<<

MNRAS 000,1–12 (2021) Preprint 25 August 2021 Compiled using MNRAS LATEX style file v3.0

Stellar kinematics of dwarf from multi-epoch spectroscopy: application to II ★

Rachel Buttry1†, Andrew B. Pace1, Sergey E. Koposov2,3,1, Matthew G. Walker1, Nelson Caldwell4, Evan N. Kirby5, Nicolas F. Martin6,7, Mario Mateo8, Edward W. Olszewski9, Else Starkenburg10, Carles Badenes11, and Christine Mazzola Daher11 1McWilliams Center for , Carnegie Mellon University, 5000 Forbes Ave, Pittsburgh, PA 15213, USA 2Institute for , University of Edinburgh, Royal Observatory, Blackford Hill, Edinburgh EH9 3HJ, UK 3Institute of Astronomy, University of Cambridge, Madingley Rd, Cambridge, CB3 0HA, UK 4Center for Astrophysics | Harvard & Smithsonian, 60 Garden Street, Cambridge, MA 02138, USA 5California Institute of Technology, 1200 E. California Blvd., MC 249-17, Pasadena, CA 91125, USA 6Observatoire astronomique de Strasbourg, Université de Strasbourg, CNRS, UMR 7550, 11 rue de l’Université, F-67000 Strasbourg, France 7Max-Planck-Institut für Astronomie, Königstuhl 17, D-69117 Heidelberg, Germany 8Department of Astronomy, University of Michigan, Ann Arbor, MI 48109, USA 9Steward Observatory, The University of Arizona, 933 N. Cherry Avenue, Tucson, AZ 85721, USA 10Kapteyn Astronomical Institute, University of Groningen,Postbus 800, 9700 AV, Groningen, the Netherlands 11Department of Physics and Astronomy and Pittsburgh Particle Physics, Astrophysics and Cosmology Center (PITT PACC), University of Pittsburgh, 3941 O‘Hara Street, Pittsburgh, PA 15260, USA

Accepted XXX. Received YYY; in original form ZZZ

ABSTRACT We present new MMT/Hectochelle spectroscopic measurements for 257 observed along the line of sight to the ultra-faint dwarf Triangulum II. Combining with results from previous Keck/DEIMOS spectroscopy, we obtain a sample that includes 16 likely members of Triangulum II, with up to 10 independent redshift measurements per . To this multi-epoch kinematic data set we apply methodology that we develop in order to infer binary orbital parameters from sparsely sampled curves with as few as two epochs. For a previously-identified (spatially unresolved) binary in Tri II, we +3.8 +0.41 − ± −1 infer an orbital solution with period 296.0−3.3 days , semi-major axis 1.12−0.24 AU, and a systemic velocity 380.0 1.7 km s that we then use in the analysis of Tri II’s internal kinematics. Despite this improvement in the modeling of , the current data remain insufficient to resolve the velocity dispersion of Triangulum II. We instead find a 95% confidence upper −1 limit of 휎푣 . 3.4 km s . Key words: (stars:) binaries: spectroscopic – galaxies: kinematics and dynamics

1 INTRODUCTION matter, as they probe the small-scale structure (< 1 Mpc) regime of ΛCDM cosmology (Bullock & Boylan-Kolchin 2017). Dwarf galaxies are of great importance for astrophysics. From a galaxy formation perspective, dwarf galaxies are among the oldest In order to place dwarf galaxies into their proper cosmological con-

arXiv:2108.10867v1 [astro-ph.GA] 24 Aug 2021 and least chemically evolved objects (Mateo 1998; Tolstoy et al. 2009; text, we must obtain accurate estimates of their dark matter content. McConnachie 2012). From a dark matter perspective, they include The simplest dynamical mass estimators, based on the assumption the most dark matter dominated systems known, with published dy- of dynamic equilibrium, are functions of the effective radius and namical mass-to-light ratios reaching as high as 104 in solar units line-of-sight velocity dispersion measured for the stellar component (Simon 2019, and references therein). In this vein, dwarf galaxies (e.g., Illingworth 1976; Walker et al. 2009; Wolf et al. 2010; Errani are believed to be key components in unpacking the mystery of dark et al. 2018). Yet measurements of the stellar velocity dispersion can be challenging. One reason is the small number and low of stellar tracers in especially the ‘ultrafaint’ dwarf galaxies. Another ★ New observations reported here were obtained at the MMT Observatory, challenge is the existence of unresolved binary stars, whose orbital a joint facility of the Smithsonian Institution and the University of Arizona. motions add a time-dependent component to the velocities measured Some of the data presented herein were obtained at the W. M. Keck Obser- for individual stars, and—if unaccounted for—can thereby inflate vatory, which is operated as a scientific partnership among the California Institute of Technology, the University of California and the National Aero- measurements of velocity dispersions. Binary orbital nautics and Space Administration. The Observatory was made possible by motions alone can generate apparent velocity dispersions of a few km −1 the generous financial support of the W. M. Keck Foundation. s (McConnachie & Côté 2010). While this effect is negligible for † E-mail: [email protected] (RB) the more luminous dwarf spheroidals, which have intrinsic velocity

© 2021 The Authors 2 R. Buttry et al. dispersions of ∼ 10 km s−1 (Olszewski et al. 1996), it can potentially as the marginalization over inclination rather than semi-amplitude < −1 contribute significantly to the ∼ 3 km s dispersions observed for (allowing for the calculation of semi-major axis), the ability to take the least luminous galaxies (McConnachie & Côté 2010; Minor et al. non-trivial priors over binary parameters, and using parameter sam- 2010). ples for hierarchical models of binary populations. Various strategies have been used to account for the effect of binary In the next section we discuss the MMT and Keck catalogues used stars on velocity dispersions and the dynamical masses derived there- in this analysis as well as the calculation of a zero-point correction from. When multi-epoch spectroscopy is available, one can identify between the two instruments. In section3 we present our methodol- probable binary systems via their observed accelerations (e.g., Ol- ogy for modeling of binary and non-binary star systems as well as the szewski et al. 1996; Martinez et al. 2011; Koposov et al. 2011; Minor galaxy kinematics. We then detail the resulting orbital parameter for et al. 2019); indeed modeling of multi-epoch spectroscopic data sets the Tri II and the overall Tri II kinematics in section for luminous dwarf spheroidals suggests typical binary fractions near 4. Lastly, we discuss the findings from our work in section5 and we ∼ 50% (Minor 2013, Spencer et al. 2017, Spencer et al. 2018), con- suggest a hierarchical model building off the methods used. sistent with studies of Galactic binaries that indicate relatively high multiplicity fractions at low (Badenes et al. 2018). In some cases the removal of suspected binary stars has significant im- 2 DATA pact on the measured velocity dispersion (e.g., Kirby et al. 2017; Venn et al. 2017). 2.1 MMT Hectochelle The Triangulum II (Tri II) ultra-faint dwarf galaxy provides an During Dec. 2015 and Oct. - Nov. 2016, we acquired new spectra of interesting case study. The original kinematic study of Tri II, based on stars in Tri II using the Hectochelle spectrograph (Szentgyorgyi et al. single-epoch spectroscopy of six member stars, measured a velocity 2011) at the 6.5-m MMT Observatory on Mt. Hopkins, Arizona. . +4.0 −1 dispersion of 5 1−1.4 km s , suggesting a dynamical mass-to-light Hectochelle deploys up to 240 optical fibers, each with aperture 1.5 +3500 ◦ ratio of 3600−2100 in solar units, and an extremely high dark matter arcsec, over a field of diameter 1 . We observed using the ‘RV31’ . +8.1 / 3 filter, isolating the wavelength range of 5150 − 5300 at resolution density of 4 8−3.5 M pc (Kirby et al. 2015). An independent study by Martin et al.(2016) obtained a spectroscopic sample of R ≈ 34, 000. 13 member stars, finding complicated kinematics in which a central We observed five different Hectochelle fiber configurations, each +2.8 −1 centered on the published center of Tri II, allowing us to observe velocity dispersion of 휎푣 = 4.4− km s gives way to a larger value 2.0 ∼ . +5.8 −1 up to 500 unique targets, with many stars included in multiple of 14 1−4.2 km s at large radius. Both studies found evidence for nonzero metallicity dispersion, supporting the conclusion that Tri II targeting configurations. is a dwarf galaxy embedded in a massive dark matter halo, and not a Targets were selected using metallicity-sensitive CaHK narrow- self-gravitating . band photometry to prioritize low metallicity stars that are more likely to be members of the very metal-poor Tri II. We obtained a However, follow-up spectroscopy soon provided a time domain series of deep broadband 푔 and 푖 observations with the ∼ 1-deg2 and revealed the presence of at least one star with significant velocity imager MegaCam on the Canada France Hawaii Telescope for total variability. From high-resolution spectra obtained primarily to an- integration times of 4,050 s and 3,150 s, respectively, along with alyze chemical abundances, Venn et al.(2017) measured a change CaHK images totaling 4,800 s. All were observed in service mode in velocity for one star (Star46) of ∼ 25 km s−1 with respect to the by the CFHT observing staff between 2015-07-18 and 2016-02-13. initial epoch measured by Martin et al.(2016). Kirby et al.(2017) Images were then processed with a version of the CASU pipeline added additional epochs for this star, independently confirming its (Irwin & Lewis 2001) tailored to the processing of MegaCam images velocity variability and finding that, when they excluded the likely (e.g., McConnachie et al. 2018). binary from their analysis, the velocity dispersion was unresolved. To isolate metal-poor point sources extracted from the resulting These circumstances leave the case for a dominant dark matter halo stacks, we follow the same strategy put in place for the Pristine in Tri II resting on the indirect argument provided by its metallicity survey (Starkenburg et al. 2017). Most stars in the field are metal- spread (Venn et al. 2017). rich and therefore follow a well-defined stellar locus whereas metal- Here we add to the saga of Tri II in two ways. First, we present poor stars deviate from this locus. In the case of Tri II, members new spectroscopic data acquired with the Hectochelle spectrograph at previously confirmed via spectroscopy are clearly offset by 0.2–0.5 the 6.5-m MMT. Second, we combine with the previously-published magnitudes from the locus of contaminating Milky Way stars. Given spectroscopic data in order to obtain a multi-epoch data set that the large number of available Hectochelle fibers we were able to then lets us model the orbital parameters of the likely binary star. prioritize metal-poor stars and remove obviously metal-rich stars, but Our orbital solution includes an inference for the binary system’s we remained generous in our selection. This strategy worked well for center-of-mass motion, allowing us properly to include this star in previous observing campaigns (Longeard et al. 2020) to ensure that our analysis of Tri II’s . we did not miss member stars with potentially higher . To date, only one within a dwarf spheroidal galaxy has In addition to known Tri II members, this selection includes potential a full orbital solution, based on 34 independent velocity measure- members out to ∼ 0.5 deg from the center of the dwarf galaxy. ments taken over a two-year baseline (Koch et al. 2014). Here we We processed all raw Hectochelle spectra using the CfA pipeline develop methodology for inferring orbital solutions with as few as (HSRED v2.11). Following the procedure described in detail by Walker two velocity epochs. The problem of finding orbital parameters for et al.(2015), we then analyzed each individual spectrum by fitting a binary system given a small number of radial velocity (RV) mea- a model based on a library of synthetic template spectra that span surements has been undertaken previously by Price-Whelan et al. a regular grid in effective temperature, surface and [Fe/H] (2017) to create the JOKER. Like the JOKER, the binary model metallicity. In addition to the stellar-atmospheric parameters, we we present in this paper takes the approach of performing rejection sampling with likelihood function marginalized over some orbital parameters. However, our method has the added modifications such 1 https://bitbucket.org/saotdc/hsred/

