016: 10-15\% (an understimate due to faint bias for blue stars). Another source of
1017: overall systematic uncertainty in distances is the normalization of
1018: Equation~\ref{Mr_geo} adopted from \citet{wwh05}. This normalization corresponds
1019: to nearby ($<$100 pc) metal-rich stars, while most stars in our sample are
1020: distances of the order 1 kpc. The disk metallicity gradient discussed by I08a
1021: implies systematic distance overestimate of about 10-20\%, partially cancelling
1022: the above underestimate. These systematic uncertainties propagate as systematic
1023: uncertainties of derived semi-major axes discussed in Section~\ref{spatial}. 
1024: 
1025: \subsection{The Color Distribution of Wide Binaries}
1026: \label{color}
1027: 
1028: The luminosity of a main-sequence star, and thus its color via photometric
1029: parallax relation, can be used as a proxy for stellar mass. The color-color
1030: distribution of wide binaries, therefore, provides constraints on the
1031: distribution of stellar masses in wide binary systems. To find the color
1032: distribution of wide binaries, we select a volume-complete ($0.7 < d/kpc < 1.0$)
1033: subsample of geometrically-selected candidate binaries with $|\delta|<0.4$, and
1034: plot their distribution in the $(g-i)_2$ vs.~$(g-i)_1$ color-color diagram in
1035: Figure~\ref{counts_0.7_1.0} ({\em top}). The sample is complete in
1036: the $0.4 < g-i < 2.8$ color and 4,200 AU $<a<$ 10,000 AU semi-major axis range
1037: (see Section~\ref{spatial}). Even though the $|\delta|<0.4$ cut increases the
1038: fraction of true binaries, about 14\% of candidates (in the $0.7 < d/kpc < 1.0$
1039: range) are still random pairs that contaminate the map. To remove the
1040: contamination, first we select pairs from the random sample (see the end of
1041: Section~\ref{initial_select}) with $|\delta|<0.4$ and $0.7<d/kpc<1.0$. The
1042: $|\delta|<0.4$ cut on the random sample will not increase the fraction of true
1043: binaries ($\epsilon$) above $\sim1\%$ because $\epsilon$ decreases rapidly with
1044: $\theta$ (see Figure~\ref{logZ_plots} ({\em middle left}) in
1045: Section~\ref{spatial}), and the pairs in the random sample have
1046: $\theta>20\arcsec$. The $(g-i)_2$ vs.~$(g-i)_1$ distribution of this random
1047: sample is shown in Figure~\ref{counts_0.7_1.0} ({\em middle}). The maps are
1048: essentially probability density maps as pixels sum to 1. To correct for the
1049: contamination in the top map, we multiply each pixel in the random map with 0.14
1050: (that being the contamination in the candidate binary sample), and subtract two
1051: maps. The corrected map, presented in Figure~\ref{counts_0.7_1.0}
1052: ({\em bottom}), shows that the color-color distribution of true binary systems
1053: is fairly uniform, has a local maximum around $(g-i)_{1,2}\sim2.5$, and reflects
1054: the underlying luminosity function which peaks for red stars
1055: (c.f.~Figure~\ref{Fig:DMvsColor}). 
1056: 
1057: The map shown in Figure~\ref{counts_0.7_1.0} ({\em bottom}) describes the
1058: probability density, $P[(g-i)_1,(g-i)_2]$, of a wide binary system with
1059: components that have $(g-i)_1$ and $(g-i)_2$ colors falling into a given pixel.
1060: This probability density can be expressed as a product
1061: \begin{equation}
1062: P[(g-i)_1,(g-i)_2] = P[(g-i)_B|(g-i)_A] \, P[(g-i)_A]
1063: \label{eq_wide_binary_prob}
1064: \end{equation}
1065: where $P[(g-i)_B|(g-i)_A]$ is the conditional probability density of having one
1066: component with $(g-i)_B$ color in a wide binary system where the other component
1067: has $(g-i)_A$, and $P[(g-i)_A]$ is the probability density that a star with
1068: $g-i=(g-i)_A$ color is in a wide binary system. These probability densities may
1069: also vary with Galactic coordinates (e.g., with the height above the Galactic
1070: plane), but we cannot study such effects directly because the samples are
1071: volume-complete only in the $0.7 < d/kpc < 1.0$ range.
1072: 
1073: The conditional probability density, $P[(g-i)_B|(g-i)_A]$, can be extracted from
1074: Figure~\ref{counts_0.7_1.0} ({\em bottom}) map by selecting pixels where either
1075: $(g-i)_1=(g-i)_A$, or $(g-i)_2=(g-i)_A$. The resulting $P[(g-i)_B|(g-i)_A]$ for
1076: several values of $(g-i)_A$ are shown in Figure~\ref{wide_binary_prob}. Red
1077: stars ($(g-i)_A \ga 2.0$) are more likely to be associated with another red star
1078: than with a blue star, while for blue stars the companion color distribution is
1079: flat. The best-fit analytic functions that describe the observed trends are
1080: given in Table~\ref{tbl_wide_bin_prob}.
1081: 
1082: The probability density, $P[(g-i)_A]$, that a star with $g-i=(g-i)_A$ is in a
1083: wide binary system can be derived by comparing the $g-i$ color distribution of
1084: stars in wide binary systems with the $g-i$ color distribution of all the stars
1085: in the same volume. As shown in Figure~\ref{prob_gi} ({\em top}), the $g-i$
1086: color distribution of stars in the volume-complete wide binary sample roughly
1087: follows the $g-i$ color distribution of all the stars in the same volume. The
1088: ratio of the two distributions (renormalized to an area of 1) gives the
1089: $P[(g-i)_A]$, and is shown in the bottom panel.
1090: 
1091: The probability for a star to be in a wide binary system ($P[(g-i)_A]$) is
1092: independent of its color. Given this color, the companions of red components
1093: seem to be drawn randomly from the stellar luminosity function, while blue
1094: components have a larger blue-to-red companion ratio than expected from
1095: luminosity function. These results are consistent with recent results by
1096: \citet{lb07}. The overall fraction of stars in wide binary systems is discussed
1097: in the next section.
1098: 
1099: \subsection{The Spatial Distribution of Wide Binaries}
1100: \label{spatial}
1101: 
1102: If the semi-major axis distribution function, $f(a)$, is known, the number of
1103: stars in wide binary systems can be determined by integrating $f(a)$ from some
1104: lower cutoff, $a_1$, to the maximum semi-major axis, $a_2$. The power-law
1105: frequency distribution, $f(a)\propto a^\beta$, $\beta=-1$, is known in the
1106: context of wide binaries as the \"Opik distribution (OD; \citealt{oep24}). When
1107: semi-major axis distribution of wide binaries follows the OD, the frequency
1108: distribution of $\log(a)$ is a straight line with a slope of zero \citep{pah07}.
1109: Alternatively, an equivalent representation of OD is the cumulative distribution
1110: $N[<\log(a)]\propto \log(a)$. In this form, OD is a straight line with a
1111: positive slope. We use the cumulative representation, instead of differential,
1112: because it reduces the counting noise in sparsely populated bins (though the
1113: errors become correlated between bins). 
1114: 
1115: We utilize geometrically-selected candidate binaries (see
1116: Section~\ref{limitations} for a discussion of the kinematic sample), but do not
1117: limit the selection to $\theta < 4\arcsec$, as we did in
1118: Section~\ref{geo_select}. Since $a \propto \theta$, the removal of upper limit
1119: on $\theta$ allows us to probe an extended range of semi-major axes. The
1120: downside is that random pairs dominate at large $\theta$ and a careful
1121: accounting for contamination as a function of $\theta$ is required before the
1122: the $f(a)$ distribution can be constrained. Since we only know the projected
1123: separation of our pairs, we use a statistical relation to calculate the average
1124: semi-major axis, $\langle a\rangle$, as $\langle a\rangle=1.411\, \theta\, d$,
1125: where $d$ is the heliocentric distance \citep{cou60}. Hereafter, we drop the
1126: brackets and simply note the average semi-major axis as $a$.
1127: 
1128: Figure~\ref{loga_cum} ({\em top left}) shows the cumulative distribution of
1129: $\log(a)$ for candidate wide binaries with $|\delta|<0.2$ selected from the
1130: $0.7< Z/kpc < 1.0$ range. The cumulative distribution does not follow a straight
1131: line, as predicted by the OD, but actually increases its slope with $\log(a)$.
1132: We assume that this is due to an increasing fraction of random pairs at high
1133: $\log(a)$, and proceed to verify this assumption.
1134: 
1135: Figure~\ref{loga_cum} ({\em top right}) shows the differential distribution of
1136: angular separation for the selected sample. For $\theta>\theta_{max}$ the
1137: observed and random distributions closely match, demonstrating that random pairs
1138: dominate at high $\theta$ (or high $\log(a)$). To calculate how the fraction of
1139: true binaries (or random pairs) changes as a function of $\theta$, we fit
1140: $f_{rnd}(\theta)= C \, \theta$ to the observed histogram, and calculate the
1141: fraction or true binaries, $\epsilon$, using Equation~\ref{eff2} (see
1142: Section~\ref{geo_select}). The calculated $\epsilon$ values, as well as the
1143: best-fit second-degree polynomial, $\epsilon(\theta)$, are shown in
1144: Figure~\ref{loga_cum} ({\em middle left}). As an independent test, the selection
1145: efficiency was calculated using Equation~\ref{eff} (i.e., from the $\delta$
1146: distribution) for three $\theta$-selected subsamples, and the obtained values
1147: agree with $\epsilon(\theta)$ at a level of 1\%. The angular separation for
1148: which $\epsilon$ falls below $\sim5\%$ is defined as $\theta_{max}$. The
1149: fraction of true binaries ($\epsilon(\theta)$) also changes as a function of
1150: $Z$, and is determined separately for different distance bins.
1151: 
1152: Since the candidates are restricted in $Z$ ($Z_{min}=0.7$ kpc to $Z_{max}=1.0$
1153: kpc in this example) and $\theta$ ($3\arcsec$ to $\theta_{max}$), to ensure a
1154: uniform selection in the $Z$ vs.~$a$ space we define
1155: \begin{equation}
1156: a_{min} = 3\arcsec \cdot 1.411 \cdot 1000 \, Z_{max} \label{a_min}
1157: \end{equation}
1158: and
1159: \begin{equation}
1160: a_{max} = \theta_{max} \cdot 1.411 \cdot 1000 \, Z_{min}/\sin(45\arcdeg)\label{a_max}
1161: \end{equation}
1162: as the minimum and maximum probed semi-major axis, shown as the selection box in
1163: Figure~\ref{loga_cum} ({\em middle right}). The $\sin(45\arcdeg)$ factor is to
1164: account for the fact that the candidates are restricted to high ($b>45\arcdeg$)
1165: Galactic latitudes.
1166: 
1167: To correct the cumulative distribution of $\log(a)$, we assign a probability
1168: $\epsilon(\theta)$ to each candidate binary in the $a_{min}$ to $a_{max}$ range,
1169: and add the probabilities (instead of counting candidates) when making the
1170: cumulative $\log(a)$ distribution. The corrected cumulative distribution, shown
1171: in Figure~\ref{loga_cum} ({\em bottom left}), follows a straight line up to the
1172: turnover point, $a_{break}$. We define $a_{break}$ as the average semi-major
1173: axis for which the straight line fit to the cumulative distribution deviates by
1174: more than $1.5\%$. In addition to $a_{break}$, we also measure the slope of the
1175: cumulative distribution where it follows the straight line. It can be shown that
1176: the slope of the cumulative distribution is equal to the constant of
1177: proportionality, $N_0$, in \"Opik distribution, $f(a) = N_0/a$. The number of
1178: binaries can be calculated by integrating $f(a)$ from $a_1$ to $a_2$, and we
1179: obtain $N_{bin}=N_0 \, \log(a_2/a_1)$. For integration limits we choose
1180: $a_2=a_{break}$ where we assume that systems with semi-major axes greater than
1181: $a_{break}$ are no longer bound, and $a_1 = 100$ AU (since $a_2 \gg a_1$, the
1182: results are not very sensitive to the choice of $a_1$).
1183: 
1184: The uncertainty in $a_{break}$, shape of $f(a)$ (or power-law index $\beta$),
1185: and number of binaries ($N_{bin}$) are estimated using Monte Carlo simulations.