MNRAS 000,1–12 (2021) Triangulum II 3

fit for line-of-sight velocity as well as several free parameters that 2.3 Zero-point offset specify the continuum shape and correct for wavelength-dependent In this section, we describe the calculation of the zero-point velocity velocity shifts. offset between Keck and MMT. Though our complete set of Tri II With respect to the procedure documented by Walker et al.(2015), Keck observations is a concatenation of the Kirby et al.(2017) and for present purposes we update our estimation of systematic er- Martin et al.(2016) datasets, the zero-point offset between the two rors associated with line-of-sight velocity and metallicity. For this appear consistent with zero (see Section 3.2 of Kirby et al. 2017). task we use our entire catalog of MMT/Hectochelle observations of For this analysis, we have defined the offset as the MMT zero-point dwarf galaxies and globular clusters, including observations span- minus the Keck zero-point. There are 8 objects in our dataset around ning the years 2005 to 2020. This sample includes 12517 indepen- the Tri II galactic center with radial velocity measurements in both dent observations of 7906 unique stars, including 2501 stars with MMT and Keck catalogues. The measurements for a given object up to 13 individual measurements. We model the pair-wise veloc- and instrument are combined into a single weighted mean value. We ity and metallicity differences as a mixture of a Gaussian with an then calculate the offset between the sets of combined instrument outlier model (see Section 4.1 of Li et al. 2019 and Section 2.2 velocities while assuming Gaussian errors. of Pace et al. 2021).The final uncertainty (휎 ) is treated as a 푣,calib The methods described here and in section3 are based on a Gaus- systematic error (휎 ) plus a scaling parameter (푘 ) and 푣, systematic 푣 sian likelihood function, which is described in terms of model pre- 휎2 = 휎2 + (푘 휎 )2. We find 푘 = 1.03 ± 0.02 푣,calib 푣, systematic 푣 푣,mcmc 푣 diction 휇, model error 휎, given data in the form of the observed −1 and 휎푣, systematic = 0.35 ± 0.02 km s for the velocity systematic velocity, 푣: errors, and 푘 [Fe/H] = 1.33±0.01 and 휎[Fe/H], systematic = 0.0±0.01 for metallicity systematic errors. ! 1 (푣 − 휇)2 G(푣|휇, 휎) = exp − (1) √︃ 2 2 2 2 2(푣error + 휎 ) 2휋(푣error + 휎 ) This is the general likelihood for a given velocity prediction and the total likelihood function is the product of the velocity likelihoods. 2.2 Keck DEIMOS In the context of finding a zero-point offset, we apply this likeli- hood to a set of velocity differences {푣푖 }, where 푣푖 represents the Our sample also includes spectroscopy obtained with the Deep Extra- difference between the 푖th star’s MMT weighted mean velocity and galactic Imaging Multi-Object Spectrograph (DEIMOS; Faber et al. its Keck weighted mean velocity, 푣푖 = 푣¯푖,MMT − 푣¯푖,Keck. We assume 2003). First, Kirby et al.(2015); Kirby et al.(2017) observed six that these velocity differences are consistent with the zero-point off- slitmasks in 2015 and 2016. They used the 1200G grating, which set between instruments 훿푣 and that there is no dispersion in 훿푣 , ∼ achieves a spectral resolution of 푅 7000 at 8500 Å, in the spectral only observational errors. Thus the total likelihood is the product of vicinity of the Ca ii triplet. Second, Martin et al.(2016) ob- Î푁 G(푣 , 훿 , ) individual object likelihoods, 푖=1 푖 푣 0 . served two DEIMOS slitmasks with a similar spectral configuration However, some objects have a very large velocity difference be- as Kirby et al. tween the instruments, possibly due to low signal-to-noise or un- Kirby et al.(2015) selected stars using Keck/LRIS (Oke et al. confirmed binarity. A convenient way to account for these objects 1995) photometry. They chose a generous selection region in the is to construct a mixture model to treat them as outliers, a similar CMD around the branch as defined by the ridgeline of the model is used in Li et al.(2019). The new likelihood for a given M92. Martin et al.(2016) used a similar selection object is the sum original Gaussian likelihood and outlier model technique with photometry (Laevens et al. 2015) from the Large likelihood weighted by the probability the object is an outlier. The Binocular Camera. In general, the field of Tri II is sparse enough new likelihood is written as... that most candidate member stars in the field of the slitmask could 푁 be observed. As a result, the samples have little selection bias due to Ö 퐿({푣 }|훿 , 훾, 푝) = [훾G(푣 |0, 푝) + (1 − 훾)G(푣 |훿 , 0)] (2) color (or stellar age or metallicity). Kirby et al.(2017) were mainly 푖 푣 푖 푖 푣 푖=1 interested in quantifying radial velocity variability, not in finding new members. As a result, they designed their slitmasks to target where 훾 represents the outlier fraction and 훿푣 represents the offset stars already identified as members by Kirby et al.(2015) and Martin correction. The outlier model is applied to the difference in the com- et al.(2016). bined weighted mean velocities for each instrument and is taken to be The velocities were measured in slightly different ways. Kirby a Gaussian with a large standard deviation. The priors are as follows: et al.(2015); Kirby et al.(2017) reduced the spectra with custom −1 −1 • offset 훿푣 : uniform(−10 km s , 10 km s ) modifications to the spec2d data reduction pipeline (Cooper et al. • outlier fraction 훾: uniform(0, 0.5) 2012). They matched empirical spectral templates observed with • outlier model standard deviation 푝: uniform(4 km s−1, DEIMOS to the Tri II spectra and varied the velocity until 휒2 was 20 km s−1) minimized. Martin et al.(2016) used their own custom pipeline (Ibata et al. 2011) to reduce the spectra. They determined radial velocities We define the posterior as the Gaussian likelihood multiplied by from the mean wavelengths of Gaussian fits to the Ca ii triplet. our set of priors and look for a 훿푣 that maximizes this posterior. Slit imaging spectrographs can experience radial velocity zero- The error on the offset is taken as the width of this offset pos- point shifts if the star is not perfectly centered in the slit. This effect terior distribution. We derive the offset value from the resulting can be mitigated by observing the wavelengths of telluric absorption posterior, sampled using Markov Chain Monte Carlo (MCMC) by lines (e.g., Sohn et al. 2007). The slit centering correction is taken to the Metropolis–Hastings algorithm via the emcee python package be the deviation of the wavelengths of these lines from the geocen- (Foreman-Mackey et al. 2013). Applying this procedure to the 8 stars tric rest frame. All of the studies used in this work performed such a in the overlapping MMT-Keck datasets observed around the Tri II −1 correction. galactic center, we find an offset value of 훿푣 = −0.11 ± 1.02 km s .

MNRAS 000,1–12 (2021) 4 R. Buttry et al.

While this value could be determined using only our Tri II dataset, 60 −1 2 the large error means that the offset is not resolved within 1 km s 40 and the introduction of a large offset error can make it more difficult 20 to resolve the Tri II velocity dispersion. 1 To improve the error on our zero-point offset between the in- 150 struments, we also include Keck/DEIMOS (Pace et al. 2020) and Members 0 MMT/Hectochelle (Spencer et al. 2018) observations from an addi- star 31 10 tional 288 Ursa Minor objects. The combined Tri II and Ursa Minor star 46 (binary) −1 Tri II 1 dataset results in an offset correction of 훿푣 = −1.33 ± 0.33km s . 1 0 1 2 5

We opt to use the offset that is calculated while including Ursa ] 1

Minor observations, as the larger amount of data gives a more precise r value for the offset. Thus, we bring all observations onto a common y 0 s zero-point by subtracting a fixed amount of 1.33 ± 0.33 km s−1 from a

m 5 [ each velocity measurement obtained with Keck/DEIMOS. We do not propagate the error on the offset as the total error is dominated by the combination of instrument-specific systematics and random errors. 10

15 2.4 Membership 20 . 휎 10 5 0 5 10 15 200 50 To determine Tri II membership, we sigma clip our dataset at 3 5 1 from both the galaxy mean line-of-sight velocity and . * [mas yr ] This step removes objects whose observations are differ from the Figure 1. Proper motion distribution of stars within 10 0(∼ 4 half-light radii) Tri II measurements > 3.5 times the root sum of the squared of the √︃ of Tri II center from Gaia eDR3 after a parallax cut (blue). The green point 휎 = 휎2 + 휎2 galaxy and measurement errors, Tri II obs. We use proper marks the Tri II proper motion value that is used to determine membership. motions from Gaia eDR3 (Gaia Collaboration et al. 2021, Figure1). The black triangle and star are the two stars with uncertain membership in All members are consistent with zero parallax (i.e, 휛 − 3휎휛 < 0). Kirby et al.(2017). The red dots are the remaining members that have Gaia The Tri II mean velocity is taken as −381.7 km s−1 (Kirby et al. 2017) proper motion measurements. −1 and the galaxy proper motion as 휇훼★ = 0.58 ± 0.06 mas yr , 휇 훿 = −1 0.11 ± 0.07 mas yr (Pace et al. in prep ). While the velocity 40 −1 dispersion is unresolved, we use 휎Tri II = 4 km s (Figure2). Our resulting sample consists of 16 member stars, one of which is 20 the previously confirmed binary star (which we refer to as Star46 or MIC2016-46 in this work). There is one star that is observed by MMT 10 only, two stars in the overlap of MMT and Keck, and the remaining star 46(binary star) 13 are observed by Keck only. star 31 Figure3 shows the location of our final sample of member stars 0 Tri II MMT obs on a color- diagram as compared to all the stars within Keck mem obs the Tri II half light radius after a star-galaxy separation cut observed 1 MMT mem obs by the Subaru/Hyper Suprime-Cam (HSC; Miyazaki et al. 2018) and ]

Pan-STARRS1 (Flewelling et al. 2020), the isochrone parameters are H / 2 taken from Carlin et al.(2017). In Figures1,2, and3, we indicate the e F locations of the binary star (Star46) and another member, referred [ 1.0 to as Star31 (or MIC2016-31) in this analysis, with black markers 3 1.5 of different shapes. Star31 is specifically marked because of its un- Star31 2.0 certainty as a member in Kirby et al.(2017) (also called in 4 their work). Its membership is brought into question due to it being 2.5

(1) far from the the galactic center, (2) the most metal rich star in the 3.0 5 sample (to the point of driving Tri II’s [Fe/H] dispersion), and (3) 400 300 200 100 0 1000 20 3.5 1 390 385 380 375 370 the only star whose velocity is > 1휎 from the galaxy mean velocity. vlos [km s ] However, the now available Gaia eDR3 indicate its proper motion is consistent with Tri II, bolstering its case as a member.

Figure 2. Mean metallicities and velocities of all stars observed in MMT 3 METHODS around the Tri II field (black). The green area represents the The Tri II mean velocity from Kirby et al.(2017) used in membership cuts. Our final sample 3.1 Binary Stars of Tri II members is plotted with two colors/markers to distinguish between Keck (blue circles) and MMT (red squares) measurements. The binary star, When a star is a part of a binary system, the orbital motion can give alone, is plotted with the systemic velocity found in this work rather than the rise to periodic variability in the line-of-sight velocity. If one star is weighted mean of the velocity observations. much brighter than its companion (e.g., a red giant with a main se- quence companion) then a single-epoch spectrum may not show any

MNRAS 000,1–12 (2021) Triangulum II 5

10 FeH=-2.0, Age=13.0Gyr star 46 (binary star) star 46(binary star) star 31 star 31 36.4 12 Keck members Keck members MMT members

MMT members rh HSC 2r 14 h panstarrs MMT obs 36.3

16 0

i 18 36.2 Dec [deg] 20

36.1 22

24 36.0

26 0.0 0.2 0.4 0.6 0.8 1.0 1.2 33.6 33.5 33.4 33.3 33.2 33.1 g0 i0 RA [deg]

0 Figure 3. Left: Color-magnitude diagram of Tri II stars within ∼ 4 half-light radii (푟ℎ =2.5 ) taken by Hyper Suprime-Cam (HSC) and Panstarrs with a MIST isochrone overlaid. The location of the Tri II binary star and Star31 is indicated by a star and triangle, respectively. The isochrone is defined by the following values: 푚 − 푀 = 17.44, [Fe/H]=−2.2, and Age=13 Gyr. Right: Sky positions of members and stars observed with MMT. The dotted lines represents the 1 and 2 half-light radii of Triangulum II. In both plots, the members are plotted with either a red square or blue circle for MMT and Keck observations respectively.