1186: We find that the uncertainty in $a_{break}$ is less than 0.1 dex, and the error
1187: on the power-law index ($\beta$) is $\la0.1$. The uncertainty in measuring
1188: $N_{bin}$ is about $10\%$. The corrected cumulative $\log(a)$ distribution
1189: obtained from one of these simulations is shown in Figure~\ref{loga_cum}
1190: ({\em bottom right}). The semi-major axis distribution of ``true'' binaries in
1191: the simulation sample is $f(a)\propto a^{-0.8}$, and is valid between 100 AU and
1192: $a_{break}=10,000$ AU. The turnover in the distribution happens because there
1193: are no ``true'' binaries above 10,000 AU, only random pairs, similar to what we
1194: observe in real data. This similarity is a strong warning not to over-interpret
1195: the slope of $f(a)$ beyond $a_{break}$.
1196: 
1197: To estimate the dependence of $\beta$ (shape of $f(a)$), $a_{break}$, and $N_0$
1198: on color, we divide the $0.7 < Z/kpc < 1.0$ sample into three color subsamples
1199: using $(g-i)_1=1.8$ and $(g-i)_2=1.5$ lines. We find that $f(a)$ follows OD in
1200: all three subsamples ($\beta = -1$), and that the average $a_{break}$ is 3.99,
1201: with a 0.07 root-mean-square scatter. The $a_{break}$ for the full
1202: $0.7<Z/kpc<1.0$ sample is 4.02. These results suggest that $a_{break}$ and the
1203: shape of $f(a)$ are {\em independent of color of binaries}. The $N_0$ value, and
1204: subsequently the number of binaries, will depend on the sample's color range.
1205: For the full $0.7<Z/kpc<1.0$ sample, the number of binaries is
1206: \begin{equation}
1207: N_{bin} = (N_0^1 + N_0^2 + N_0^3) \log_{10}(a_2/a_1),
1208: \end{equation}
1209: where $N_0^i$, $i=1,2,3$, are $N_0$ values measured for each color subsample.
1210: Therefore, the number of binaries calculated for a distance bin will change as
1211: the color range changes. Assuming that the $g-i$ color distribution of binaries
1212: does not change with $Z$, we can use the $g-i$ color distribution for the
1213: $0.7<Z/kpc<1.0$ sample (solid line in Figure~\ref{prob_gi} ({\em top})), to
1214: correct for color incompleteness. We also assume that the fraction of binaries
1215: outside the $0.4 < g-i < 2.8$ color range is small. The correct number of
1216: binaries is then
1217: \begin{equation}
1218: N_{bin} = N_0/A[(g-i)_{min},(g-i)_{max}] \,\log_{10}(a_2/a_1),
1219: \end{equation}
1220: where $A[(g-i)_{min},(g-i)_{max}]$ is the area underneath the solid line 
1221: histogram in Figure~\ref{prob_gi} ({\em top}), between $(g-i)_{min}$ and
1222: $(g-i)_{max}$ ($g-i$ color range for a given distance bin).
1223: 
1224: The estimated systematic error in $a_{break}$ due to the choice of the
1225: $|\delta|$ cut is measured using $|\delta|<0.1$ and $|\delta|<0.4$ samples. We
1226: find that $a_{break}$ changes by $\la0.03$ dex between these samples. This
1227: result suggests that $a_{break}$ is not sensitive to the choice of the
1228: $|\delta|$ cut. Similarly, the change in $a_{break}$ is less than 0.03 dex if
1229: the estimate of $\epsilon(\theta)$ is off by $\pm0.1$ ($\sim10\%$ change) from
1230: the best-fit $\epsilon(\theta)$.
1231: 
1232: To establish whether semi-major axis distribution follows the OD in other $Z$
1233: bins, we repeat the $f(a)$ and $a_{break}$ measuring procedure on 8 $Z$ bins,
1234: and show the corrected cumulative distributions with best-fit straight lines in
1235: Figure~\ref{loga_cum_fits}. In general, the corrected cumulative distributions
1236: follow a straight line, and then start to deviate from it at $a_{break}$. In the
1237: $0.1<Z/kpc<0.4$ bin we do not see a turnover due to a narrow range of probed
1238: projected separations ($\theta_{max}=16\arcsec$ limits the range to 3193 AU, see
1239: Figure~\ref{theta_eff_0.1_0.4_geo}), and only determine the upper limit on
1240: $a_{break}$.
1241: 
1242: As the average height above the Galactic plane increases, the $a_{break}$ moves 
1243: to higher values. We investigate this correlation in more detail in
1244: Figure~\ref{logZ_plots} ({\em top left}). The data follow a straight line
1245: $\log(a_{break})=k \, \log(Z[pc])+l$, where $k=0.72\pm0.05$ and
1246: $l=1.93\pm0.15$, or approximately, $a_{break}[AU] = 12,300 \, Z[kpc]^{0.7}$ in
1247: the $0.3 < Z/kpc < 3.0$ range.
1248: 
1249: It is possible that $a_{break}$ also depends on the cylindrical radius, $R$,
1250: with the Sun at $R_\odot$=8 kpc, and perhaps on the local density of stars,
1251: $\rho$. Because the sample is dominated by stars at high Galactic latitudes, it
1252: is hard to disentangle the $Z$ dependence from the other two effects (the $R$
1253: range is small, and $\rho$ varies strongly with $Z$). We attempt to do so using
1254: the volume-complete $0.7<Z/kpc<1.0$ sample. First we divide this sample into
1255: three subsamples with median Galactic latitudes, $\langle b \rangle$, of
1256: $35\arcdeg$, $49\arcdeg$, and $80\arcdeg$ and determine $a_{break}$ for each
1257: subsample. The best-fit $a_{break}$ varies by $\sim$0.3 dex between the
1258: low-latitude and high-latitude subsample, despite the same median $Z$. When the
1259: $0.7<Z/kpc<1.0$ sample is divided into the Galactic anticenter
1260: ($90\arcdeg < l < 270\arcdeg$) and the Galactic center ($l > 270\arcdeg$ or
1261: $l<90\arcdeg$) subsamples, the best-fit $a_{break}$ varies by $\sim$0.1 dex.
1262: These variations suggest that the best-fit $Z$ dependence does not fully
1263: capture the behavior of $a_{break}$. Nevertheless, they are smaller ($\la0.3$
1264: dex) than the observed variation of $a_{break}$ ($\sim$1 dex). 
1265: 
1266: The spatial distribution of wide binaries can now be compared to the number
1267: density of all stars as a function of height above the Galactic plane. In
1268: Figure~\ref{logZ_plots} ({\em bottom left}) we show that wide binaries closely
1269: follow the spatial distribution of stars, with exponential decline in the number
1270: density as a function of $Z$. The fraction of binaries relative to the number of
1271: all stars, shown in the bottom right panel, changes by only a factor of 2 over a
1272: range of 3 kpc, starting from $0.9\%$ at $Z=500$ pc and declining to $0.5\%$ at
1273: $Z=3000$ pc.
1274: 
1275: \subsection{The Limitations of the Kinematic Sample}
1276: \label{limitations}
1277: 
1278: It would be informative to repeat the $f(a)$ and $a_{break}$ analysis using
1279: kinematically-selected binaries, but unfortunately, the apparent incompleteness
1280: of SDSS-POSS proper motion data at $\theta<9\arcsec$ prevents us in doing so. As
1281: shown in Figure~\ref{theta_fit_500pc}, the number of common proper motion pairs
1282: drops sharply below $\theta=9\arcsec$, probably due to blending of close sources
1283: in the POSS data. Because of this $\theta$ cut-off, for the same range in $Z$,
1284: the effective $a_{min}$ for the kinematic sample is three times that of the
1285: geometric sample (where the lower limit on $\theta$ is $3\arcsec$). In the case
1286: of $0.1<Z/kpc<0.4$ sample observed here, the smallest probed semi-major axis
1287: ($a_{min}$) is at 5079 AU, well above the $a_{break}$ value of 4534 AU predicted
1288: by the $a_{break} \propto Z^{0.7}$ relation. Since we are outside the range
1289: where OD is valid, we cannot measure where the turnover in $f(a)$ happens, and
1290: cannot determine $a_{break}$ or $N_{bin}$. In all the other $Z$ bins, $a_{min}$
1291: is also above the predicted $a_{break}$ value, and therefore outside the \"Opik
1292: regime.
1293: 
1294: \section{Discussion and Conclusions}
1295: \label{discussion}
1296: 
1297: We have presented a novel approach to photometric parallax estimation based on
1298: samples of candidate wide binaries selected from the Sloan Digital Sky Survey
1299: (SDSS) imaging catalog. Our approach uses the fact that binary system's
1300: components are at equal distances and estimates the photometric parallax
1301: relation for main-sequence stars by minimizing the difference of their distance 
1302: moduli. While this method is similar to constraints on photometric parallax
1303: relation obtained from globular clusters in that it does not require absolute
1304: distance estimates, it has the advantage that it extends to redder colors than
1305: available for globular clusters observed by the SDSS, and it implicitly accounts
1306: for the metallicity effects.
1307: 
1308: The derived best-fit photometric parallax relations represent
1309: metallicity-averaged relations and thus provide an independent confirmation of
1310: relations proposed by J08 in their study of the Galactic structure. An important
1311: result of this work is our estimate of the expected error distribution for
1312: absolute magnitudes determined from photometric parallax relations (a
1313: root-mean-square scatter of $\sim$0.3 mag, see Section~\ref{scatter}), which is
1314: in good agreement with modeling assumptions adopted by J08. The mildly
1315: non-Gaussian error distribution is consistent with both the impact of unresolved
1316: binary stars, and the variation of photometric parallax relation with
1317: metallicity; we are unable to disentangle these two effects. 
1318: 
1319: The best-fit photometric parallax relations enabled the selection of
1320: high-efficiency samples of disk wide binaries with $\sim22,000$ candidates, that
1321: include about 14,000 true binary systems (efficiency of $\sim2/3$). Using the
1322: photometric measurements and angular distance of the two components, samples
1323: with efficiency exceeding 80\% can be constructed (see Section~\ref{samples}).
1324: Such samples could be used as a starting point to further increase the selection
1325: efficiency with the aid of radial velocity measurements. Spectral observations
1326: of systems where the brighter component is an F/G star, for which it is easy to
1327: estimate metallicity, could be used to calibrate both spectroscopic and
1328: photometric methods for estimating metallicity of cooler K and M dwarfs.
1329: Compared to the state-of-the-art catalogs of wide binaries by \citet{cg04} and 
1330: \citet{lb07}, the samples discussed here represent a significant increase in the
1331: number of potential binaries, and probe larger distances (to $\sim4$ kpc). To
1332: facilitate further studies of wide binaries, we make the catalog publicly
1333: available\footnote{The catalog can be downloaded from
1334: \url[HREF]{http://www.astro.washington.edu/bsesar/SDSS\_wide\_binaries.tar.gz}}.
1335: 
1336: Using the high-efficiency subsamples, we analyzed their dynamical and physical 
1337: properties. We find that the spatial distribution of wide binaries follows the
1338: distribution of single stars to within a factor of 2, and that the probability
1339: for a star to be in a wide binary system is independent of its color. However, 
1340: given this color, the companions of red components seem to be drawn randomly
1341: from the stellar luminosity function, while blue components have a larger
1342: blue-to-red companion ratio than expected from luminosity function (see
1343: Section~\ref{color}). These results are consistent with recent results by
1344: \citet{lb07}, and provide strong constraints for the scenarios describing the
1345: formation of such systems (e.g., \citealt{gie06} and references therein;
1346: \citealt{cla07}; \citealt{hur07}).
1347: 
1348: We also study the semi-major axis distribution of wide binaries in the
1349: $2,000-47,000$ AU range (see Section~\ref{spatial}). The observed distribution
1350: is well described by the \"Opik distribution, $f(a)\propto 1/a$, for
1351: $a<a_{break}$, where $a_{break}$ increases roughly linearly with the height
1352: above the Galactic plane ($a_{break}\sim 12,300$ AU at $Z=1$ kpc).
1353: Alternatively, the $a_{break}$ correlates with the local number density of stars
1354: as $a_{break} \propto \rho^{-1/4}$, but we are unable to robustly identify the
1355: dominant correlation ($Z$ and $\rho$ are highly correlated). 