evidence of variability. Conventional methods of combining obser- to a constant called the time of periastron passage, 푇, by a factor of vations, such as a weighted mean, can misrepresent the true systemic 2휋/푃. We can relate 휈 and Φ0 using the eccentricity anomaly, 휁. velocity if the system is sparsely observed and/or result in inflated error in the combined velocity. Thus, binary stars can directly in- √︂  1  1 + 푒  1  flate the calculated velocity dispersion if not accounted for properly. tan 휈 = tan 휁 (5) − 푒 Though the true line-of-sight velocity will vary with time, it will 2 1 2 oscillate around the center-of-mass (systemic) velocity. 휈 2휋 2휋 In terms of true anomaly (angular position in orbital plane from (푡 − 푇) = 푡 + Φ = 휁 − 푒 sin 휁 (6) periastron direction), argument of periapsis 휔 (angle in orbital plane 푃 푃 0 between ascending node and periastron) and systemic velocity 푣0, a We furthermore include a jitter parameter that acts as secondary member of a binary stars system has line-of-sight velocity that varies error to account for velocity variability not due to periodic orbital with time according to... motion. We note that Hekker et al.(2008) found a correlation between velocity jitter and surface gravity for red-giant branch stars, where stars with log 푔 . 1 can have excess jitter ∼ 1 km s−1. However, the Star46 푉 = 퐾(cos(휔 + 휈) + 푒 cos 휔) + 푣0 (3) majority of our members, including , exist on the lower red giant branch where log 푔 & 1. This is commonly the case for stars in ultra-faint dwarfs, so this is little concern that internal mechanisms (Murray & Correia 2010). We can further expand the semi-amplitude will inflate the galaxy velocity dispersion. 퐾 in terms of the binary parameters period 푃, semi-major axis 푎, We continue to assume a Gaussian error on velocity measurements, inclination sin 푖, and eccentricity 푒 as thus allowing us to use the previously defined Gaussian likelihood function, equation1. We generate samples of the posterior on the orbital binary parameters using rejection sampling. This sampling 2휋 푎 sin 푖 5 퐾 = √ . (4) is done over the course of 10 iterations with each iteration starting 푃 1 − 푒2 with 106 points before rejection. After rejection, the surviving sam- ples across all iterations are combined. The choice to use rejection This is the model that we use to fit to the observed radial velocity sampling instead of Monte-Carlo-Markov-Chain (MCMC) is due to curves. It is worth mentioning that instead of using 휈 as a free pa- the potential multi-modality posterior. A sparsely observed binary rameter, we sample in Φ0, representing the phase, which is related can have multiple possible orbital period solutions, which have cor-

MNRAS 000,1–12 (2021) 6 R. Buttry et al. responding total mass and systemic velocity values, represented as 331086418824762240 \ KCS2015-76 \ MIC2016-23

N(-386.96,3.49) peaks in the posterior. This specific scenario can lead to untouched 382 0.150 N(-386.95,6.30) 0.125 Gaussian Fit areas of the posterior when using MCMC. ] 384 1 Binary Model

The priors are quite uninformative, ensuring that we can fully s 386 0.100 m

k 388 0.075

sample all the possible binary parameters. Note that there is a pre- [

s o l 390 0.050 defined mass range, such that the prior in total mass is loguniform v posterior PDF 392 from log(0.1 M ) to log(10 M ), which is used to bound the semi- 0.025 Keck 394 0.000 major axis prior. The priors on all the angles are uniform. We adopt 2015-09-18 2016-01-14 2016-05-12 2016-09-07 450 425 400 375 350 325 300 Time [YYYY-MM-DD] 1 the uniform prior on cos(푖) corresponding to random orientation of v0 [km s ] binary and close to uniform between 0 and 1 beta distribution of eccentricities, Beta(1.5, 1.5). For compactness, we explicitly state Figure 4. Left: RV curve of an example member star with more than one the priors on each of the orbital parameters alongside the resulting observation. Right: Corresponding posterior on the mean velocity resulting Star46 binary parameter posteriors in Table1. from both the Gaussian fit and binary model. The binary model posterior We are able to make further assumptions about the system by samples have had both modifications that remove short period solutions to reduce the posterior scatter. modifying the surviving posterior samples. We remove nonphysical solutions corresponding to close binaries where the pericenter is less than the stellar radius or where the mass of the companion previously defined Gaussian likelihood function, equation1, and a star is negative. These stellar radii and masses are calculated using set of priors in true velocity and velocity error to sample from the true the Isochrones python package (Morton 2015) to determine stellar velocity posterior. The sampling is done using MCMC sampling and parameters for the binary star of age 13 Gyr, [Fe/H] of -2.2, and is performed with 24 walkers, 1000 steps each, and 100 step burn- distance modulus of 17.24 using the MESA Isochrones and Stellar in. For a set of line-of-sight velocity observations {푣푖 } with errors Tracks (MIST) grid (Dotter 2016; Choi et al. 2016). Figure3 shows {휖푖 }, we define a uniform prior in true velocity that is non-zero from the relevant isochrone fit on a color-magnitude diagram with Tri II min({푣푖 }) − 4 max({휖푖 }) to max({푣푖 }) + 4 max({휖푖 }). We also use members plotted. For the binary star Star46, we determine a mass Jeffrey’s prior for a the standard deviation of a Gaussian distribution of 푀bin = 0.777 M and stellar radius of 푟bin = 0.027 AU from an (1/휎 or uniform in log 휎) for the true velocity error. The error on the isochrone fit to Tri II stars. We equivalently derive a mass and radius combined velocity measurement is taken as one standard deviation for the other Tri II members by finding the point on the isochrone from the posterior on the mean. We use the mixture sub package found closest to each member star using the same steps. in scikit-learn (Pedregosa et al. 2011) to fit a Gaussian mixture model Similarly, Raghavan et al.(2010) looked at 259 confirmed solar- (GMM) of a singular Gaussian to the resulting posterior samples. type binaries in the Hipparcos catalog and found the periods fit are Figure4 shows the radial velocity curve and resulting posteriors distributed according to a log-normal distribution with 휇log 푃 = 5.03 using this method, referred to as the Gaussian fit method, and binary and standard deviation of 휎log 푃 = 2.28, where the 푃 is in days. This model method for a star with multiple velocity observations. is a more informative prior on the period than our default period This Gaussian Fit method has the advantage in that it does not prior. We refer to this lognormal distribution as the "Raghavan prior" assume binarity, but does retain increased error with larger variability for the remainder of this paper. We present this distribution with in the RV curve, more than the weighted mean. For comparison, we the minor caveat that most of the detectable member stars of dwarf also calculate the combined velocities under the weighted mean and galaxies are not solar-type () stars, but rather red giant binary model methods (Figure5). Applying the binary model to branch stars. the, assumed non-binary, member stars takes up significantly more We are able to infer the posterior under this prior by re-sampling computation time and results in a larger uncertainty in the combined from the current surviving samples with non-uniform weights. These velocity. Over short time-scales, the observations can be consistent weights are proportional to the ratio of the Raghavan period distribu- with either a constant velocity model or the RV curve of a long period tion to the default prior. All of these assumptions shift the posterior binary. away from shorter orbital periods, but as we’ll see in section 4.2, the application of this prior on Star46 significantly alters the parameter posterior. As such, we present the results both before and after the 3.3 Mean Velocity and Velocity Dispersion application of the Raghavan prior. The Tri II velocity dispersion can be inferred from the posterior us- ing the same Gaussian likelihood function as in previous sections 3.2 Non-Binary Stars (Equation1). We define another Gaussian model in terms of the mean velocity 푣¯ and velocity dispersion 휎푣 of the galaxy. The to- In the case of non-binary stars, we expect line-of-sight velocity ob- tal likelihood is the product of the likelihood of the model on the servations to be consistent with a constant velocity model with some stellar velocity of each member. For stars with multiple velocity scatter in each measurement proportional to the observational error. measurements, the measured velocities are combined using one of Our goal is to determine a star’s true velocity by combining multi- the methods presented in the previous sections. In the case of a multi- ple epochs of radial velocity measurements into one stellar velocity. modal posterior, such as for Star46, we must fit a one-dimensional A common approach to this is calculating the weighted mean—i.e., GMM to the systemic velocity posterior and the likelihood for that summing the measurements weighted by the inverse of the squared object becomes... errors. This approach has the benefits of being analytic and thus not 푛 computationally intensive. However, it has the downside of under- ∑︁ 퐿 = 푤 G(휇 |푣,¯ 휎 ) (7) estimating the error when dealing with potentially variable systems. 푗 푗 푣 푗 In this section we offer another method to calculate the true stellar velocity by finding its posterior distribution. where 휇 푗 and 푤 푗 represent the mean and weight of the 푗th Gaussian Once again, the assumption of Gaussian error allows us to use the in the mixture model with 푛 Gaussians. The error on 휇 푗 is taken

MNRAS 000,1–12 (2021) Triangulum II 7

Binary Star Orbital Parameters

Parameter Prior Short Per. Long Per.

푃 푃 10 . +2.8 . +3.8 Period linearly decreasing in log( ) from 1 day to 10 days 148 1−1.4 days 296 1−3.3 days

푒 . +0.10 . +0.06 Eccentricity Beta(1.5,1.5) 0 29−0.07 0 50−0.04

푣 −1 − . +0.8 −1 − . +1.8 −1 Systemic velocity 0 Normal(0,474 km s ) 387 6−1.2 km s 380 0−1.7 km s

Φ 휋 . +2.21 . +2.04 Phase 0 uniform(0,2 ) 3 12−2.10 3 25−2.23

휔 휋 . +0.35 . +0.28 Argument of periapsis uniform(0,2 ) 4 04−0.65 3 43−0.24

−1, −1) . +1.4 −1 . +1.1 −1 Jitter Normal(0 k ms 2 k ms 1 3−0.9 km s 0 9−0.6 km s

푎 . +0.26 . +0.41 Semi-major axis loguniform, bounds determined from mass 0 71−0.15 AU 1 12−0.24 AU

푖 휋 푖 . +0.08 . +0.11 Inclination sin uniform(0,2 ) in cos 0 31−0.09 0 39−0.11

Table 1. Priors used in binary modeling and resulting orbital parameters for the Tri II binary star. The posterior is separated into short and long period solutions at the 0.6 years (219.2 days) boundary.

331089343699707776 \ KCS2015-128 331088381627372800 \ MIC2016-31 MIC2016-29 374 382 385 376 v = -384.6, 1.0 v = -390.4, 6.8 384 B = -383.2, 5.8 390 v = -378.2, 2.0 B = -387.6, 25.4 = -384.7, 2.1 378 B = -376.7, 5.0 = -392.0, 13.9 = 0.992 = -378.3, 2.5 395 = 5.458 386 380 = 1.836 400 388 382 405 384 2015-10-07 2016-01-27 2016-05-18 2016-09-07 2015-09-18 2016-01-14 2016-05-12 2016-09-07 2015-09-18 2015-10-17 2015-11-15 2015-12-14

MIC2016-27 MIC2016-24 331086418824762240 \ KCS2015-76 \ MIC2016-23

370 382.5 360

380 385.0 v = -385.6, 15.8 v = -387.1, 3.5 B = -383.8, 31.1 370 B = -385.6, 6.4 390 v = -383.5, 4.7 387.5 = -387.2, 3.7 = -387.6, 16.1 B = -378.6, 26.9 = 11.973 = 2.972 = -382.2, 24.2 390.0 380 400 = 6.338 392.5 410 390 2015-09-18 2015-11-01 2015-12-15 2016-01-29 2015-09-18 2015-11-01 2015-12-15 2016-01-29 2015-09-18 2016-01-14 2016-05-12 2016-09-07

329588128075263872 \ MIC2016-22 331089378059445504 \ KCS2015-116 \ MIC2016-21 331089545560979072 \ KCS2015-91 \ MIC2016-20 375.0 360 377.5 377.5 380.0 370 v = -383.2, 3.6 380.0 v = -381.2, 4.0 B = -381.7, 6.0 v = -383.0, 2.6 382.5 B = -379.9, 5.4 380 = -383.3, 3.0 382.5 B = -381.4, 4.6 = -381.9, 2.9 = 3.623 = -382.9, 1.8 385.0 = 4.110 385.0 = 2.574 390 387.5 387.5 390.0 400 390.0 392.5 2015-09-18 2016-01-14 2016-05-12 2016-09-07 2015-09-18 2016-01-14 2016-05-12 2016-09-07 2015-09-18 2016-01-14 2016-05-12 2016-09-07

MIC2016-9 331085117451894528 331089446778920576 \ KCS2015-106 \ MIC2016-40 380 378 382 385 v = -382.1, 0.5 B = -382.0, 3.1 383 380 390 = -382.2, 0.3 = 0.836 v = -400.8, 11.1 384 395 B = -394.8, 31.7 v = -386.0, 1.3 = -398.3, 17.1 385 B = -386.0, 7.8 382 400 = 8.557 = -385.5, 1.6 386 = 1.313 405 384 387 410

2015-09-18 2015-11-01 2015-12-15 2016-01-29 2016-10-21 2016-10-30 2016-11-09 2016-11-18 2015-09-18 2016-02-08 2016-06-29 2016-11-19

Figure 5. Radial velocity curves of non-binary member stars with multiple epochs. The different values are results from different methods of combining observations. 푣¯: weighted mean; 퐵: fitting a Gaussian distribution to the binary model 푣0 posterior; G: assuming the observations are taken from a Gaussian distribution centered on a true systemic velocity. For comparison, the standard deviation of the observations is also listed (휎). The stars are identified by their Gaia ID when applicable. All values listed and on the y-axis are in km s−1. The dates are in the format Year-Month-Day.

MNRAS 000,1–12 (2021) 8 R. Buttry et al.

Raghavan prior: False N(-379.71,2.10), w=0.70 0.20 N(-387.71,2.46), w=0.30 0.15 1.0 0.10 0.8

e 0.6

0.4 0.05 0.2 posterior PDF 0.00 400 395 390 385 380 375 370 2.0 1 1.6 v0[km s ] 1.2 a[AU] 0.8 Raghavan prior: True 0.4 N(-379.98,1.38), w=1.00 1.0 0.4 0.8 0.6 sini 0.3 0.4 0.2 0.2 ] 1 78.0 0.1 y 79.5 U posterior PDF A

[ 81.0 0

0.0 v 400 395 390 385 380 375 370 82.5

0.2 0.4 0.6 0.8 1.0 0.4 0.8 1.2 1.6 2.0 0.2 0.4 0.6 0.8 1.0 1 0.45 0.60 0.75 0.90 82.5 81.0 79.5 78.0 v0[km s ] 1 P [y] e a[AU] sini v0[AU y ]

Figure 6. The posterior on the systemic velocity of the binary star and the Figure 7. Orbital parameters’ posterior for the binary star derived from our Gaussian mixture model fit under different conditions. These conditions are a binary model. There are two distinct modes corresponding to re-weight of the samples such that the period prior is the Raghavan distribution and removing non-physical samples. to less overlap in number of stars. Second, we have explored including them and found that they do not impact the resulting orbital parameter as as the standard deviation of the Gaussian 휎푗 and the number of posterior. Gaussians in the GMM is determined by manually inspecting the After an initial pass with our model on Star46, we find that there 푣0 posterior. Figure6 shows the GMM fits to the Star46 systemic are two noticeable peaks in the posterior corresponding to periods of velocity posterior before and after the application of the Raghavan ∼ 0.4 and ∼ 0.8 years. We are able to eliminate the sub-harmonics prior. (e.g., 0.2, 0.1 years) from our restrictions on pericenter and Kepler’s The velocity dispersion prior is log-uniform in from third law both limiting the allowed semi-major axis values. Because ( −1) ( −1) log10 0.05km s to log10 100km s , which is equivalent to Jef- the period modes are distinct, we re-run the sampling with a more frey’s prior. The minimum value corresponds to the case where the restricted period prior bound at 0.3 and 0.95 years to ensure a larger galactic dynamics are fully determined by visible matter (See Ap- number of surviving posterior samples. The resulting binary orbital pendixA). We have chosen to find the posterior of log10 휎푣 rather parameters presented in this section are derived from the posterior than 휎푣 to better sample smaller velocity dispersion values. Once after removing samples that correspond to a negative companion ({ }) − ({ }) again, we use a uniform prior from min 푣푖 4 max 휖푖 to mass, but before the application of the Raghavan prior. Figure7 is ({ }) + ({ }) max 푣푖 4 max 휖푖 as our prior on mean velocity. a corner plot of the posterior of select parameters and shows two distinct peaks corresponding to a long and short period solution. We separate the posterior samples at the 0.6 year boundary and present 4 RESULTS the orbital parameter values taken from the truncated posteriors in Table1. Figure8 shows the RV curve for this star as well as the 4.1 Binary Orbital Parameters orbital solutions that correspond to surviving samples. Within our data set, we are able to spectroscopically identify one The semi-major axis posteriors for both solutions are consistent 푎 = star, Star46, as a binary. Attempting to fit a constant velocity model with values less than 2AU, indicating that this is a close binary. to the observations of this object, we find a reduced chi-squared of The resulting overall binary parameter space is that of a common binary star system. The companion mass posterior is not distinct 휒red = 157.2. Previous analyses by Venn et al.(2017) and Kirby et al.(2017) also confirmed this star to be part of a binary system. enough from the prior to be informative, but all of the samples For comparison, the member with the second highest reduced chi- correspond to stellar masses potentially on the main sequence. For comparison, the only other binary star in an ultra-faint dwarf squared is MIC2016-27 with 휒red = 4.4. While this value could suggest possible binarity for MIC2016-27, its velocity variations are spheroidal, Her-3 in the dwarf, has a smaller eccentric- 푒 = . 푃 = . ± . more consistent with a constant velocity model than Star46. ity ( Her−3 0 18) and period ( Her−3 135 28 0 33days) In this analysis, we use the 4 velocity measurements of this star values (Koch et al. 2014). Plugging into equation4 we find the corresponding semi-amplitudes, 퐾 = 16.24+1.59 km s−1 and from MMT (Table B1) and 5 from Keck (Table B2). There are 2 short −1.17 퐾 = . +2.67 −1 additional velocity measurements made in Geminin/GRACES data long 18 72−1.59 km s which are only slightly larger than to −1 (Venn et al. 2017; Ji et al. 2020) that we have opted to not use for Herc-3’s solution, 퐾Her−3 = 14.48 ± 0.82 km s . two reasons. First, the inclusion of this data would add an additional The truncation in companion mass slightly shifts the posterior to- unknown zero-point correction that is more difficult to quantify due wards longer periods, but no more than 0.2 km s−1, much smaller than