1356: 
1357: The distribution of semi-major axes for wide binaries was also discussed by
1358: \citet{cg04}. They used a sample of wide binaries selected using common proper 
1359: motion from the rNLTT catalog \citep{gs03}, and found $f(a)\propto 1/a^{1.6}$,
1360: with no evidence of a turnover at $a \la 3000$. Their sample extended to larger
1361: angular separations than ours, and probed smaller distances. On the other hand, 
1362: \citet{pah07} used wide binaries from the same \citet{cg04} sample, and detected
1363: \"Opik distribution, $f(a)\propto 1/a$, for $a<3,000$, consistent with the
1364: result of \citet{cg04}. In a recent study,  L\'epine \& Bongiorno searched for 
1365: faint common proper motion companions of Hipparcos stars and detected a turnover
1366: from \"Opik distribution to a steeper distribution around $a\sim3,000$ AU. Their
1367: sample also probed much smaller distances than ours. We compare these results in
1368: Figure~\ref{Fig:compare}. As evident, the variation of $a_{break}$ with distance
1369: from the Galactic plane detected here (approximately with distance, as shown in
1370: Figure~\ref{Fig:compare}, since stars in our sample are mostly at high galactic
1371: latitudes), is consistent with the above results that are based on more local
1372: samples. In particular, this comparison of different studies suggests that the
1373: flattening of $f(a)$ for small $a$ that ``puzzled'' Chanam\'e \& Gould (see
1374: their section 4.3) is probably due to a combination of selection effects and the
1375: approach of the domain where \"Opik distribution is valid in their sample. 
1376: 
1377: The \"Opik distribution suggests that the process of star formation produces
1378: multiple stars, which evolve towards binaries after ejecting one or more single
1379: stars \citep{pah07}. The departure from the \"Opik distribution may be evidence
1380: for disruption of wide binaries over long periods of time by passing stars,
1381: giant molecular clouds, massive compact halo objects (MACHOs), or disk and
1382: Galactic tides \citep{heg75,wsw87,ycg04}. However, we note that the
1383: $a_{break} \propto \rho^{-1/4}$ correlation (see Figure~\ref{logZ_plots}) is
1384: outside the expected range discussed by \citet{ycg04}
1385: ($a_{break}\propto \rho^{-2/3}$ for close strong encounters, and
1386: $a_{break}\propto \rho^{-1}$ for weak encounters). 
1387: 
1388: The samples presented here can be further refined and enlarged. First, the SDSS
1389: covers only a quarter of the sky. Upcoming next-generation surveys, such as the
1390: SkyMapper \citep{kel07}, the Dark Energy Survey \citep{fla07}, Pan-STARRS
1391: \citep{kai02} and the Large Synoptic Survey Telescope (\citealt{ive08b}, LSST
1392: hereafter), will enable the construction of such samples over most of the sky.
1393: Due to fainter flux limits (especially for the Pan-STARRS and LSST), the samples
1394: will probe a larger distance range and will reach the halo-dominated parts of
1395: the Galaxy. Furthermore, due to improved photometry and seeing (e.g., for the
1396: LSST, by about a factor of two), the selection will be more robust. We scale the
1397: 20,000 candidates discussed here, assuming $log(N) = C + 0.4\,r$, to the LSST
1398: depth that enables accurate photometric metallicity ($r<23$; I08a) and predict a
1399: minimum sample size of $\sim$400,000 candidate wide binary systems in 20,000
1400: deg$^2$ of sky. It is likely that the sample would include more than a million
1401: systems due to the increase of the stellar counts close to the Galactic plane.
1402: 
1403: Another important development will come from the Gaia mission
1404: \citep{per01,wil05}, which will provide direct trigonometric distances for stars
1405: with $r<20$. With trigonometric distances, accurate photometric parallax
1406: relation can be used to provide strong constraints on the incidence and color
1407: distribution of unresolved multiple systems. Until then, a radial velocity 
1408: survey of candidate binaries assembled here could help with pruning the sample 
1409: from random associations, and with better characterization of various
1410: selection effects. 
1411: 
1412: \acknowledgments
1413: 
1414: This work was supported by the NSF grant AST-0707901, and the NSF grant
1415: AST-0551161 to the LSST for design and development activity. M.~J.~gratefully
1416: acknowledges support from the Taplin Fellowship and from the NSF grant
1417: PHY-0503584. We are grateful to Nick Cowan and Eric Agol (UW) for help with the
1418: Markov chain Monte Carlo code. Early motivation for this analysis came from a
1419: preliminary study by Taka Sumi.
1420: 
1421: Funding for the SDSS and SDSS-II has been provided by the Alfred P. Sloan
1422: Foundation, the Participating Institutions, the National Science Foundation, the
1423: U.S. Department of Energy, the National Aeronautics and Space Administration,
1424: the Japanese Monbukagakusho, the Max Planck Society, and the Higher Education
1425: Funding Council for England. The SDSS Web Site is http://www.sdss.org/.
1426: 
1427: The SDSS is managed by the Astrophysical Research Consortium for the
1428: Participating Institutions. The Participating Institutions are the American
1429: Museum of Natural History, Astrophysical Institute Potsdam, University of Basel,
1430: University of Cambridge, Case Western Reserve University, University of Chicago,
1431: Drexel University, Fermilab, the Institute for Advanced Study, the Japan
1432: Participation Group, Johns Hopkins University, the Joint Institute for Nuclear
1433: Astrophysics, the Kavli Institute for Particle Astrophysics and Cosmology, the
1434: Korean Scientist Group, the Chinese Academy of Sciences (LAMOST), Los Alamos
1435: National Laboratory, the Max-Planck-Institute for Astronomy (MPIA), the
1436: Max-Planck-Institute for Astrophysics (MPA), New Mexico State University, Ohio
1437: State University, University of Pittsburgh, University of Portsmouth, Princeton
1438: University, the United States Naval Observatory, and the University of
1439: Washington.
1440: 
1441: \appendix
1442: 
1443: \section{SQL Queries}
1444: \label{appendix}
1445: 
1446: The following SQL queries were used to select initial samples of candidate
1447: binaries through the SDSS CasJobs interface. When running these queries, the
1448: database context must be set to ``DR6'' or higher.
1449: 
1450: \begin{verbatim}
1451: select -- geometric selection of candidate binaries
1452: 
1453: round(p1.ra,6) as ra1, round(p1.dec,6) as dec1, round(p1.extinction_r,3) as rExt1,
1454: round(p1.psfMag_u,3) as psf_u1, round(p1.psfMag_g,3) as psf_g1,
1455: round(p1.psfMag_r,3) as psf_r1, round(p1.psfMag_i,3) as psf_i1,
1456: round(p1.psfMag_z,3) as psf_z1, round(p1.psfMagErr_u,3) as psfErr_u1,
1457: round(p1.psfMagErr_g,3) as psfErr_g1, round(p1.psfMagErr_r,3) as psfErr_r1,
1458: round(p1.psfMagErr_i,3) as psfErr_i1, round(p1.psfMagErr_z,3) as psfErr_z1,
1459: p1.objid as objid1,
1460: 
1461: round(p2.ra,6) as ra2, round(p2.dec,6) as dec2, round(p2.extinction_r,3) as rExt2,
1462: round(p2.psfMag_u,3) as psf_u2, round(p2.psfMag_g,3) as psf_g2,
1463: round(p2.psfMag_r,3) as psf_r2, round(p2.psfMag_i,3) as psf_i2,
1464: round(p2.psfMag_z,3) as psf_z2, round(p2.psfMagErr_u,3) as psfErr_u2,
1465: round(p2.psfMagErr_g,3) as psfErr_g2, round(p2.psfMagErr_r,3) as psfErr_r2,
1466: round(p2.psfMagErr_i,3) as psfErr_i2, round(p2.psfMagErr_z,3) as psfErr_z2,
1467: p2.objid as objid2,
1468: 
1469: round(NN.distance*60,3) as theta
1470: 
1471: into mydb.binaryClose
1472: 
1473: from Neighbors as NN join star as p1 on p1.objid = NN.objid
1474: join star as p2 on p2.objid = NN.neighborobjid
1475: where NN.mode = 1 and NN.neighbormode = 1
1476: and NN.type = 6 and NN.neighbortype = 6
1477: 
1478: and p1.psfMag_r between 14 and 20.5
1479: and (p1.flags_g & '229802225959076') = 0 and (p1.flags_r & '229802225959076') = 0
1480: and (p1.flags_i & '229802225959076') = 0 and (p1.flags_g & '268435456') > 0
1481: and (p1.flags_r & '268435456') > 0 and (p1.flags_i & '268435456') > 0
1482: 
1483: and p2.psfMag_r between 14 and 20.5
1484: and (p2.flags_g & '229802225959076') = 0 and (p2.flags_r & '229802225959076') = 0
1485: and (p2.flags_i & '229802225959076') = 0 and (p2.flags_g & '268435456') > 0
1486: and (p2.flags_r & '268435456') > 0 and (p2.flags_i & '268435456') > 0
1487: 
1488: and (p1.psfMag_r-p1.extinction_r) < (p2.psfMag_r-p2.extinction_r)
1489: and (p1.psfMag_g-p1.extinction_g - p1.psfMag_i+p1.extinction_i)<
1490: (p2.psfMag_g-p2.extinction_g - p2.psfMag_i+p2.extinction_i)
1491: 
1492: and NN.distance*60 between 3 and 4
1493: \end{verbatim}
1494: 
1495: \begin{verbatim}
1496: select -- kinematic selection of candidate binaries
1497: 
1498: round(p1.ra,6) as ra1, round(p1.dec,6) as dec1, round(p1.extinction_r,3) as ext1,
1499: round(p1.psfMag_u,3) as u1, round(p1.psfMag_g,3) as g1,
1500: round(p1.psfMag_r,3) as r1, round(p1.psfMag_i,3) as i1,
1501: round(p1.psfMag_z,3) as z1, round(p1.psfMagErr_u,3) as uErr1,
1502: round(p1.psfMagErr_g,3) as gErr1, round(p1.psfMagErr_r,3) as rErr1,
1503: round(p1.psfMagErr_i,3) as iErr1, round(p1.psfMagErr_z,3) as zErr1,
1504: (case when ((p1.flags & '16') = 0) then 1 else 0 end) as ISOLATED1,
1505: NN.objid as objid1,
1506: 
1507: round(p2.ra,6) as ra2, round(p2.dec,6) as dec2, round(p2.extinction_r,3) as ext2,
1508: round(p2.psfMag_u,3) as u2, round(p2.psfMag_g,3) as g2,
1509: round(p2.psfMag_r,3) as r2, round(p2.psfMag_i,3) as i2,
1510: round(p2.psfMag_z,3) as z2, round(p2.psfMagErr_u,3) as uErr2,
1511: round(p2.psfMagErr_g,3) as gErr2, round(p2.psfMagErr_r,3) as rErr2,
1512: round(p2.psfMagErr_i,3) as iErr2, round(p2.psfMagErr_z,3) as zErr2,
1513: (case when ((p2.flags & '16') = 0) then 1 else 0 end) as ISOLATED2,
1514: NN.neighborobjid as objid2,
1515: 
1516: round(NN.distance*60,3) as theta,
1517: round(s1.pmL,3) as pmL1, round(s1.pmB,3) as pmB1,
1518: round(s2.pmL,3) as pmL2, round(s2.pmB,3) as pmB2
1519: 
1520: into mydb.binaryPM
1521: 
1522: from Neighbors as NN join star as p1 on p1.objid = NN.objid
1523: join star as p2 on p2.objid = NN.neighborobjid
1524: join propermotions as s1 on s1.objid = NN.objid
1525: join propermotions as s2 on s2.objid = NN.neighborobjid
1526: 
1527: where NN.mode = 1 and NN.neighbormode = 1
1528: and NN.type = 6 and NN.neighbortype = 6
1529: 
1530: and p1.psfMag_r between 14 and 19.5
1531: and (p1.flags_g & '229802225959076') = 0 and (p1.flags_r & '229802225959076') = 0
1532: and (p1.flags_i & '229802225959076') = 0 and (p1.flags_g & '268435456') > 0
1533: and (p1.flags_r & '268435456') > 0 and (p1.flags_i & '268435456') > 0
1534: 
1535: and p2.psfMag_r between 14 and 19.5
1536: and (p2.flags_g & '229802225959076') = 0 and (p2.flags_r & '229802225959076') = 0
1537: and (p2.flags_i & '229802225959076') = 0 and (p2.flags_g & '268435456') > 0
1538: and (p2.flags_r & '268435456') > 0 and (p2.flags_i & '268435456') > 0
1539: 
1540: and (p1.psfMag_r-p1.extinction_r) < (p2.psfMag_r-p2.extinction_r)
1541: and (p1.psfMag_g-p1.extinction_g - p1.psfMag_i+p1.extinction_i)<
1542: (p2.psfMag_g-p2.extinction_g - p2.psfMag_i+p2.extinction_i)
1543: 
1544: and s1.match = 1 and s2.match = 1
1545: and s1.sigra < 350 and s1.sigdec < 350
1546: and s2.sigra < 350 and s2.sigdec < 350
1547: and sqrt(power(s1.pmL - s2.pmL,2) + power(s1.pmB - s2.pmB,2)) < 5
1548: and (case when sqrt(power(s1.pmL,2) + power(s1.pmB,2)) >
1549: sqrt(power(s2.pmL,2) + power(s2.pmB,2)) then
1550: sqrt(power(s1.pmL,2) + power(s1.pmB,2)) else
1551: sqrt(power(s2.pmL,2) + power(s2.pmB,2)) end) between 15 and 400
1552: \end{verbatim}
1553: 
1554: \section{  The Limitations of the Reduced Proper Motion Diagram   }
1555: 
1556: Recent analysis of metallicity and kinematics for halo and disk stars by I08a
1557: provides sufficient information to understand the behavior of the reduced proper
1558: motion diagram in quantitative detail (including both the sequence separation
1559: and their widths), and to demonstrate that its efficiency for separating halo
1560: and disk stars deteriorates at distances beyond a few kpc from the Galactic
1561: plane. As Equation~\ref{Eq:rpm2} shows, for a population of stars with the same
1562: $v_t$, the reduced proper motion is a measure of their absolute magnitude. For
1563: two stars with the same color that is sensitive to the effective temperature
1564: (such as the $g-i$ color), but with different metallicities and tangential
1565: velocities, the difference in their reduced proper motions is
1566: \begin{equation}
1567: \label{Eq:rpm3}
1568:  \Delta r_{RPM} = r^H_{RPM} - r^D_{RPM} =  \Delta M_r + 5\log{\left({v^H_t \over v^D_t}\right)},
1569: \end{equation}
1570: where $H$ and $D$ denote the two stars. In the limit that the {\it shape} of the
1571: photometric parallax relation does not depend on metallicity, $\Delta M_r$ does
1572: not depend on color, and is fully determined by the metallicity difference of
1573: the two stars (or populations of stars). Using metallicity distributions for
1574: disk and halo stars obtained by I08a, and their expression for
1575: $\Delta M_r([Fe/H])$ (Equation~A2), we find that the expected offset between
1576: $M_r$ for halo and disk stars with the same $g-i$ color varies from 0.6 mag for
1577: stars at 1 kpc from the Galactic plane to 0.7 mag at 5 kpc from the plane, where
1578: the variation is due to the vertical metallicity gradient for disk stars. The
1579: finite width of halo and disk metallicity distributions induces a spread of
1580: $M_r$ (root-mean-square scatter computed using interquartile range) of 0.15 mag
1581: for disk stars and 0.18 mag for halo stars. 