MNRAS 000,1–12 (2021) Triangulum II 9

Raghavan prior: False Raghavan prior: True 1.0 1 1 log10( v/km s ) = 0.970(16th) 1.0 log10( v/km s ) = 0.966(16th) 1 1 360 log10( v/km s ) = 0.271(50th) log10( v/km s ) = 0.259(50th) 0.8 1 1 log10( v/km s ) = 0.319(84th) 0.8 log10( v/km s ) = 0.320(84th)

0.6 0.6 380 ] 1 0.4 0.4 s

0.2 m 0.2 400 k [ s o

l 0.0 0.0 1 0 1 2 1 0 1 2 v 1 1 log10( v/km s ) log10( v/km s ) 420

Figure 9. Velocity dispersion posterior with and without the use of the Ragha- MMT Star46 440 Keck van prior on . The 16th, 50th, and 84th percentiles are marked by the blue, orange and green lines respectively. 2015-09-11 2015-12-30 2016-04-18 2016-08-05 2016-11-23 Time [YYYY-MM-DD]

Figure 8. Possible fits to the radial velocity curve of the binary star, Star46, calculated using our binary model. The orange and purple lines correspond to solutions with orbital periods of approximately 0.4 years and 0.8 years respectively. The dates are in the format Year-Month-Day. the width of the posterior peaks (∼ 2 km s−1). Re-weighting the sam- ples such that the prior is the Raghavan period distribution weights the posterior in favor of the longer period solutions and the corre- sponding larger systemic velocities, removing the multi-modality of the posterior (Figure6). Even before applying the Raghavan prior, )

there is a strong preference towards the long period solutions as the 1 ratio of the long period posterior to short period posterior is approxi- 1.2 s mately 5:2. After applying the Raghavan prior, the posterior becomes almost entirely in favor of the long period solution as the ratio be- m 0.6 k comes 80:1, effectively removing the short period mode. As such, we / v present the parameters corresponding to the long period mode as the 0.0 ( orbital solution, but we also explore the effect using the Raghavan 0 prior has on the velocity dispersion. 1 0.6 g

This binary star is only 1 of the 16 stars in our sample. Moe o l 1.2 et al.(2019) and Mazzola et al.(2020) found that intrinsic close bi- 1.2 0.6 0.0 0.6 1.2 nary fraction is anti-correlated with metallicity such that Tri II’s 387.5 385.0 382.5 380.0 377.5 1 1 metallicity value of [Fe/H]=-2.1 implies a close binary fraction v [km s ] log10( v/km s ) 푓bin,close ≈ 0.5. This suggests that there are more close binary mem- bers, but the current data is not enough to determine binarity for Figure 10. Complete posterior Tri II mean velocity and velocity dispersion any other binaries in the galaxy. Though, with growing observation when treating the binary star under the binary model and other non-binary power, they may one day be observed sufficiently. members under the Gaussian fit with the Raghavan prior.

4.2 Triangulum II Dynamics While we have briefly explored applying the binary model to the non-binary members and applying the Gaussian fit model to Star46, The posterior has a small peak at log10 휎푣 = 0.2 (which corre- −1 we unsurprisingly find that the best variation is to treat confirmed sponds to 휎푣 = 1.6km s ), but a large probability tail that stretches the binary star under the binary model and the remaining members back to the minimum value. We see that applying the Raghavan under the Gaussian fit model. This option assumes variability for only prior only slightly reduces this tail (Figure9). Figure 10 shows the Star46. These additional variations made it more difficult to resolve complete posterior with the binary under the Raghavan prior. This the dispersion and instead result in a posterior with a maximum at remaining tail of probability means we were unable to resolve the the 휎푣 prior’s minimum value. Thus, we focus our final results on velocity dispersion of Tri II, and we instead present upper limits on the best variation. 휎푣 . Before applying the Raghavan prior, we find a 90% confidence The mean velocity posterior is consistent with the previously found limit of 2.7 km s−1 and 95% confidence limit of 3.5 km s−1. Af- mean velocity of Tri II, h푣i = −381.7 ± 1.1 km s−1 (Kirby et al. ter applying the Raghavan prior, we find a 90% confidence limit of 2017) within 2휎, where 휎 is the error on the mean velocity from the 2.6 km s−1 and 95% confidence limit of 3.4 km s−1. Due to the posterior. We present a mean velocity of h푣i = −382.3 ± 0.7 km s−1 differences between these values being so small, we again choose from this analysis. However, even in this best variation, were we to present the values determined using the more informative period unable to completely resolve the velocity dispersion. prior.

MNRAS 000,1–12 (2021) 10 R. Buttry et al.

5 CONCLUSIONS U.S. National Science Foundation (NSF) grants AST-1312997, AST- 1726457 and AST-1815403. ES acknowledges funding through VIDI We have found orbital parameters for the Tri II binary star member, grant "Pushing Galactic Archaeology to its limits" (with project num- making it the second binary star within a ultra-faint dwarf with an ber VI.Vidi.193.093) which is funded by the Dutch Research Council orbital solution after Her-3 in Hercules Dwarf (Koch et al. 2014). In (NWO). CB and CMD are supported by NSF grant AST-1909022. doing so, we have demonstrated a method of accounting for binary This work has made use of data from the European Space star systems in velocity dispersion calculations that does not demand Agency (ESA) mission Gaia (https://www.cosmos.esa.int/ the removal of such systems. We also see that using our binary gaia), processed by the Gaia Data Processing and Analysis Consor- model when unnecessary, such as with non-binary stars, not only tium (DPAC, https://www.cosmos.esa.int/web/gaia/dpac/ adds lengthy computation time, but can lead to an overestimation of consortium). Funding for the DPAC has been provided by national velocity error. institutions, in particular the institutions participating in the Gaia The classification of Tri II as either an ultra-faint dwarf or globular Multilateral Agreement. cluster remains an open question. Though the use of this model has The Pan-STARRS1 Surveys (PS1) and the PS1 public science offered an improved analysis of the dwarf galaxy, we were unable to archive have been made possible through contributions by the In- resolve a velocity dispersion that would confirm the existence of a stitute for Astronomy, the University of Hawaii, the Pan-STARRS dark matter halo in Tri II. However, Kirby et al.(2017) found a metal- Project Office, the Max-Planck Society and its participating in- licity dispersion of 휎([Fe/H]) = 0.53+0.38 dex when including all −0.12 stitutes, the Max Planck Institute for Astronomy, Heidelberg and available metallicity measurements for the dataset. They make clear the Max Planck Institute for Extraterrestrial Physics, Garching, The that this value relies on the membership two stars (labelled Star31 Johns Hopkins University, Durham University, the University of Ed- and Star46 in this analysis), one of which was the now confirmed inburgh, the Queen’s University Belfast, the Harvard-Smithsonian binary star (Star46). The Gaia proper motions of both stars confirm Center for Astrophysics, the Las Cumbres Observatory Global Tele- that they are Tri II members (Figure1). We briefly tested performing scope Network Incorporated, the National Central University of Tai- the same 휎([Fe/H]) calculation while including MMT metallicities. wan, the Space Telescope Science Institute, the National Aeronau- In doing so, we assumed the existence of another zero-point offset tics and Space Administration under Grant No. NNX08AR22G is- which was determined using the non-outlier offset model described sued through the Planetary Science Division of the NASA Science in Section 2.3 with the two stars that are in the overlap of the MMT Mission Directorate, the National Science Foundation Grant No. and Keck catalogues. We found that the inclusion of MMT measure- AST-1238877, the University of Maryland, Eotvos Lorand Univer- ments did not significantly impact the resulting dispersion where the sity (ELTE), the Los Alamos National Laboratory, and the Gordon difference after inclusion was . 0.03 dex from the previous value. and Betty Moore Foundation. This metallicity dispersion greatly strengthens the case for Tri II’s This research has made use of NASA’s Astrophysics Data Sys- classification as a dwarf galaxy, indicating the system is embedded tem Bibliographic Services. This paper made use of the Whole Sky in a dark matter halo (Willman & Strader 2012). (wsdb) created by Sergey Koposov and maintained at the The choice to apply our binary model to only one star means Institute of Astronomy, Cambridge by Sergey Koposov, Vasily Be- that we are operating with a hard binary fraction of 푓 ≈ 0.06. bin lokurov and Wyn Evans with financial support from the Science & A more informed model would be to assume that there is some Technology Facilities Council (STFC) and the European Research non-zero fraction of the “non-binary” member stars that are actually Council (ERC). binaries. Each star would have an associated binary and non-binary Observations reported here were obtained at the MMT Observa- likelihoods that are calculated from marginalizing over the binary tory, a joint facility of the University of Arizona and the Smithsonian parameter posterior. These likelihoods would allow for the sampling Institution. The authors wish to recognize and acknowledge the very of binary fraction as a parameter alongside with mean velocity and significant cultural role and reverence that the summit of Maunakea velocity dispersion, forming a hierarchical model that builds from has always had within the indigenous Hawaiian community. We are the set of binary posteriors. This likelihood would be functionally most fortunate to have the opportunity to conduct observations from similar to the mixture model used to determine the offset between this mountain. the instrumental zero-point velocities. However, this approach intro- duces complications as marginalization becomes less straightforward when modifying the posterior. For instance, removing binary sam- ples based on a parameter, such as pericenter, makes the posterior no DATA AVAILABILITY longer normalized to the calculated binary/non-binary likelihoods. The likelihoods must be re-normalized after truncation for every ad- The processed MMT/Hectochelle spectra for Triangulum II targets ditional condition that is imposed. Though it is outside the scope of are publicly available at the Zenodo database, http://. this paper, we find the hierarchical model to be an interesting problem and important step to explored in future works.