1582: 
1583: The effect of metallicity on the separation of halo and disk sequences in the
1584: reduced proper motion diagram is smaller than the effect of different tangential
1585: velocity distributions. Assuming for simplicity that stars are observed towards
1586: a Galactic pole, and that the median heliocentric tangential velocities are 30
1587: km s$^{-1}$ for disk stars and 200 km s$^{-1}$ for halo stars, the induced
1588: separation of their reduced proper motion sequences is $\sim$4.1 mag (the
1589: expected scatter in the reduced proper motion due to finite velocity dispersion
1590: is $\sim$1-1.5 mag). Together with the $\sim0.7$ mag offset due to different
1591: metallicity distributions, the separation of $\sim$5 mag between the two
1592: sequences makes the reduced proper motion diagram a promising tool for
1593: separating disk and halo stars.
1594: 
1595: However, the reduced proper motion diagram is an efficient tool only for stars
1596: within 1-2 kpc from the Galactic plane. The main reason for this limitation is
1597: the decrease of rotational velocity for disk stars with distance from the
1598: Galactic plane, with a gradient of about $-30$ km s$^{-1}$ kpc$^{-1}$ (see
1599: Section 3.4.2 in I08a). As the difference in rotational velocity between halo
1600: and disk stars diminishes with the distance from the plane, the separation of
1601: their reduced proper motion sequences decreases, too. A mild increase in the
1602: velocity dispersion of disk stars, as well as a decrease of their median
1603: metallicity with the distance from the plane, also decrease the sequence
1604: separation, but the dominant cause is the rotational velocity gradient.
1605: 
1606: To illustrate this effect, we select a sample of $\sim$60,000 stars with
1607: $14<r<20$ and $0.2<g-r<0.4$, that are observed towards the north Galactic pole
1608: ($b>70\arcdeg$). In this color range it is possible to separate disk and halo
1609: stars using photometric metallicity estimator from I08a, and we further select a
1610: sample of $\sim$16,000 likely disk stars with $[Fe/H] > -0.9$, and a sample of
1611: $\sim$34,400 likely halo stars with $[Fe/H] < -1.1$ (see Figure~9 in I08a for
1612: justification). Their proper motion distributions as functions of distance from
1613: the Galactic plane, $Z$, are shown in the top left panel in
1614: Figure~\ref{Fig:App1}. Because of the gradient in the rotational velocity for
1615: disk stars, their median proper motion becomes constant at $\sim$8 mas yr$^{-1}$
1616: beyond $Z\sim2$ kpc, while the median proper motion for halo stars is roughly
1617: proportional to $1/Z$, with a value of $\sim$11 mas yr$^{-1}$ at $Z=5$ kpc.
1618: 
1619: The top right panel in Figure~\ref{Fig:App1} shows the positions and widths of
1620: the reduced proper motion sequences for disk and halo stars as functions of $Z$,
1621: and the two bottom panels show the sequence cross-sections for stars with
1622: $Z=1-1.5$ kpc and $Z=3.5-4$ kpc. At distances beyond $\sim$2 kpc from the plane,
1623: the reduced proper motion diagram ceases to be an efficient tool for separating
1624: halo and disk stars because the two sequences start to significantly overlap.
1625: This increasing overlap is a result of the rotational velocity gradient for disk
1626: stars, and the finite width of halo and disk velocity distributions, and would
1627: be present even for {\it infinitely accurate} measurements (with the proper
1628: motion errors of $\sim$3 mas yr$^{-1}$ per coordinate, \citealt{mun04}, the
1629: sequence widths of $\sim$1.0-1.5 mag are dominated by velocity dispersions).
1630: Hence, beyond $\sim$2 kpc from the plane, metallicity measurements are necessary
1631: to reliably separate disk and halo populations.
1632: 
1633: The above analysis is strictly valid only for fields towards the north Galactic
1634: pole. \citet{sg03} found that the position of disk and halo reduced proper
1635: motion sequences, relative to their positions at the north Galactic pole, varies
1636: with galactic latitude as 
1637: \begin{equation}
1638: \label{Eq:rpm4}
1639:  \Delta r_{RPM}(b) = 5\log(v_t/v_t^{NGP}) = -1.43 \left(1-\sin(|b|)\right),
1640: \end{equation}
1641: where $v_t^{NGP}$ is the median value of $v_t$ for stars observed towards the
1642: north Galactic pole. This result is a bit unexpected because it does not contain
1643: longitudinal variation due to projection effects of the rotational motion of the 
1644: local standard of rest. We show the variation of $\Delta r_{RPM}$, for stars
1645: with $0.2<g-r<0.4$, as a function of galactic coordinates in
1646: Figure~\ref{Fig:App2}. We use photometric metallicity to separate stars into
1647: disk and halo populations. As figure demonstrates, the longitudinal dependence
1648: is present for halo sample, but not for disk samples. We have generated
1649: simulated behavior of $\Delta r_{RPM}$ using kinematic model from I08a, and
1650: reproduced the observed behavior to within the measurement noise. It turns out
1651: that the vertical gradient of rotational velocity for disk stars is fully
1652: responsible for the observed strong dependence of $\Delta r_{RPM}$ on latitude,
1653: and which masks the dependence on longitude. Hence, the $\sin(|b|)$ term
1654: proposed by \citet{sg03} is an indirect discovery of the vertical gradient of
1655: rotational velocity for disk stars! These empirical models also show that a
1656: linear dependence of $\Delta r_{RPM}(b)$ on $\sin(|b|)$ is only approximately
1657: correct, and that it ignores the dependence on distance. While a more involved
1658: best-fit expression is possible (full two-dimensional consideration of proper
1659: motion also helps to better separate disk and halo stars), we find that halo
1660: stars can always be efficiently rejected at $|b|>30\arcdeg$, if the separator
1661: shown in Figure~\ref{Fig:1wd} is shifted upwards by 1 mag. 
1662: 
1663: \section{The Modeling of Unresolved Binaries in the Samples of Wide Binaries}
1664: \label{model}
1665: 
1666: One major uncertainty when using a photometric parallax relation is the lack of
1667: information whether the observed ``star'' is a single star, or a binary
1668: (multiple) system. If the observed ``star'' is a binary system, its luminosity
1669: will be underestimated, with the magnitude of the offset depending on the actual
1670: composition of the binary. To model this offset, or to correct for it, one would
1671: ideally like to have a probability density map that gives the probability of a
1672: magnitude offset, $\Delta M_r$, as a function of the {\em observed} binary
1673: system's color.
1674: 
1675: To construct such a map, we have generated a sample of 100,000 unresolved binary
1676: systems by randomly pairing stars drawn from the \citet{ktg90} luminosity
1677: function. By independently drawing the luminosities of each component to 
1678: generate unresolved binary systems, we implicitly assume that the formation of
1679: each component is unaffected by the presence of the other. While there are other
1680: proposed mechanisms for binary formation (\citealt{cla07}, and references
1681: therein), we have chosen this one because it was easy to implement.
1682: 
1683: For every unresolved binary system we calculate the total $r$ band luminosity,
1684: and the $r-i$ and $g-i$ color of the system. The magnitude offset, $\Delta M_r$,
1685: caused by unresolved binarity, is obtained as the difference between the true
1686: $r$ band absolute magnitude, and the absolute magnitude for the pair's joint
1687: $r-i$ color calculated using Equation~\ref{Mr_bright}. The probability density
1688: map is then simply the number of unresolved binary systems (normalized with the
1689: total number of systems at a given color) as a function of $\Delta M_r$ and
1690: pair's joint $g-i$ color, shown in Figure~\ref{dMr_gi}.
1691: 
1692: It is worth noting that, with the adopted binary formation mechanism, the
1693: magnitude offset is the smallest ($\Delta M_r<0.1$ mag) for the bluest stars,
1694: and greatest ($\Delta M_r>0.7$ mag) for the reddest stars. Because of this, the
1695: scatter due to unresolved binarity in the $\delta$ distribution should be more
1696: pronounced in a sample of red stars ($g-i>2.0$), than in a sample of blue stars.
1697: 
1698: The map shown in Figure~\ref{dMr_gi} can be parametrized as a Gaussian
1699: distribution $P(\Delta M_r|\mu, \sigma)$, where
1700: \begin{equation}
1701: \mu = 0.037+0.10(g-i)+0.09(g-i)^2-0.012(g-i)^3
1702: \end{equation}
1703: is the median $\Delta M_r$, and
1704: \begin{equation}
1705: \sigma = 0.041+0.03(g-i)+0.15(g-i)^2-0.057(g-i)^3
1706: \end{equation}
1707: is the scatter (determined from the interquartile range). To verify the validity
1708: of this parametrization, we subtract $\Delta M_r$ and $\mu$, normalize the
1709: difference with $\sigma$, find the distribution of such values, and fit a
1710: Gaussian to it. As shown in Figure~\ref{dMr_chi_gauss}, the distribution of
1711: normalized residuals is well described by a Gaussian with $\sigma = 0.9$. The
1712: two peaks in the distribution are due to highly asymmetric distributions of
1713: $\Delta M_r$ values around the median $\Delta M_r$ for the bluest ($g-i\sim0.1$)
1714: and reddest ($g-i\sim2.9$) systems.
1715: 
1716: To create a sample of wide binaries where some of the stars are unresolved
1717: binary systems, first we select pairs with $20\arcsec<\theta<30\arcsec$ from
1718: the initial sample of stellar pairs. Following the procedure described in
1719: Section~\ref{algorithm}, we create the ``true'' wide binaries by changing the
1720: $r_2$ magnitude using Equation~\ref{r2_mag}, and add 0.15 mag of Gaussian noise
1721: to simulate the scatter in the photometric parallax due to photometric errors.