REFERENCES

Badenes C., et al., 2018, ApJ, 854, 147 ACKNOWLEDGEMENTS Bullock J. S., Boylan-Kolchin M., 2017, ARA&A, 55, 343 Carlin J. L., et al., 2017,AJ, 154, 267 This work is supported in part by NSF grants AST-1813881 and Choi J., Dotter A., Conroy C., Cantiello M., Paxton B., Johnson B. D., 2016, AST-1909584. E.N.K. is supported by NSF grant AST-1847909, ApJ, 823, 102 and he gratefully acknowledges support from a Cottrell Scholar Cooper M. C., Newman J. A., Davis M., Finkbeiner D. P., Gerke B. F., 2012, award administered by the Research Corporation for Science Ad- spec2d: DEEP2 DEIMOS Spectral Pipeline (ascl:1203.003) vancement. EO is partially supported by NSF grant AST-1815767. Dotter A., 2016, ApJS, 222, 8 NC is supported by NSF grant AST-1812461. MM is supported by Errani R., Peñarrubia J., Walker M. G., 2018, MNRAS, 481, 5073

MNRAS 000,1–12 (2021) Triangulum II 11

Faber S. M., et al., 2003, in Iye M., Moorwood A. F. M., eds, Society of Photo- Walker M. G., Olszewski E. W., Mateo M., 2015, MNRAS, 448, 2717 Optical Instrumentation Engineers (SPIE) Conference Series Vol. 4841, Willman B., Strader J., 2012,AJ, 144, 76 Instrument Design and Performance for Optical/Infrared Ground-based Wolf J., Martinez G. D., Bullock J. S., Kaplinghat M., Geha M., Muñoz R. R., Telescopes. pp 1657–1669, doi:10.1117/12.460346 Simon J. D., Avedo F. F., 2010, MNRAS, 406, 1220 Flewelling H. A., et al., 2020, ApJS, 251, 7 Foreman-Mackey D., Hogg D. W., Lang D., Goodman J., 2013, PASP, 125, 306 Gaia Collaboration et al., 2021, A&A, 649, A1 APPENDIX A: LOWER LIMIT ON log10 휎푉 PRIOR Hekker S., Snellen I. A. G., Aerts C., Quirrenbach A., Reffert S., Mitchell From the virial theorem, we know that the mass contained within D. S., 2008, in Journal of Physics Conference Series. p. 012058 the half-light radius of a dwarf galaxy in virial equilibrium can be (arXiv:0710.4134), doi:10.1088/1742-6596/118/1/012058 described by... Ibata R., Sollima A., Nipoti C., Bellazzini M., Chapman S. C., Dalessandro E., 2011, ApJ, 738, 186 2 푅1/2휎푣 Illingworth G., 1976, ApJ, 204, 73 푀(< 푅1/2) = 퐶 (A1) Irwin M., Lewis J., 2001, New Astron. Rev., 45, 105 퐺 Ji A. P., Simon J. D., Frebel A., Venn K. A., Hansen T. T., 2020, VizieR Where 휎푣 is the galaxy velocity dispersion and 퐺 is the gravi- Online Data Catalog, p. J/ApJ/870/83 tational constant. 퐶 is a proportionality constant. Assuming that all Kirby E. N., Cohen J. G., Simon J. D., Guhathakurta P., 2015, The Astro- of the mass is accounted for by the total , the following physical Journal, 814, L7 relation becomes true. Kirby E. N., Cohen J. G., Simon J. D., Guhathakurta P., Thygesen A. O., Υ퐿 Duggan G. E., 2017, ApJ, 838, 83 푀(< 푅1/2) = (A2) Koch A., Hansen T., Feltzing S., Wilkinson M. I., 2014, ApJ, 780, 91 2 Koposov S. E., et al., 2011, ApJ, 736, 146 Where Υ is the stellar mass-to-light ratio and L is the total . Laevens B. P. M., et al., 2015, The Astrophysical Journal, 802, L18 Combining these, we can solve for 휎푣 . Li T. S., et al., 2019, Monthly Notices of the Royal Astronomical Society, √︄ 490, 3508 퐺 Υ퐿 Longeard N., et al., 2020, MNRAS, 491, 356 휎푣 = (A3) 2퐶 푅1/2 Martin N. F., et al., 2016, The Astrophysical Journal, 818, 40 Martinez G. D., Minor Q. E., Bullock J., Kaplinghat M., Simon J. D., Geha It is estimated that Tri II has a total luminosity 퐿 ∼ 400 L and M., 2011, ApJ, 738, 55 a half light radius 푅1/2 ∼ 34 pc (Laevens et al. 2015). Stellar mass Mateo M. L., 1998, ARA&A, 36, 435 −1 to light ratios usually exist in the rage Υ ∼ 0.5 to 3.0 M L and Mazzola C. N., et al., 2020, MNRAS, 499, 1607 the proportionality constant is in the range 퐶 ∼ 2 to 4. Taking the McConnachie A. W., 2012,AJ, 144, 4 Υ McConnachie A. W., Côté P., 2010, ApJ, 722, L209 lower limit for and the upper limit for 퐶, we find that 휎푣 can go −1 McConnachie A. W., et al., 2018, ApJ, 868, 55 as small as 0.05 km s . This value will serves as the lower limit of Minor Q. E., 2013, ApJ, 779, 116 our velocity dispersion prior and corresponds to a system where the Minor Q. E., Martinez G., Bullock J., Kaplinghat M., Trainor R., 2010, ApJ, dynamics are fully determined by visible matter. 721, 1142 Minor Q. E., Pace A. B., Marshall J. L., Strigari L. E., 2019, MNRAS, 487, 2961 APPENDIX B: DATA TABLES Miyazaki S., et al., 2018, PASJ, 70, S1 Moe M., Kratter K. M., Badenes C., 2019, ApJ, 875, 61 Morton T. D., 2015, isochrones: Stellar model grid package (ascl:1503.010) Murray C. D., Correia A. C. M., 2010, Keplerian and Dynamics of . pp 15–23 Oke J. B., et al., 1995, PASP, 107, 375 Olszewski E. W., Pryor C., Armandroff T. E., 1996,AJ, 111, 750 Pace A. B., et al., 2020, MNRAS, 495, 3022 Pace A. B., Walker M. G., Koposov S. E., Caldwell N., Mateo M., Ol- szewski E. W., Bailey John I. I., Wang M.-Y., 2021, arXiv e-prints, p. arXiv:2105.00064 Pedregosa F., et al., 2011, Journal of Machine Learning Research, 12, 2825 Price-Whelan A. M., Hogg D. W., Foreman-Mackey D., Rix H.-W., 2017, ApJ, 837, 20 Raghavan D., et al., 2010, ApJS, 190, 1 Simon J. D., 2019, ARA&A, 57, 375 Sohn S. T., et al., 2007, ApJ, 663, 960 Spencer M. E., Mateo M., Walker M. G., Olszewski E. W., McConnachie A. W., Kirby E. N., Koch A., 2017,AJ, 153, 254 Spencer M. E., Mateo M., Olszewski E. W., Walker M. G., McConnachie A. W., Kirby E. N., 2018,AJ, 156, 257 Starkenburg E., et al., 2017, MNRAS, 471, 2587 Szentgyorgyi A., et al., 2011, PASP, 123, 1188 Tolstoy E., Hill V., Tosi M., 2009, ARA&A, 47, 371 Venn K. A., Starkenburg E., Malo L., Martin N., Laevens B. P. M., 2017, MNRAS, 466, 3741 Walker M. G., Mateo M., Olszewski E. W., Peñarrubia J., Evans N. W., Gilmore G., 2009, ApJ, 704, 1274