1722: A fraction of stars is then randomly converted to unresolved binary systems by
1723: subtracting a $\Delta M_r$ value from the $r$ band (apparent) magnitude, where
1724: the $\Delta M_r$ is drawn from a $g-i$ color-dependent
1725: $P(\Delta M_r|\mu, \sigma)$ distribution.
1726: 
1727: Figure~\ref{gauss_fit_red} shows the $\delta$ distribution for such a mock
1728: sample, where the components are redder than $g-i=2.0$ and have a 40\%
1729: probability to be unresolved binary systems. Different configurations of single
1730: stars and unresolved binaries that contribute to the observed $\delta$
1731: distribution can be easily identified. Wide binaries where both components are
1732: single stars contribute the central narrow Gaussian, with its width due to
1733: photometric errors. If the brighter component is an unresolved binary system,
1734: its absolute magnitude is underestimated, and the result is an offset in
1735: $\delta$ in the negative direction. A similar outcome happens if the fainter
1736: component is an unresolved binary system, but the offset is positive. Single
1737: star-unresolved binary configurations, therefore, contribute the left and the
1738: right Gaussians. If both components are unresolved binary systems, the $\delta$
1739: will be centered on zero and will be $\sigma_0 \sqrt2$ wide, where $\sigma_0$ is
1740: the width of the $(\Delta M_r - \mu)$ distribution. This behavior is consistent
1741: with the $\delta$ distributions observed in Figure~\ref{delta_hists}.
1742: 
1743: \begin{thebibliography}{}
1744: \bibitem[Adelman-McCarthy et al.(2008)]{amc08}Adelman-McCarthy, J. K. et al.
1745: 2008, \apjs, 175, 297
1746: \bibitem[Allen, Poveda, \& Herrera(2000)]{aph00}Allen, C., Poveda, A., \&
1747: Herrera, M. 2000, \aap, 356, 529
1748: \bibitem[Allen, Poveda, \& Hern\'andez-Alc\'antara(2007)]{aph07}Allen, C.,
1749: Poveda, A., \& Hern\'andez-Alc\'antara, A. 2007, in IAU Symp. 240, Binary Stars
1750: as Critical Tools \& Test in Contemporary Astrophysics, ed. B. Hartkopf,
1751: E. Guinan, \& P. Harmanec, 405
1752: \bibitem[Andersen(1991)]{and91}Andersen, J. 1991, \araa, 3, 91
1753: \bibitem[Bahcall \& Soneira(1981)]{bs81}Bahcall, J. N. \& Soneira, R. M. 1981,
1754: \aj, 246, 122
1755: \bibitem[Bochanski et al.(2008)]{boc08}Bochanski, J. et al, 2008, in prep
1756: \bibitem[Chanam\'e(2007)]{cha07}Chanam\'e, J. 2007, in IAU Symp. 240, Binary
1757: Stars as Critical Tools \& Tests in Contemporary Astrophysics, ed. B. Hartkopf,
1758: E. Guinan, \& P. Harmanec, 1
1759: \bibitem[Chanam\'e \& Gould(2004)]{cg04}Chanam\'e, J. \& Gould, A. 2004,
1760: \apj, 601, 289
1761: \bibitem[Clarke(2007)]{cla07}Clarke, C. J. 2007, in IAU Symp. 240, Binary Stars
1762: as Critical Tools \& Tests in Contemporary Astrophysics, ed. B. Hartkopf,
1763: E. Guinan, \& P. Harmanec, 337
1764: \bibitem[Covey et al.(2007)]{cov07}Covey, K. R., et al. 2007, \aj, 134, 2398
1765: \bibitem[Couteau(1960)]{cou60}Couteau, P. 1960, J. des Observateurs, 43, 41
1766: \bibitem[Croll(2006)]{cro06}Croll, B. 2006., \pasp, 118, 1351
1767: \bibitem[Duquennoy \& Mayor(1991)]{dm91}Duquennoy, A. \& Mayor, M. 1991, \aap,
1768: 248, 485
1769: \bibitem[Finlator et al.(2000)]{fin00}Finlator, K. et al. 2000, \aj, 120, 2615
1770: \bibitem[Fischer \& Marcy(1992)]{fm92}Fischer, D. A. \& Marcy, G. W. 1992, \apj,
1771: 396, 178
1772: \bibitem[Flaugher et al.(2007)]{fla07}Flaugher, B. \& the Dark Energy Survey
1773: Collaboration 2007, \baas, 209, 22.01
1774: \bibitem[Ford(2005)]{for05}Ford, E. B. 2005, \aj, 129, 1706
1775: \bibitem[Gelman \& Rubin(1992)]{gr92}Gelman, A. \& Rubin, D. 1992, Stat. Sci.,
1776: 7, 457
1777: \bibitem[Giersz(2006)]{gie06}Giersz, M. 2006, \mnras, 371, 484
1778: \bibitem[Gould(1995)]{gou95}Gould, A. 1995, \apj, 440, 510
1779: \bibitem[Gould \& Salim(2003)]{gs03}Gould, A., \& Salim, S. 2003, \apj, 582,
1780: 1001
1781: \bibitem[Gunn et al.(1998)]{gun98}Gunn, J. E. et al. 1998, \aj, 116, 3040
1782: \bibitem[Gunn et al.(2006)]{gun06}Gunn, J. E. et al. 2006, \aj, 131, 2332
1783: \bibitem[Hastings(1970)]{has70}Hastings, W. K. 1970, Biometrika, 57, 97
1784: \bibitem[Hawley et al.(2002)]{haw02}Hawley, S. L. et al. 2002, \aj, 123, 3409
1785: \bibitem[Heggie(1975)]{heg75}Heggie, D. C. 1975, \mnras, 173, 729
1786: \bibitem[Helmi et al.(2003)]{hel03}Helmi, A. et al. 2003, \apj, 586, 195
1787: \bibitem[Hogg et al.(2002)]{hog02}Hogg, D. W., Finkbeiner, D. P., Schlegel,
1788: D. J. \& Gunn, J.E. 2002, \aj, 122, 2129
1789: \bibitem[Hurley, Aarseth \& Shara(2007)]{hur07}Hurley, J. R., Aarseth, S. J., \&
1790: Shara, M. M. 2007, \apj, 665, 707
1791: \bibitem[Ivezi\'{c} et al.(2003)]{ive03}Ivezi\'{c}, \v{Z}. et al. 2003, \memsai,
1792: 74, 978
1793: \bibitem[Ivezi\'{c} et al.(2004)]{ive04}Ivezi\'{c}, \v{Z}. et al. 2004,
1794: Astron. Nachr., 325, 583
1795: \bibitem[Ivezi\'c et al.(2005)]{ive05}Ivezi\'c, \v{Z}., Vivas, A. K., Lupton,
1796: R. H., \& Zinn, R. 2005, \aj, 129, 1096
1797: \bibitem[Ivezi\'c et al.(2008a)]{ive08a}Ivezi\'c, \v{Z}. et al. 2008a, accepted to
1798: \apj (also astro-ph/0804.3850)
1799: \bibitem[Ivezi\'c et al.(2008b)]{ive08b}Ivezi\'c, \v{Z}. et al. 2008b, astro-ph/0805.2366
1800: \bibitem[Juri\'c et al.(2008)]{jur08}Juri\'c, M. et al. 2008, \apj, 673, 864
1801: \bibitem[Kaiser et al.(2002)]{kai02}Kaiser, N. et al. 2002, \procspie, 4836, 154
1802: \bibitem[Keller et al.(2007)]{kel07}Keller, S.C. et al. 2007, 24, 1
1803: \bibitem[Kroupa, Tout \& Gilmore(1990)]{ktg90}Kroupa, P., Tout, C. A. \&
1804: Gilmore, G. 1990, \mnras, 244, 76
1805: \bibitem[Laird, Carney \& Latham(1988)]{lcl88}Laird, J. B., Carney, B. W \&
1806: Latham, D. W. 1988, \aj, 95, 1843
1807: \bibitem[Lenz et al.(1998)]{len98}Lenz, D. D., Newberg, J., Rosner, R.,
1808: Richards, G. T., \& Stoughton, C. 1998, \apjs, 119, 121
1809: \bibitem[L\'{e}pine \& Bongiorno(2007)]{lb07}L\'{e}pine, S. \& Bongiorno, B.
1810: 2007, \aj, 133, 889
1811: \bibitem[Lupton et al.(2002)]{lup02}Lupton, R. H., Ivezi\'{c}, \v{Z}., Gunn,
1812: J. E., Knapp, G. R., Strauss, M. A. \& Yasuda, N. 2002 \procspie, 4836, 350
1813: \bibitem[Luyten(1979)]{luy79}Luyten, W. J. 1979, New Luyten Catalog of Stars
1814: with Proper Motions Larger than Two Tenths of an Arcsecond (Minneapolis: Univ.
1815: of Minnesota Press)
1816: \bibitem[Minowski \& Abel(1963)]{ma63}Minkowski, A. \& Abell, G. O. 1963, in
1817: Stars and Stellar Systems, Vol. 3
1818: \bibitem[Metropolis et al.(1953)]{met53}Metropolis, N., Rosenbluth, A. W.,
1819: Rosenbluth, M. N., Teller, A. H., \& Teller, E., J. Chem. Phys., 21, 1087
1820: \bibitem[Monet et al.(2003)]{mon03}Monet, D. G. et al. 2003, \aj, 125, 984
1821: \bibitem[Munn et al.(2004)]{mun04}Munn, J. A. et al. 2004, \aj, 127, 3034
1822: \bibitem[\"Opik(1924)]{oep24}\"Opik, E. J. 1924, Tartu Obs. Publ. 25, No. 6
1823: \bibitem[Perryman et al.(2001)]{per01}Perryman, M. A. C. et al. 2001, \aap,
1824: 369, 339
1825: \bibitem[Pier et al.(2003)]{pie03}Pier, J. R., Munn, J. A., Hindsley, R. B.,
1826: Hennesy, G. S., Kent, S. M., Lupton, R. H. \& Ivezi\'{c}, \v{Z}. 2003, \aj,
1827: 125, 1559
1828: \bibitem[Poveda et al.(1994)]{pov94}Poveda, A. et al. 1994, Revista Mexicana de
1829: Astronom\'{i}a y Astrofis\'{i}ca, 28, 43
1830: \bibitem[Poveda, Allen \& Hern\'andez-Alc\'antara(2007)]{pah07}Poveda, A.,
1831: Allen, C. \& Hern\'andez-Alc\'antara, A. 2007, in IAU Symp. 240, Binary Stars as
1832: Critical Tools \& Test in Contemporary Astrophysics, ed. B. Hartkopf, E. Guinan,
1833: \& P. Harmanec, 417
1834: \bibitem[Press et al.(1992)]{pre92}Press, W. H., Teukolsky, S. A., Vetterling,
1835: W. T., \& Flannery, B. P. 1992, Numerical Recipes in C (2nd ed.; New York:
1836: Cambridge Univ. Press)
1837: \bibitem[Reid et al.(1991)]{rei91}Reid, I. N. et al. 1991, \pasp, 103, 661
1838: \bibitem[Ribas(2006)]{rib06}Ribas, I. 2006, ASP Conf.~349, 55
1839: \bibitem[Richards et al.(2002)]{ric02}Richards, G. T. et al. 2002, \aj, 123,
1840: 2945
1841: \bibitem[Salim \& Gould(2003)]{sg03}Salim, S. \& Gould, A. 2003, \apj, 582, 1011
1842: \bibitem[Schlegel, Finkbeiner \& Davis(1998)]{SFD98}Schlegel, D., Finkbeiner,
1843: D. P., \& Davis, M. 1998, \apj 500, 525
1844: \bibitem[Scranton et al.(2002)]{scr02}Scranton, R. et al. 2002, \apj, 579, 48
1845: \bibitem[Sesar et al.(2007)]{ses07}Sesar, B. et al. 2007, \aj, 134, 2236
1846: \bibitem[Siegel et al.(2002)]{sie02}Siegel, M. H., Majewski, S. R., Reid, I. N.