MNRAS 000,1–12 (2021) 12 R. Buttry et al.

This paper has been typeset from a TEX/LATEX file prepared by the author. 305141958Tu 347973.908 478.286 -386.9 8.67 2457683.72 36.0950281 33.4272917 True 331085117451894528 309478256164 re3.198 6199253482.8-382.8 -386.4 27.68 -405.1 2.88 -405.9 2457374.8 -400.4 5.08 2457711.69 -383.4 6.78 36.17939 2457688.77 36.1659431 -384.4 8.84 4.74 2457683.72 36.1659431 33.3189583 33.33975 -385.5 2457374.8 6.07 36.1659431 2457685.74 33.33975 36.1659431 10.68 True 2457711.69 36.0950281 33.33975 True 2457688.77 36.0950281 33.4272917 33.33975 True 36.0950281 33.4272917 True 40 True 33.4272917 46 True True 46 True 46 106 46 65 65 331089446778920576 65 331086526201161088 65 331086526201161088 331086526201161088 331086526201161088 331085117451894528 331085117451894528 331085117451894528 309478256164 re3.198 619925637 16 -381.8 21.62 2457683.72 36.17939 33.3189583 True 40 106 331089446778920576 309478256164 re3.198 619925687 38 -381.9 23.86 2457688.77 36.17939 33.3189583 True 40 106 331089446778920576 309478256164 re3.198 619925716 31 -381.9 13.11 2457711.69 36.17939 33.3189583 True 40 106 331089446778920576 309478256164 re3.198 619925657 28 -382.2 12.85 2457685.74 36.17939 33.3189583 True 40 106 331089446778920576 aaI C21 I21 ebrR IR)Dc(CS J dy)S/N (days) HJD (ICRS) Dec (ICRS) RA Member MIC2016 KCS2015 ID Gaia al 1 niiulosrain fTiI ebr ae ihMTHcohle hstbei usto h ulMTcatalogue. MMT full the of subset a is table This Hectochelle. MMT with taken members II Tri of observations Individual B1: Table 푣 los ( ms km ± ± ± ± ± ± ± ± ± ± ± ± ± . -3.0 0.7 . -3.5 -0.8 0.6 -2.0 2.1 -2.9 1.0 -2.5 0.8 -3.0 0.7 -3.2 1.7 -3.2 1.7 0.9 . -3.3 0.5 . -3.3 0.5 . -3.6 0.7 . -3.2 0.8 − 1 ) [Fe/H] ± ± ± ± ± ± ± ± ± ± ± ± ± . 4760 0.3 . 4367 6811 0.1 5416 0.9 4535 0.7 4823 0.3 4783 0.3 4772 0.7 4684 0.6 0.4 . 4507 0.2 . 4505 0.1 . 4253 0.2 . 4597 0.3 T eff ± ± ± ± ± ± ± ± ± ± ± ± ± 2 0.7 226 5 2.2 2.0 855 0.9 588 2.0 215 1.7 159 0.9 495 1.4 453 255 3 0.7 135 9 1.1 195 70.9 87 51.0 95 20.8 82 log ± ± ± ± ± ± ± ± ± ± ± ± ± 0.5 1.0 0.7 0.4 0.4 0.9 0.7 0.5 0.2 0.3 0.2 0.3 0.5 푔

MNRAS 000,1–12 (2021) Triangulum II 13

−1 KCS2015 MIC2016 RA (ICRS) Dec (ICRS) HJD (days) 푣los (km s ) N/A 25 33.3214149 36.1205826 2457284.07 −364.1 ± 5.6 N/A 8 33.2591667 36.2090836 2457284.07 −388.4 ± 7.7 128 N/A 33.3093333 36.1641944 2457303.10 −386.3 ± 3.2 - - - - 2457416.70 −385.4 ± 2.1 - - - - 2457416.80 −384.2 ± 2.0 - - - - 2457639.10 −383.8 ± 2.1 N/A 31 33.4694167 36.2233611 2457370.70 −377.6 ± 1.8 - - - - 2457416.70 −381.5 ± 2.6 - - - - 2457639.10 −377.0 ± 2.8 - - - - 2457284.07 −377.1 ± 3.1 N/A 29 33.3789583 36.1988889 2457370.70 −387.5 ± 4.7 - - - - 2457284.07 −398.4 ± 7.8 N/A 27 33.3389583 36.1414167 2457370.70 −373.5 ± 5.2 - - - - 2457416.70 −389.9 ± 8.3 - - - - 2457284.07 −402.7 ± 6.6 N/A 26 33.3534583 36.1727222 2457370.70 −376.9 ± 11.2 N/A 24 33.3416667 36.1738611 2457416.70 −371.8 ± 17.1 - - - - 2457284.07 −384.4 ± 4.9 76 23 33.3358750 36.1629167 2457303.10 −391.0 ± 3.0 - - - - 2457370.70 −384.6 ± 3.1 - - - - 2457639.10 −384.0 ± 3.0 - - - - 2457284.07 −389.2 ± 3.6 N/A 22 33.3028750 36.1470556 2457370.70 −381.5 ± 3.4 - - - - 2457416.70 −381.4 ± 3.3 - - - - 2457416.80 −384.9 ± 3.7 - - - - 2457569.10 −381.1 ± 23.5 - - - - 2457639.10 −376.6 ± 6.6 - - - - 2457284.07 −388.3 ± 3.8 116 21 33.3165000 36.1710556 2457303.10 −378.9 ± 3.7 - - - - 2457370.70 −383.3 ± 2.3 - - - - 2457416.70 −387.0 ± 3.2 - - - - 2457416.80 −382.5 ± 3.6 - - - - 2457639.10 −380.6 ± 3.1 - - - - 2457284.07 −384.1 ± 3.1 91 20 33.3305000 36.1925833 2457303.10 −387.3 ± 3.1 - - - - 2457370.70 −379.1 ± 1.9 - - - - 2457569.10 −388.2 ± 3.9 - - - - 2457639.10 −379.2 ± 2.4 - - - - 2457284.07 −380.0 ± 2.9 N/A 9 33.3639183 36.2251930 2457416.70 −388.9 ± 7.7 - - - - 2457284.07 −406.0 ± 5.1 65 46 33.33975 36.1659431 2457282.50 −373.8 ± 2.4 - - - - 2457303.10 −375.8 ± 1.7 - - - - 2457416.70 −386.7 ± 1.6 - - - - 2457569.10 −376.3 ± 4.0 - - - - 2457639.10 −384.6 ± 1.6

−1 Table B2. Individual observations of Tri II members from Keck DEIMOS used in this analysis after adding a 훿푣 = −1.33 km s zero-point offset correction. The first epochs for each star are listed with the star’s ID numberings and sky position while any additional epochs for the same star are in subsequent rows with "-" in the ID and sky position columns.

MNRAS 000,1–12 (2021)