1847: \& Thompson, I. B. 2002, \apj, 578, 151
1848: \bibitem[Sirko et al.(2004)]{sir04}Sirko, E. et al. 2004, \aj, 127, 899
1849: \bibitem[Smith et al.(2002)]{smi02}Smith, J. A. et al. 2002, \aj, 123, 2121
1850: \bibitem[Stoughton et al.(2002)]{sto02}Stoughton, C. et al. 2002, \aj, 123, 485
1851: \bibitem[Tegmark et al.(2004)]{teg04}Tegmark, M. et al. 2004, \prd, 69, 103501
1852: \bibitem[Tucker et al.(2006)]{tuc06}Tucker, D. et al. 2006, Astron. Nachr., 327,
1853: 821
1854: \bibitem[Weinberg, Shapiro \& Wasserman(1987)]{wsw87}Weinberg, M. D, Shapiro,
1855: S. L.,  \& Wasserman, I. 1987, \apj, 312, 367
1856: \bibitem[West, Walkowicz, \& Hawley(2005)]{wwh05}West, A. A., Walkowicz, L. M.
1857: \& Hawley, S. L. 2005, \pasp, 117, 706
1858: \bibitem[Wilkinson et al.(2005)]{wil05}Wilkinson, M. I. et al. 2005, \mnras,
1859: 359, 1306
1860: \bibitem[Yanny et al.(2000)]{yan00}Yanny, B. et al. 2000, \apj, 540, 825
1861: \bibitem[Yoo, Chanam\'e, \& Gould(2004)]{ycg04}Yoo, J., Chanam\'e, J. \& Gould,
1862: A. 2004, \apj, 601, 311
1863: \end{thebibliography}
1864: 
1865: \clearpage
1866: 
1867: %% TABLES %%
1868: 
1869: \begin{deluxetable*}{crrrr}
1870: \tabletypesize{\scriptsize}
1871: \tablecolumns{5}
1872: \tablewidth{0pc}
1873: \tablecaption{The centers, widths, and areas for best-fit Gaussian distributions\label{gauss}}
1874: \tablehead{
1875: \multicolumn{1}{c}{} & \multicolumn{2}{c}{Geometrically-selected sample} & \multicolumn{2}{c}{Kinematically-selected sample} \\
1876: \cline{1-5} \\
1877: \colhead{ } & \colhead{Narrow Gaussian} & \colhead{Wide Gaussian} &
1878: \colhead{Narrow Gaussian} & \colhead{Wide Gaussian} 
1879: }
1880: \startdata
1881: Center                            &  -0.01 &   -0.03 &  -0.05 &   0.01 \\
1882: Width                             &   0.12 &    0.54 &   0.11 &   0.51 \\
1883: Area\tablenotemark{a}             &   0.26 &    0.74 &   0.34 &   0.66
1884: \enddata
1885: \tablenotetext{a}{Areas of the narrow and wide Gaussians sum to 1}
1886: \end{deluxetable*}
1887: 
1888: \clearpage
1889: 
1890: \begin{deluxetable*}{crrr}
1891: \tabletypesize{\scriptsize}
1892: \tablecolumns{4}
1893: \tablewidth{0pc}
1894: \tablecaption{The conditional probability density functions $P[(g-i)_B|(g-i)_A]=a+b(g-i)+c(g-i)^2$\label{tbl_wide_bin_prob}}
1895: \tablehead{
1896: \multicolumn{1}{c}{} & \multicolumn{3}{c}{Best-fit parameters} \\
1897: \cline{1-4} \\
1898: \colhead{$(g-i)_A$ bin} & \colhead{a} & \colhead{b} & \colhead{c} 
1899: }
1900: \startdata
1901: $0.4 < (g-i)_A < 0.8$ & 0.38 &     0 &    0 \\
1902: $0.8 < (g-i)_A < 1.2$ & 0.46 &     0 &    0 \\
1903: $1.2 < (g-i)_A < 1.6$ & 0.37 &     0 &    0 \\
1904: $1.6 < (g-i)_A < 2.0$ & 0.37 &     0 &    0 \\
1905: $2.0 < (g-i)_A < 2.4$ & 0.08 &  0.14 & 0.04 \\
1906: $2.4 < (g-i)_A < 2.8$ & 0.23 & -0.50 & 0.38
1907: \enddata
1908: \end{deluxetable*}
1909: 
1910: \clearpage
1911: 
1912: %% FIGURES %%
1913: 
1914: \begin{figure}
1915: \epsscale{0.5}
1916: \plotone{f1a_color.eps}
1917: 
1918: \plotone{f1b_color.eps}
1919: 
1920: \plotone{f1c_color.eps}
1921: \caption{
1922: {\em Top:} A comparison of observed ($f_{obs}$, solid histogram) and random
1923: ($f_{rnd}$, dashed line, see text) distributions of angular separation $\theta$.
1924: {\em Middle:} Ratio $f_{obs}/f_{rnd}$ as a function of angular separation
1925: $\theta$.
1926: {\em Bottom:} Fraction of true binary systems, $\epsilon$, as a function of
1927: angular separation $\theta$.
1928: \label{ang_sep}}
1929: \end{figure}
1930: 
1931: \clearpage
1932: 
1933: \begin{figure}
1934: \epsscale{0.3}
1935: \plotone{f2a.eps}
1936: 
1937: \plotone{f2b.eps}
1938: 
1939: \plotone{f2c.eps}
1940: \caption{
1941: Distribution of counts for the geometrically-selected candidate sample
1942: ({\em top}), random sample ({\em middle}), and the ratio of two maps
1943: ({\em bottom}) in the $\Delta r=r_2-r_1$ vs.~$\Delta(g-i)=(g-i)_2 - (g-i)_1$
1944: diagram, binned in $0.05\times0.1$ mag bins. The average candidate-to-random
1945: ratio in the region outlined by the dashed lines (Eq.~\ref{color_cut1}
1946: and~\ref{color_cut2}) is $\sim1.7$, implying that $>40\%$ of candidates are true
1947: binaries.
1948: \label{counts}}
1949: \end{figure}
1950: 
1951: \clearpage
1952: 
1953: \begin{figure}
1954: \epsscale{0.4}
1955: \plotone{f3.eps}
1956: \caption{
1957: The $r$ vs.~$g-i$ distribution of brighter ({\em top}) and fainter
1958: ({\em bottom}) components from the geometrically-selected sample of candidate
1959: binaries, shown with linearly spaced contours.
1960: \label{r_vs_gi}}
1961: \end{figure}
1962: 
1963: \clearpage
1964: 
1965: \begin{figure}
1966: \epsscale{0.55}
1967: \plotone{f4.eps}
1968: \caption{The color-coded map, with the legend shown in the top right corner,
1969: shows the logarithm of the volume number density (stars/kpc$^3$/mag) of
1970: $\sim$2.8 million stars with $14<r<21.5$ observed towards the north Galactic
1971: pole ($b>70\arcdeg$), as a function of their distance modulus and the $g-i$
1972: color (the density variation in the horizontal direction represents luminosity
1973: function, and the variation in the vertical direction reflects the spatial
1974: volume density profiles of disk and halo stars). The absolute magnitudes are
1975: computed using expressions A3 and A7 from I08a, and the displayed distance range
1976: is 100 pc to 25 kpc. Stars are color-selected from the main stellar locus
1977: (dominated by main-sequence stars) using criteria 3-5 from Section 2.3.1 in
1978: I08a. The metallicity correction is applied using photometric metallicity for
1979: stars with $g-i<0.7$ (based on Eq.~4 from I08a), and by assuming $[Fe/H]=-0.6$
1980: for redder stars. As illustrated above the $g-i$ axis using the MK spectral type
1981: vs.~$g-i$ color table from \citet{cov07}, this color roughly corresponds to G5.
1982: The two vertical arrows mark the turn-off color for disk stars, and the red edge
1983: of M dwarf color distribution (there are redder M dwarfs detected by SDSS, but 
1984: their volume number density, i.e., the luminosity function, falls precipitously 
1985: beyond this limit; J. Bochanski, priv. comm.). The two diagonal dashed lines 
1986: show the apparent magnitude limits, $r=14$ and $r=21.5$. The dot-dashed
1987: diagonal line corresponds to $r=20$, which approximately describes the 50\% 
1988: completeness limit for stars with cataloged proper motions \citep{mun04}. Around
1989: the marked distance range of 3-4 kpc, the counts of halo stars begin to dominate
1990: disk stars (see Fig.~6 in I08a), and the distance range around 1 kpc offers the
1991: largest color completeness.
1992: \label{Fig:DMvsColor}}
1993: \end{figure}
1994: 
1995: \clearpage
1996: 
1997: \begin{figure}
1998: \epsscale{1.0}
1999: \plotone{f5.eps}
2000: \caption{The reduced proper motion diagrams for two subsamples of stars shown in
2001: Fig.~\ref{Fig:DMvsColor}. The color-coded maps show the logarithm of the number
2002: of stars per pixel, according to the legends. The left panel corresponds to a
2003: sample of $\sim$446,000 stars with proper motions in the range 15-50 mas/yr, and
2004: the right panel to a sample of 43,000 stars from the range 50-400 mas/yr. The
2005: requirement of larger proper motions introduces bias towards closer, and thus
2006: redder stars. Two two long-dashed lines in each panel correspond to photometric
2007: parallax relation from I08a, evaluated for $[Fe/H]=-0.6$ and with tangential
2008: velocity of 55 km/s (top curve) and 120 km/s (bottom curve). This variation of
2009: tangential velocity is consistent with the rotational velocity gradient
2010: discussed by I08a. The dot-dashed line is evaluated for $[Fe/H]=-1.5$ and with
2011: tangential velocity of 300 km/s. The short-dashed line (second from the bottom)
2012: separates disk and halo stars, and is evaluated for $[Fe/H]=-1.5$ and with
2013: tangential velocity of 180 km/s.
2014: \label{Fig:1wd}}
2015: \end{figure}
2016: 
2017: \clearpage
2018: 
2019: \begin{figure}
2020: \epsscale{0.85}
2021: \plotone{f6.eps}
2022: \caption{Analogous to Fig.~\ref{Fig:DMvsColor}, for subsamples selected using
2023: proper motion measurements. Out of 2.8 million stars shown in
2024: Fig.~\ref{Fig:DMvsColor}, 1.24 million are brighter than $r=19.5$ and have
2025: proper motion measurements. Of those, 498,000 have proper motion in the range
2026: 15-400 mas/yr (only 10\% of selected stars have proper motions greater than 50
2027: mas/yr). The color-coded map in the top left panel shows the fraction of such
2028: stars, as a function of distance and the $g-i$ color. At a distance of $\sim$1
2029: kpc, about half of all stars have proper motion larger than 15 mas/yr. The top
2030: right panel shows the counts of candidate disk stars, selected as stars above
2031: the separator shown in Fig.~\ref{Fig:1wd}, and the bottom left panel shows halo
2032: stars selected from below the separator. The bottom right panel shows the counts
2033: of halo stars, as a fraction of all stars selected using the reduced proper
2034: motion diagram. Note that beyond 3 kpc, the sample is dominated by halo stars.
2035: \label{Fig:2wd}}
2036: \end{figure}
2037: 
2038: \clearpage
2039: 
2040: \begin{figure}
2041: \epsscale{0.45}
2042: \plotone{f7_color.eps}
2043: \caption{
2044: {\em Top:} The photometric metallicity vs.~distance from the Galactic plane
2045: diagram for candidate binaries selected from the geometric sample using
2046: $|\delta|<0.4$ and $0.2<(g-r)_1<0.4$. The $|\delta|<0.4$ cut is used to reduce
2047: the contamination by random pairs (see Section~\ref{scatter}). Note that the
2048: fraction of low-metallicity halo binaries ($[Fe/H]<-1$) becomes significant only
2049: at $Z>2$ kpc. {\em Middle:} Analogous to the top panel, except that binaries
2050: from the kinematic sample are shown. Dots correspond to binaries with reduced
2051: proper motions characteristic of disk binaries, and triangles to candidate halo
2052: binaries. Note that binaries with disk-like metallicity ($[Fe/H]>-1$) at large
2053: distances ($Z>2$ kpc) are misclassified as halo binaries. {\em Bottom:} The
2054: comparison of the $(u-g)_1$ color distributions, and corresponding photometric
2055: metallicity distributions, for binaries from the top two panels. The metallicity
2056: vs.~$u-g$ color transformation is taken from I08a. The distribution for binaries
2057: from the geometric sample is shown by the thick solid line, and the
2058: distributions for binaries from the kinematic sample are shown by the thin solid
2059: line (disk candidates) and dotted line (halo candidates). 
2060: \label{plot_ug}}
2061: \end{figure}
2062: 
2063: \clearpage
2064: 
2065: \begin{figure}
2066: \epsscale{0.55}
2067: \plotone{f8a_color.eps}
2068: 
2069: \plotone{f8b_color.eps}
2070: \caption{
2071: Distribution of $\delta=(M_{r2}-M_{r1})-(r_2-r_1)$ values for a mock sample of
2072: candidate binaries ({\em solid line}) when $M_r=M_r(r-i|\mathbf{p_0})$
2073: ({\em top}), and for a $M_r$ different from $M_r(r-i|\mathbf{p_0})$
2074: ({\em bottom}). The fraction of random pairs (the contamination) in the sample
2075: is $80\%$. The $\delta$ distribution for ``true'' binaries ({\em dots}) is
2076: obtained by subtracting the $\delta$ distribution of random pairs ({\em open
2077: circles}) from the candidate binary $\delta$ distribution. The best-fit Gaussian
2078: for the ``true'' binaries $\delta$ distribution is centered on 0 and 0.1 mag
2079: wide in the top panel, and centered on -0.02 and 0.13 mag wide in the bottom
2080: panel.
2081: \label{delta_hist_mock}}
2082: \end{figure}
2083: 
2084: \clearpage
2085: 
2086: \begin{figure}
2087: \epsscale{0.44}
2088: \plotone{f9a.eps}
2089: 
2090: \plotone{f9b.eps}
2091: \caption{
2092: The dependence of median $\delta$, $\langle \delta \rangle$, values on $r-i$
2093: colors of the brighter and fainter component for the geometrically- ({\em top})
2094: and kinematically-selected ({\em bottom}) samples of candidate binaries with
2095: $|\delta|<0.4$. The $r-i$ color axes are interpolated from $g-i$ axes using
2096: Eq.~\ref{interp}. Sources are binned in $0.1\times0.1$ mag $g-i$ color pixels
2097: (minimum of 6 sources per pixel), and the median values are color-coded
2098: according to the legends given at the top of each panel. Inset histograms show
2099: the distribution of $\langle \delta \rangle$. The $\langle \delta \rangle$
2100: distribution medians are 0 to within $<0.01$ mag, and the scatter (determined
2101: from the interquartile range) is 0.07 mag for both samples.
2102: \label{medians}}
2103: \end{figure}
2104: 
2105: \clearpage
2106: 
2107: \begin{figure}
2108: \epsscale{1.0}
2109: \plotone{f10_color.eps}
2110: \caption{
2111: Comparison of Eq.~\ref{Mr_faint} ({\em dot-dashed line}) and Eq.~\ref{Mr_bright}
2112: ({\em dashed line}) photometric parallax relations from J08 (their
2113: Eqs.~1 and 2) with Eq.~\ref{Mr_geo} ({\em solid line}) and Eq.~\ref{Mr_kin}
2114: ({\em dotted line}) photometric parallax relations determined in this work. The
2115: inset shows the magnitude difference, $\Delta=M_{J08}-M_{S08}$, between the
2116: Eq.~\ref{Mr_bright} photometric parallax relation, and Eqs.~\ref{Mr_geo}
2117: ({\em solid line}) and~\ref{Mr_kin} ({\em dotted line}) from this work. The rms
2118: scatter between Eqs.~\ref{Mr_geo} and~\ref{Mr_kin}, and Eq.~\ref{Mr_bright}, is
2119: 0.13 mag. The rms scatter between Eqs.~\ref{Mr_geo} and~\ref{Mr_kin}
2120: ({\em dashed line}) is also 0.13 mag.
2121: \label{Mr_ri}}
2122: \end{figure}
2123: 
2124: \clearpage
2125: 
2126: \begin{figure}
2127: \epsscale{0.5}
2128: \plotone{f11.eps}
2129: \caption{
2130: A comparison of $(g-i)_2$ vs.~$(g-i)_1$ color-color distributions of
2131: geometrically-selected ({\em top}) and kinematically-selected disk binaries
2132: ({\em bottom}) with $|\delta|<0.4$. The fraction of binaries in a pixel is
2133: color-coded according to legends. The pixels are $0.2\times0.2$ mag wide in
2134: $g-i$ color, and the $r-i$ color axes are interpolated from $g-i$ axes using
2135: Eq.~\ref{interp}.
2136: \label{geo_kin_gi1_gi2_comp}}
2137: \end{figure}
2138: 
2139: \clearpage
2140: 
2141: \begin{figure}
2142: \epsscale{0.55}
2143: \plotone{f12a_color.eps}
2144: 
2145: \plotone{f12b_color.eps}
2146: \caption{
2147: Distribution of $\delta$ values for the geometrically- ({\em top}) and
2148: kinematically-selected ({\em bottom}) samples of candidate binaries, with
2149: absolute magnitudes, $M_r$, calculated using Eqs.~\ref{Mr_geo} and~\ref{Mr_kin},
2150: respectively. The $\delta$ distribution for true binaries ({\em open circles})
2151: is obtained by subtracting the $\delta$ distribution of random pairs
2152: ({\em triangles}) from the $\delta$ distribution for candidate
2153: binaries ({\em thick solid line}). The $\delta$ distribution for true binaries
2154: is a non-Gaussian distribution ({\em dashed line}), that can be described as a
2155: sum of two Gaussian distributions. The centers, widths, and areas for the
2156: best-fit narrow ({\em dotted line}) and wide ({\em thin solid line}) Gaussian
2157: distributions are given in Table~\ref{gauss}. The integrals (areas) of $\delta$
2158: distributions for random pairs and candidate binaries are
2159: $A_{random}$ and $A_{observed}$, respectively.
2160: \label{delta_hists}}
2161: \end{figure}
2162: 
2163: \clearpage
2164: 
2165: \begin{figure}
2166: \epsscale{0.7}
2167: \plotone{f13_color.eps}
2168: \caption{
2169: Distribution of $\delta$ values normalized by the expected formal errors,
2170: $\sigma_{\delta}$, for the kinematically-selected sample of candidate binaries.
2171: The $\delta/\sigma_{\delta}$ distribution for true binaries ({\em open circles})
2172: is obtained by subtracting the $\delta/\sigma_{\delta}$ distribution of random
2173: pairs ({\em triangles}) from the $\delta/\sigma_{\delta}$ distribution for
2174: candidate binaries ({\em thick solid line}). The $\delta/\sigma_{\delta}$
2175: distribution for true binaries is a non-Gaussian distribution ({\em dashed
2176: line}), that can be described as a sum of two Gaussian distributions. The
2177: best-fit narrow Gaussian ({\em dotted line}) is 0.75 wide and centered on -0.10,
2178: while the best-fit wide Gaussian ({\em thin solid line}) is 4.04 wide and
2179: centered on -0.14.
2180: \label{delta_norm}}
2181: \end{figure}
2182: 
2183: \clearpage
2184: 
2185: \begin{figure}
2186: \epsscale{1.0}
2187: \plotone{f14_color.eps}
2188: \caption{
2189: Dependence of median PSF magnitude errors on magnitude for the brighter
2190: ({\em left}) and fainter ({\em right}) components in the geometrically-
2191: ({\em dots}) and kinematically-selected ({\em triangles}) samples of candidate
2192: binaries. The vertical bars show the rms scatter in each bin (not the error of
2193: the median which is much smaller). The fainter components of
2194: geometrically-selected candidate binaries have overestimated median PSF
2195: magnitude errors when compared to the kinematically-selected binaries.
2196: \label{photo_errors}}
2197: \end{figure}
2198: 
2199: \clearpage
2200: 
2201: \begin{figure}
2202: \epsscale{0.29}
2203: \plotone{f15a.eps}
2204: 
2205: \plotone{f15b.eps}
2206: 
2207: \plotone{f15c.eps}
2208: \caption{
2209: The fraction of $|\delta|<0.4$ binaries in $0.7 < d/kpc < 1.0$ volume-complete
2210: geometrically-selected ({\em top}) and random ({\em middle}) samples that have
2211: $(g-i)_1$ and $(g-i)_2$ as the colors of the brighter and fainter component.
2212: The pixels are $0.2\times0.2$ mag wide in $g-i$ color, and the $r-i$ color axes
2213: are interpolated from $g-i$ axes using Eq.~\ref{interp}. The pixels in maps sum
2214: to 1. The bottom plot shows the difference,
2215: $f_{cand}[(g-i)_1,(g-i)_2]-C*f_{rand}[(g-i)_1,(g-i)_2]$, between the two maps,
2216: where $C=0.14$ is the fraction of random pairs estimated
2217: using Eq.~\ref{eff} for the $|\delta|<0.4$, $0.7 < d/kpc < 1.0$
2218: geometrically-selected sample. The pixels with negative values are not shown and
2219: the map is renormalized so that the pixels sum to 1.
2220: \label{counts_0.7_1.0}}
2221: \end{figure}
2222: 
2223: \clearpage
2224: 
2225: \begin{figure}
2226: \epsscale{1.0}
2227: \plotone{f16_color.eps}
2228: \caption{
2229: Conditional probability density of having one component with $(g-i)_B$ color in
2230: a wide binary system where the other component has $(g-i)_A$. The conditional
2231: probability density for $(g-i)_A < 2.0$ ({\em top} and {\em middle}) is
2232: independent of $(g-i)_B$, while for $(g-i)_A > 2.0$ ({\em bottom}) it changes as
2233: a square of $(g-i)_B$. The best-fit functions describing these conditional
2234: probabilities are given in Table~\ref{tbl_wide_bin_prob}.
2235: \label{wide_binary_prob}}
2236: \end{figure}
2237: 
2238: \clearpage
2239: 
2240: 
2241: \begin{figure}
2242: \epsscale{0.5}
2243: \plotone{f17.eps}
2244: \caption{
2245: {\em Top:} A comparison of $g-i$ color distribution of stars in the
2246: $|\delta|<0.4$, $0.7 < d/kpc < 1.0$ volume-complete, geometrically-selected,
2247: wide binary sample ({\em solid line}), and of all stars in the same volume
2248: ({\em dashed line}). The distributions are normalized to an area of 1, and the
2249: error bars show the Poisson noise. {\em Bottom:} The probability density for
2250: finding a star with $g-i$ color in a wide binary system,
2251: $P[(g-i)_A] = P_{wide binary}$, calculated as a ratio of the two distributions
2252: from the top panel, and renormalized to an area of 1. The equal probability
2253: distribution is shown as the dashed line. 
2254: \label{prob_gi}}
2255: \end{figure}
2256: 
2257: \clearpage
2258: 
2259: \begin{figure}
2260: \epsscale{0.38}
2261: \plotone{f18_color.eps}
2262: \caption{
2263: {\em Top left:} The cumulative distribution of $\log(a)$ for
2264: geometrically-selected candidate binaries with $|\delta|<0.2$ and
2265: $0.7< Z/kpc < 1.0$, where $a$ is the average semi-major axis. {\em Top right:}
2266: The differential distribution of angular separation, $\theta$, for
2267: geometrically-selected candidate binaries with $|\delta|<0.2$ and
2268: $0.7< Z/kpc < 1.0$. The distribution of random pairs ({\em dashed line}) is
2269: obtained by fitting a linear function $f_{rnd}(\theta) = C \, \theta$ to the
2270: observed histogram for $\theta > 18\arcsec$. $\theta_{max}$ is defined as the
2271: angular separation for which the fraction of true binaries falls below
2272: $\sim5\%$. {\em Middle left:} The fraction of true binaries, $\epsilon$
2273: ({\em solid line}), calculated from the $\theta$ distribution using
2274: Eq.~\ref{eff2} (see Section~\ref{geo_select}) for the $0.7< Z/kpc < 1.0$ sample,
2275: is modeled as a second-degree polynomial, $\epsilon(\theta)$
2276: ({\em dashed line}). For three $\theta-$selected subsamples
2277: ($4\arcsec-5\arcsec$, $5\arcsec-6\arcsec$, and $7\arcsec-8\arcsec$), the
2278: fraction of true binaries was also calculated using Eq.~\ref{eff} (i.e., from
2279: the $\delta$ distribution) and is shown with symbols. {\em Middle right:} The
2280: box ({\em dashed lines}) shows the allowed range in $a$ defined by $Z_{min}$,
2281: $Z_{max}$, and $\theta_{max}$ (see Eqs.~\ref{a_min} and~\ref{a_max}). Only
2282: binaries within this $a$ range are considered when plotting the corrected
2283: cumulative distribution of $\log(a)$. {\em Bottom left:} The cumulative
2284: distribution of $\log(a)$ for candidate binaries with $|\delta|<0.2$ and
2285: $0.7< Z/kpc < 1.0$ ({\em dashed line}), corrected using $\epsilon(\theta)$ to
2286: account for the decreasing fraction of true binaries at large
2287: $\theta\propto a/d$ separations. The vertical lines show $\log(a)$ for which the
2288: straight line fit ({\em dot-dashed line}) to the cumulative distribution
2289: deviates by more than $1.0\%$ ($\log(a_{low})$), $1.5\%$ ($\log(a_{break})$),
2290: and $2.0\%$ ($\log(a_{high})$). {\em Bottom right:} The corrected cumulative
2291: distribution of $\log(a)$ for mock candidate binaries created using the
2292: $f(a)\propto a^{-0.8}$ distribution limited to $a_1 = 100$ AU and $a_2 = 10000$
2293: AU.
2294: \label{loga_cum}}
2295: \end{figure}
2296: 
2297: \clearpage
2298: 
2299: \begin{figure}
2300: \epsscale{0.9}
2301: \plotone{f19_color.eps}
2302: \caption{
2303: Similar to Fig.~\ref{loga_cum} ({\em bottom}) plot, but for different $Z$
2304: (height above the Galactic plane) bins ranging from $0.1< Z/kpc < 0.4$
2305: ({\em top left}) to $2.6< Z/kpc < 3.6$ ({\em bottom right}). The sampled range
2306: of average semi-major axes and angular separations is given for each panel. In
2307: the $0.1< Z/kpc < 0.4$ bin ({\em top left}), the upper limit on
2308: $\log(a_{break})$ is 3.50.
2309: \label{loga_cum_fits}}
2310: \end{figure}
2311: 
2312: \clearpage
2313: 
2314: \begin{figure}
2315: \epsscale{1.0}
2316: \plotone{f20_color.eps}
2317: \caption{
2318: The fraction of true binaries ($\epsilon$) in the $0.1 <Z/kpc < 0.4$,
2319: $|\delta|<0.2$ geometrically-selected sample as a function of angular
2320: separation. The fraction goes below $\sim5\%$ at $\theta_{max}=16\arcsec$, and
2321: puts the upper limit on probed semi-major axes to $\sim3,200$ AU.
2322: \label{theta_eff_0.1_0.4_geo}}
2323: \end{figure}
2324: 
2325: \clearpage
2326: 
2327: \begin{figure}
2328: \epsscale{0.7}
2329: \plotone{f21_color.eps}
2330: \caption{
2331: {\em Top left:} The dependence of $\log(a_{break})$ values
2332: (c.f.~Fig.~\ref{loga_cum_fits}) on $\log(Z)$ ({\em dots}) is modeled as
2333: $\log(a_{break})=k \, \log(Z)+l$, where $k=0.72\pm0.05$ and $l=1.93\pm0.15$,
2334: or approximately, $a_{break}[AU] = 12,302 \, Z[kpc]^{0.72}$. The symbol size
2335: shows the range between $\log(a_{low})$ and $\log(a_{high})$. The arrow
2336: indicates that the $\log(a_{break})$ in the $0.1< Z/kpc < 0.4$ bin
2337: ($\log(Z)\sim2.4$) is an upper limit. {\em Top right:} The dependence of
2338: $\log(a_{break})$ on $\log(\rho)$, where $\rho$ is the local number density of
2339: stars, is modeled as $\log(a_{break})=k \, \log(\rho)+l$, where $k=-0.24\pm0.02$
2340: and $l=3.35\pm0.07$, or $a_{break} \propto \rho^{-1/4}$. {\em Bottom left:} The
2341: dependence of local number density, $\ln(\rho)$, of binaries ({\em dots}) and
2342: stars ({\em circles}) on the height above the Galactic plane, where the density
2343: of stars is normalized to match the density of binaries at 1 kpc.
2344: {\em Bottom right:} The fraction of binaries relative to the total number of
2345: stars as a function of the height above the Galactic plane. The arrow shows the
2346: predicted fraction of binaries in the $0.1< Z/kpc < 0.4$ bin, if the $a_{break}$
2347: value follows the $a_{break}\propto Z^{0.72}$ relation.
2348: \label{logZ_plots}}
2349: \end{figure}
2350: 
2351: \clearpage
2352: 
2353: \begin{figure}
2354: \epsscale{1.0}
2355: \plotone{f22_color.eps}
2356: \caption{
2357: The distribution of angular separation for the $0.1 < Z/kpc < 0.4$,
2358: $|\delta|<0.2$ kinematically-selected sample of candidate binaries. The data
2359: ({\em solid line}) extend to $\theta=500\arcsec$, though the plotted range is
2360: restricted for clarity. The distribution of random pairs ({\em dashed line})
2361: was obtained by fitting $f_{rnd}(\theta) = C \, \theta$ to the observed
2362: histogram for $\theta > 200\arcsec$. The sharp drop-off in the observed
2363: distribution for $\theta\la9$ is probably due to blending of close pairs in the
2364: POSS data.
2365: \label{theta_fit_500pc}}
2366: \end{figure}
2367: 
2368: \clearpage
2369: 
2370: \begin{figure}
2371: \epsscale{1.0}
2372: \plotone{f23_color.eps}
2373: \caption{A comparison of results for the turnover in the distribution of
2374: semi-major axes, $a_{break}$, as a function of distance modulus, of wide binary
2375: systems determined here (symbols with error bars; the horizontal bars mark the
2376: range of probed semi-major axes, and the vertical bars mark the width of the
2377: distance bins; the lowest point is only a lower limit, for the sake of
2378: comparison with other results we ignore the difference between distance from us
2379: and distance from the Galactic plane because our sample is dominated by high
2380: galactic latitude stars), determined by L\'{e}pine \& Bongiorno (2007; the
2381: dashed rectangle indicates  constraint on $a_{break}$ and the probed distance
2382: range), and determined by Chanam\'{e} \& Gould (2004; big arrows, indicating
2383: upper limits on $a_{break}$ and the probed distance range; the point at larger
2384: distance modulus corresponds to halo binaries). The diagonal dashed lines are
2385: lines of constant angular scale, $\theta$, for values of $3\arcsec$, $4\arcsec$,
2386: $5\arcsec$, $10\arcsec$, $20\arcsec$ and $30\arcsec$ (from left to right).
2387: \label{Fig:compare}}
2388: \end{figure}
2389: 
2390: \clearpage
2391: 
2392: \begin{figure}
2393: \epsscale{0.7}
2394: \plotone{f24.eps}
2395: \caption{The top left panel shows the proper motion distribution as a function
2396: of distance from the Galactic plane ($Z$) for a sample of $\sim$16,000 likely
2397: disk stars (red dots) and a sample of $\sim$34,400 likely halo stars (blue
2398: dots). All stars have $14<r<20$ and $0.2<g-r<0.4$, and are separated using
2399: photometric metallicity. The triangles show the median values in 500 pc wide $Z$
2400: bins for each sample (lower symbols: disk, upper symbols: halo). Note that the
2401: median proper motion for disk stars becomes constant beyond $Z\sim2$ kpc due to
2402: the vertical gradient of rotational velocity for disk stars. The top right panel
2403: shows the median position (symbols) and widths (lines; $\pm1\sigma$ envelope
2404: around the medians) of the reduced proper motion sequences for disk (red dots
2405: and dashed lines) and halo (blue squares and dot-dashed lines) stars, as
2406: functions of $Z$. The two bottom panels show the cross-sections of the reduced
2407: proper motion sequences for stars with $Z=1-1.5$ kpc (bottom left; red histogram
2408: for disk stars and blue for halo stars) and $Z=3.5-4$ kpc. The histograms are
2409: normalized by the total number of stars in each subsample. The disk-to-halo star
2410: count ratio is 4.3 in the left panel, and 0.38 in the right panel. Note the
2411: significant overlap of the two sequences for large $Z$.
2412: \label{Fig:App1}}
2413: \end{figure}
2414: 
2415: \clearpage
2416: 
2417: \begin{figure}
2418: \epsscale{0.8}
2419: \plotone{f25.eps}
2420: \caption{An illustration of the offsets in the position of reduced proper motion
2421: sequences as a function of distance, position on the sky and population. Each
2422: panel shows the median value of $5\log(v_t/v_t^{NGP})$, where $v_t$ is the
2423: heliocentric tangential velocity, and $v_t^{NGP}$ is its value at the north
2424: Galactic pole, in Lambert projection of northern galactic hemisphere. The maps
2425: are color-coded according to the legend shown in the middle of the figure
2426: (magnitudes), and are constructed using stars with $0.2<g-r<0.4$. Stars are 
2427: separated into halo and disk populations using photometric metallicity (for
2428: details see I08a). The top left panel shows results for halo stars with
2429: distances in the 3-4 kpc range. The other three panels correspond to disk stars
2430: in the distance range 3-4 kpc, 2-2.5 kpc, and 1-1.5 kpc.}
2431: \label{Fig:App2}
2432: \end{figure}
2433: 
2434: \clearpage
2435: 
2436: \begin{figure}
2437: \epsscale{0.7}
2438: \plotone{f26.eps}
2439: \caption{
2440: The number of unresolved binary systems (normalized with the total count in a
2441: given $g-i$ bin) with a magnitude offset $\Delta M_r=M_r(assumed)-M_r(true)$ as
2442: a function of the system's $g-i$ color. The assumed absolute magnitude for a
2443: system with a $g-i$ color, $M_r(assumed)$, was calculated using
2444: Eq.~\ref{Mr_bright} (Eq.~1 from J08), while the true absolute magnitude,
2445: $M_r(true)$ was calculated by adding up luminosities of components. The mean,
2446: median, and the rms scatter of $\Delta M_r$ are shown with the dotted, solid,
2447: and dashed lines, respectively.
2448: \label{dMr_gi}}
2449: \end{figure}
2450: 
2451: \clearpage
2452: 
2453: \begin{figure}
2454: \epsscale{0.7}
2455: \plotone{f27.eps}
2456: \caption{
2457: Distribution of differences between the magnitude offset, $\Delta M_r$, and the
2458: median magnitude offset, $\mu$, normalized with rms scatter, $\sigma$,
2459: ({\em solid line}) can be modeled as a 0.9 wide Gaussian ({\em dotted line}).
2460: \label{dMr_chi_gauss}}.
2461: \end{figure}
2462: 
2463: \clearpage
2464: 
2465: \begin{figure}
2466: \epsscale{1.0}
2467: \plotone{f28_color.eps}
2468: \caption{
2469: The distribution of $\delta$ values for the mock sample of wide binaries with
2470: both components redder than $g-i=2.0$ ({\em open circles}). In this sample, a
2471: star has a 40\% probability to be an unresolved binary system. Single
2472: star-single star configurations contribute the central narrow Gaussian
2473: ({\em dotted line}), unresolved binary-unresolved binary configurations
2474: contribute the central wide Gaussian ({\em thin solid line}), while the single
2475: star-unresolved binary configurations contribute the left and the right
2476: Gaussians ({\em dot-dashed lines}). The centers, widths, and areas of Gaussians
2477: are: $N_1(0.00,0.15,0.34)$, $N_2(0.06,0.35,0.28)$, $N_3(-0.64,0.18,0.18)$,
2478: $N_4(0.71,0.17,0.19)$ for the narrow, wide, left, and right Gaussians,
2479: respectively.
2480: \label{gauss_fit_red}}
2481: \end{figure}
2482: 
2483: \end{document}
2484: 
2485: First, we select pairs from the random
2486: sample with $0.7 < Z/kpc < 1.0$, and generate new $\theta$ values using
2487: \begin{equation}
2488: \theta = 30\arcsec \, \sqrt{\zeta}
2489: \end{equation}
2490: where $\zeta$ is a random number between 0 and 1. This effectively creates
2491: random pairs (their differential distribution of $\theta$ is linear with
2492: $\theta$). The size of the mock sample is restricted to be the same as of the
2493: $0.7 < Z/kpc < 1.0$ sample used above. For a fraction of pairs, semi-major axes
2494: values are generated using
2495: \begin{equation}
2496: a = a_{min} \, \exp(\zeta \, \ln(a_{break}/a_{min}))
2497: \end{equation}
2498: where $a_{min} = 100$ AU and $a_{break}=10,000$ AU, and $\theta=a/(1.411 d)$. We
2499: apply the $f(a)$ and $a_{break}$ analysis to the mock sample, and find 
2500: